26.06.2013 Views

GHY 4530/5530 - Andean Mountain Geography - Appalachian State ...

GHY 4530/5530 - Andean Mountain Geography - Appalachian State ...

GHY 4530/5530 - Andean Mountain Geography - Appalachian State ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>GHY</strong> <strong>4530</strong>/<strong>5530</strong> - <strong>Andean</strong> <strong>Mountain</strong> <strong>Geography</strong> – Summer Session II, 2013<br />

Faculty: Dr. Baker Perry (with Dr. Anton Seimon)<br />

Office: 363 Rankin Science West<br />

Phone: (828) 262-7058<br />

Email: perrylb@appstate.edu<br />

Course Description: The course will begin with discussion of the significance of mountains in geographical inquiry,<br />

continue with an overview of the important physical processes (i.e. mountain forming processes, mountain<br />

meteorology, vegetation) and also include study of the human dimensions of mountain environments (i.e.<br />

mountain peoples and cultures, sustainable development), paying particular attention to the Andes <strong>Mountain</strong>s and<br />

the country of Peru.<br />

Objectives: Upon completion of this course, students should have a better understanding of <strong>Andean</strong> mountain<br />

environments and peoples and be able to think critically about contemporary mountain issues. Students will<br />

demonstrate an understanding of the role of global climate change in changing the physical and human geography<br />

of mountain regions.<br />

Grading: Field Journal 50%<br />

Discussion 20%<br />

Response Paper 15%<br />

Exam 15%<br />

Journal: A detailed field journal comprises 50 percent of your overall grade for this course. You are expected to<br />

keep daily entries that 1) summarize your activities, hypotheses, data collection, and field methods, 2) process and<br />

reflect upon discussions, conversations, and/or observations, 3) draw connections between your<br />

experiences/observations and the readings and discussions. The journal is due on July 29. Please note: This is not a<br />

personal journal.<br />

Discussion: Facilitating a discussion on one of the readings is 20 percent of your final grade in this course. Outside<br />

research to provide additional information on the topic is expected. The student discussant should provide a<br />

general overview of the topic, introduce the authors of the article, briefly summarize the major findings and<br />

conclusions, and come up with several questions to guide group discussion.<br />

Response Paper: A response paper will account for 15 percent of your final grade. Papers should be between three<br />

and five pages in length and respond to one of the issues discussed in the readings, seminars, or field visits.<br />

Students should meet with the instructor to choose and appropriate topic. The response paper is due on July 29.<br />

Exam: The exam will be administered at the conclusion of the trip and is worth 15 percent of your grade.<br />

Graduate Students (<strong>GHY</strong> <strong>5530</strong>-145): In addition to all assignments described above, graduate students enrolled in<br />

<strong>GHY</strong> <strong>5530</strong> will be expected to do a significant amount of additional work and all assignments will be graded at a<br />

higher level of expectations. Graduate students will lead three discussions of articles assigned and will provide a<br />

typed summary of each presentation to hand out to the class. Each of those summaries should incorporate at<br />

least five additional sources from scientific journals and must provide full bibliographic listings. Graduate students<br />

will also write a term paper upon return from the field portion of the course. The presentations, journal, exam, and<br />

term paper will each account for a quarter of your final grade. Papers should be between twelve and fifteen pages<br />

in length and focus on a topic agreed upon with the professor. A handout will provide additional details. The term<br />

paper and journal are due on July 29.<br />

Selected Reading List for <strong>Andean</strong> <strong>Mountain</strong> <strong>Geography</strong> – <strong>GHY</strong> <strong>4530</strong>/<strong>5530</strong><br />

Background Readings<br />

Burger, R.L., and L.C. Salazar. 2004. Machu Picchu: Unveiling the Mystery of the Incas. New Haven, CT: Yale<br />

University Press.<br />

Halloy, S.R.P., A. Seimon, K. Yager and A. Tupayachi. 2005. Multi-dimensional (climatic,<br />

biodiversity, socioeconomic, and agricultural) context of changes in land use in the Vilcanota watershed, Peru.


In Land Use Changes and <strong>Mountain</strong> Biodiversity. E.M. Spehn, M. Liberman and C. Körner (eds.), CRC Press,<br />

Boca Raton.<br />

Ives, J.D. B. Messerli, and E. Spiess. 1997. <strong>Mountain</strong>s of the World—a global priority. In <strong>Mountain</strong>s of the World: A<br />

Global Priority, eds. Bruno Messerli and Jack D. Ives, 1-15. New York, NY: Parthenon.<br />

Schroeder, K., C. Wood, S. Galiardi, and J. Koehn. 2009. First, do no harm: Ideas for mitigating negative community<br />

impacts of short-term study abroad. Journal of <strong>Geography</strong> 108: 141-147.<br />

Starn, O., C.I. Degregori, and R. Kirk. 2005 (2 nd Edition). The Peru Reader: History, Culture, Politics. Durham, NC:<br />

Duke University Press.<br />

Discussion Articles (In Chronological Order)<br />

Lauer, W. 1993. Human development and environment in the Andes: a geoecological overview. <strong>Mountain</strong><br />

Research and Development 13:157-166.<br />

Chepstow-Lusty, A.J., M.R. Frogley, B.S. Bauer, M.J. Leng, K.P. Boessenkool, C. Carcaillet, A.A. Ali, and A. Gioda.<br />

2009. Putting the rise of the Inca Empire within a climatic and land management context. Climates of the Past<br />

5: 375-388.<br />

Harden, C. 2006. Human impacts on headwater fluvial systems in the northern and central Andes. Geomorphology<br />

79: 249–263.<br />

Wright, K.R., G.D. Witt, A.V. Zegarra. 1997. Hydrogeology and paleohydrology of ancient Machu Picchu. Ground<br />

Water 35: 660-666.<br />

Larson, L.R., N.C. Poudyat. 2012. Developing sustainable tourism through adaptive resource management: a case<br />

study of Machu Picchu, Peru. Journal of Sustainable Tourism 20: 917-938.<br />

Young, K.R., and J.K. Lipton. 2006. Adaptive governance and climate change in the tropical highlands of western<br />

South America. Climatic Change 78: 63-102.<br />

Hole, D.G., K.R. Young, A. Seimon, C. Gomez, D. Hoffmann, K. Schutze, S. Sanchez, D. Muchoney, H.R. Grau, E.<br />

Ramirez. 2010. Adaptive management for biodiversity conservation under climate change – a tropical <strong>Andean</strong><br />

perspective. In, Herzog, S.K., R. Martínez, P.M. Jørgensen & H. Tiessen (Eds.). Climate change effects on the<br />

biodiversity of the tropical Andes: an assessment of the status of scientific knowledge. Inter-American Institute<br />

of Global Change Research (IAI) and Scientific Committee on Problems of the Environment (SCOPE), São José<br />

dos Campos and Paris.<br />

Seimon, A., T.A. Seimon (co-lead authors), P. Daszak, S. Halloy, P. Sowell, B. Konecky, L.M. Schloegel, C.A. Aguilar, J.<br />

Simmons, and A. Hyatt. 2007. Upward range extension of <strong>Andean</strong> anurans to extreme elevations in response<br />

to tropical deglaciation. Global Change Biology 13:288-299.<br />

Hardy, D., M.W. Williams, and C. Escobar. 2001. Near-surface faceted crystals, avalanches, and climate in highelevation,<br />

tropical mountains of Bolivia. Cold Regions Science and Technology 33: 291-302.<br />

Bush, M.B., J.A. Hanselman, and W.D. Gosling. 2010. Nonlinear climate change and <strong>Andean</strong> feedbacks: an<br />

imminent turning point? Global Change Biology 16: 3223-3232.<br />

<strong>State</strong>ment of Student Engagement with Courses<br />

In its mission statement, <strong>Appalachian</strong> <strong>State</strong> University aims at “providing undergraduate students a rigorous liberal<br />

education that emphasizes transferable skills and preparation for professional careers” as well as “maintaining a<br />

faculty whose members serve as excellent teachers and scholarly mentors for their students.” Such rigor means<br />

that the foremost activity of <strong>Appalachian</strong> students is an intense engagement with their courses. In practical terms,<br />

students should expect to spend two to three hours of studying for every hour of class time. For this study abroad<br />

course, students should expect to spend the bulk of that time, or approximately 50 to 75 hours, completing the<br />

readings and other assignments prior to departure. Approximately 25-38 hours of journaling, reading, and study<br />

will be expected during the time abroad, and an additional 25-38 hours of research and writing will be expected in<br />

order to complete the final paper for the course upon return.<br />

2


<strong>GHY</strong> <strong>4530</strong>/<strong>5530</strong> (<strong>Andean</strong> <strong>Mountain</strong> <strong>Geography</strong>)<br />

Author and Year Title Date Responsible<br />

Lauer 1993 Human development and environment in the Andes: a geoecological overview 8-Jul<br />

Chepstow-Lusty et<br />

al. 2009<br />

Harden 2006<br />

Putting the rise of the Inca Empire within a climatic and land management<br />

context<br />

Human impacts on headwater fluvial systems in the northern and central<br />

Andes<br />

Wright et al. 1997 Hydrogeology and paleohydrology of ancient Machu Picchu 12-Jul<br />

Larson & Poudyat Developing sustainable tourism through adapative resource management: a<br />

2012<br />

case study from Machu Picchu, Peru<br />

Adaptive governance and climate change in the tropical highlands of western<br />

Young & Lipton 2006<br />

South America<br />

Hole et al. 2010<br />

Adaptive management for biodiversity conservation under climate change -- a<br />

tropical <strong>Andean</strong> perspective<br />

Seimon et al. 2007<br />

Upward range extension of <strong>Andean</strong> anurans to extreme elevations in response<br />

to tropical deglaciation<br />

Hardy et al. 2001<br />

Near-surface faceted crystals, avalanches, and climate in high-elevation,<br />

tropical mountains of Bolivia<br />

8-Jul<br />

9-Jul<br />

12-Jul<br />

13-Jul<br />

13-Jul<br />

17-Jul<br />

18-Jul<br />

Bush et al. 2010 Nonlinear climate change and <strong>Andean</strong> feedbacks: an imminent turning point? 19-Jul


This article was downloaded by: [<strong>Appalachian</strong> <strong>State</strong> University]<br />

On: 1 September 2009<br />

Access details: Access Details: [subscription number 907136658]<br />

Publisher Routledge<br />

Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,<br />

37-41 Mortimer Street, London W1T 3JH, UK<br />

Journal of <strong>Geography</strong><br />

Publication details, including instructions for authors and subscription information:<br />

http://www.informaworld.com/smpp/title~content=t770943818<br />

First, Do No Harm: Ideas for Mitigating Negative Community Impacts of Short-<br />

Term Study Abroad<br />

Kathleen Schroeder a ; Cynthia Wood b ; Shari Galiardi c ; Jenny Koehn c<br />

a Department of <strong>Geography</strong> and Planning, <strong>Appalachian</strong> <strong>State</strong> University, Boone, North Carolina, USA b<br />

Sustainable development, <strong>Appalachian</strong> <strong>State</strong> University, Boone, North Carolina, USA c <strong>Appalachian</strong> <strong>State</strong><br />

University, Boone, North Carolina, USA<br />

Online Publication Date: 01 May 2009<br />

To cite this Article Schroeder, Kathleen, Wood, Cynthia, Galiardi, Shari and Koehn, Jenny(2009)'First, Do No Harm: Ideas for<br />

Mitigating Negative Community Impacts of Short-Term Study Abroad',Journal of <strong>Geography</strong>,108:3,141 — 147<br />

To link to this Article: DOI: 10.1080/00221340903120866<br />

URL: http://dx.doi.org/10.1080/00221340903120866<br />

PLEASE SCROLL DOWN FOR ARTICLE<br />

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf<br />

This article may be used for research, teaching and private study purposes. Any substantial or<br />

systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or<br />

distribution in any form to anyone is expressly forbidden.<br />

The publisher does not give any warranty express or implied or make any representation that the contents<br />

will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses<br />

should be independently verified with primary sources. The publisher shall not be liable for any loss,<br />

actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly<br />

or indirectly in connection with or arising out of the use of this material.


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

First, Do No Harm: Ideas for Mitigating Negative Community Impacts of<br />

Short-Term Study Abroad<br />

Kathleen Schroeder, Cynthia Wood, Shari Galiardi, and Jenny Koehn<br />

ABSTRACT<br />

This article presents the results from a<br />

research project on the host community<br />

impact of college students participating<br />

in university-sponsored international<br />

experiences. It finds that little reliable<br />

data is available on the impact that<br />

our students have on host communities.<br />

The article concludes that nondamaging<br />

international experiences require a substantial<br />

amount of planning, experienced<br />

group facilitation, and solid debriefing<br />

of students and community members.<br />

We recommend that geographers with<br />

a critical perspective and extensive<br />

foreign expertise should help guide<br />

the development of these experiences<br />

and urge their universities to screen<br />

study abroad for unintended negative<br />

outcomes on local communities.<br />

Key Words: short-term study abroad,<br />

impact on host communities<br />

Kathleen Schroeder is an associate professor at<br />

<strong>Appalachian</strong> <strong>State</strong> University, Boone, North<br />

Carolina, USA, in the Department of <strong>Geography</strong><br />

and Planning.<br />

Cynthia Wood is an associate professor at<br />

<strong>Appalachian</strong> <strong>State</strong> University, Boone, North<br />

Carolina, USA, in sustainable development.<br />

Shari Galiardi is the Director of Service-<br />

Learning at <strong>Appalachian</strong> <strong>State</strong> University,<br />

Boone, North Carolina, USA.<br />

Jenny Koehn is the Associate Director<br />

of Student Programs at <strong>Appalachian</strong> <strong>State</strong><br />

University, Boone, North Carolina, USA.<br />

INTRODUCTION<br />

In the United <strong>State</strong>s, internationalization is an important component of the<br />

mission statements of many colleges and universities, and the numbers reflect<br />

this commitment as student participation in study abroad has grown 150 percent<br />

over the past decade (Institute of International Education 2007). In the President’s<br />

Column of the Association of American Geographers (AAG) newsletter, Victoria<br />

Lawson (2005) comments on this trend and the contributions that geographers<br />

are well positioned to make to an internationalized research and teaching agenda<br />

across campuses.<br />

With an increased emphasis on internationalization, U.S. institutions are<br />

looking to expand their offerings beyond the typical model where language<br />

students spend a semester in a European country. Students across campuses<br />

are encouraged to go abroad in a growing variety of models that can range from<br />

one-week alternative spring break service projects to year-long exchanges. Demand<br />

for study-abroad opportunities is increasing and the Lincoln Commission<br />

recently proposed sending one million Americans abroad by 2016 (Commission<br />

on the Abraham Lincoln Study Abroad Fellowship Program 2005).<br />

Where are all these students going? Presently, the United Kingdom remains<br />

the number-one-ranked destination for U.S. students, but China is now<br />

a top-ten destination as are Mexico and Costa Rica. Countries on the list of<br />

the top-twenty destinations include South Africa, Brazil, and Ecuador (Institute<br />

of International Education 2007). International offerings at a typical mid-sized<br />

public institution in the United <strong>State</strong>s include programs to locations as diverse as<br />

Ecuador, India, New Zealand, and Ghana. The Lincoln Commission recommends<br />

that these nontraditional locations become an increasing percentage of studyabroad<br />

destinations. Following the recommendations of the Lincoln Commission,<br />

the Simon Act currently awaiting approval in Congress mandates “diversifying<br />

locations of study abroad, particularly in developing countries” (NAFSA<br />

2008).<br />

What has been virtually ignored as short-term study abroad has grown on<br />

college campuses is a critical examination of the impact of these programs on the<br />

communities that students visit. 1 The most recent edition of the Forum on Education<br />

Abroad’s Standards of Good Practice for Education Abroad (2008, 19), for example,<br />

makes no recommendations on mitigating the effects of study abroad on local communities,<br />

suggesting only that the organization sending students “respect the cultures<br />

and values of the countries in which it operates.” The implications of a group<br />

going to London are significantly different than going to a remote village in China,<br />

but increasingly universities are offering and students are choosing programs that<br />

take them far off the beaten track, with unexamined and unintended consequences<br />

for host communities. Given the drive to increase the number of students going<br />

abroad, it is urgent that these consequences be considered by both institutions<br />

and faculty in the design and implementation of short-term study-abroad<br />

programs. 2<br />

In this article, we draw attention to the numerous and often unforeseen ways<br />

that students might impact local communities. We have developed this inquiry<br />

through an examination of the relevant bodies of literature and from ongoing<br />

research and discussions with a wide variety of stakeholders involved in shortterm<br />

study abroad on and off our campus. We conclude with six recommendations<br />

that we encourage faculty to take to their home institutions.<br />

Journal of <strong>Geography</strong> 108: 141–147<br />

C○2009 National Council for Geographic Education 141


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

Schroeder et al.<br />

WHAT ARE THE IMPACTS ON LOCAL COMMUNITIES?<br />

Considering how difficult it is to collect information on<br />

the topic, it should not be surprising that we have virtually<br />

no idea of how students’ impact places—particularly the<br />

small, rural, research sites that many geographers are<br />

likely to take our students. David Zurick (1992) and other<br />

geographers have examined sustainable tourism, but not<br />

with a specific eye towards the impact that our own<br />

students are having.<br />

Many of the potential negative impacts of foreign visitors<br />

are highlighted by the literature on tourism, especially<br />

its economic, social, and cultural effects (Archer, Cooper,<br />

and Ruhanen 2005; McLaren 2006). Shaw and Williams<br />

(2002) and Lew, Hall, and A. Williams (2004) provide<br />

comprehensive examinations of tourism from a geographic<br />

perspective. While we believe that there are significant differences<br />

between study abroad and tourism, in considering<br />

the potential negative impacts on host communities there<br />

are many similarities. Like tourism, study abroad “creates<br />

impacts and consequences; we cannot prevent these, but<br />

need to plan and manage to minimize the negative impacts<br />

and emphasize the positive impacts” (Archer, Cooper, and<br />

Ruhanen 2005, 79). And as with tourism, these impacts<br />

occur because study abroad “brings about an intermingling<br />

of people from diverse social and cultural backgrounds,<br />

and also a considerable spatial redistribution of spending<br />

power, which has a significant impact on the economy of<br />

the destination” (Archer, Cooper, and Ruhanen 2005, 79).<br />

It is possible that even more than most tourism, study<br />

abroad is “by its very nature. . . attracted to unique and<br />

fragile environments and societies and. . . in some cases the<br />

economic benefits [to host communities] may be offset by<br />

adverse and previously unmeasured environmental and<br />

social consequences” (Archer, Cooper, and Ruhanen 2005,<br />

79).<br />

The tourism literature suggests that short-term studyabroad<br />

programs may do damage to their host communities,<br />

and points to an array of questions that we should ask<br />

in order to evaluate the potential for such negative impacts,<br />

including the following:<br />

Where does the food/water/housing for our students<br />

come from? Do we impose any hardship on<br />

local people, such as water shortages? What about<br />

garbage disposal and pollution? Is land being used<br />

for visitors rather than local needs?<br />

Does the economic impact of study abroad promote<br />

economic inequality in the community? Do<br />

foreigners or local elites own or manage the<br />

hotels that students occupy during their visit? If<br />

“home stays” are part of the study abroad, do<br />

students live in middle-class homes, so that poorer<br />

people do not receive any economic benefit and<br />

income inequality is worsened? Are guides and<br />

drivers outsiders or wealthier members of the<br />

community? Do local prices go up as a result of the<br />

student visit? The giving of gifts can contribute to<br />

142<br />

similar inequalities, however well-intentioned—<br />

can nonmaterial gifts be given instead, or gifts to<br />

the community as a whole?<br />

Do student visits contribute to economies of dependency<br />

on outsiders, orienting those economies<br />

to pleasing or providing pleasure for wealthy<br />

foreigners rather than to local needs?<br />

Is there a “season” for foreign visitors to come to<br />

the area, such that student visits contribute to a<br />

“boom and bust” cycle in the local economy? Is<br />

there any way to mitigate this effect?<br />

Do students’ patterns of consumption (both during<br />

and before the visit) contribute to problems in<br />

the community? The “demonstration effect” of<br />

students bringing high-end travel gear, lots of<br />

clothes, spending money easily on restaurants, giving<br />

gifts, etc. may create resentment, the perception<br />

of American students as wealthy consumers with<br />

no responsibilities at home (McLaren 2006), or the<br />

desire in local people (especially youth) to leave<br />

the community so that they can make money to<br />

buy similar goods and services. Even traveling on<br />

an airplane or simply traveling away from home<br />

can create these problems among people who do<br />

not have that option.<br />

Are local people excluded from any of the areas<br />

where students are encouraged or allowed to go?<br />

Are students well-behaved and respectful in terms<br />

of the local culture? Do they dress inappropriately,<br />

or otherwise commit cultural offenses that will<br />

anger, distress, or shock people in the local community?<br />

Do students see culture, indigeneity, and<br />

the “authenticity” of local people as commodities<br />

to be consumed? What other cultural impacts<br />

result from student visits? Cultural differences in<br />

themselves are likely sources of confusion and<br />

conflict if unanticipated.<br />

Do students smoke, drink, or do drugs during their<br />

visit? The effect of these behaviors can range from<br />

being poor role models for local youth to bringing<br />

new addictions to the community.<br />

Are other expressions of privilege demonstrated<br />

by students during their visit, such as doing<br />

things “our” way, eating “our” food, playing<br />

“our” music, requiring things to be done on “our”<br />

schedule?<br />

How well are students prepared to understand<br />

the community they are visiting? Do they bring<br />

damaging stereotypes with them that can be countered<br />

before, during, and after the program? These<br />

stereotypes might be as narrow as “Bolivians,” but<br />

for most students are more likely to be broader,<br />

such as “poor people,” “indigenous people,” or<br />

“people in developing countries,” as well as racist


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

and exoticizing images of people in out-of-the-way<br />

places.<br />

Are there human rights issues already present that<br />

are exacerbated by the presence of foreign visitors?<br />

Does anything about the students’ presence or<br />

activities reinforce a negative self-image for local<br />

people, for example that Americans are smarter,<br />

more competent, more attractive? Is there any way<br />

their presence could promote a positive self-image<br />

instead?<br />

There are many other questions that can and should be<br />

asked when considering the effects of short-term study<br />

abroad on local communities. Many of these are placespecific<br />

and evaluation requires local knowledge. Consulting<br />

local people on these questions may be helpful, but is<br />

unlikely to give a complete picture. The economic stake of<br />

having visitors return may be very high, so there is incentive<br />

to give positive reviews of the local experience of students<br />

and the impacts of their presence. Politeness compels most<br />

people to respond favorably when asked if student visits<br />

have had a positive effect on the community. And of course,<br />

most people are not trained to detect or analyze the effect<br />

of visitors on local communities. Direct observation can<br />

also be helpful, but must be considered from a critical<br />

perspective as well. Local people may be observed to smile<br />

and appear happy when they are genuinely happy, but also<br />

when they have little choice about it, as “being happy” is<br />

required for visitors to spend money, give gifts, or come<br />

back.<br />

RESEARCH DESIGN AND METHODS<br />

These issues prompted the development of a research<br />

team (composed of faculty members, student development<br />

staff, and a student) at <strong>Appalachian</strong> <strong>State</strong> University to<br />

look at the impact of our students on the places that they<br />

visited. In the course of our research, we have examined<br />

one-week alternative spring break service projects, as well<br />

as other short-term study-abroad programs, especially nontraditional<br />

ones, through a review of data on program<br />

destination and continuity, surveys of students, interviews<br />

with host community partners, and focus groups with<br />

faculty and staff leading study-abroad programs.<br />

The one-week international alternative spring break<br />

service study-abroad programs offer one hour of academic<br />

credit, but are not faculty-led. With the guidance of the<br />

campus-based volunteer and service-learning program,<br />

students develop the programs, recruit other students to<br />

participate and find a faculty partner to accompany the<br />

group. Unlike many alternative spring break programs<br />

across the country, there is academic coursework and<br />

substantial preparation and reflection required of students<br />

before and during the program abroad. From 2006 to 2008<br />

ten alternative spring break programs were implemented<br />

in the Dominican Republic, Costa Rica, Panama, Jamaica,<br />

and Belize. A total of 125 students participated over that<br />

Ideas for Mitigating Negative Community Impacts of Short-Term Study Abroad<br />

time period; some have returned for more than one year. Of<br />

these 125, twenty have subsequently enrolled in a semesterlong<br />

international program. This enrollment is significantly<br />

higher than for the population of students at large and may<br />

support the argument that international alternative spring<br />

breaks and other short-term study-abroad programs can<br />

serve as gateways to longer study-abroad experiences.<br />

In addition to alternative spring breaks, there are a significant<br />

number of other short-term study-abroad programs<br />

at our university, which are designed and implemented by<br />

faculty and professional staff. These programs all carry academic<br />

credit, with coursework completed before, during,<br />

and after the program abroad. The “abroad” components<br />

of such programs occur at various times depending on<br />

the program, including winter and spring break, but most<br />

commonly are completed during the summer.<br />

The research team collected data from four sources<br />

hoping to catch a glimpse of how students impact the<br />

communities that they visited:<br />

1. In 2007 we conducted semistructured interviews<br />

with the faculty and professional staff partners who<br />

participated on alternative spring break programs.<br />

2. That same year we conducted semistructured interviews<br />

with host agency personnel. These interviews<br />

were conducted by the faculty and staff partners on<br />

each program.<br />

3. We also analyzed data from a required student<br />

survey from the international alternative spring<br />

breaks.<br />

4. In 2008 we conducted a series of focus groups with<br />

a wider range of faculty and professional staff who<br />

have led study-abroad courses (not all of which<br />

included a service-learning component).<br />

In these group discussions, we hoped to see if participants<br />

were concerned about the impacts that their students<br />

are having on host communities and, if so, what they<br />

did to mitigate potential damage. In this research we<br />

were especially concerned about issues of equity and,<br />

in particular, who benefited locally from the students’<br />

presence in the community.<br />

RESULTS<br />

Faculty and Professional Staff Interviews<br />

The faculty and professional staff who served as partners<br />

for the alternative spring break programs in 2007 included<br />

three people who had considerable international experience<br />

and one who had no previous international travel experience.<br />

One faculty member was a geography professor. Two<br />

of the four participants expressed some skepticism about<br />

the positive impact on communities that they expected from<br />

the program before they departed.<br />

After they returned, all four faculty and professional<br />

staff partners were interviewed. All reported favorable<br />

143


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

Schroeder et al.<br />

impressions of the student impact on the communities they<br />

served. When asked why they thought that students were<br />

well received in their host communities, their responses<br />

varied. One explained that they had worked closely with a<br />

successful U.S. Peace Corps volunteer that was living in the<br />

host community as they were setting up their program.<br />

The volunteer had done an excellent job of preparing<br />

both the students and the host community for the visit.<br />

Others commented that the agencies that they were volunteering<br />

with played similar roles in preparing both<br />

students and the communities for the exchanges. One<br />

faculty member commented that the host community was<br />

so inundated with volunteers that he doubted that his<br />

particular group could have any negative impact.<br />

With regard to issues of how equitably the benefits of the<br />

exchange were shared within the communities, two faculty<br />

and professional staff partners noticed significant effort by<br />

the host agency to make sure that the tangible benefits of the<br />

students’ presence were shared within the community. One<br />

commented about the particular effort of the family that<br />

was hosting them to make sure that the students stopped<br />

by the stores that were owned and run by other families.<br />

However, another noticed that the project that they were<br />

working with seemed to be exacerbating inequality.<br />

Host Agency Interviews<br />

In 2007 faculty and professional staff partners on the<br />

alternative spring break programs were asked to interview<br />

representatives of their host agency to try to determine<br />

the possible negative impacts that student groups could<br />

have, phrased in terms of how the programs might be improved.<br />

These interviews were shared with members of the<br />

research team when the groups returned to campus. In all<br />

cases, the representatives of the host agencies said that<br />

student groups provided needed assistance and, in some<br />

cases, a significant revenue stream for their projects. This<br />

result was expected and not terribly helpful. Several of<br />

the faculty and professional staff partners pushed their<br />

interview subjects a bit more by asking about the negative<br />

impacts of other student groups in the past. One host agency<br />

representative complained about the late night drinking<br />

activities of other student groups.<br />

Student Participant Surveys<br />

Completion of a post-experience survey is a requirement<br />

for alternative spring break participants. The research<br />

group collected fifty-four completed surveys from the<br />

2007 international alternative spring break participants.<br />

Students were asked if they believed that their group had<br />

an impact in the community in which they worked. All<br />

students responded either “strongly agree” or “agree” on a<br />

five-point Likert scale. When prompted to provide details<br />

of the intangible impacts of their programs (impacts other<br />

than money and physical labor), some students felt that<br />

they reduced local people’s negative stereotypes about<br />

Americans and that they helped local people feel more<br />

144<br />

pride in their communities. Student surveys were not<br />

particularly good instruments for prompting reflection on<br />

the possible negative impact that their presence might cause<br />

(for example, demonstration effects and exacerbation of existing<br />

inequality). However, one professional staff partner<br />

commented that serious discussion of these concerns was<br />

shared during time set aside for reflection while they were<br />

in the host community.<br />

Focus Groups with Faculty and Staff<br />

In order to expand our research beyond the experiences<br />

of alternative spring break programs, four one-time focus<br />

groups were completed in 2008 with faculty and staff<br />

who had led or accompanied short-term study-abroad<br />

programs through the university’s Office of International<br />

Education and Development in the previous five years. All<br />

programs are organized by the group’s facilitator—vendors<br />

are not used on our campus. Of the forty-one faculty and<br />

staff invited to participate, twenty-six participated. Staff<br />

participants were student development professionals from<br />

a variety of offices across campus and faculty were from numerous<br />

departments, with representatives from all colleges<br />

at the university. The research team intentionally sought<br />

out participants who had led programs to “vulnerable”<br />

places: all who had led or facilitated programs in Latin<br />

America, Africa, and Asia were asked to participate, as well<br />

as one that visited the Maori in New Zealand. However,<br />

many who had been to less vulnerable places in Europe<br />

and Australia were also asked. When there was a choice,<br />

faculty and staff who had led multiple programs with<br />

larger numbers of students were asked to participate over<br />

those with fewer numbers and less experience. Participants<br />

ranged in experience from second-year assistant professors<br />

who had led one program to full professors who have been<br />

leading programs for over twenty years.<br />

One striking finding was that most participants had<br />

not seriously considered the negative impacts that their<br />

students might have abroad when they were planning<br />

their program, even those who were able to articulate<br />

potential negative effects during the focus groups. This was<br />

attributed to the real needs of pressing program logistics,<br />

including planning the syllabus. Of those who did plan<br />

for potentially negative impacts of their program, the most<br />

common measure was developing group guidelines or<br />

setting up penalties (including sending students home)<br />

for destructive behavior related to alcohol consumption.<br />

Participants were nearly unanimous in their view that<br />

alcohol abuse contributed substantially to poor student<br />

behavior and increased the likelihood that the students<br />

would have a negative impact on the community they<br />

visited.<br />

However, few participants went beyond thinking about<br />

the negative impact that alcohol can have to consider other<br />

unintended negative consequences. Few expressed concern<br />

over exacerbating income inequality, though some commented<br />

on the demonstration effects of bringing wealthy<br />

students into poor communities. In general, virtually all


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

participants assumed or asserted that economic repercussions<br />

were positive, and had considered few or no<br />

cultural or social effects of the study-abroad program<br />

they led. Even senior faculty members with extensive<br />

international experience and doctorates in a social science<br />

discipline had not thought through the implications for<br />

host communities of their visits. However, there were a<br />

few faculty and staff (not always the most experienced)<br />

who did discuss negative effects, some they had actually<br />

seen, others that they feared. Once these were mentioned<br />

in the focus groups, most participants were interested in<br />

hearing about issues they had not considered, and all were<br />

actively concerned about the effects of student visits on host<br />

communities.<br />

The few who mentioned negative impacts were concerned<br />

primarily with economic issues that result in<br />

inequity or dependency in host communities, though there<br />

were also concerns about cultural impacts. Home stays<br />

in middle-class housing, unequal distribution of gifts, the<br />

demonstration effect of student consumption, and reliance<br />

upon relatively wealthy visitors to solve local problems<br />

were all discussed. One participant had put a hold on<br />

taking students to an area he had visited for decades,<br />

because he saw major signs of dependency, on him and<br />

the relationships his programs had established in the<br />

community, for addressing local concerns, especially in<br />

providing money. One had stopped going to areas that were<br />

geographically and culturally isolated because he feared<br />

the effect of his programs on those communities, which he<br />

perceived as particularly vulnerable. On the other hand,<br />

one faculty member commented that ten more students<br />

from our university going to Madrid were not going to<br />

have much of an impact on local inhabitants.<br />

These focus groups indicate that faculty and staff who<br />

lead short-term study-abroad programs are generally unaware<br />

of possible negative impacts on host communities<br />

and do not consider the effects of their programs, with the<br />

exception of bad behavior resulting from student abuse of<br />

alcohol. However, once potential negative impacts came<br />

up in discussion, virtually all were receptive to considering<br />

these impacts and thinking about how to mitigate<br />

them.<br />

RECOMMENDATIONS AND CONCLUSIONS<br />

Although this research project is still in its early stages,<br />

we feel comfortable drawing some basic conclusions and<br />

making some general recommendations. As researchers,<br />

we have been pleased with how receptive stakeholders<br />

are to making improvements to study abroad in terms of<br />

its effects on host communities. We are hopeful, therefore,<br />

that with time and effort many of these recommendations<br />

will become standard practice at universities and colleges<br />

engaged in such programs.<br />

Institutional Commitment<br />

Colleges and universities must make institutional commitments<br />

to evaluating and mitigating the negative im-<br />

Ideas for Mitigating Negative Community Impacts of Short-Term Study Abroad<br />

pacts of short-term study abroad on host communities.<br />

The commitment of individual leaders of such programs<br />

is necessary, but not sufficient to achieve this goal, as<br />

institutions must provide training, support, and review of<br />

programs to ensure that host communities are not harmed<br />

by our students’ education abroad. As research in this<br />

previously unexplored aspect of study abroad develops,<br />

institutions can help program leaders learn about potential<br />

negative impacts of study abroad. In those institutions that<br />

use vendors to provide short-term study abroad, this<br />

commitment must be transmitted to vendors, and<br />

contracts issued only to those that meet this new<br />

criteria.<br />

Knowledgeable Program Leaders<br />

Program leaders and administrators must become as<br />

knowledgeable as possible about host communities and the<br />

ways they may be harmed by short-term study abroad, in<br />

order to predict and evaluate potential negative impacts.<br />

This process is likely to be helped tremendously by<br />

the involvement of experienced faculty. This does not<br />

necessarily mean that such faculty must lead every program<br />

since they are not necessarily sensitive or trained to observe<br />

the impacts on communities. A committed program leader<br />

may be able to reduce such effects through serious study<br />

and consultation with faculty as well as members of the<br />

community. Whenever possible, however, the expertise of<br />

faculty knowledgeable about the community or the area<br />

to which students are traveling should be involved in<br />

the planning of the program, and in evaluating potential<br />

impacts. Superficial observation is no substitute for<br />

knowledge and analysis resulting from training in the<br />

field.<br />

Student and Community Preparation<br />

Students must be prepared and guided during the program<br />

so that they become active participants in evaluating<br />

and preventing negative impacts on host communities.<br />

Preparation and work during the program has three<br />

components.<br />

First, knowledge of the community must be built so<br />

that students will understand, insofar as possible, how and<br />

why things work the way they do in that area, including<br />

the likely reaction to Americans and effects of students’<br />

presence there. Through guest lectures, readings, films,<br />

and pre-program group presentations, much of this can be<br />

accomplished and will allow service to be well-integrated<br />

with cultural learning.<br />

Second, the group must engage in systematic analyses of<br />

the many ways outsiders can affect any community, how<br />

power dynamics are likely to come into play within and<br />

between the host community and the student group, the<br />

special vulnerabilities of the community being visited, and<br />

ways to minimize negative impacts. This analysis must<br />

include the (gentle!) lesson that good intentions do not<br />

necessarily prevent harm—those intentions must be armed<br />

with knowledge, sensitivity, humility, and commitment.<br />

145


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

Schroeder et al.<br />

Facilitating power/privilege simulation activities, conducting<br />

panel discussions with previous participants, engaging<br />

in frank group discussions, and requiring reflection/reaction<br />

papers from the students before departure<br />

can all assist with teaching these important lessons before<br />

engaging with the host community.<br />

Third, group cohesion and a shared commitment to<br />

respect and share with the host community in a spirit<br />

of mutual learning, and an exchange of equals must be<br />

developed. This kind of cohesion and commitment must<br />

be created through activities, sharing and reflection both<br />

before and during the program. Before the program,<br />

students might engage in group projects about the culture,<br />

meet with students who have been on the program before,<br />

and discuss their likely reactions to the culture(s) they will<br />

be experiencing. Building a group contract about behaviors<br />

on the program that will mitigate potential negative impacts<br />

will also make for a more successful program generally.<br />

Once abroad, home stays to encourage cultural immersion,<br />

making time in the program to be with local people in<br />

contexts that allow for sharing, using local languages when<br />

possible, structuring evening reflection time that allows<br />

for open discussion, and encouraging a group journal<br />

and/or individual journals are among the practices that<br />

are fundamental to creating groups that will be sensitive<br />

to and work to prevent negative impacts on the community<br />

being visited, as well as encouraging good cultural<br />

exchange.<br />

Teaching students how to journal prior to departure<br />

is both an important academic and life skill. Instructors<br />

should develop structured journal questions that<br />

allow reflection to take a deeper, more critical perspective,<br />

rather than just reporting on the experience.<br />

Varying the reflection processes throughout the experience<br />

will accommodate various learning styles and<br />

encourage deeper, more powerful learning. Many ideas<br />

for reflection can be found on the Campus Compact<br />

(http://www.compact.org) and the National Service-Learning<br />

Clearinghouse (http://www.servicelearning.org) Web sites.<br />

Finally, consult with and prepare the community, in<br />

advance of the visit. Members of the host community may<br />

have good suggestions for minimizing negative effects,<br />

but not necessarily. They are as unlikely as students to<br />

have an understanding of foreigners, however interested<br />

they may be. If members of the community have some<br />

understanding of the students’ culture, some potential<br />

conflicts and distress may be reduced. If a local university<br />

or agency is part of the study-abroad program, they should<br />

also be brought into discussions of the effects of the program<br />

on local people, something that they may not have thought<br />

about any more than our institutions have. Ultimately, host<br />

communities should have control over if and how student<br />

groups should visit and study.<br />

Consider Not Going<br />

Consider not going on certain programs and reducing the<br />

numbers of students going overseas. Given the potential<br />

146<br />

environmental, social, and cultural costs of study abroad,<br />

there are some places there is just no responsible way to<br />

visit. Programs with the highest probability of harm to<br />

the host communities should simply not be developed,<br />

despite the recommendation of the Lincoln Commission<br />

and student demand. These locations may not always<br />

be obvious. The environmental impact of visiting some<br />

locations may be profound even when the cultural effect<br />

might be minimal. Some places may be visited so much that<br />

it seems they cannot be harmed—but the cumulative impact<br />

of many visitors to a small place may be cause to hold back<br />

from going there. Attention to the cumulative impact as<br />

well as the higher investment in time of responsible study<br />

abroad also suggests that the number of programs as well as<br />

the number of students on each program should be reduced.<br />

With greater understanding of student learning outcomes<br />

and how they can be applied, this reduction in numbers<br />

might be used to improve the quality of short-term study<br />

abroad.<br />

Long-Term Relationships<br />

Establish long-term commitments to specific communities.<br />

Most of the recommendations above are ultimately<br />

dependent on this; while many of the negative effects of<br />

study abroad can be anticipated with sufficient preparation,<br />

others cannot. Only by knowing people and learning about<br />

a community over time can these unanticipated impacts<br />

be detected and mutual trust created such that problems<br />

can be discussed and addressed. This is not a panacea—<br />

a long-term relationship may in itself become a problem<br />

if it becomes a dependent one. But the dangers of “drive<br />

by” study-abroad programs are too clear to be continued,<br />

once the effects on host communities are acknowledged.<br />

We should be clear that “drive by” is not the same as a<br />

program with a short duration. Programs that last as little as<br />

a week can be long-term provided the institution maintains<br />

a relationship with the host community. Conversely, simply<br />

because a group stays in a particular place for an extended<br />

period of time does not ensure an equitable long-term<br />

relationship.<br />

Institutional Review of Study Abroad<br />

U.S. colleges and universities should develop institutional<br />

review boards (IRBs) to screen international experiences<br />

for unintended negative impacts that shortterm<br />

study abroad might have. IRBs already exist across<br />

campuses to ensure that faculty research does not endanger<br />

the health or safety of our research subjects.<br />

The same consideration should be given to communities<br />

that host our students abroad. These IRBs should be<br />

composed of faculty with academic expertise in vulnerable<br />

areas as well as those with experience in study-abroad<br />

programs.<br />

Though the effects of short-term study abroad pale in<br />

comparison to the overall impact of other forms of travel<br />

from the U.S., academic institutions bear a special responsibility<br />

to engage in ethical relationships with communities


Downloaded By: [<strong>Appalachian</strong> <strong>State</strong> University] At: 16:54 1 September 2009<br />

hosting our students during study abroad. Understanding<br />

and working to mitigate the negative impacts of study<br />

abroad on host communities must become part of how we<br />

understand what we do when our colleges and universities<br />

sponsor such programs.<br />

We are all too aware that this research raises as many<br />

questions as it answers. How exactly does one prevent<br />

dependency from developing in a long-term study-abroad<br />

relationship? How do we determine the cultural impact<br />

over time of students’ study abroad? How do we spend<br />

money to feed and house students without increasing<br />

income inequalities? These are difficult questions, and<br />

the answers to some of them may not even be possible.<br />

But we believe that if these questions are not asked, if the<br />

impacts on local communities of study abroad continue to<br />

be ignored, that we will without a doubt be engaged in the<br />

pursuit of academic goals at the expense of people who<br />

have no choice about the matter. Many of the answers to<br />

these questions are held within our collective knowledge<br />

of the places with which we have deep relationships. Our<br />

long-term commitments to the places we have studied<br />

provide the starting point for what we hope will be long<br />

and fruitful conversation.<br />

ACKNOWLEDGMENTS<br />

The authors are indebted to our student researcher Kara<br />

Brown who did much to get this research project started<br />

and provided thoughtful critiques in its early stages. Dr.<br />

Sarah Banks, a sustainable tourism expert, also contributed<br />

substantially to this project in its formative stages.<br />

NOTES<br />

1. Research on the effects of study abroad on students<br />

remains underdeveloped. How long do students<br />

need to be abroad before they start to gain crosscultural<br />

skills? What experiences and assignments<br />

are necessary to meet learning objectives? Of the<br />

few studies that examine the impact on students,<br />

some are very disturbing. Gmelch (2004) studied the<br />

journals of students enrolled in anthropology classes<br />

at the University of Innsbruck. Students attended<br />

classes Monday through Thursday and traveled on<br />

the weekends. Analysis of student journals found<br />

that they were surprisingly shallow, naïve, and<br />

simplistic; students gained very little from simply<br />

traveling.<br />

2. Short-term study abroad is the largest growing<br />

proportion of international programs at this time,<br />

and, in our view, much more likely to have negative<br />

impacts on host communities than individual or<br />

small groups of students who go on semester or yearlong<br />

study abroad at a foreign university.<br />

Ideas for Mitigating Negative Community Impacts of Short-Term Study Abroad<br />

REFERENCES<br />

Archer, B., C. Cooper, and L. Ruhanen. 2005. The positive<br />

and negative impacts of tourism. In Global Tourism, ed.<br />

W. F. Theobold, pp. 79–102. New York: Elsevier.<br />

Campus Compact. 2008. Resources. http://www.compact.<br />

org/syllabi/syllabus.php?viewsyllabus=691 (accessed<br />

September 23, 2008).<br />

Commission on the Abraham Lincoln Study Abroad<br />

Fellowship Program. 2005. Global competence and national<br />

needs: One million Americans studying abroad.<br />

http://www.nafsa.org/public policy.sec/public policy<br />

document/study abroad 1/lincoln commission report<br />

(accessed May 14, 2008).<br />

Forum on Education Abroad. 2008. Standards of good practice<br />

for education abroad. http://www.forumea.org/<br />

standards-standards.cfm (accessed October 3, 2008).<br />

Gmelch, G. 2004. Let’s go Europe: What student tourists<br />

really learn. In Tourists and Tourism: A Reader, ed. S. B.<br />

Gmelch, pp. 419–432. Long Grove, Illinois: Waveland<br />

Press.<br />

Institute of International Education. 2007. Opendoors 2007<br />

fast facts. http://opendoors.iienetwork.org/ (accessed<br />

February 19, 2008).<br />

Lawson, V. A. 2005. Internationalizing geography at the<br />

dawn of the twenty-first century. AAG Newsletter 40:2,<br />

3.<br />

Lew, A., C. Hall, and A. Williams, eds. 2004. A Companion to<br />

Tourism. Malden, Massachusetts: Blackwell Publishers.<br />

McLaren, D. 2006. Rethinking Tourism and Ecotravel. Bloomfield,<br />

Connecticut: Kumarian Press.<br />

NAFSA: Association of International Educators. 2008.<br />

Public policy: Senator Paul Simon Study<br />

Abroad Foundation Act. http://www.nafsa.org/<br />

public policy.sec/commission on the abraham (accessed<br />

October 3, 2008).<br />

National Service-Learning Clearinghouse. 2008. Library<br />

services. http://servicelearning.org/library/lib cat/<br />

index.php?action=detailed&item=4272 (accessed September<br />

23, 2008).<br />

Shaw, G., and A. Williams. 2002. Critical Issues in Tourism: A<br />

Geographical Perspective. 2nd ed. Oxford, UK: Blackwell<br />

Publishers.<br />

Zurick, D. 1992. Adventure travel and sustainable tourism<br />

in the peripheral economy of Nepal. Annals of<br />

the Association of American Geographers 82 (4): 608–<br />

628.<br />

147


<strong>Mountain</strong>s of the World - A Global 1<br />

Priority<br />

Jack D. lues, Bruno Messerli, and Ernst Spiess<br />

INTRODUCTION - CHAPTER 13<br />

(1992) AND THE EARTH SUMMIT<br />

REVIEW (1997)<br />

In 1992, the Rio Earth Summit (UNCED) not only<br />

established the UN Commission on Sustainable<br />

Development (CSD), it set in motion the mechanism<br />

for review of the progress made within the<br />

first five years following the Summit (Rio Plus Five<br />

or Earth Summit Review, 1997). In turn, the CSD<br />

appointed Task Managers for each of the chapters<br />

of Agenda 21. Task Managers were related to Focal<br />

Points located within the existing structures of the<br />

relevant UN agencies. As mentioned in the Preface,<br />

the UN Food and Agricultural Organisation (FAO)<br />

accepted the role of Task Manager for Chapter 13<br />

(managing fragile ecosystems - sustainable mountain<br />

development) (see also Chapter 18).<br />

The main approach of the Rio Plus Five review<br />

has involved each Task Manager unit, through its<br />

own committee structure, working in close association<br />

with the CSD. The ensuing comprehensive<br />

report is to be submitted by the CSD to a Special<br />

Session of the UN General Assembly in New York,<br />

scheduled for June 1997. This report should contain<br />

the following elements:<br />

(1) overall assessment of the economic, social, and<br />

environmental situation five years after<br />

UNCED, and prospects for the future;<br />

(2) main achievements following UNCED;<br />

(3) most significant problems in Agenda 21 implementation;<br />

(4) major challenges and priorities for the period<br />

following 1997;<br />

(5) suggestions for institutional changes following<br />

1997.<br />

This general framework has been applied in standardised<br />

fashion to each Task Manager and the<br />

space allotted for the individual chapter reports has<br />

been tightly constrained. Nevertheless, beyond the<br />

precise specifications for the chapter reports, the<br />

Task Managers have had considerable freedom of<br />

action. In the case of Chapter 13, there has been a<br />

virtually unprecedented level of collaboration<br />

amongst UN agencies, national governments, international<br />

organisations, NGOs, and research institutions.<br />

This book, and the companion document:<br />

<strong>Mountain</strong>s of the World: Challenges for the 21st<br />

Century (<strong>Mountain</strong> Agenda, 1997), of course, is<br />

only one element of the FAO Chapter 13 Task<br />

Manager initiative. Its preparation, while supported<br />

financially by several governmental and UN<br />

agencies, especially by the Swiss Agency for<br />

Development and Cooperation, has been a strictly<br />

open academic and independent process.<br />

There is no doubt that the Rio Earth Summit<br />

with Agenda 21 has produced a much heightened<br />

awareness of the essential role played by the mountains<br />

as an integral part of the global biophysical<br />

and socio-economic system. It has also stimulated<br />

a large number of initiatives at all governmental<br />

and NGO levels. Whether or not this activity has<br />

moved from simply further increasing mountain<br />

awareness to actually changing conditions for the<br />

better, will be part of the task of the Special Session<br />

of the UN General Assembly to answer. In practice,<br />

we believe that much longer than five years will be<br />

required and the question will need a specific answer<br />

in a case-by-case evaluation for individual<br />

mountain regions. Thus, the present Rio Plus Five<br />

exercise, while it will go part way toward a general<br />

assessment, should be regarded as a first step in<br />

what must be framed as a long-term undertaking.<br />

In addition, the review process should aid in the<br />

identification of any trends, based on selected case<br />

studies, it should point to areas of weakness in the<br />

post-Rio process, and layout, much more precisely,<br />

the challenges and priorities that must be faced as<br />

we enter the next millennium.<br />

To provide a baseline for future mountain evaluation<br />

it should be worthwhile to consider the<br />

introductory statement from Chapter 13:<br />

1


MOUNTAINS - A GLOBAL PRIORITY<br />

<strong>Mountain</strong>s are an important source of water,<br />

energy and biological diversity. Furthermore,<br />

they are a source of such key resources as<br />

minerals, forest products, and agricultural<br />

products, and of recreation. As a major<br />

ecosystem representing the complex and<br />

interrelated ecology of our planet, mountain<br />

environments are essential to the survival of<br />

the global ecosystem. <strong>Mountain</strong> ecosystems<br />

are, however, rapidly changing. They are<br />

susceptible to accelerated soil erosion, landslides,<br />

and rapid loss of habitat and genetic<br />

diversity. On the human side, there is widespread<br />

poverty among mountain inhabitants,<br />

and loss of indigenous knowledge. As a<br />

result, most global mountain areas are<br />

experiencing environmental degradation.<br />

Hence, the proper management of mountain<br />

resources and socio-economic development<br />

of the people deserves immediate attention.<br />

(Agenda 21, Chapter 13, final version,<br />

adopted by the Plenary Session in Rio de<br />

Janeiro on 14 June, 1992).<br />

We submit that the above quotation is an imperfect<br />

statement: it was a compromise that had to be<br />

pushed through the UNCED Preparatory Committees<br />

in the face of considerable opposition. This<br />

opposition challenged the stand that mountains<br />

constituted a viable unitary system, or formulation,<br />

and argued that the mountain issues were already<br />

covered in many of the other chapters that were<br />

being developed to become Agenda 21 (Table 1.1).<br />

This led to the contention that a special chapter for<br />

mountains (i.e. Chapter 13) was not necessary.<br />

While it is impressive to see how many chapters<br />

of Agenda 21 have elements that relate directly to<br />

mountains, the same can be said for tropical<br />

rainforests, and the other 'fragile ecosystems': and<br />

at the moment of potential challenge in Rio, Chapter<br />

13 was approved by acclamation and without<br />

controversy. However, we expect that the issue of<br />

mountain authenticity will continue to surface<br />

during these times of limited resources and international<br />

uncertainties. Thus we will contend that,<br />

while the mountains do contain elements of all the<br />

other major (and more readily defined) ecosystems<br />

- tropical rain forests, arid zones, polar-type<br />

regions, coastal zones, grasslands, wetlands, and<br />

the various other forest ecosystems - their unique<br />

characteristic, verticality, overwhelmingly warrants<br />

their priority ranking. To this argument can<br />

be added the warning that many of the past<br />

2<br />

development projects failed, at least in part, because<br />

mountain specificities were either ignored or<br />

not fully understood (Chapter 14).<br />

Our argument, however, will only be justified to<br />

the extent that an integrated and comprehensive<br />

approach to mountain development, and its<br />

supporting research structure, is recognised and<br />

attained. By this we mean, amongst other requirements,<br />

a strong linkage between the human and<br />

biophysical elements of the mountain environment<br />

must be perfected, and that the mountain problems<br />

must be integrated with all related global issues<br />

(Figure 1.1).<br />

DEFINITION OF THE MOUNTAIN<br />

ZONE<br />

As with many aspects of application of mountain<br />

knowledge to policy formulation in the interests of<br />

sustainable development, a basic problem remains:<br />

how do we define mountain? The inability of<br />

mountain scholars to produce a rigorous definition<br />

Table 1.1 Agenda 21 chapters with relevance to<br />

mountains<br />

Chapter 2<br />

Chapter 3<br />

Chapter 6<br />

Chapter 7<br />

Chapter 8<br />

Chapter 11<br />

Chapter 12<br />

Chapter 14<br />

Chapter 15<br />

Chapter 18<br />

Chapter 24<br />

Chapter 26<br />

Chapter 27<br />

Chapter 28<br />

Chapter 32<br />

Chapter 33<br />

Chapter 34<br />

Chapter 35<br />

Chapter 36<br />

Chapter 37<br />

Chapter 39<br />

Chapter 40<br />

International Cooperation<br />

Combating Poverty<br />

Protecting and Promoting Human<br />

Health<br />

Sustainable Human Settlements<br />

Making Decisions for Sustainable<br />

Development<br />

Combating Deforestation<br />

Combating Desertification<br />

Sustainable Agriculture and Rural<br />

Development<br />

Conservation of Biological Diversity<br />

Protecting and Managing Fresh Water<br />

Women in Sustainable Development<br />

Strengthening the Role of Indigenous<br />

People<br />

Partnerships with NGOs<br />

Local Authorities<br />

Strengthening the Role of Farmers<br />

Financing Sustainable Development<br />

Technology Transfer<br />

Sciencefor Sustainable Development<br />

Education, Training, and Public<br />

Awareness<br />

Creating Capacity for Sustainable<br />

Development<br />

International Law<br />

Information for Decision Making<br />

(and so on)


MOUNTAINS - A GLOBAL PRIORITY<br />

<strong>Mountain</strong>s are an important source of water,<br />

energy and biological diversity. Furthermore,<br />

they are a source of such key resources as<br />

minerals, forest products, and agricultural<br />

products, and of recreation. As a major<br />

ecosystem representing the complex and<br />

interrelated ecology of our planet, mountain<br />

environments are essential to the survival of<br />

the global ecosystem. <strong>Mountain</strong> ecosystems<br />

are, however, rapidly changing. They are<br />

susceptible to accelerated soil erosion, landslides,<br />

and rapid loss of habitat and genetic<br />

diversity. On the human side, there is widespread<br />

poverty among mountain inhabitants,<br />

and loss of indigenous knowledge. As a<br />

result, most global mountain areas are<br />

experiencing environmental degradation.<br />

Hence, the proper management of mountain<br />

resources and socio-economic development<br />

of the people deserves immediate attention.<br />

(Agenda 21, Chapter 13, final version,<br />

adopted by the Plenary Session in Rio de<br />

Janeiro on 14 June, 1992).<br />

We submit that the above quotation is an imperfect<br />

statement: it was a compromise that had to be<br />

pushed through the UNCED Preparatory Committees<br />

in the face of considerable opposition. This<br />

opposition challenged the stand that mountains<br />

constituted a viable unitary system, or formulation,<br />

and argued that the mountain issues were already<br />

covered in many of the other chapters that were<br />

being developed to become Agenda 21 (Table 1.1).<br />

This led to the contention that a special chapter for<br />

mountains (i.e. Chapter 13) was not necessary.<br />

While it is impressive to see how many chapters<br />

of Agenda 21 have elements that relate directly to<br />

mountains, the same can be said for tropical<br />

rainforests, and the other 'fragile ecosystems': and<br />

at the moment of potential challenge in Rio, Chapter<br />

13 was approved by acclamation and without<br />

controversy. However, we expect that the issue of<br />

mountain authenticity will continue to surface<br />

during these times of limited resources and international<br />

uncertainties. Thus we will contend that,<br />

while the mountains do contain elements of all the<br />

other major (and more readily defined) ecosystems<br />

- tropical rain forests, arid zones, polar-type<br />

regions, coastal zones, grasslands, wetlands, and<br />

the various other forest ecosystems - their unique<br />

characteristic, verticality, overwhelmingly warrants<br />

their priority ranking. To this argument can<br />

be added the warning that many of the past<br />

2<br />

development projects failed, at least in part, because<br />

mountain specificities were either ignored or<br />

not fully understood (Chapter 14).<br />

Our argument, however, will only be justified to<br />

the extent that an integrated and comprehensive<br />

approach to mountain development, and its<br />

supporting research structure, is recognised and<br />

attained. By this we mean, amongst other requirements,<br />

a strong linkage between the human and<br />

biophysical elements of the mountain environment<br />

must be perfected, and that the mountain problems<br />

must be integrated with all related global issues<br />

(Figure 1.1).<br />

DEFINITION OF THE MOUNTAIN<br />

ZONE<br />

As with many aspects of application of mountain<br />

knowledge to policy formulation in the interests of<br />

sustainable development, a basic problem remains:<br />

how do we define mountain? The inability of<br />

mountain scholars to produce a rigorous definition<br />

Table 1.1 Agenda 21 chapters with relevance to<br />

mountains<br />

Chapter 2 International Cooperation<br />

Chapter 3 Combating Poverty<br />

Chapter 6 Protecting and Promoting Human<br />

Health<br />

Chapter 7 Sustainable Human Settlements<br />

Chapter 8 Making Decisions for Sustainable<br />

Development<br />

Chapter 11 Combating Deforestation<br />

Chapter 12 Combating Desertification<br />

Chapter 14 Sustainable Agriculture and Rural<br />

Development<br />

Chapter 15 Conservation of Biological Diversity<br />

Chapter 18 Protecting and Managing Fresh Water<br />

Chapter 24 Women in Sustainable Development<br />

Chapter 26 Strengthening the Role of Indigenous<br />

People<br />

Chapter 27 Partnerships with NGOs<br />

Chapter 28 Local Authorities<br />

Chapter 32 Strengthening the Role of Farmers<br />

Chapter 33 Financing Sustainable Development<br />

Chapter 34 Technology Transfer<br />

Chapter 35 Sciencefor Sustainable Development<br />

Chapter 36 Education, Training, and Public<br />

Awareness<br />

Chapter 37 Creating Capacity for Sustainable<br />

Development<br />

Chapter 39 International Law<br />

Chapter 40 Information for Decision Making<br />

(and so on)


MOUNTAINS - A GLOBAL PRIORITY<br />

Plate 1.1 Nevado Sajama, 18° South, Bolivia's<br />

highest mountain and site of the world's highest forests<br />

tPolylepis trees close to 5000 m). (Photograph:<br />

B. Messerli)<br />

Scottish Cairngorms, and large sections of the<br />

mountains in Norway and Sweden, at least rise well<br />

above both the climatic and the anthropomorphic<br />

timberlines; the Mittelgebirge (Middle <strong>Mountain</strong>s)<br />

of Central Europe rarely do so. Yet they must all<br />

be included as mountains.<br />

We can trace the beginnings of intellectual realisation<br />

of the significance of mountains in the<br />

'Western context' to Alexander von Humboldt<br />

(1769-1859). His travels in the Andes led him to<br />

categorise the three-dimensional nature of mountains<br />

and to recognise the influence of latitude on<br />

the elevation of the distinctive altitudinal belts<br />

(Figure 1.2). These altitudinal belts primarily identify<br />

the upslope progression of changes in vegetation<br />

cover and landforms. He also recorded the<br />

adaptation of the <strong>Andean</strong>s to these altitudinal 'life<br />

zones' in terms of their subsistence and patterns of<br />

movement and exchange. Much later, this very<br />

4<br />

early initiative evolved into John Murra's (1972)<br />

concept of the 'vertical archipelago', itself derived<br />

from Carl Troll's (1954, 1975, 1978) elaboration<br />

of Humboldt's legacy (Bromme, 1851) into his<br />

more sophisticated concept of altitudinal belts.<br />

Troll went on to attempt a natural science classification<br />

of mountains with a clear distinction between<br />

Hochgebirge (High <strong>Mountain</strong>s, such as the<br />

Alps) and Mittelgebirge (Middle <strong>Mountain</strong>s, such<br />

as the Black Forest), and between the glaciated and<br />

non-glaciated mountains.<br />

Nevertheless, since Troll's interests extended to<br />

embrace mountains world-wide, he quickly identified<br />

the dilemma facing any attempt to produce a<br />

universal classification. For instance, he referred to<br />

mountain areas in the humid equatorial zone, such<br />

as those extending above 3000 m in Indonesia, as<br />

'high mountains without a high mountain landscape'.<br />

This problem facing attempts to produce a<br />

classification of mountains becomes even more<br />

acute, for instance, when we try to incorporate the<br />

high lands of Ethiopia and East Africa. These highland<br />

areas traditionally have been much more<br />

favourable to human settlement than the surrounding<br />

arid and semi-arid lowlands. Further difficulties<br />

arise with mountains in the extreme arid zone, such<br />

as the Eastern Pamir and the Tibesti and Hoggar of<br />

the Sahara that have neither forest belts nor landforms<br />

produced by former glaciation. And for<br />

mountains in general, the most intensely utilised<br />

terrain is that located below the high mountain<br />

stage (upper timberline) (Figure 1.3): here are to be<br />

found the more serious intensities of human landscape<br />

intervention. More recently, these schematic<br />

representations have combined physical and<br />

human attributes of verticality (Lauer, 1993; Figure<br />

1.4) and have been employed to demonstrate<br />

responses to climatic change over thousands of<br />

years (Messerli, et al., 1993; Figure 1.5).<br />

Concurrent with, and subsequent to, but also<br />

using the Trollean altitudinal belt classification as<br />

a starting point, Uhlig (1984, 1995), Grotzbach<br />

(1984, 1988), Kreutzmann (1995), and others,<br />

developed a tentative cultural geography of mountain<br />

regions (see Chapter 2). This depicts a fine line<br />

between the impacts of increasing altitude (the<br />

verticality issue) and the overlay of culture and,<br />

more recently, the penetration of the world market<br />

economy with the accelerated spread of modern<br />

communication systems. Uhlig, in particular,<br />

studied the altitudinal limits of domestic crops and<br />

forms of human subsistence, such as Almwirtschaft


(mixed mountain farming), transhumance, and nomadism<br />

(Uhlig, ed. Kreutzmann, 1995). Crotzbach<br />

(1988) proposed the first cultural geographic<br />

typology of mountains.<br />

From the foregoing discussion it should be<br />

apparent that the world's mountains do not lend<br />

themselves to a unifying definition and classification<br />

that goes beyond the simple combination of<br />

'steepness of slope' and 'altitude'. <strong>Mountain</strong>s,<br />

obviously, are regions of accentuated relief and<br />

MOUNTAINS OF THE WORLD<br />

altitude, which influence climate, soil fertility,<br />

vegetation, slope instability, and accessibility.<br />

From the humid sub-tropical and temperate zones<br />

polewards, all land-use activities are disadvantaged<br />

compared with the subjacent and neighbouring<br />

more densely populated lowlands. This comparative<br />

mountain disadvantage has been recognised in<br />

the recent extant and pending legislation of<br />

European countries, such as Switzerland, Germany,<br />

Norway, and others (Chapters 4 and 5). But<br />

Plate 1.2 <strong>Mountain</strong>s in high latitude, even of modest elevation, display a 'high mountain' (Hochgebirge)<br />

landscape, comparable to the Alps cut off at timberline and reduced to sea level. Inugsuin Pinnacles, Baffin<br />

Island, Canada, in latitude 70 0N. (Photograph: J. D. Ives)<br />

5


MOUNTAINS - A GLOBAL PRIORITY<br />

11 percent exceeds 2000 m; 5 percent lies above<br />

3000 m; and 2 percent above 4000 m. In certain<br />

regions of the world, such as Western Europe, the<br />

humid inner tropics, and the Pacific high islands,<br />

500 m is already a significant altitude. However,<br />

much of Kansas, Nebraska, and eastern Colorado<br />

lies appreciably higher than this and, to the inhabitants<br />

of these flat prairie lands, mountains begin<br />

much further westward, along the Rocky <strong>Mountain</strong><br />

front at about 1500 m. Similar comparisons<br />

can be made for other parts of the world.<br />

To demonstrate that nearly half of the world's<br />

land surface lies above 500 m tends to support our<br />

conclusion that the search for a unitary definition<br />

of mountain is to chase a chimera. It follows that<br />

several definitions, which are region-specific, are<br />

needed. While we believe it has been necessary to<br />

pursue this rather tortuous path in our opening<br />

chapter, we also contend that to pursue it further<br />

becomes academic. We therefore return to our<br />

long-standing guestimate that 'mountains occupy<br />

about a fifth of the world's land surface and provide<br />

the direct life-support base for about a tenth<br />

of humankind'. We will return to this 'truism' after<br />

a global survey of mountains and a discussion of<br />

the one outstanding aspect - verticality - mountains<br />

as high-energy environments.<br />

A BRIEF SURVEY OF THE WORLD'S<br />

MOUNTAINS<br />

<strong>Mountain</strong>s are found on every continent, from the<br />

equator polewards as far as land persists. Taken<br />

together, on account of their three-dimensional<br />

nature, as a single great landscape category, or<br />

ecosystem in the broadest sense, they encompass<br />

the most extensive array of topography, climate,<br />

flora and fauna, as well as human cultural differentiation,<br />

known to humankind. Furthermore,<br />

geologically and tectonically, mountains comprise<br />

the most complex elements of the earth's underlying<br />

structures.<br />

<strong>Mountain</strong>s and uplands incorporate the inhuman<br />

and extremely cold and sterile high ice<br />

plateaus of Antarctica and Greenland, and the<br />

high, dry, hypoxic, and almost inhospitable ranges<br />

of Central Asia and the south-central High Andes.<br />

They also include the richly varied, and even<br />

luxuriant, ridge and valley systems of the humid<br />

subtropics and tropics, such as the Himalaya, the<br />

Hengduan <strong>Mountain</strong>s, Mt Cameroon, sections of<br />

the Northern Andes, and parts of New Guinea. In<br />

East Africa and Ethiopia the flanks of the high<br />

8<br />

mountains have long been a preferred human habitat<br />

compared with the more arid lowland areas<br />

that surround them. Additionally, there must be<br />

included an enormous compass of other mountains.<br />

Thus, there are the high volcanoes, for instance,<br />

of Indonesia, the Caribbean, and Hawai'i,<br />

where humans have long benefited from access to<br />

rich soils and have been exposed to extreme fiery<br />

hazards. There are also what are usually called<br />

middle mountains (German: Mittelgebirge), ranging<br />

from Tasmania to South Africa, and from<br />

Central and Northern Europe to the Urals and<br />

Siberia. While these latter contrast with the Alps<br />

(the epitome of the Hochgebirge) because of<br />

their more subdued relief, their other mountain<br />

attributes demand special policies to ensure sustained<br />

resource use.<br />

<strong>Mountain</strong>s and uplands contain the largest<br />

number of environmentally protected areas,<br />

including biosphere reserves, national forests,<br />

national and international parks, and World Heritage<br />

Sites (see Chapter 11), of any of the world's<br />

major landscape categories. Even when we exclude<br />

the two big ice sheets, mountains provide more<br />

than half the world's fresh water (see Chapter 7),<br />

as well as significant proportions of its timber (see<br />

Chapter 13), minerals (see Chapter 9), and grazing<br />

lands (see Chapter 15). They also serve as the abode<br />

of the deities of many of the world's religions (see<br />

Chapter 3) and provide an over-arching spirituality,<br />

aesthetic, and source of myth, legend, and<br />

psychological balm and aspiration for society at<br />

large (see Chapter 12).<br />

When the intrinsic and spiritual resources are<br />

considered together, we are compelled to assert that<br />

mountains, in addition to providing the life-support<br />

base for about ten percent of the world population,<br />

are vital for the well-being of more than half<br />

of humankind. Their priority ranking, as world<br />

governance begins to grapple with sustainable<br />

development for the twenty-first century, would<br />

appear secure from the foregoing introductory remarks<br />

alone. Yet we must also take into account<br />

the actual and perceived threats to the lowlands if<br />

mismanagement of mountain resources continues<br />

unabated. In this sense, mountains are not only<br />

suppliers of many products, they are protection<br />

watersheds for the lowlands. The converse of 'protection'<br />

is that they are potential destroyers of the<br />

life-support systems of the hundreds of millions of<br />

people on the plains. Even when the threats, for<br />

instance, of deforestation as a cause of catastrophic<br />

flooding of Gangetic India and Bangladesh, are


Plate 1.3 Mount Kasbegh (5047 rn), Caucasus,<br />

once the place where an angry Zeus chained the<br />

miscreant Prometheus for giving fire to mankind.<br />

Today, Georgians, Ossetians, Chechens, and Russians<br />

devastate both the mountain environment and<br />

each others' homelands. (Photograph: J. D. Ives)<br />

only perceived rather than proven, there are both<br />

serious implications as well as prospects for profligate<br />

expenditures to solve problems that have not<br />

been correctly analysed (Ives and Messerli, 1989).<br />

In many parts of the world, mountains form<br />

international and provincial frontiers, and many of<br />

these today are in contention. We need only<br />

mention the entire Hindu Kush-Karakorum­<br />

Himalayan region; the frontiers of Chile, Argentina,<br />

Bolivia, Peru, and Ecuador in the Andes; the<br />

Caucasus; and the Balkans. The majority of today's<br />

most pernicious and destructive armed conflicts,<br />

guerrilla activities (Chapter 6), as well as the devastating<br />

drug wars, hinge on our mountains.<br />

But there are also positive aspects that need to<br />

bereinforced. <strong>Mountain</strong>s harbour by far the largest<br />

number of distinct ethnic groups, varied remnants<br />

of cultural traditions, environmental knowledge,<br />

and habitat adaptations; they host some of the<br />

world's most complex agro-cultural gene pools and<br />

traditional management practices (ef Peruvian<br />

Andes: Zimmerer, 1996). They offer primary challenges<br />

to research and scholarship and will likely<br />

prove vital field laboratories for early detection of<br />

some of the first indications of climate change<br />

(Chapter 17). The positive benefits that are inher-<br />

MOUNTAINS OF THE WORLD<br />

ent in the proliferation of trans-border parks-forpeace,<br />

inspirational enrichment, recreation, and<br />

spiritual and physical challenge, make mountains<br />

an indispensable element of the human heritage.<br />

These attributes will become increasingly important<br />

if the enormous problems of the next century<br />

are to be faced and overcome.<br />

Perhaps the only generalisation that can be<br />

made about mountains, however, is that, except for<br />

the three-dimensional landscape characteristic and<br />

the relative marginality and inaccessibility, generalisation<br />

should be avoided. In the popular eye,<br />

until recently (and still a prevailing trait), mountains<br />

have been perceived as vast, rugged, and<br />

remote landscapes, seemingly inured to human<br />

environmental impacts. Despite this, by the 1960s<br />

another, contrasting, popular view emerged. In this<br />

view, over-development, winter sports, mass tourism<br />

in general, rapid population growth, deforestation,<br />

soil erosion, accelerated run-off and devastating<br />

downstream effects, have been perceived as<br />

leading to a world super-crisis. Yet these viewpoints<br />

are overwhelmingly those from centres of<br />

population and power from outside the mountains.<br />

At the same time as this great mountain contradiction<br />

was emerging (the one viewpoint producing<br />

complacency and benign neglect, the other leading<br />

to hysteria) bilateral and international development<br />

agencies tended to treat mountains as<br />

unimportant two-dimensional adjuncts to be<br />

accommodated as fringe attachments to the big<br />

development projects on the surrounding and<br />

much more densely populated plains. It seems that<br />

there was (is) a prevailing conviction that, dollar<br />

for dollar, better investment value would be obtained,<br />

in terms of economic successes, in the lowlands.<br />

Here accessibility is superior, infrastructure<br />

better developed, or less costly to establish and<br />

maintain. Traditionally, therefore, most political<br />

and economic interactions between highlands and<br />

lowlands have been initiated from the lowlands:<br />

policy decisions made in the lowlands and imposed<br />

on the highlands. The inhabitants of the mountain<br />

regions have been obliged to suffer exploitation, or<br />

to defend themselves, or to react to outside pressures;<br />

they have had little chance to act out their<br />

own destinies.<br />

PHYSICAL PROCESSES IN MOUNTAINS<br />

In providing an overview of the dynamic geomorphic<br />

processes operating in mountains, we must<br />

emphasise that by far the majority of twentieth<br />

9


MOUNTAINS - A GLOBAL PRIORITY<br />

century research in this field has concentrated on<br />

the Hochgebirge. The reasons for this are obvious<br />

and yet are worthy of emphasis here. First, there is<br />

the personal attitude of individuals from 'Western'<br />

affluent societies. The high mountains are dramatically<br />

appealing and there is no doubt that an<br />

adventure element has influenced many mountain<br />

researchers. Second, there is also a pragmatic<br />

element. With the quantification of geophysical<br />

mountain research in the two decades following<br />

World War II, steeper and relatively unvegetated<br />

slopes greatly facilitated the acquisition of 'precisely'<br />

surveyed process data in a relatively short<br />

time period, simply because the processes are more<br />

dynamic under these types of locale. (Downslope<br />

movement proceeds much more rapidly and, therefore,<br />

can be measured to a much higher standard<br />

of accuracy with the available instruments in unit<br />

time.) A third factor could well be that of the<br />

relative inaccessibility of the research site, whereby<br />

markers and instruments could be left in place over<br />

several years with much reduced risk of interference<br />

by curious villagers or passers-by.<br />

Today, a vast literature exists on this general<br />

field of enquiry. Some of the elements of high<br />

mountain process studies are mentioned here because<br />

the results must be taken into consideration<br />

whenever human interventions are to be perpetrated<br />

on mountain slopes at any altitude. The<br />

Plate 1.4 Unteraargletscher near Grimselpass,<br />

Switzerland: frost weathering, glaciated landforms<br />

and retreating glaciers. (Photograph: J. D. Ives)<br />

10<br />

primary English language references are: Rapp's<br />

(1960) pioneer study of mountain slope processes;<br />

Price's (1981) synthesis of mountain physical geography;<br />

Gerrard's (1990) more recent synthesis of<br />

research on the same topic; Barsch (1984); and<br />

Hewitt's attempts to link our growing knowledge<br />

of catastrophic processes and their interactions<br />

with humans and their property, in terms of natural<br />

hazard research (see Chapter 16).<br />

Itcan be argued that more geomorphic work can<br />

be accomplished during single catastrophic events<br />

than from all mass-transfer precisely measured for<br />

a decade, or observed over centuries (Ives, 1989).<br />

The effects of a hundred-year or even several<br />

hundred years' event may also be dwarfed by even<br />

more spectacular occurrences: for example, a very<br />

large landslide that occurred in the Langtang<br />

Himal, Nepal, 25 000 or more years ago. This<br />

landslide displaced approximately 10 km 3 of debris<br />

through a vertical distance of 2000 m. Something<br />

similar occurred at Kofels, in the Otztal Alps<br />

of Austria, about 9500 years ago, when a landslide<br />

displaced 3 km ' of debris through up to 1000<br />

vertical metres (Heuberger et al., 1984).<br />

Some other examples may be mentioned here.<br />

First, in 1970, a debris flow, resulting from an<br />

earthquake-induced avalanche near the summit of<br />

Huascaran (6768 m) in the Peruvian Cordillera<br />

Blanca, covered 150 km 2 with debris, annihilated<br />

the town of Yungay, and killed 18000 people<br />

(Patzelt, 1983). Second, in October, 1968, rainfall,<br />

varying between 600 and 1200 mm fell on the<br />

Darjeeling area of the Himalaya, during a three-day<br />

period at the end of the summer monsoon period<br />

when the ground was already saturated. Some<br />

20000 people were killed, injured, or displaced<br />

(Ives, 1970; Starkel, 1972). Third, in southeast<br />

Iceland 155 mm of rain in twenty-four hours on 29<br />

July 1982 caused several dozen debris flows in a<br />

small area of Skaftafell National Park. One of us<br />

has photographed and observed this area qualitatively<br />

from 1952 to 1994 and estimates that (with<br />

the exception of the very high rates of glacial<br />

erosion, deposition, and associated jokulhlaup or<br />

glacial lake outburst flood and fluvio-glacial activity)<br />

this single geomorphic event accounted for<br />

more 'work' than all other processes combined<br />

over the last 100 years. Fourth, on July 19 and 20,<br />

1993, an extraordinary flood event took place in<br />

eastern and central Nepal with catastrophic effects:<br />

several districts were hit by floods and hundreds of<br />

landslides, and many people died or became homeless.<br />

Due to the very high level of sedimentation,


MOUNTAINS - A GLOBAL PRIORITY<br />

knowledge and long-term experience relating to the<br />

use of mountain resources. External impacts, such<br />

as mass-tourism and new traffic systems in<br />

industrialised countries, and economic decline and<br />

migration in developing countries, have disturbed<br />

this balance. The following three examples<br />

illustrate how mountain areas under different<br />

climatic-ecological, socio-cultural and economicdemographic<br />

conditions have been confronted<br />

with very different problems and processes on their<br />

development path from the past into the future:<br />

(1) Let us first consider Kilimanjaro, a volcanic<br />

massif in the tropical zone close to the equator:<br />

in addition to its cultural and spiritual significance,<br />

this mountain is the primary source of<br />

much of the water, food, fuel, and building<br />

material for the people of north-central Tanzania.<br />

Unfortunately, the capacity of Mount<br />

Kilimanjaro to continue to provide these vital<br />

products and services is being threatened by<br />

inappropriate and, in some cases, over-exploitation<br />

of many of its natural resources. Much<br />

of the current stress on natural resources is a<br />

result of the dramatic increase in human population<br />

on the slopes of Mount Kilimanjaro: it<br />

has more than tripled during the last forty<br />

years, and the mountain area now has the<br />

highest rural density of any administrative<br />

district in Tanzania (Newmark, 1991).<br />

By contrast with mountain areas in other<br />

climatic zones, the tropical mountains offer<br />

very favourable ecological conditions with<br />

fertile soils, especially on volcanoes, and sufficient<br />

precipitation for agricultural production.<br />

But will it be possible to find a balance<br />

between the new cultural, economic, and ecological<br />

forces?<br />

(2) Our second example is the Middle <strong>Mountain</strong>s<br />

of Nepal in the subtropical climatic zone with<br />

a seasonal monsoon precipitation regime:<br />

parts of Nepal are facing a serious problem in<br />

being unable to maintain soil fertility in<br />

agriculture and forestry. In order to meet the<br />

food, animal feed, and fuelwood demands of<br />

the rapidly growing population, there has been<br />

a gradual transformation from single to multiple<br />

annual crop rotation. This practice has<br />

developed to the point where three crops are<br />

grown annually in all areas of the Middle<br />

<strong>Mountain</strong>s where irrigation water is available.<br />

Soil fertility has changed and soils have become<br />

very acidic. All soil fertility conditions<br />

12<br />

are marginal and put into question the longterm<br />

sustainability of current levels of production<br />

(Schreier et al., 1994). How can a new<br />

balance be struck between ecological conditions<br />

and economic needs? Can the needed<br />

alteration in cropping intensity and the introduction<br />

of nitrogen-fixing trees and crops<br />

constitute sufficient responses for promoting<br />

sustainability?<br />

(3) Third, let us consider the Alps, situated in the<br />

temperate climatic zone, surrounded by highly<br />

urbanised and industrialised areas. Recent<br />

research projects have found that today about<br />

44 percent of the Alpine population is living<br />

in so-called Alpine towns, concentrated in the<br />

main valleys along the traffic routes (pers.<br />

comm. Paul Messerli; Batzing et aI., 1996).<br />

These approximately 150 towns could assume<br />

new functions or new responsibilities in the<br />

future development of Alpine regions. But it is<br />

not yet clear, whether these urban populations<br />

could preserve a special Alpine-oriented cultural<br />

identity, and whether there is the necessary<br />

interest and capacity to create a new<br />

balance between rural and urban areas within<br />

the Alps. Could this represent an opportunity<br />

for new forms of co-operation between mountain<br />

farming and off-farm employment in the<br />

urban areas close by?<br />

These three examples, illustrate three very different<br />

situations on the long and difficult path to sustainable<br />

development. New questions are being raised<br />

everywhere, and new solutions must be found in<br />

connection with all the internal and external forces.<br />

Solutions could include: economic diversification,<br />

appropriate agriculture, adapted forms of tourism,<br />

controlled migration, long-term investment in<br />

infrastructure, improved social conditions, better<br />

education, and more political autonomy.<br />

CONCLUSION<br />

The complexities of the high mountain stage<br />

(Hochgebirge), in terms of the physical processes<br />

that are occurring today and have occurred in the<br />

past, have been reviewed briefly above. It is equally<br />

apparent, however, that inhabited mountain<br />

regions are even more highly complex fields for<br />

study and, therefore, for management. Thus, international<br />

collaborative efforts are needed, including<br />

standardisation of objectives and methods, shared<br />

data banks, and identification of minimum needs.


MOUNTAINS OF THE WORLD<br />

Plate 1.6 The Yulongxue Shan (Jade Dragon Snow <strong>Mountain</strong>s 5596 m) from the black Dragon Park,<br />

Lijang City, NW Yunnan, China. (Photograph:]. D. Ives)<br />

In these closing years of the twentieth century that<br />

are characterised by increasing scarcity of research<br />

funds, this recommendation for expansion of<br />

effort, even for the establishment of a world network<br />

of mountain research activities, may appear<br />

impractical. In this respect it is worth considering<br />

the fate of some highly expensive development<br />

projects in mountain regions. For instance, the<br />

engineer's estimate of the useful life of a large<br />

hydroelectric scheme is often proved to have been<br />

greatly over-optimistic because no attention was<br />

paid to the dynamics of the mountain catchment in<br />

which the infrastructure was placed. Reservoirs<br />

that rapidly fill with sediment should not be contemplated,<br />

let alone constructed; and this occurs so<br />

frequently and at great cost to all concerned because<br />

of the lack of attention paid to geomorphic<br />

and hydrologic conditions. Similarly, poorly<br />

designed roads lead to instability of extensive<br />

mountain slopes and serious downslope damage.<br />

Resource mismanagement in mountain lands in<br />

general, often due to ignorance, or to disregard for<br />

long-term objectives and concentration on the<br />

short term, has the potential for devastating losses<br />

in the surrounding and densely populated lowlands.<br />

The problem is exacerbated by the extreme<br />

unreliability of much of the available data and the<br />

tendency to generalisation instead of careful local<br />

assessment.<br />

The need for special attention being given to the<br />

mountains and their people, therefore, is understandable.<br />

If aid and development must take on a<br />

capitalist concern with immediate returns on investment<br />

then, obviously, it would be rational to<br />

neglect the mountains. But this is the short-term<br />

view. The long-term implications are becoming<br />

increasingly clear and are inextricably entwined in<br />

the growing concern that continuing development<br />

must be sustainable. This overwhelmingly reflects<br />

the remarkable initiatives taken during the Rio<br />

Earth Summit (UNCED). The present volume has<br />

the intention of aiding the clarification of the longterm<br />

viewpoint; this chapter to lay the general<br />

framework for that clarification.<br />

13


MOUNTAINS - A GLOBAL PRIORITY<br />

References<br />

Agenda 21 (1992). A Guide to Agenda 21 - A Global<br />

Partnership, UNCED: Geneva, p. 115<br />

Aulitzky, H. (1974). Endangered Alpine Regions and<br />

Disaster Prevention Measures, Nature and Environment<br />

Series 6, Council of Europe, Strasbourg, p. 103<br />

Batzing, W., Perlik, M. and Deklova, M. (1996).<br />

Urbanization and depopulation in the Alps (with 3<br />

colored rnaps). <strong>Mountain</strong> Research and Development,<br />

16(4): 335-50<br />

Barsch, D. (ed.) (1984). High <strong>Mountain</strong> Research.<br />

Special issue of <strong>Mountain</strong> Research and Development,<br />

4 (4): 286-374<br />

Bromme, T. (1851). Atlas zu Alexander von Humboldt.<br />

Kosmos: Stuttgart<br />

Dow, V., Kienholz, H., Plam, M. and Ives, ]. D.<br />

(1981). <strong>Mountain</strong> hazards mapping: Development of<br />

a prototype combined hazards map, Monarch Lake<br />

Quadrangle, Colorado, USA (with fold-in map).<br />

<strong>Mountain</strong> Research and Development, 1(1): 55-64<br />

Grotzbach, E. F. (1984). Mobility of labour in high<br />

mountains and the socio-economic integration of peripheral<br />

areas. <strong>Mountain</strong> Research and Development,<br />

4 (3): 229-35<br />

Grotzbach, E. F. (1988). High <strong>Mountain</strong>s as Human<br />

Habitat. In Allan, N.]. R., Knapp, G. W. and Stadel,<br />

C. (eds.) Human Impact on <strong>Mountain</strong>s. Rowman and<br />

Littlefield: Totowa, N], pp. 24-35<br />

Gerrard, A.]. (1990). <strong>Mountain</strong> Environments: An<br />

examination ofthe physical geography ofmountains.<br />

MIT Press: Cambridge, MA, p. 317<br />

Heuberger, H., Masch, L., Preuss, E. and Schroecher,<br />

A. (1984). Quaternary Landslides and Rock Fusion in<br />

Central Nepal and in the Tyrolean Alps. <strong>Mountain</strong><br />

Research and Development, 4 (4): 345-62<br />

Heuberger, H. and Ives,]. D. (eds.) (1994). <strong>Mountain</strong><br />

Hazards Geomorphology. Special issue of <strong>Mountain</strong><br />

Research and Development, 14 (4): 271-363<br />

Ives, ]. D. (1970). Himalayan Highway. Canadian<br />

Geographical Journal, 80: 26-31<br />

Ives, ]. D. (1989). <strong>Mountain</strong> Environments. In<br />

Marini-Bettolo, G. B. (ed.) A Modern Approach to the<br />

Protection of the Environment. Citra del Vaticano:<br />

14<br />

Pontificiae Academiae Scientiarum Scripta Varia,<br />

pp.289-345<br />

Ives,]. D. and Messerli, B. (1981). <strong>Mountain</strong> hazards<br />

mapping in Nepal: Introduction to an applied mountain<br />

research project. <strong>Mountain</strong> Research and<br />

Development, 1 (3/4): 223-30<br />

Ives, ]. D. and Messerli, B. (1989). The Himalayan<br />

Dilemma: Reconciling Development and Conservation.<br />

Routledge: London and New York, p. 295<br />

Kienholz, H., Hafner, H., Schneider, G. and Tamrakar,<br />

R. (1983). <strong>Mountain</strong> hazards mapping in<br />

Nepal's Middle <strong>Mountain</strong>s, with maps of land use and<br />

geomorphic damages (Kathmandu-Kakani area).<br />

<strong>Mountain</strong> Research and Development, 3 (3): 195-220<br />

Kreutzrnann, H. (1995). Globalization, spatial integration,<br />

and sustainable development in northern<br />

Pakistan. <strong>Mountain</strong> Research and Development, 15<br />

(3): 213-27<br />

Lauer, W. (1993). Human development and environment<br />

in the Andes: A geoecological overview. <strong>Mountain</strong><br />

Research and Development, 13 (2): 157-66<br />

Liu, Dong-sheng and Sun, Honglie (eds.) (1981). Geological<br />

and Ecological Studies of Qinghai-Xizang<br />

Plateau. Science Press, Gordon and Breach, Science<br />

Publishers Inc.: Beijing and New York, 2 vols., p. 2138<br />

Messerli, B., et al. (1993). Climate change and natural<br />

resource dynamics of the Atacama Altiplano during<br />

the last 18,000 years: A preliminary synthesis. <strong>Mountain</strong><br />

Research and Development, 13 (2): 117-27<br />

<strong>Mountain</strong> Agenda (1997). <strong>Mountain</strong>s of the World:<br />

Challenges for the 21st Century. Geographical Institute<br />

University of Bern: Bern, p. 36<br />

Murra, ]. V. (1972). El 'vertical control' de un<br />

maximo de pisos ecologicos en la economia de<br />

las sociedades Andinas. In Ortiz de Zuniga, 1. (ed.),<br />

Visita de la Provincia de Leon de Huanuco (1562).<br />

Torno 2, Universidad Hermilio Valdizan: Huanuco,<br />

pp.429-76<br />

Newmark W. D. (ed.) (1991). The Conservation of<br />

Mount Kilimanjaro. IUCN Tropical Forest Programme:<br />

Gland and Cambridge<br />

Patzelt, G. (ed.) (1983). Die Berg- und Gletscberstiirze<br />

von Huascaran, Cordillera Blanca, Peru. Innsbruck,<br />

Hochgebirgsforschung, no. 6, p. 110


3523_book.fm Page 321 Thursday, September 29, 2005 10:38 AM<br />

24<br />

Multidimensional (Climatic,<br />

Biodiversity, Socioeconomic),<br />

Changes in Land Use in the<br />

Vilcanota Watershed, Peru<br />

Stephan Halloy, Anton Seimon, Karina Yager,<br />

and Alfredo Tupayachi<br />

INTRODUCTION<br />

To investigate the dynamic changes affecting<br />

biodiversity across the vertical gradient of the<br />

Vilcanota watershed in Peru, we utilize the<br />

major vertical profile of the Vilcanota–Urubamba<br />

Valley (the Sacred Valley of the<br />

Incas at its center). The area combines features<br />

of interest for our research, such as a tropical<br />

location in a major biodiversity hotspot, which<br />

has also been a cultural vortex with thousands<br />

of years of occupation and development of<br />

resilient sustainable land uses; the point of origin<br />

of many indigenous agricultural staples,<br />

some of which are now important agricultural<br />

crops at a global level; and a unique annually<br />

resolved climatic record of more than 500 years<br />

in the Quelccaya ice cap to the southeast of the<br />

watershed (Thompson et al. 1985; Seimon and<br />

Halloy, this volume). As it descends, the Vilcanota–Urubamba<br />

changes its cross section (Figure<br />

24.1), topography, and mesoclimates, traversing<br />

an extreme range of climates and<br />

environments. These have been described and<br />

classified by many researchers (e.g. Brisseau,<br />

1981; Galiano Sánchez, et al., 1995; Gentry,<br />

1993; Sibille, 1997). The watershed starts in the<br />

permanent snow and glaciers of the steep peaks<br />

above 6300 m (Ausangate), where mean temperatures<br />

are below 0°C. We recently recorded<br />

(in 2002) the highest vascular plants at 5510 m,<br />

close behind the retreating glaciers in this area.<br />

High-<strong>Andean</strong> vegetation develops rapidly down<br />

from this level. Around 4900 m, llama and<br />

alpaca grazing signal the rising level of human<br />

occupation. The highest human occupation<br />

found is the house of Pedro Godofredo above<br />

Murmurani, at ~5050 m.<br />

The undulating altiplano between 4900 and<br />

4200 m gives way to steep incised valleys as<br />

the rivers cut their way down to the Amazon.<br />

As in the altiplano, human occupation has<br />

developed in these valleys over the centuries,<br />

cultivating the valley floors and terracing the<br />

steep valley slopes to expand production areas.<br />

Apart from the valley topography and gradual<br />

increase in temperature, an important environmental<br />

factor is the drying of the climate<br />

towards the valley floors as a climatic effect of<br />

valley wind circulation (Troll, 1968). About 350<br />

km down from its source, the valley finally<br />

opens into the foothills of the Andes and the<br />

Amazonian lowland forests and savannahs,<br />

where mean annual temperatures are around 23<br />

to 29°C, and annual rainfall is around 1700 to<br />

2000 mm. Due to the strong orographic gradients,<br />

all climate parameters vary in short distances.<br />

For example, rainfall slightly to the<br />

southeast of the Urubamba at San Gabán and<br />

Quince Mil (600 m) reaches 3000 to 6000 mm<br />

per year.<br />

Data on species richness will be reviewed,<br />

and we will examine information on present<br />

impacts affecting the natural and managed<br />

biodiversity and the manner in which the latter<br />

is distributed. Given the region’s rich biodiver-<br />

321


3523_book.fm Page 322 Thursday, September 29, 2005 10:38 AM<br />

322<br />

Altitude (m)<br />

7000<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

0 0 50 100 150<br />

sity and the reported past levels of prosperity<br />

at a time (>500 years BP) when resource use<br />

has been claimed to be more sustainable in the<br />

long term, the question that comes to the fore<br />

is: Why human populations now suffer extreme<br />

poverty and environments undergo rapid degradation?<br />

We examine the temporal dynamics of<br />

various components in this three-dimensional<br />

space and explore possible drivers in view of<br />

human pressures and climate change. Several<br />

questions that arise are: Is loss of biodiversity<br />

through land use change a consequence of poverty?<br />

Is poverty related to a failure to incorporate<br />

traditional biodiversity stewardship into<br />

modern agricultural systems? Do market pressures<br />

tend to decrease the use of traditional<br />

agricultural management (e.g. Swinton and<br />

Quiroz, 2003; Halloy et al., 2004)?<br />

METHODS<br />

We surveyed, collated, and calculated the information<br />

and literature on land use and biodiversity<br />

for the Vilcanota–Urubamba watershed.<br />

Political (and hence, census) boundaries are not<br />

drawn along watershed boundaries, so we<br />

selected 33 representative districts along the<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

200<br />

km from source<br />

250 300 350 400<br />

FIGURE 24.1 Topographic profile of the Vilcanota Valley, lengthwise from SSE to NNW with five cross<br />

sections approximately W–E to show the changing valley configuration. The Vilcanota is represented by the<br />

altitudes of 33 district capitals (dots), some of which are located away from the valley center, hence the higher<br />

points. Four additional points complete the profile: village of Santa Barbara (4000 m), outlet of Sibinacocha<br />

Lake (4850 m), Rititica summit (5250 m), and the summit of Vizcachani (near the source of the Vilcanota<br />

above 6200 m). The five cross sections (full lines) are taken at the level of the capitals (from left to right)<br />

Sicuani, Pisac-Cusco, Ollantaytambo, Machupicchu, and Quellouno.<br />

main axis of the valley. To approach biodiversity<br />

at this regional scale, we use proxies (which<br />

are more or less relevant and debatable, and<br />

provide insights into the system) such as percentages<br />

of land use and rates of change (e.g.<br />

deforestation, cultivated crops, and grazing),<br />

each of which has its own impacts on biodiversity.<br />

Cultivated area of each species of crops<br />

was collated from all districts, a necessary<br />

caveat being that census data are sensitive to<br />

human reporting and data-gathering techniques.<br />

Many smaller crops and crop areas are not<br />

reported, thus biasing the data toward larger<br />

areas and crops. However, this is not unlike the<br />

bias that occurs in any biodiversity study toward<br />

larger, more abundant, and more visible species.<br />

Table 24.1 shows the seven provinces of the<br />

Cusco Department, along with some portions<br />

in the Vilcanota Valley. Further details on the<br />

33 districts are in Appendix I. Cusco Department<br />

has a total area of 71,987 km2,<br />

slightly<br />

larger than the island of Tierra del Fuego. The<br />

area of the 33 districts studied here is 29,337<br />

km2,<br />

or almost half of the department.<br />

Diversity was evaluated as simple species<br />

richness, following the Shannon–Weaver information<br />

index of diversity ( H = ln , where<br />

p<br />

i<br />

p<br />

i


3523_book.fm Page 323 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru<br />

TABLE 24.1<br />

Provinces of the Cusco Department with districts used in this study, together with their<br />

population and area<br />

Population,<br />

Density (inhabitants<br />

Province Capital Projection 2002 Area (km2)<br />

km-2)<br />

Total departments Cusco 1,208,689 71,987 16.8<br />

Acomayo Acomayo 34,652 948.22 36.5<br />

Calca Calca 65,330 4,414 14.8<br />

Canchis Sicuani 107,012 3,999 26.8<br />

Cusco Cusco 319,422 617 517.7<br />

La Convención Quillabamba 194,395 30,062 6.5<br />

Quispicanchi Urcos 89,264 7,565 11.8<br />

Urubamba Urubamba 56,352 1,439 39.1<br />

Source: From the 1993-1994 Census, Instituto Nacional de Estadística e Informática, Peru, (INEI 2003).<br />

pi<br />

= (abundance of species i)/total abundance;<br />

[Shannon and Weaver, 1949]), and as frequency<br />

distributions (Williams, 1964).<br />

We integrate this study with ongoing<br />

research at the regional altitudinal limits of life<br />

in the Lake Sibinacocha area. As part of a global<br />

network to monitor the effects of global change<br />

on biodiversity, we established in 2002 a Global<br />

Research Initiative in Alpine Environments<br />

(GLORIA) site at 5250 m. This follows a standardized<br />

methodology of inventories and temperature<br />

measurements for long-term comparisons<br />

(Pauli et al., 2002) and is logged as a<br />

Global Terrestrial Observation Site (Halloy and<br />

Tupayachi, 2004).<br />

VERTICAL DISTRIBUTION OF<br />

DIVERSITY<br />

Braun et al. (2002) calculated the number of<br />

species of seed plants in an altitudinal profile<br />

of Peru from Brako and Zarucchi (1993) (Figure<br />

24.2). They found that the number of species<br />

in the Andes above 500 m is more than the<br />

total number of Amazonian species in Peru. At<br />

the highest levels, over 250 species of seed<br />

plants are recorded above 4500 m for the whole<br />

of Peru. At the eastern headwaters of the Vilcanota,<br />

at the Rititica GLORIA site, we found<br />

24 vascular plants and 28 nonvascular plants<br />

(bryophytes and lichens) in a 274-m2<br />

sampling<br />

area at 5250 m in midwinter 2002. Higher up,<br />

323<br />

flowering plants were found to 5510 m, right<br />

up to the receding ice cliff edge above Rititica.<br />

Gentry (1993) noted that although 43% of<br />

Peruvian seed plant species are from lowland<br />

Amazonia, 34% grow in lower-<strong>Andean</strong> forests<br />

between 500 and 1500 m, and a remarkable<br />

57% are recorded from <strong>Andean</strong> cloud forests.<br />

The high-<strong>Andean</strong> region above 3500 m contains<br />

approximately 14% of the Peruvian flora.<br />

LAND<br />

USE<br />

IMPACT<br />

Land-based agriculture contributes 25.4% of<br />

the gross domestic product (GDP) and provides<br />

47.5% of employment in the Cusco Department<br />

(MAP, 2003). The proportion of total land area<br />

that is dedicated to cultivation averages 8% for<br />

the whole valley, ranging from less than 1% for<br />

Pitumarca and Checacupe districts (limiting<br />

ecological conditions near the altitudinal limits<br />

of cultivation) to 33% for Quellouno (recent<br />

major increase in export crops, principally coffee).<br />

Grazing affects almost all lands accessible<br />

to stock within the valley. Based on a generous<br />

assumption (with present management prac-<br />

tices) of one stock unit<br />

1<br />

per hectare, and calcu-<br />

lating from all stock censused in the six valley<br />

provinces (Sibille, 1997), we obtain that most<br />

1 Stock unit is equivalent to a 45 kg ewe suckling a lamb<br />

or a 55 kg pregnant ewe. This amounts to around 0.02 stock<br />

units per 1 kg of live weight; 1 stock unit requires 520 kg<br />

of dry matter of feed per year.


3523_book.fm Page 324 Thursday, September 29, 2005 10:38 AM<br />

324<br />

Altitude (m) (max level)<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

0<br />

1 10 100<br />

n species seed plants<br />

provinces carry stock requiring 60% (Calca,<br />

Quispicanchi) to 150% (Canchis) and 190%<br />

(Cusco) of their total land area. Only La Convención<br />

requires a minor 3.5% of its land area<br />

to feed existing stock. Because only a certain<br />

fraction of their total land area is suitable for<br />

natural pastures (e.g. 64% for Canchis, 40% for<br />

Cusco, and less than 24% for the remaining<br />

provinces, INEI in MAP [2003]), the overstocking<br />

becomes even more notorious. These are<br />

indications of unsustainable levels of overgrazing<br />

that exceed the carrying capacity of the<br />

land. Fallow and harvested lands also fulfill a<br />

role in providing feed for grazing stock, but this<br />

is not quantified in censuses.<br />

Although some level of grazing can<br />

enhance biodiversity by reducing competition<br />

(Fowler 2002), intense overgrazing as suggested<br />

by these data leads to depletion of palatable<br />

species, reduction of ground cover, and<br />

erosion (Duncan et al., 2001). Depending on<br />

management, livestock, as do cultivated plants,<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

1000 10000<br />

FIGURE 24.2 Number of seed plants at each altitudinal level in Peru, combined from Braun.<br />

will carry with it a variety of commensal/accompanying<br />

species including their parasites,<br />

as well as transport seed plants that are<br />

abundant near their main grazing areas. A 2001<br />

survey around Lake Sibinacocha found that<br />

rodent diversity increased around llama and<br />

alpaca corrals at an altitude of 4900 m as an<br />

effect of anthropogenic enhancement.<br />

The steep terrain of most of the central valley<br />

implies high erosion risk: 85% of areas cultivated<br />

in the higher areas (310,000 ha) are on<br />

steep to moderately steep slopes. They are susceptible<br />

to erosion but most are not subject to<br />

any soil protection practices at this time (MAP,<br />

2003), unlike ancient mitigation practices of<br />

terracing, irrigation, managing soil organic<br />

matter, etc.<br />

Deforestation for agricultural land and firewood<br />

is claiming large areas of the central valley.<br />

For the center of the Valle Sagrado, Galiano<br />

Sánchez et al. (1995) quote deforestations levels<br />

of 90% of original forests for valley bottom


3523_book.fm Page 325 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru<br />

forests (2700 to 3300 m), 60% for mixed forests<br />

of the slopes (3300 to 3700 m), and 20% of the<br />

Polylepis forests from 3700 to 4800 m. The<br />

Ministerio de Agricultura (MAP, 2003) estimated<br />

that 50% of the best forests of the department<br />

were cut down by 1995, including 15%<br />

of the humid lowland forest, more of which is<br />

being cut at a rate of 20,000 ha per year. Land<br />

use conversion has opened up 630,000 ha in the<br />

22 years from 1972 to 1994, representing an<br />

increase of 29.5%.<br />

Introduced species constitute an insufficiently<br />

evaluated risk in the area. Weeds of temperate<br />

regions are widespread in the middle<br />

reaches of the valley, although many weeds in<br />

turn have their uses (see subsection titled Land<br />

Use Impact). Irreversible changes are being<br />

mediated by exotic species: large areas are<br />

reforested with eucalyptus, bringing considerable<br />

changes to the landscape and ecosystem,<br />

including scenic aspects, soils, erosion, availability<br />

of firewood, and capability of native species<br />

(including animals and medicinal plants)<br />

to survive under their canopy. Other invasive<br />

species that have probably had a major impact<br />

in this area include trout, widely introduced for<br />

subsistence and recreational fishing.<br />

Mining at high altitudes, as well as the<br />

impact of large oil deposits found in lowlands<br />

(Camisea, Sibille, 1997), provide an incentive<br />

and a subsidy to develop roads and infrastructure<br />

that then allow penetration into vast new<br />

areas, in addition to their direct impacts on<br />

devegetation and toxic wastes.<br />

Factors slowing the expansion of land use<br />

impacts include difficult access and legislation.<br />

Although steepness and lack of roads has provided<br />

some protection to more remote parts of<br />

the valley, the only formally protected area in<br />

the Vilcanota Valley is the Santuario Histórico<br />

de Machu Picchu in the Province of Urubamba.<br />

With 32,592 ha, it represents almost 23% of the<br />

area of that province but only 1% of the area<br />

of the 33 districts considered in this study. For<br />

comparison, in its land use capability classification,<br />

INRENA (2000; in MAP 2003) considers<br />

that 66% of Departmental lands should be<br />

classified as protection land, with only 33%<br />

suitable for agriculture (3% arable, 0.4% permanent<br />

crops, 14% suitable for forestry plantations,<br />

and 16% suitable for rangeland manage-<br />

325<br />

ment). Yet in the 1994 census of the 33 districts<br />

of the Vilcanota, arable and permanent crops<br />

alone already cover 8% of the land area, implying<br />

that expansion is unsustainable.<br />

RESOURCE DISTRIBUTION IN<br />

HUMAN POPULATIONS<br />

The distribution of economic resources can<br />

determine the magnitude and type of land use<br />

and its effect on biodiversity. Resource distribution<br />

is explored from the point of view of<br />

land size distribution, distribution of the abundance<br />

of crops, and distribution of wealth<br />

(social indicators of poverty).<br />

CULTIVATED<br />

DIVERSITY<br />

LAND<br />

DISTRIBUTION<br />

AND<br />

The distribution of access to productive land<br />

depends on the distribution of cultivated parcel<br />

sizes. This overlooks the issue of spatial distribution<br />

but is, nevertheless, a large-scale proxy<br />

for overall distribution. Plots around a peasant<br />

community tend to be of relatively small (typically,<br />

much less than 0.5 ha) and even sizes<br />

(e.g. for similar cultural landscapes in Peru and<br />

Bolivia, see Liberman Cruz, 1987; Pietilä and<br />

Jokela, 1988). These areas close to villages produce<br />

the mainstay of daily sustenance and hold<br />

the highest crop and native plant diversity (Zimmerer,<br />

1997; Ramirez, 2002). In the 17 higher<br />

districts (>3000 m, more highly populated) of<br />

the Cusco Department, Peru, 93% of properties<br />

are less than 5 ha, the mean parcel size is 0.37<br />

ha, and the average cultivated area per person<br />

in the overall population is 0.14 ha (INEI,<br />

2003). The distribution of plot sizes controlled<br />

by a single family tends to a classic lognormal<br />

pattern with occasional large outliers, indicating<br />

an imbalance (Halloy et al., 2004). Larger<br />

cultivated areas are developed further from<br />

houses and are hence tied to the availability of<br />

transportation and farm machinery. In the two<br />

lower, more market-oriented districts (~650 m),<br />

only 22% of properties are less than 5 ha, the<br />

mean property size is 1.3 ha, and the cultivated<br />

area per person is 0.74 ha.<br />

Larger plot sizes are driven mainly by largescale<br />

cultivation of commercial crops (e.g. coffee<br />

and cocoa in lowlands; maize, wheat,


3523_book.fm Page 326 Thursday, September 29, 2005 10:38 AM<br />

326<br />

ulluco,<br />

and potato in highlands). Much larger<br />

cultivated sizes in tropical lowlands are an<br />

effect of dynamic colonial expansion into the<br />

lowlands and are contrary to ecological expectations<br />

(i.e. higher potential yields mean that<br />

smaller plots are sufficient for equivalent<br />

yields). Older, more established societies tend<br />

to produce lognormal distributions of the cultivated<br />

areas of crops (e.g. Halloy, 1994; Halloy,<br />

1999), whereas younger colonizing societies<br />

have distributions that depart strongly from the<br />

lognormal. In the Vilcanota, we can see this, in<br />

particular, in the lowering of diversity index ( H)<br />

values in La Convención (below 1.8), despite<br />

high species numbers (60 to 75) (Figure 24.3).<br />

Many central and highland areas, despite species<br />

numbers well below 50, maintain a relatively<br />

high diversity ( H between 1.6 and 2.4),<br />

thanks to a more even species distribution.<br />

However, some highland areas have very low<br />

diversity where crop cultivation becomes ecologically<br />

marginal.<br />

H diversity cultivated plants<br />

2.6<br />

2.4<br />

2.2<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8 0 100 200<br />

km from source<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

WEALTH<br />

DISTRIBUTION<br />

AND NUTRITION<br />

Despite a wealth of biodiversity and productive<br />

land, the 1993 census recorded that 60% of<br />

children were chronically malnourished and<br />

infant mortality was 91.8 per thousand for the<br />

Cusco Department (Table 24.2).<br />

Fecundity (number of children per woman)<br />

typically declines with development. The more<br />

highly developed Cusco Province shows a rating<br />

of 2.8, but poorer and less educated provinces<br />

show much higher values (e.g. Quispicanchi<br />

5.8, Urubamba 5.0; Sibille [1997]. In a<br />

paradox that is repeated around the world, the<br />

areas richest in cultivated plants are the poorest<br />

and most malnourished. However, we note that<br />

this is not a linear relation, as improved quality<br />

of life was found at even higher diversity in<br />

traditionally cultivated areas (Halloy et al.<br />

2004).<br />

300 400<br />

FIGURE 24.3 Shannon–Weaver index of diversity for cultivated plants across 33 districts of the Vilcanota<br />

Valley.


3523_book.fm Page 327 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru<br />

TABLE 24.2<br />

Social indicators vs. cultivated plant diversity in some Cusco Provinces, 1993 census<br />

Area Cusco Department Canchis Province Calca Province Cusco Province<br />

Chronically<br />

malnourished children<br />

(%)<br />

60.0 59.2 65.5 42.0<br />

Infant mortality rate per<br />

1000<br />

91.8 114.2 86.7 47.7<br />

Number of species of<br />

cultivated plants per<br />

1000 inhabitants<br />

3.8 3.6 3.8 0.6<br />

Source: 2003 Censos. INEI Instituto Nacional de Estadística e Informática (Peru). In: http://www.inei.gob.pe.<br />

SPECIES<br />

CULTIVATED<br />

RICHNESS<br />

AND<br />

SPECIES<br />

DISTRIBUTION<br />

OF<br />

A total of 157 categories of cultivated plants<br />

were recorded in the 1993 agricultural census.<br />

Several census categories represent mixed bags<br />

of species in which there may be only one or<br />

several species ( Vergel Hortícola Plátano [vegetable<br />

plots planted with bananas], Vergel Frutí-<br />

cola<br />

[fruit orchards],<br />

Flores<br />

[flowers], etc.;<br />

Table 24.3). Hence, estimates of species richness<br />

based on the census are underestimates.<br />

This species richness is not fixed in time; the<br />

actual varieties and species that are grown are<br />

continuously changing with a rapid turnover<br />

rate (e.g. Halloy, 1999; Ramirez, 2002).<br />

Census data of cultivated crops represents<br />

only a fraction of total cultivated plants. For<br />

example, for the total area above 3500 m, the<br />

INEI 1993 census data records 52 species of<br />

cultivated plants in a total of 6679 ha. However,<br />

in a small area of 686 ha above 3500 m in Calca<br />

Province, Ramirez (2002) recorded 76 species.<br />

In addition, a large number of adventive or<br />

“weedy” species accompany cultivation, and<br />

additional native species “tolerate” and persist<br />

in cultivated areas along road edges, hedges,<br />

gullies, etc. Many such species are also used by<br />

local populations (Rapoport et al. 1998). For<br />

example, Vieyra-Odilon and Vibrans (2001)<br />

report 74 weed species found in maize fields in<br />

Mexico that were useful as forage, potherb,<br />

medicinal, or ornamental plants. In the high<br />

327<br />

Andes of neighboring Bolivia, Hensen (1992)<br />

reports the use of 204 species of plants in the<br />

community of Chorojo, Cochabamba, from<br />

3500 to 3800 m, most with forage and medicinal<br />

uses. Of these, 24 species were used as<br />

food. In every relevé in fallow terrain near La<br />

Paz, de Morales (1988) reports that 6 to 12<br />

weedy species are found. Detailed recordings<br />

of plant use in the Andes are available in a range<br />

of publications (e.g. Brücher, 1989; NRS, 1989;<br />

Zimmerer, 1997).<br />

Sibille (1997) (following INEI, 1986)<br />

quotes 193 plant products (including 142 arable<br />

crops, 37 permanent crops, and 14 grasses) for<br />

the whole of Cusco Department, whereas<br />

Galiano Sánchez et al. (1995) quote 96 useful<br />

species (including this time forestry species)<br />

and 685 vascular plant species in a 50 km<br />

2<br />

area<br />

of the Sacred Valley, ranging from 2715 to 5300<br />

m. They also recorded 40 nonvascular cryptogams.<br />

The present total of 157 cultivated species<br />

in the Vilcanota Valley and 193 for the whole<br />

Cusco Department can be compared to 160 species<br />

claimed to have been commonly used for<br />

food, medicine, and other purposes in precolonial<br />

times for Peru (Tapia and Torre, 2003).<br />

It is of some concern for conservation that<br />

most of the rarest cultivated plants are natives,<br />

whereas many of the common species are<br />

exotic.


AU: What is<br />

this?<br />

3523_book.fm Page 328 Thursday, September 29, 2005 10:38 AM<br />

328<br />

TEMPORAL DYNAMICS<br />

HISTORICAL<br />

PERSPECTIVE<br />

It is interesting to compare the present situation<br />

with that recorded by the Spaniards in the early<br />

1500s. The area that was then the center of the<br />

Inca dominions was praised by chroniclers as<br />

a place where “no one ever went hungry” and<br />

where “purposely made storage areas were<br />

overflowing with vegetables and roots to feed<br />

the people and also herbs” (Peró Sancho quoted<br />

in Murra, 1975).<br />

Indeed, traditional land use management<br />

practices were able to support the livelihoods<br />

of households and communities for several millennia<br />

and were sufficient for the rise of complex<br />

civilizations centuries prior to Spanish<br />

occupation. The ample increase in production<br />

under the Inca empire may have, in part,<br />

depended on their careful environmental husbandry<br />

(including tactics of soil conservation,<br />

water management and irrigation, management<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

TABLE 24.3<br />

Most Commonly Cultivated Plants over 33 Districts according to 1993 Agricultural<br />

Census<br />

Census Name<br />

Species/Variety<br />

English Name Scientific Name Area (ha) Number of Districts<br />

Café or cafeto Coffee<br />

Coffea arabica<br />

25,511 7<br />

Maiz amiláceo Starch maize Zea mays<br />

11,375 33<br />

Papa Potato<br />

Solanum tuberosum<br />

8,132 33<br />

Cacao Cocoa<br />

Theobroma cacao<br />

6,581 7<br />

Achiote<br />

Annatto Bixa orellana<br />

4,462 6<br />

Coca Coca<br />

Erythroxylum coca<br />

3,705 8<br />

Haba Broad bean<br />

Vicia faba<br />

3,282 30<br />

Yuca Cassava<br />

Manihot esculenta<br />

3,003 8<br />

Cebada grano Barley grain Hordeum vulgare<br />

2,428 27<br />

Trigo Wheat<br />

Triticum aestivum<br />

2,029 29<br />

Vergel<br />

Vegetable<br />

Multispecies 1,006 32<br />

hortícola–plátano garden–banana<br />

Maiz amarillo Yellow maize Zea mays<br />

1,916 30<br />

Vergel frutícola Fruit orchard Multispecies 1,435 29<br />

Arveja (alverjón) Green pea<br />

Pisum sativum<br />

529 29<br />

Olluco Ulluco<br />

Ullucus tuberosus<br />

1,466 29<br />

Oca Oca, NZ yam Oxalis tuberosa<br />

100 28<br />

Note:<br />

The two right columns show total area cultivated (first ten are the species with largest cultivated areas) and number<br />

of districts where the crop is recorded (following six are species with high number of districts but lower area).<br />

Source:<br />

2003 Censos.<br />

Instituto Nacional de Estadística e Informática (Peru). In: http://www.inei.gob.pe.<br />

of domesticated plant and animal diversity, and<br />

protection of natural vegetation and fauna)<br />

(Halloy et al., 2004).<br />

It is possible that habitat degradation<br />

induced by ancient hunter–gatherers and pastoral<br />

nomads may have contributed — together<br />

with population increase, extended annual<br />

occupation, rise of social stratification, and the<br />

need to increase production for both social and<br />

livelihood needs — to the development of civilizations<br />

incorporating the conservation measures<br />

in force at the time of arrival of the Spanish<br />

(Kessler, 1998).<br />

Despite such measures, it seems likely that<br />

considerable destruction of the high-altitude<br />

Polylepis<br />

forests took place long before the<br />

arrival of the conquistadores in 1532 (Gade,<br />

1999; Kessler et al., 1998). Before the arrival<br />

of the Spanish, the <strong>Andean</strong> landscape had<br />

already experienced significant levels of transformation<br />

and degradation. Gade estimates that<br />

some 65% of the natural forest had been


3523_book.fm Page 329 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru<br />

depleted before the Spanish arrival, shortly after<br />

which 90% became depleted (Gade 1999). The<br />

Spanish conquest resulted in increasing deforestation<br />

rates as they consumed large amounts<br />

of wood for constructions and the smelting of<br />

ores in mining activities.<br />

Upon arrival, the Spanish implemented<br />

agroforestry measures in an attempt to compensate<br />

for excessive consumption levels. Unfortunately,<br />

such measures did not suffice, and the<br />

landscape became mostly depleted of trees.<br />

After the decimation of the indigenous population<br />

in the 16th century, a majority of the rural<br />

landscape was abandoned. Denevan argues that<br />

much of the natural landscape was able to<br />

recover as a result of the population decline and<br />

may have contributed to the early 19th-century<br />

misconceptions of the “pristine” landscape<br />

(Denevan, 1992). However, the introduction of<br />

nonnative species also became commonplace.<br />

The Spanish experimented early with the nonnative<br />

poplar and capuli trees (Gade 1975), but<br />

the most influential species to be introduced in<br />

the late 19th century with unprecedented fruition<br />

was Eucalyptus globulus.<br />

CLIMATE<br />

CHANGE<br />

Given the context of intense human and environmental<br />

heterogeneity and fluctuations over<br />

time, encountering a signal of climate change<br />

effects is not a simple matter (e.g. see meta<br />

analyses as in Parmesan and Yohe (2003), but<br />

such an approach is still to be realized in Peru).<br />

However, there are some observations pointing<br />

towards vegetation and land use advancing<br />

towards higher altitudes in recent decades.<br />

Toward the middle of the last century, Troll<br />

(1968) observed that in the Central Andes of<br />

Peru and Bolivia, maize could be grown up to<br />

3500 m, whereas tuberiferous plants (potato,<br />

oca,<br />

isaño,<br />

and ulluco)<br />

and introduced wheat<br />

and barley reached their upper limit at 4100 m.<br />

Mitchell (1976), followed by Price (1981), also<br />

placed the altitudinal limit of cropping at 4100<br />

m. Higher up, the grasslands were grazed by<br />

llamas, alpacas, and wild vicuñas. Uninterrupted<br />

plant cover ended at around 4700 m,<br />

where nightly frosts began. The climatic snowline<br />

was indicated at 5300 m.<br />

329<br />

Interpretation of such reports is problematic,<br />

given their nonspecificity in terms of location<br />

or dates, as well as issues of time lag<br />

between observations and publication. However,<br />

these and other authors had extensive<br />

experience in geography, and it is unlikely that<br />

their observations would be far off the mark.<br />

More recently, Tapia and Torre (2003) quote<br />

several crops grown up to 4000 m, two species<br />

grown up to 4100 m (maca and kañiwa), and<br />

one (Papa amarga: Solanum juzepczukii)<br />

grown<br />

up to 4200 m. Potato cultivation today in the<br />

Vilcanota headwaters occasionally reaches<br />

4580 m (Chillca, our observations in 2004).<br />

Recent attempts to cultivate oats and potato<br />

have even been made at 5050 m above Murmurarni,<br />

although these were unsuccessful.<br />

Interestingly, archaeological remains show<br />

that these and higher areas were cultivated in the<br />

more remote past. Archaeological remains of cultivation<br />

higher than today have also been noted<br />

in the Cordillera Blanca (Cardich, 1985). In 1985,<br />

Cardich also observed that since he began making<br />

observations, “the limit of cultivation has been<br />

moving upward and crops are now grown at<br />

higher elevations than during previous decades.<br />

Simultaneously, there has been an accelerated<br />

recession of glaciers in the high cordilleras, as<br />

well as disappearance of snow and consequent<br />

opening of passes connecting the Pacific and<br />

Atlantic slopes.” In summary, 2002 cultivation<br />

levels are higher than in the past decades and<br />

recent centuries, but still not as high as maximum<br />

levels reached at some time in the past, presumably<br />

before the Little Ice Age. The first post–Little<br />

Ice Age settlers in the Sibinacocha area moved<br />

into the valley in 1906 (Pedro Godofredo, personal<br />

communication, 2003). Today, there are a<br />

number of corrals and settlements.<br />

POPULATION<br />

AND LAND<br />

USE<br />

An indication of the growing population impact<br />

is its increasing concentration in urban centers,<br />

from 25% of the department’s population in<br />

1940 to 46.5% in the 1993 census (Figure 24.4).<br />

For the whole Department, infant mortality has<br />

declined from 149 per 1000 births in 1979 to<br />

1980 to 101 in 1990 to 1991 (Sibille, 1997).<br />

Because of political–economic change, there is<br />

also a strong net outmigration, principally


3523_book.fm Page 330 Thursday, September 29, 2005 10:38 AM<br />

330<br />

towards centers offering employment and natural<br />

resources (Lima, Arequipa, and Madre de<br />

Dios). Emigration rates are rapidly increasing.<br />

From 8% of the total Departmental population<br />

in 1961, emigration climbed to 16% in 1972,<br />

18% in 1981, and 21% in 1993 (Sibille, 1997).<br />

Emigration from rural areas contributes to<br />

important land use changes with mixed impacts<br />

on biodiversity: lack of maintenance of terraces<br />

and irrigation leading to erosion, lack of control<br />

of animals leading to grazing and overgrazing,<br />

lack of cultivation leading to weedy successional<br />

phases, then back to vegetation that is<br />

more diverse, etc.<br />

Sibille (1997) also indicates that agricultural<br />

land is decreasing significantly in several areas<br />

due to urban encroachment. Thus, Cusco province<br />

lost 62% of its arable land in the 10 years<br />

from 1985 to 1995, whereas Urubamba lost<br />

25%. At the same time, he reports a loss of some<br />

of the more traditional crops to livestock grazing<br />

and intensification of farming (e.g. irrigated land<br />

has increased 89% in 22 years from 1972).<br />

Many irrigation schemes disregard impacts on<br />

the overall social and natural web of interactions<br />

(Liberman Cruz, 1987), leading to further loss<br />

of arable land and native biodiversity.<br />

The advance of the agricultural frontier is<br />

particularly evident in the lowland areas, where<br />

Standardized to 1 for first value<br />

2.00<br />

1.50<br />

1.00<br />

0.50<br />

0.00 1940<br />

1950 1960<br />

Urban population<br />

Arable land, cusco prov.<br />

Remaining forests<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

it is marked by large-scale deforestation. However,<br />

although less evident, use pressure is<br />

growing in the highlands as well, as manifest<br />

by the increasing altitude at which crops and<br />

livestock are grown and the increasing intensity<br />

and density of cultivation.<br />

THREATENED<br />

SPECIES<br />

Although there is no data specific to the Cusco<br />

Department, Pulido (2001) reports threatened<br />

animal species for Peru have increased from<br />

162 to 222 from 1990 to 1999. Such increases<br />

have been recorded around the world in what<br />

is often more a matter of increased monitoring<br />

and perception than of real change of status in<br />

such short times. Amphibian decline noted<br />

around the world is also being observed in the<br />

Vilcanota region, with local people reporting<br />

the apparent reduction or total disappearance of<br />

three to four species of frogs in areas close to<br />

4000 m. Although a causal relation has yet to<br />

be found, it is of concern that recent sampling<br />

above 4400 m found evidence of deadly chytrid<br />

fungus infections (believed to be implicated in<br />

global decline) in remote populations of the<br />

aquatic Telmatobius marmoratus (DeVries et<br />

al., 2004).<br />

1970 1980 1990<br />

Year<br />

Total population<br />

Irrigated land<br />

Infant mortality<br />

Utilised land<br />

FIGURE 24.4 Combined social and environmental changes in Cusco Department, Peru (data from INEI in<br />

MAP, 2003 and Sibille, 1997).


3523_book.fm Page 331 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru<br />

DISCUSSION AND<br />

CONCLUSION:<br />

MACROECONOMIC DRIVERS<br />

There is considerable understanding of the<br />

small-scale effects on biodiversity of land use<br />

changes (e.g. other chapters in this book).<br />

Although we recognize the rich tapestry of<br />

ecology, farm- and people-scale dynamic processes<br />

that underpin the large scale, there is also<br />

a need to understand the large spatial scale and<br />

drivers. The increasing complexity on larger<br />

scales creates particular research difficulties,<br />

including reliance on secondary information,<br />

the importance of historical information, consistency<br />

of information across scales and across<br />

disciplines, and the translation of methods and<br />

language between disciplines.<br />

The remarkable diversity of environments<br />

in the Vilcanota Valley arising from the combination<br />

of topographic conditions, altitude, and<br />

rainfall, together with the mosaic of natural disturbances<br />

and dynamic human management<br />

strategies, has led to a high-energy system with<br />

high biodiversity and high flows of materials<br />

among its landscape components.<br />

The main threats to biodiversity in the Vilcanota<br />

presently involve land use changes. Mitigating<br />

the effects of those changes on biodiversity<br />

requires identifying and understanding<br />

the drivers of change. This chapter is a contribution<br />

to identify the next level of causal interactions<br />

between biodiversity, land use changes,<br />

and socioeconomic drivers (the macroeconomic<br />

system). There is added value in that patterns<br />

observed in the Vilcanota are comparable and,<br />

hence, can be extrapolated (with careful consideration<br />

of differences) to a large range of<br />

similar valleys along the Central Andes, and<br />

likely provide insights for similar valleys<br />

worldwide.<br />

SUMMARY<br />

We explore the multidimensional environment–biodiversity–human–time<br />

complex of an<br />

important cultural and ecological hub in the<br />

Central Andes of Peru: the Vilcanota–Urubamba<br />

river catchment (Sacred Valley<br />

of the Incas and Cordillera de Vilcanota). The<br />

331<br />

watershed begins at the upper borderline of the<br />

biosphere where glaciers are retreating, vegetation<br />

and local fauna are rising, and humans are<br />

cultivating crops and herding camelids at<br />

increasingly higher altitudes. Maximum altitude<br />

of potato cropping is now reaching 4580<br />

m, whereas the highest vascular plants were<br />

found at 5510 m. The number of cultivated species<br />

and varieties censused is up to 34 above<br />

3600 m. The midaltitude Valle Sagrado has<br />

been occupied for millennia, but is presently<br />

undergoing dramatic changes (deforestation at<br />

60 to 90%, fire, exotic invasions, large eucalyptus<br />

plantations, etc) resulting from socioeconomic<br />

pressures (poverty, malnutrition in 60%<br />

of children, outmigration at >21%, market pressures<br />

leading to monocultures, and intensive<br />

cropping), causing a restructuring of spatially<br />

and temporally integrated land use patterns. The<br />

number of cultivated plants can be up to 49 in<br />

one district around 3500 m, whereas total number<br />

of vascular plants above 3500 m in Peru is<br />

>1800. The lower part of the valley reaches the<br />

Amazonian lowlands, providing a pathway for<br />

intensive exchanges with the higher regions. Up<br />

to 75 species of cultivated plants are censused<br />

in one lower district, with 6500 species of vascular<br />

plants recorded below 500 m. <strong>Andean</strong> valleys<br />

such as the Vilcanota, reaching from glacial<br />

peaks to tropical rainforests, provide unique<br />

opportunities for understanding the complex<br />

interactions between landscape, biodiversity,<br />

human cultures, and land use change at a large<br />

scale. The complex mix of macroeconomics,<br />

culture, and ecology are key causes of land use<br />

change and its effect on biodiversity.<br />

References<br />

Brako, L. and Zarucchi, J.L. (1993). Catalogue of<br />

the Flowering Plants and Gymnosperms of<br />

Peru.<br />

Missouri Botanical Garden, St. Louis<br />

MO, USA.<br />

Braun, G., Mutke, J., Reder, A., and Barthlott, W.<br />

(2002). Biotope patterns, phytodiversity and<br />

forestline in the Andes, based on GIS and<br />

remote sensing data. In Körner, C. and<br />

Spehn, E.M. (Eds.), <strong>Mountain</strong> Biodiversity:<br />

A Global Assessment. Parthenon Publishing,<br />

London, pp. 75–89.


3523_book.fm Page 332 Thursday, September 29, 2005 10:38 AM<br />

332<br />

Brisseau, J. (1981). Le Cuzco dans sa Région: Étude<br />

de l'aire d'influence d'une ville andine.<br />

IFEA/CEGET, Bordeaux.<br />

Brücher, H. (1989). Useful Plants of Neotropical<br />

Origin and their Wild Relatives. Springer-<br />

Verlag, Berlin.<br />

Cardich, A. (1985). The fluctuating upper limits of<br />

cultivation in the Central Andes and their<br />

impact on Peruvian prehistory. In Wendorf,<br />

F. and Close, A.E. (Eds.), Advances in World<br />

Archaeology, Vol. 4. Academic Press,<br />

Orlando, pp. 293–333.<br />

de Morales, C.B. (1988). Manual de Ecología. Instituto<br />

de Ecología, Universidad Mayor de San<br />

Andrés (UMSA), La Paz, Bolivia.<br />

Denevan, W. (1992). The Pristine Myth: The Landscape<br />

of the Americas in 1492. Annals of the<br />

Association of American Geographers, 82:<br />

369–385.<br />

DeVries, T.A., Hoernig, G., Sowell, P., Halloy,<br />

S.R.P., and Seimon, A. (2005). Identification<br />

of Chytridiomycosis in Telmatobius marmoratus<br />

at 4450m in the Cordillera Vilcanota<br />

of Southern Peru. In Lavella, E.O. de la Riva,<br />

I. (Eds.). Studies on the <strong>Andean</strong> Frogs of hte<br />

General Telmatobius and Batrachophrynus.<br />

Monografias de Herpetologia 7, Valencia.<br />

Duncan, R.P., Webster, R.J., and Jensen, C.A. (2001).<br />

Declining plant species richness in the tussock<br />

grasslands of Canterbury and Otago,<br />

South Island, New Zealand. New Zealand<br />

Journal of Ecology, 25: 35–47.<br />

Fowler, N.L. (2002). The joint effects of grazing,<br />

competition, and topographic position on six<br />

savanna grasses. Ecology, 83: 2477–2488.<br />

Gade, D. (1999). Nature and Culture in the Andes.<br />

University of Wisconsin Press, Wisconsin.<br />

Gade, D.W. (1975). Plants, Man and the Land in the<br />

Vilcanota Valley of Peru. Dr. W. Junk, The<br />

Hague.<br />

Galiano Sánchez, W., de Olarte Estrada, J., Tupayachi,<br />

A., and Ardiles Jara, A. (1995). Conservación<br />

de Recursos Fitogenéticos y Análisis<br />

de una Microcuenca Hidrográfica en el Valle<br />

Sagrado: Calca-Urubamba. Consejo de<br />

Investigación de la Universidad Nacional de<br />

San Antonio Abad del Cusco. Informe No 5:<br />

1–51.<br />

Gentry, A.H. (1993). Overview of the Peruvian flora.<br />

In Brako, L. and Zarucchi, J.L. (Eds.), Catalogue<br />

of the Flowering Plants and Gymnosperms<br />

of Peru. Missouri Botanical Garden,<br />

St. Louis MO, USA, pp. 29–40.<br />

Land Use Change and <strong>Mountain</strong> Biodiversity<br />

Halloy, S.R.P. (1994). Long term trends in the relative<br />

abundance of New Zealand agricultural<br />

plants. In Fletcher, D.J. and Manly, B.F.J.<br />

(Eds.), Statistics in Ecology and Environmental<br />

Monitoring, Vol. Otago Conference<br />

Series 2. Otago University, Dunedin, New<br />

Zealand, pp. 125–142.<br />

Halloy, S.R.P. (1999). The dynamic contribution of<br />

new crops to the Agricultural economy: is it<br />

predictable? In Janick, J. (Ed.), Perspectives<br />

on New Crops and New Uses. ASHS Press,<br />

Alexandria, Virginia, USA, pp. 53–59.<br />

Halloy, S.R.P. and Tupayachi, A. (2004). Rititica,<br />

Cordillera Vilcanota (code: PE-SIB-RIT). In<br />

http://www.fao.org/gtos/tems/tsite_show.jsp<br />

?TAB=4&TSITE_ID=3249.<br />

Halloy, S.R.P., Ortega Dueñas, R., Yager, K., and<br />

Seimon, A. (2005). Traditional <strong>Andean</strong> Cultivation<br />

Systems and Implications for Sustainable<br />

Land Use. Acta Horticulturae, in<br />

670: 31–55.<br />

Hensen, I. (1992). La flora de la comunidad de<br />

Chorojo — Su uso, taxonomía científica y<br />

vernacular. Agruco, Serie Técnica, 28: 1–37.<br />

INEI (2003). Censos. Instituto Nacional de Estadística<br />

e Informática (Perú). In<br />

http://www.inei.gob.pe/.<br />

Kessler, M. (1998). Forgotten forests of the high<br />

Andes. Plant Talk, 15: 25–28.<br />

Kessler, M., Bach, K., Helme, N., Beck, S.G., and<br />

Gonzales, J. (1998). Floristic diversity of<br />

<strong>Andean</strong> dry forests in Bolivia — an overview.<br />

In Breckle, S.W., Schweizer, B., and<br />

Arndt, U. (Eds.), Results of Worldwide Ecological<br />

Studies. Proceedings of the 1st Symposium<br />

of the A.F. Schimper-Foundation est.<br />

by H. and E. Walter. Günter Heimbach, Stuttgart,<br />

Hohenheim, Germany, pp. 219–234.<br />

Liberman Cruz, M. (1987). Uso de la tierra en el<br />

altiplano norte de Bolivia como base para la<br />

evaluación del impacto ambiental de un<br />

proyecto de desarrollo rural. Rivista di Agricoltura<br />

Subtropicale e Tropicale, Firenze,<br />

Anno LXXXI: 207–235.<br />

MAP (2003). Portal Agrario, Ministerio de Agricultura<br />

del Perú (MAP). In http://www.portalagrario.gob.pe/polt_cusco22.shtml.<br />

Mitchell, W.P. (1976). Irrigation and community in<br />

the central Peruvian highlands. Am. Anthrop,<br />

78: 25–44.<br />

Murra, J.V. (1975). Formación Económica y Política<br />

del Mundo Andino. Instituto de Estudios<br />

Peruanos, Lima.


3523_book.fm Page 333 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru 333<br />

NRS (ed.) (1989). Lost Crops of the Incas. Littleknown<br />

plants of the Andes with promise for<br />

worldwide cultivation. National Research<br />

Council, National Academy Press, Washington,<br />

D.C., USA.<br />

Parmesan, C. and Yohe, G. (2003). A globally coherent<br />

fingerprint of climate change impacts<br />

across natural systems. Nature, 421: 37–42.<br />

Pauli, H., Gottfried, M., Hohenwallner, D., Reiter,<br />

K., and Grabherr, G. (Eds.) (2002). The<br />

GLORIA Field Manual — Multi-Summit<br />

Approach, 4th version (draft), December<br />

2002 edn. Global Observation Research Initiative<br />

in Alpine Environments — A contribution<br />

to the Global Terrestrial Observing<br />

System (GTOS), Vienna.<br />

Pietilä, L. and Jokela, P. (1988). Cultivation of minor<br />

tuber crops in Peru and Bolivia. Journal of<br />

Agricultural Science in Finland, 60: 87–92.<br />

Price, L.W. (1981). <strong>Mountain</strong>s and Man — A Study<br />

of Process and Environment. University of<br />

California Press, Berkeley, California.<br />

Pulido, V. (2001). El Libro Rojo de la fauna silvestre<br />

del Perú. Normal Legal: DS 013-99-Agricultura,<br />

Lima.<br />

Ramirez, M. (2002). On farm conservation of minor<br />

tubers in Peru: the dynamics of oca (Oxalis<br />

tuberosa) landrace management in a peasant<br />

community. Plant Genetic Resources Newsletter,<br />

132: 1–9.<br />

Rapoport, E.H., Ladio, A., Raffaele, E., Ghermandi,<br />

L., and Sanz, E.H. (1998). Malezas comestibles<br />

— hay yuyos y yuyos … Ciencia Hoy,<br />

9: 30–43.<br />

Shannon, C.E. and Weaver, W. (1949). The Mathematical<br />

Theory of Communication. University<br />

of Illinois Press, Urbana, IL.<br />

Sibille, O. (1997). La Cuenca del Vilcanota y sus<br />

Áreas de Influencia. Coordinación Intercentros<br />

de Investigación, Desarrollo y Educación<br />

(COINCIDE), Cusco, p. 56.<br />

Swinton, S.M. and Quiroz, R. (2003). Is poverty to<br />

blame for soil, pasture and forest degradation<br />

in Peru’s Altiplano? World Development,<br />

31: 1903–1919.<br />

Tapia, M.E. and Torre, ADl (2003). La mujer campesina<br />

y las semillas andinas: Género y el<br />

manejo de los recursos genéticos. In<br />

http://www.fao.org/DOCREP/x0227s/x022<br />

7s00.htm#TopOfPage.<br />

Thompson, L.G., Mosley-Thompson, E., Bolzan,<br />

J.F., and Koci, B.R. (1985). A 1500 year<br />

record of tropical precipitation in ice cores<br />

from the Quelccaya Ice Cap, Peru. Science,<br />

229: 971–973.<br />

Troll, C. (1968). The Cordilleras of the tropical<br />

Americas — aspects of climatic, phytogeographical<br />

and agrarian ecology. In Troll, C.<br />

(Ed.), Geoecology of the <strong>Mountain</strong>ous<br />

Regions of the Tropical Americas, Vol. 9.<br />

Ferd. Dümmlers Verlag, Bonn, pp. 15–56.<br />

Vieyra-Odilon, L. and Vibrans, H. (2001). Weeds as<br />

crops: the value of maize field weeds in the<br />

valley of Toluca, Mexico. Economic Botany,<br />

55: 426–443.<br />

Williams, C.B. (1964). Patterns in the Balance of<br />

Nature and Related Problems in Quantitative<br />

Ecology. Academic Press, London and<br />

New York.<br />

Zimmerer, K. (1997). Changing Fortunes: Biodiversity<br />

and Peasant Livelihood in the Peruvian<br />

Andes. University of California Press, Berkeley.


3523_book.fm Page 334 Thursday, September 29, 2005 10:38 AM<br />

334 Land Use Change and <strong>Mountain</strong> Biodiversity<br />

APPENDIX I<br />

Altitude Population,<br />

Capital Projection Area<br />

Province District Capital (m) 2002 (km2 Density<br />

(inhabitants<br />

) km-2 ) Educationa Housingb ha Percentage Ratio of<br />

cultiv. of Land in Permanent<br />

n cult. Cultivated per Parcels


3523_book.fm Page 335 Thursday, September 29, 2005 10:38 AM<br />

Climatic, Biodiversity & Socio-Economic Changes in Land Use in the Vilcanota Watershed, Peru 335<br />

Canchis Pitumarca Pitumarca 3570 7,692 1,118 6.9 24 56.7 20 674.6 0.09 56.2% 0.0000 2.60 2.96<br />

Urubamba Chinchero Chinchero 3762 10,166 95 107.5 49.2 21.3 34 3138 0.31 64.7% 0.0001 3.34 1.08<br />

Acomayo Mosoc Llacta Mosco Llacta 3802 1,750 44 40.1 43.3 32.6 6 106.5 0.06 64.7% 0.0000 3.43 5.63<br />

Note: Districts of the Cusco Department with parts in the Vilcanota watershed, ranked by altitude of the district capital. Collated and calculated from 1993–1994 census, (INEI, 2003).<br />

a Proportion of total population above 15 year with complete or partial primary schooling.<br />

b Percentage of housing with no services (water, electricity, and sewage).


3523_book.fm Page 336 Thursday, September 29, 2005 10:38 AM


Human Development and Environment in the Andes: A Geoecological Overview<br />

Author(s): Wilhelm Lauer<br />

Source: <strong>Mountain</strong> Research and Development, Vol. 13, No. 2, <strong>Mountain</strong> Geoecology of the<br />

Andes: Resource Management and Sustainable Development (May, 1993), pp. 157-166<br />

Published by: International <strong>Mountain</strong> Society<br />

Stable URL: http://www.jstor.org/stable/3673633<br />

Accessed: 03/05/2010 10:26<br />

Your use of the JSTOR archive indicates your acceptance of JSTOR's Terms and Conditions of Use, available at<br />

http://www.jstor.org/page/info/about/policies/terms.jsp. JSTOR's Terms and Conditions of Use provides, in part, that unless<br />

you have obtained prior permission, you may not download an entire issue of a journal or multiple copies of articles, and you<br />

may use content in the JSTOR archive only for your personal, non-commercial use.<br />

Please contact the publisher regarding any further use of this work. Publisher contact information may be obtained at<br />

http://www.jstor.org/action/showPublisher?publisherCode=intms.<br />

Each copy of any part of a JSTOR transmission must contain the same copyright notice that appears on the screen or printed<br />

page of such transmission.<br />

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of<br />

content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms<br />

of scholarship. For more information about JSTOR, please contact support@jstor.org.<br />

http://www.jstor.org<br />

International <strong>Mountain</strong> Society is collaborating with JSTOR to digitize, preserve and extend access to<br />

<strong>Mountain</strong> Research and Development.


MOUNTAIN RESEARCH AND DEVELOPMENT, VOL. 13, NO. 2, 1993, PP. 157-166<br />

HUMAN DEVELOPMENT AND ENVIRONMENT IN THE ANDES:<br />

A GEOECOLOGICAL OVERVIEW<br />

WILHELM LAUER<br />

Geographisches Institut der Universitdt Bonn<br />

Meckenheimer Allee 166<br />

5300 Bonn 1, Germany<br />

ABSTRACT The natural environment and human development and culture are closely related. The basic physical features of relief,<br />

climate, and soil and the biological resources of flora and fauna were important factors that influenced early human evolution,<br />

expansion, and development. In Latin America, the geoecological features gave rise to horizontal and vertical differentiation. Two<br />

examples are examined.<br />

In Mexico between 1600 B.C. and A.D. 1500 climatic changes resulted in population fluctuations, varying degrees of environmental<br />

damage, and cultivation change from dryland farming to irrigated fields.<br />

The system of subsistence of the Callawaya people in Bolivia is adapted to the environment and each family owns land in all<br />

altitudinal belts between 3,000 and 4,300 m, so that they are almost self-sufficient and risks of food shortage are minimized.<br />

Today, population pressure, on the one hand, and migration to the cities, on the other, have undermined traditional land use.<br />

Socioeconomic change in rural mountain areas is inevitable and a balance should be maintained between preservation of traditional,<br />

though less productive, systems and the modernization and improvement of the standards of living of mountain people.<br />

RESUMt Perspective geologique de l'homme et de l'environnement dans les Andes latines-americaines. Le milieu naturel, le developpment<br />

humain et la culture sont etroitement lies. Le relief, le climat, le sol et les ressources biologiques de la flore et de la faune ont<br />

beaucoup influence l'evolution humaine, l'expansion et le developpement. En Amerique latine, les caracteristiques geoecologiques<br />

ont donne lieu a une differentiation horizontale et verticale. Deux exemples sont examines dans cet article.<br />

Au Mexique, au cours de la periode comprise entre 1.600 ans av. J.-C. et 1.500 ans ap. J.-C., les changements climatiques ont<br />

entraine des fluctuations de la population, diff6rents degr6s d'endommagement de l'environnement et un changement du mode<br />

de culture, de la culture a sec a la culture irriguee.<br />

Le systeme de subsistance des Callawaya en Bolivie est bien adapte a l'environnement et chaque famille de paysans possede des<br />

terres dans toutes les zones altitudinales entre 3.000 et 4.300 m, de sorte qu'ils se suffisent pratiquement a eux-memes et que les<br />

risques de disette sont minimises.<br />

A l'epoque actuelle, la pression demographique et la migration vers les villes ont ebranlI les modes traditionnels d'utilisation<br />

des terres. Le changement socio-economique dans les regions rurales de montagne est inevitable, mais il est important de respecter<br />

a la fois les besoins de preservation des modes de vie traditionnels, bien que moins productifs, et ceux de la modernisation et<br />

de l'amelioration du niveau de vie de la population montagnarde.<br />

ZUSAMMENFASSUNG Mensch und Umwelt in den lateinamerikanischen Anden: Eine geookologische Ubersicht. Natfirliche Umwelt, menschliche<br />

Entwicklung und Kultur sind eng miteinander verflochten. Grundlegende, physikalische Merkmale wie Relief, Klima und Boden<br />

und die biologischen Ressourcen wie Flora und Fauna waren wichtige Faktoren, die die fruhe menschliche Evolution, Ausbreitung<br />

und Entwicklung beeinfluB3ten. In Lateinamerika verursachten die geo-6kologischen Merkmale horizontale und vertikale Differenzierung.<br />

Zwei Beispiele werden untersucht.<br />

Zwischen 1600 BC und 1500 AD ffhrten Klimaanderungen in Mexiko zu Schwankungen in der Bev6olkerungszahl, zu<br />

Umweltschaden von verschiedenen Schweregraden und zu einem Wechsel in der Landbestellung, wobei bisher unbewasserte Felder<br />

jetzt bewassert wurden.<br />

Das Subsistenzsystem des Callawaya Volkes in Bolivien ist der Umwelt angepaBt; jede Familie besitzt in allen H6hengiirteln zwischen<br />

3.000 m und 4.300 m Land, was sie beinah unabhangig macht gegenuber dem Risiko der Nahrungsmittelverknappung.<br />

Heutzutage haben Bev6olkerungsdruck einerseits und Abwanderung in die Stadte andererseits die traditionelle Landnutzung<br />

untergraben. Ein sozio-6konomischer Wandel in den landlichen Berggemeinden ist unabwendbar, und dabei sollte ein Gleichgewicht<br />

zwischen den traditionellen - wenn auch weniger produktiven- Systemen und der Modernisierung angestrebt werden, um den<br />

Lebensstandard der Bergbev6olkerung zu verbessern.<br />

INTRODUCTION<br />

The history of mankind in America is characterized by the American Cordillera from Alaska to Tierra del Fuego;<br />

relatively clear premises: during the late-Pleistocene era within a relatively short time a significant differentiation<br />

early peoples intruded into the continent from the Bering and diversification took place among the hitherto uni-<br />

Strait and continued south along the mountain range of formly structured population; with this came an adapta-<br />

? International <strong>Mountain</strong> Society and United Nations University


158 / MOUNTAIN RESEARCH AND DEVELOPMENT<br />

tion to a mountainous land with dissimilar geological<br />

formations and that included nearly all the climates that<br />

exist on earth.<br />

The natural features of this mountain region and<br />

human development and culture are closely correlated.<br />

Especially in the early days of human expansion basic<br />

In this paper the geoecological characteristics of the<br />

Latin American Andes are discussed by describing the<br />

horizontal and vertical differentiations, and the relation-<br />

ships between the people and environment are explained<br />

in terms of historical stages of development as well as<br />

spatial distinctions.<br />

When mountains are examined as living space (human<br />

habitat) it is apparent that they present both desirable<br />

and undesirable qualities. Because of favorable climates<br />

that allowed plants to be cultivated, centers of early<br />

civilization developed in mountains and here the gene-<br />

pools are concentrated, from where they spread through-<br />

out the world.<br />

Glaciated mountain regions provide water supplies and<br />

mountains of volcanic origin are rich in mineral deposits.<br />

Furthermore, areas with soils derived from volcanic<br />

activity in the intermontane valleys and basins are positive<br />

factors for human habitation. They were beneficial to the<br />

development of early civilizations in Columbia, Peru, and<br />

Bolivia, as well as the Mexican highlands and the young<br />

volcanic mountain ridge between Guatemala and Costa<br />

Rica.<br />

In non-tropical regions mountains may be seen as<br />

living spaces of an inhospitable nature. <strong>Mountain</strong>s may<br />

be barriers to human activities, areas of inhospitable<br />

climate with frequent natural hazards, such as landslides,<br />

avalanches, and solifluction, and regions of severe<br />

cold.<br />

However, in the tropics and some areas of the sub-<br />

tropical mountains, land that is hospitable to human<br />

activity is extended altitudinally due to climatic factors,<br />

and therefore living space is vertically expanded. Outside<br />

the tropics permanent settlements are restricted to lower<br />

levels of the mountains and only a few seasonal dwellings<br />

lie above. In contrast, in the tropical Andes permanent<br />

settlements lie at higher altitudes and seasonal supple-<br />

mentary areas lie on the lower slopes and even in the hot<br />

lowlands.<br />

Climatic conditions are very important factors in the<br />

development of mankind throughout the Andes. The<br />

American Cordillera encompasses all the climatic zones<br />

of the world (Figure 1). The Latin American sector lies<br />

mainly in the tropics. Only the southern end extends to<br />

the climatic belt of the westerlies. Thus, in most of the<br />

region seasons are determined not by temperature but<br />

by rainfall, and altitudinal zonation controls the natural<br />

environment. This, in turn, influences the vegetative cycle<br />

THE GEOECOLOGICAL POTENTIAL<br />

physical factors, such as climate, rock strata, and soil, and<br />

biological resources were important factors of selection.<br />

People, in groups or as individuals, had to adapt to<br />

natural conditions and to utilize the natural resources for<br />

the development of civilization.<br />

of the plants and the course of the agrarian year, and so<br />

the economic rhythm of each local area.<br />

The well-known three-dimensional scheme of temperature<br />

decrease with increasing altitude (temperature<br />

altitudinal belts) and the decrease of rainy seasons from<br />

the equator to deserts divides the <strong>Andean</strong> chain into<br />

humid and arid, as well as into warm and cold, tropics<br />

(Figure 2). In Latin America these thermal zones are<br />

called by the well-known names first used by Alexander<br />

von Humboldt (tierra caliente, tierra templada, tierrafria, and<br />

tierra helada or gelida).<br />

Towards the outer tropics the climatic as well as<br />

vegetation belts come closer together. Whereas the timberline<br />

from the inner tropics towards the marginal<br />

tropics increases slightly from 3,000 up to 4,000 m,<br />

beyond the tropics it decreases continually and relatively<br />

rapidly to low elevations; in Tierra del Fuego the timberline<br />

is only a few hundred meters above sea level (Figure<br />

3), finally reaching sea level toward the southern extremity<br />

of the landmass.<br />

The correlations between altitudinal belts of climate<br />

and vegetation, on the one hand, and the agricultural<br />

belts, on the other, are significant. The relationship<br />

between the land-use systems and the natural potential<br />

mainly depends on the climatic adaptability of the cultivated<br />

plants, so that the agricultural altitudinal belts can<br />

be described by means of climatic parameters and are<br />

characterized by cultivated crops. Therefore, the history<br />

of <strong>Andean</strong> agriculture is at the same time the history of<br />

the adaptation of rural production to the thermal and<br />

hygric altitudinal belts, so that even today economy and<br />

ecology interact closely.<br />

The physical parameters are not the only factors and<br />

the land is also characterized by human activity. The<br />

feasibility of cultivation in different regions plays an<br />

important role in the evolution of the way of life. In the<br />

Andes the development of the Inca civilization is based<br />

on the cultivation of root crops (potatoes, oca, izano, and<br />

mashua) in the altitudinal belt above 3,600 m. Such a belt<br />

does not exist in the northern hemisphere in the marginal<br />

tropics of Mexico. There maize has been the basic food<br />

since settlement although it grows only up to a height of<br />

around 3,000 m. In colonial times the cultivation belt<br />

extended up to the treeline and wheat and barley were<br />

grown. In Peru and Bolivia there is a cereal belt below<br />

the root crop belt where mainly domestic maize, wheat,<br />

and barley as well as legumes are cultivated.


.. ...: . . ... .. . .. . ...<br />

* '.. .. ..:... ar'-.<br />

. ..0.JF I<br />

....lbima tic d ren aiaion<br />

-:<br />

150- :N<br />

50O-<br />

* .fo -:^<br />

_ 10-U humid months.<br />

'- isubaShid modnth<br />

. 4-6.i humid monts<br />

.2-3-Apinid t<br />

a.rhudid M athsl:<br />

-.2.-3h.uhmid ..mo.nths;. :<br />

-. r l. *-mn . : h : s<br />

*" 6. :1 .humid. months :<br />

Sm..<br />

T(C)<br />

.<br />

^% . :.. ,.:. . z.O<br />

C.. .. 3312m<br />

... : .. . . . .. .<br />

: .. . . . . . .. % . . . . . . . .<br />

i<br />

TCC)<br />

T(CC)<br />

20<br />

10 .. . . .. . .. . ::<br />

La Pa.-l Alto. 4058m.<br />

*5.0- N:(mm):: T(:):<br />

100-<br />

: : t:<br />

. .: . x<br />

.. 1<br />

: :: j..: .: :. .. I .<br />

.. . _ ><br />

:::... ...<br />

:: .::: . [:<br />

*:. t. :'<br />

*. o:. ..<br />

... .. .X .<br />

* b ..<br />

* .: . :: / . ::<br />

Subtropics and Mid-La<br />

s a ntarctic rain-fores<br />

bro.. .leaves: a<br />

mixed :. : .._ .s.'. :::<br />

: '_ hard. eavs forest<br />

* moist high grass-land<br />

* thorn- and succulent si<br />

....<br />

:.......... . ..:.. ... . : .. .....<br />

*: :.: ** ^ P .:.:. .............: ...: : :^^^<br />

[..] deserty Puna. .. ::<br />

[* semidesert and desert.<br />

. . : ---- ---- - . . . . . . . . . . :<br />

FIGURE 1. Climatic differentiation and vegetation belts of the Andes (modified from various authors).<br />

W. LAUER / 159<br />

.' ...<br />

.....<br />

* . . . .


160 / MOUNTAIN RESEARCH AND DEVELOPMENT<br />

humid months<br />

12 11 10 9 8 716 5 4 3 2 1<br />

arid months 0 1 2 3 4 5 6 7 8 9 10 11 12<br />

IMES<br />

humid semihumid<br />

I<br />

semiarid<br />

arid<br />

wet tropics I<br />

dry tropics<br />

P>pE P=pE<br />

P annual range of temperature<br />

Evergreen montane rain forest (cloud forest)<br />

V/////// Evergreen rain forest (warm tropics)<br />

Warm<br />

tropics<br />

FIGURE 2. Three-dimensional hygrothermal belts of the tropical Andes.<br />

p,",,,,, Semi- Evergreen montane forest<br />

.\A,\,\, {(ptly. trade wind influence)<br />

rj/'/;"; Semi-Evergreen rain forest<br />

/,5,,///J ( ptly. monsoon forest)<br />

[ j Cold<br />

tropics<br />

PHASES OF DEVELOPMENT IN THE ANDEAN AGRICULTURAL LANDSCAPES<br />

In general, the Indian culture of the Andes was mainly the end of the nineteenth and the beginning of the<br />

influenced by three phases of revolution: the Hispanic twentieth centuries.<br />

Conquista; the Age of Enlightenment which led to the<br />

separation of colonies from the motherland; and the<br />

technological conquest of the 'American Way of Life" at<br />

These external influences had more disadvantages<br />

than advantages for Latin American development. Indian<br />

civilization received "transformation pushes" from the<br />

I<br />

0


W. LAUER / 161<br />

0 Lo<br />

N<br />

20 0? 20 40 60 8o 10? 120o v 16? 18o 200 220 2to 26? 28o 30<br />

I.- Kolumbien - 1- Ecuador -<br />

| Peru Peru/Bolivien- Chile<br />

tr<br />

'-j ; Pramo<br />

I 1 7Ceja de la<br />

^ .' 4 Montaia<br />

tropical<br />

''<br />

upper montane<br />

a IT' cloud forest<br />

| t l t t |tropical<br />

9' t t T montane forest<br />

. tropical<br />

LT i p rain-forest<br />

L<br />

.,* l.<br />

|,/<br />

moist Puna<br />

"{Pajonalcs<br />

~' |'<br />

raingreen mesophytic<br />

, dry shrub<br />

I ;p ;P P| raingreen moist Forst<br />

I ?I and moist savannah<br />

-- -y ? I raingreen dry forest<br />

-? |'/ and dry savannah<br />

desert Puna<br />

* " .: " dry Puna<br />

? ??4t IPolylepis woodland<br />

| ^ * 1 raingreen thorn<br />

'v f vr and succulent shrub<br />

v semidesert<br />

:v v .. ~ succulent scurb<br />

v sm v<br />

FIGURE 3. A cross-section of the vegetation belts along the western slope of the tropical Andes.<br />

European culture when America was discovered. Depend-<br />

ing on the stage of development each indigenous group<br />

was affected differently and unequal relationships of<br />

dependency remain today. The native population has<br />

received no real benefits from the influx of foreigners<br />

and their industrial world, especially in rural areas.<br />

The major social and technical changes that occurred<br />

in Europe and America did not affect the <strong>Andean</strong> states<br />

or lead to comparable innovation there. Neither the<br />

Agrarian Revolution, that began in Europe around<br />

A.D.1 700 and continues even today as the "Green Revolu-<br />

tion," nor the Industrial Revolution, which began in<br />

Europe about A.D. 1780, has had a comparative influence<br />

in South America. The establishment of universal educa-<br />

tion did not take place there and knowledge remained<br />

mainly a privilege of the elite. The advances in medical<br />

technology led to a population explosion followed by a<br />

diminishing standard of living. Also the rapid extension<br />

of the transport and communication systems in the Old<br />

World during the second half of the nineteenth century<br />

were only barely realized in the Andes.<br />

*I '. * outertropic andine<br />

bunch grass formation<br />

desert<br />

[ |.:'5~!. -Loma-vegetation<br />

***. v . . with Tillandsia<br />

Unfortunately, in light of the above, it seems unlikely<br />

that the ecological revolution which has just taken effect<br />

in the industrialized countries will affect Latin American.<br />

One of the basic requirements for environmental aware-<br />

ness is a high standard of living that is the outcome of<br />

effective reforms and control of population growth. In<br />

other words, the balance between demographic factors,<br />

economy, and ecology is disturbed. This imbalance may<br />

be derived from external factors, such as climatic change,<br />

as well as from internal ones. Disruptive factors cause<br />

radical changes that may have either positive or negative<br />

consequences.<br />

A major transition took place when societies moved<br />

from hunting and gathering to a settled way of life. It<br />

is likely that the eradication of the Pleistocene mega-<br />

fauna, for the most part by hunters, led to development<br />

of agrarian societies in the "Old" as well as in the "New"<br />

World. This transition lasted at least 2,000 years and<br />

certainly man came into conflict with nature as well as<br />

into conflict with man.


162 / MOUNTAIN RESEARCH AND DEVELOPMENT<br />

FIGURE 4. Number of the archaeological<br />

sites in different prehistoric<br />

cultural phases in the Tlaxcalaregion<br />

(after Garcia Cook, 1976),<br />

and the morphological processes<br />

caused by human activity in different<br />

regions of study area (modified after<br />

Heine, 1976).<br />

Number of the settlements<br />

-400<br />

-300<br />

-200<br />

-100<br />

C|<br />

Cultural Phases C - A 0z C<br />

S ,<br />

I<br />

(after Garcia Cook) E | 1<br />

B.C. 2000 1500 1000 500 0 500 1000 1500 AD<br />

Slopes $ t<br />

| t I , . _ ....<br />

I<br />

High-<br />

Plains<br />

and<br />

Basins<br />

North r-.- ~<br />

Middle<br />

East J.<br />

0<br />

.-<br />

j.<br />

-<br />

1<br />

*-<br />

U<br />

a I<br />

Glacier advance i,- M<br />

M v<br />

in comparison with present<br />

Archaeological and geoecological studies in Mexico<br />

indicate the interaction of landscape change with the<br />

history of humankind during the prehispanic cultural<br />

phases between 2000 B.C. and A.D. 1500. The results of<br />

these interdisciplinary studies include: that migration and<br />

changes in the location of settlements occurred within<br />

each epoch (Figure 4); that changing climate has had<br />

an influence on historical development within the past<br />

3,500 years; that population has fluctuated throughout<br />

time and with it differing degrees of environmental<br />

damage; a cultivation change from dryland farming to<br />

irrigation led from dispersal to concentration of population<br />

and from a purely agrarian society to one based on<br />

the division of labor. Furthermore, there are indications<br />

that increases in population led to intensification of<br />

production and environmental degradation.<br />

In the study area of Puebla-Tlaxcala during the late<br />

pre-classical era between 800 and 300 B.C. the increase<br />

I<br />

k/ 11.1 /4<br />

6I<br />


0<br />

t)<br />

.c<br />

b<br />

ft<br />

iz<br />

m<br />

5500<br />

5000-<br />

I O0<br />

3<br />

I *500<br />

free solifluction<br />

"'<br />

bound soliifluction<br />

." Puna belt }^<br />

Nivodo Ullo Kh<br />

561<br />

E semiarid mes ophytic cjshrub terrace cultures<br />

N;vodo Akamoni m<br />

o- 5500<br />

semihumid mesophytic cej shrub33<br />

:<br />

E tropical montane forest [j[',', :<br />

[P<br />

[:1<br />

Polylepis trees<br />

thorn and succulent shrub .,<br />

"<br />

iros<br />

"5<br />

ris belts' ^ ,<br />

PTL potential tree line<br />

PFL potential forest line<br />

FL forest line<br />

...f,rr- ...<br />

':- ''',<br />

.<br />

'f .<br />

-'' .^: pastures tor lomas, ''<br />

-5000<br />

-".<br />

i'.' :".-; . ., ~".~ .s' '.': s-:? : ,ee, 2. ''<br />

-4500<br />

alpocat and shehp. .'.:'.' ."<br />

---- ?<br />

.?~ ..`.Z:~:?Z..~.~:`~..~.:4:.`~.~;'~:!;~*.:~~~.!~;i:~: :..x' -A"~~~~~~~~~~~~~~~~~~~~~~~~~~b:'<br />

e<br />

:x=.-:./.~.'''A<br />

.~:;~:.~.`o.... ~.~. ..<br />

~~~~~~~~~~000~~~~~~~~~~~~~~~~~<br />

FIGURE 5. Agro-ecological altitudinal belts of the Charazani region.<br />

The cultivated area was reduced, small villages were<br />

abandoned and, where people concentrated within the<br />

irrigated areas, the slopes were severely eroded.<br />

It can be seen that climatic change certainly influenced<br />

The example of the Callawaya people of Bolivia exem-<br />

plifies the persistence of ancient systems of cultivation<br />

that still operate in several regions of the Andes. The<br />

ecological system and the land-use system are in balance<br />

although the economy is characterized by the common<br />

problem of overpopulation, low crop yields, and lack of<br />

modern technology. This traditional way of life is worth<br />

preserving as a relic of pre-colonial times although the<br />

indigenous people may strive for modernization.<br />

The Callawaya are an ethnic group in the northeastern<br />

Bolivian Cordillera whose way of life and land-use systems<br />

are based primarily on the traditional, colonial agro-<br />

ecological conditions. Families own land and cultivate<br />

terraces in all altitudinal belts between 3,000 and 4,300<br />

m (Figure 5). Hence the communities are nearly self-<br />

sufficient in their food requirements. The wide vertical<br />

extent of the fields reduces the risk of harvest failure and<br />

trade relationships between the three altitudinal belts<br />

provide the Callawaya with all their essential needs-from<br />

the puna belt above, characterized by cattle-breeding,<br />

and from the lower mountain forest belt of the Yungas.<br />

. ' C..'ot:.t<br />

. - : - 6.~


164 / MOUNTAIN RESEARCH AND DEVELOPMENT<br />

FIGURE 6. The subsistence system of<br />

the Callawaya mountain people.<br />

cultivation of.: 'mdnufacture of..:<br />

bitter potatoes chufio and tunta<br />

(3900. -4.300mn, from bitte'r potatoes<br />

' ' ' .; .:. . , nd:'tu rilltia from<br />

::" :-. . .:..:.:ull u co :<br />

'c :':, .<br />

llama, and alpaca, and since the Hispanic Conquista<br />

there are also sheep, goat, pigs, and cattle. During<br />

colonial times cereals were grown below the Indian<br />

agricultural belt (3,500-3,000 m) where, in addition to<br />

the main domestic crop of maize, wheat has become<br />

increasingly important; the rotation of wheat with barley<br />

and legumes (peas and beans) optimizes the use of soil<br />

nutrients so that fallow is no longer necessary. In this<br />

cereal belt, which was initiated during the Hispanic<br />

colonial time, nutritional levels of the local population<br />

improved noticeably.<br />

The Callawaya own land in the three cultivation belts<br />

in order to diversify food supply. Access to lands in more<br />

distant altitudinal belts and down to the Amazonian<br />

lowlands is made possible by special farming arrange-<br />

ments and paid for by division of labor or a proportion<br />

of the harvest. Interchange with stable trade relationships<br />

takes place beyond family limits and even over tribal<br />

boundaries. Caravans transport agricultural products to<br />

markets on the old footpaths down to the tierra caliente<br />

and upwards to the pasture belt, and also in a lateral<br />

direction to the capital of La Paz and to Peru where<br />

complementary exchanges are made (Schoop, 1982)<br />

(Figure 6).<br />

The Callawaya are an example of an indigenous pop-<br />

ulation that has been, and still is, able to use natural<br />

HUMAN ACTIVITY AND THE ANDEAN ECOSYSTEM<br />

dlmas,alpocas, heepeand pigs<br />

.....................:<br />

Isheep pand goats, pigsj<br />

(st ubble :'pa sture).?.:::<br />

The vertical system of the crop and pasture belts as well<br />

as the vertical and horizontal traffic of production,<br />

exchange, and trade secures the lifestyle of the commu-<br />

nities and tribes (the Ayllu). The term "vertical control"<br />

implies maximum use of ecological belts in the economy<br />

of the <strong>Andean</strong> society (Murra, 1975). Brush (1977, 1982)<br />

found three types of use of the different altitudinal belts<br />

by one ethnic group, the Ayllu.<br />

1. The compressed type, where the settlements in the<br />

belt of root crops and cereals are accessible within<br />

a few hours from all ecological belts and are used<br />

directly by the settlement.<br />

2. The archipelago type, where other agricultural belts<br />

can be reached from the central settlement within<br />

a few days.<br />

3. The supplementary type where the essential food<br />

supply of a settlement is supplemented by trading<br />

with producers in other belts. This is the most<br />

frequent type today (Figure 7).<br />

This entire system operates on a network of footpaths<br />

partly initiated by the Inca. The caravans consist mainly<br />

of mules although at higher altitudes llamas are also<br />

used. However, today there are trucks wherever roads are<br />

well constructed.<br />

resources by using the different altitudinal belts in an<br />

adapted system of economy which includes terrace culti-


UmLanded property <<br />

(perhaps Market-place)easonal<br />

vation and well-designed crop rotation. But environ-<br />

mental degradation caused by the indigenous people is<br />

severe, especially at the high altitude up to treeline where<br />

the main settlements lie. In the upper tierrafria destruc-<br />

tion of the natural vegetation is nearly complete. Land<br />

clearing by burning for cultivation, the unregulated<br />

taking of fire wood and commercial timber, and extensive<br />

cattle ranching are responsible for destruction of the<br />

natural vegetation especially in the puna and the<br />

grasslands.<br />

The model (Figure 8) shows human influence on the<br />

natural ecosystem and the partial destruction due to<br />

extensive cultivation and cattle breeding. In many cases,<br />

the timberline has been disturbed and lowered by several<br />

Although the systems of adaptation to the mountain<br />

environment in rural areas are impressive, the con-<br />

sequences are misuse and degradation of the natural<br />

vegetation. Not only population pressure but also aban-<br />

donment due to migration to cities leads to change.<br />

Large areas of farm land are no longer cultivated;<br />

terraces and irrigation networks are going to ruin; con-<br />

sequences of erosion leave irreversible damage due to<br />

ettlements<br />

MODERN TENDENCIES AND DANGERS<br />

W. LAUER / 165<br />

FIGURE 7. Types of vertical accessi-<br />

bility of the agrarian belts of the<br />

central Andes.<br />

hundred meters, in some places to below 3,000 m. In<br />

Ecuador our research placed the potential treeline at<br />

about 4,000 m, but this cannot be determined if the<br />

treeline comprises a closed wood. It may be indicated that<br />

climate played a part because, with the arrival of humans<br />

in the Andes during the postglacial warming, the mean<br />

annual air temperature was about 2?C higher than today.<br />

Therefore treeline would be several hundred meters<br />

higher. The gradual climatic cooling may have hastened<br />

the process of degradation of the upper treeline area in<br />

addition to human influence. Grazing by goats and sheep<br />

introduced from the Old World led to more rapid<br />

destruction in the upper treeline area and a lowering of<br />

the belt of closed forests.<br />

more frequent floods and greater sedimentation in the<br />

foothills.<br />

The future development of the <strong>Andean</strong> region requires<br />

a return to the farming practices of the prehispanic<br />

period that have been largely forgotten. The extension<br />

of a system of terracing is a priority, but modern tools<br />

may be used without incurring damage in some instances.<br />

Old tools may be more suitable but a well-adapted land-


166 / MOUNTAIN RESEARCH AND DEVELOPMENT<br />

FIGURE 8. Succession of ecosystems in the Andes<br />

(modified after Ellenberg, 1979).<br />

use system may sometimes be labor intensive. The cultiva-<br />

tion of the indigenous and well-adapted introduced<br />

species may be improved by modern technology that<br />

includes resistance to disease and pests, reduction of<br />

harvest risk, and a higher intensity of use. Industries<br />

based on native products for the local market and also<br />

for export should be encouraged. The pasture use today<br />

is characterized by overstocking by introduced European<br />

hoofed mammals such as cattle, sheep, and goats which<br />

rapidly destroy vegetation. One solution to this problem<br />

may be the cultivation of fodder crops or stabling and<br />

creation larger markets for animal products.<br />

Improvements in agricultural production should be<br />

judged not only by their economic success but also in<br />

relation to their ecological and social effects. In spite of<br />

the preservation of traditional systems adapted to the<br />

natural environment, optimum socio-cultural conditions<br />

and educational possibilities require control of popula-<br />

tion growth. Some of the present traditional rural systems<br />

in the High Andes adjust the pressure of population by<br />

migration to urban areas, but this relocates the instability<br />

to the cities where new problems arise.<br />

Nevertheless, the question is whether or not it is<br />

possible to help the people make their own decisions<br />

about their future. In recent years a world-wide increase<br />

REFERENCES<br />

Brush, S. B., 1977: <strong>Mountain</strong>, Field, and Family: The Economy and<br />

Human Ecology of an <strong>Andean</strong> Valley. University of Pennsylvania<br />

Press, Philadelphia, PA. 199 pp.<br />

, 1982: The natural and human environment of the<br />

central Andes. <strong>Mountain</strong> Research and Development, 2(1):<br />

19-38.<br />

Ellenberg, H., 1979: Man's influence on tropical mountain<br />

ecosystems in South America. Journal of Ecology, 67:<br />

401-416.<br />

Garcia Cook, A., 1976: El desarollo cultural en el norte del valle<br />

poblano. Departamento de Monumentos Prehistoricos, Serie:<br />

Arqueologia 1 (INAH), Mexico.<br />

*C<br />

-c<br />

.t<br />

I.V<br />

cIL<br />

s<br />

c)<br />

r<br />

b J<br />

o I<br />

E<br />

in urbanization has occurred, great cities have arisen, and<br />

the system of a division of labor has been introduced to<br />

increase the gross national product of a country or<br />

region. In future, the <strong>Andean</strong> mountain region will have<br />

to be incorporated into this system. Newjobs for the large<br />

numbers of rural people must be created. It may be<br />

impossible to halt socioeconomic change in rural areas<br />

without an increase in productivity-as would be possible<br />

if there were a "Green Revolution." Therefore, entry into<br />

the modern age by developing countries of Latin America<br />

will be by means of improvements in infrastructure and<br />

technical "know how." The three most important meas-<br />

ures are control of population growth, increase in the<br />

educational level, and improvement of technology. With<br />

these measures must come the protection of the natural<br />

resources of rural mountain ecosystems.<br />

Our research in the Latin American Andes posed quite<br />

a dilemma: on the one hand, we hope that the rural<br />

population will continue their traditions of working<br />

adequately in a less productive, but labor intensive,<br />

subsistence land-use system that will maintain the balance<br />

of ecology and economy; on the other hand, we cannot<br />

deny mountain people the means for improvement of<br />

their standard of living.<br />

Heine, K., 1976: Blockgletscher und Blockzungen-Generationen<br />

am Nevado de Toluca, Mexiko. Die Erde, H. 4: 330--352.<br />

Murra, J. V., 1975: El control vertical de un maximo de pisos<br />

ecolo6gicos en la economia de las sociedades andinas. In<br />

Formaciones econ6micas y politicas del mundo andino. Instituto de<br />

Estudios Peruanos, Lima, 59-115.<br />

Schoop, W., 1982: Giiteraustausch und regionale Mobilitat im<br />

Kallawaya-Tal (Bolivien). Erdkunde, 36, H. 4, 254-266.<br />

Trautmann, W., 1981: Las transformaciones en elpaisaje cultural de<br />

Tlaxcala durante la epoca colonial. Das Mexico-Projekt der<br />

deutschen Forschungsgemeinschaft, Bd. 17.


Clim. Past, 5, 375–388, 2009<br />

www.clim-past.net/5/375/2009/<br />

© Author(s) 2009. This work is distributed under<br />

the Creative Commons Attribution 3.0 License.<br />

Putting the rise of the Inca Empire within a climatic and land<br />

management context<br />

Climate<br />

of the Past<br />

A. J. Chepstow-Lusty 1,2 , M. R. Frogley 3 , B. S. Bauer 4 , M. J. Leng 5 , K. P. Boessenkool 6 , C. Carcaillet 2,7 , A. A. Ali 2 , and<br />

A. Gioda 8<br />

1 Institut Français d’Etudes Andines (IFEA), Lima, Peru<br />

2 Centre for Bio-Archaeology and Ecology, Université Montpellier 2, Montpellier, France<br />

3 Department of <strong>Geography</strong>, University of Sussex, Brighton, UK<br />

4 Department of Anthropology, The University of Illinois at Chicago, Illinois, USA<br />

5 NERC Isotope Geoscience Laboratory, Nottingham, UK<br />

6 School of Earth, Ocean and Planetary Sciences, University of Cardiff, Cardiff, UK<br />

7 Paleoenvironments and Chronoecology, Institut de Botanique, Montpellier, France<br />

8 Hydrosciences, IRD, Lima, Peru<br />

Received: 30 January 2009 – Published in Clim. Past Discuss.: 4 March 2009<br />

Revised: 7 July 2009 – Accepted: 11 July 2009 – Published: 22 July 2009<br />

Abstract. The rapid expansion of the Inca from the Cuzco<br />

area of highland Peru (ca. AD 1400–1532) produced the<br />

largest empire in the New World. Although this meteoric<br />

growth may in part be due to the adoption of innovative<br />

societal strategies, supported by a large labour force and a<br />

standing army, we argue that it would not have been possible<br />

without increased crop productivity, which was linked<br />

to more favourable climatic conditions. Here we present a<br />

multi-proxy, high-resolution 1200-year lake sediment record<br />

from Marcacocha, located 12 km north of Ollantaytambo, in<br />

the heartland of the Inca Empire. This record reveals a period<br />

of sustained aridity that began from AD 880, followed by increased<br />

warming from AD 1100 that lasted beyond the arrival<br />

of the Spanish in AD 1532. These increasingly warmer<br />

conditions would have allowed the Inca and their immediate<br />

predecessors the opportunity to exploit higher altitudes<br />

(post-AD 1150) by constructing agricultural terraces that employed<br />

glacial-fed irrigation, in combination with deliberate<br />

agroforestry techniques. There may be some important<br />

lessons to be learnt today from these strategies for sustainable<br />

rural development in the Andes in the light of future<br />

climate uncertainty.<br />

Correspondence to: A. Chepstow-Lusty<br />

(a.lusty@aliceadsl.fr)<br />

1 Introduction<br />

Published by Copernicus Publications on behalf of the European Geosciences Union.<br />

Through the use of regional-scale, multidisciplinary studies<br />

it is becoming increasingly evident that many prehistoric<br />

South American cultures were highly adaptable and<br />

able to ensure food security even under sustained periods<br />

of harsh climatic conditions. For example, immense, state<br />

level coastal societies such as the Moche (ca. AD 200–600)<br />

flourished in arid regions by developing sophisticated irrigation<br />

technologies to cope with extremes of water availability<br />

(Bawden, 1996). Highland cultures, on the other hand, not<br />

only had to deal with seasonal water supplies, but also had to<br />

combat large temperature ranges and steep terrain. A variety<br />

of innovative coping-strategies were therefore developed to<br />

maximise land-use and reduce soil erosion. These included<br />

the use of raised field systems by the Tiwanaku (ca. AD 500–<br />

1100) (Kolata 1996; Erickson 1999, 2000) and constructing<br />

major terracing and irrigation systems by the Wari (AD 500–<br />

1000) (Williams, 2006).<br />

The Inca of the south-central Andes were also able to develop<br />

a range of landscape-scale practices across a range of<br />

agro-ecological zones to support the growing numbers of<br />

individuals subsumed under their political control. Whilst<br />

these practices were in some cases innovative, some were<br />

built on already-proven, sustainable techniques developed by<br />

previous societies (including the Wari). After a prolonged<br />

period of cultural development (ca. AD 1100–1400) the Inca<br />

rapidly expanded beyond their heartland region of the Cuzco


376 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

Valley (Bauer, 2004). By the time of European contact in<br />

1532 they had become the largest empire to develop in the<br />

Americas, controlling a region stretching from what is today<br />

northern Ecuador to central Chile and supporting a population<br />

of more than 8 million. While the factors involved<br />

in their sudden collapse under the stress of the Spanish invasion,<br />

including the introduction of new diseases and civil<br />

wars have been well documented (Hemming, 1970), relatively<br />

little scholarly attention has been focused on the rapid<br />

rise of the Inca as a pan-<strong>Andean</strong> power. Although part of<br />

the reason behind this expansion has been attributed to the<br />

development of new economic institutions and revolutionary<br />

strategies of political integration (e.g. Covey, 2006), these social<br />

advances had to be underpinned by the ability to generate<br />

agricultural surplus that could sustain large populations, including<br />

the standing armies necessary for wide-ranging military<br />

campaigns.<br />

Long before the Inca state developed, much of the central<br />

Andes was occupied by the Ayacucho-based Wari state<br />

(ca. AD 500–1000) and the Lake Titicaca-centred Tiwanauku<br />

state (ca. AD 500–1100). It is widely suggested that their<br />

declines were accelerated by worsening environmental conditions<br />

that decreased opportunities for agrarian intensification<br />

(Kolata, 1996; Williams, 2002, 2006; Bauer and<br />

Kellett, 2009). At around AD 1000, many lower elevation<br />

settlements in this region were abandoned and local<br />

populations became concentrated in newly constructed,<br />

defensibly-positioned sites located along high ridges. This<br />

dramatic settlement shift is generally thought to reflect a<br />

change from reliance on lower valley terrace (Wari) and<br />

raised field systems (Tiwanaku) to more diversified agropastoral<br />

economies. These agro-pastoral economies would<br />

have served as an effective risk-reduction subsistence strategy<br />

in the wake of widespread demographic, settlement and<br />

environmental change (Stanish, 2003; Arkush, 2006, 2008).<br />

In the Cuzco region, the decline of the Wari ushered in a<br />

period of regional development with the initiation of the Inca<br />

state that later culminated in a relatively brief yet significant<br />

period of imperial expansionism (Bauer, 2004). Documenting<br />

in detail the environmental backdrop against which the<br />

Inca expanded their influence and power enhances our understanding<br />

of the social, political and economic challenges that<br />

they faced. One of the difficulties in achieving this, however,<br />

is that the earlier and more substantive part of the rise of the<br />

Inca took place over several centuries; a longer-term environmental<br />

perspective is therefore needed to understand the social<br />

and ecological contexts in which this development took<br />

place. In this study, therefore, we provide a detailed synthesis<br />

of a 1200-year palaeoenvironmental dataset derived from<br />

a climatically-sensitive, sedimentary sequence located in the<br />

Inca heartland. New geochemical (C/N and δ 13 C) and floral<br />

(plant macrofossil, macrocharcoal and Myriophyllum pollen)<br />

proxy data, chosen for their ability to reflect both palaeoclimatic<br />

and anthropogenic change, are combined here for<br />

the first time with existing palynological, faunal and sedi-<br />

Table 1. 210 Pb dates for Marcacocha (adapted from Chepstow-<br />

Lusty et al., 2007). Determinations followed Flynn et al. (1968).<br />

Depth (cm) Calendar date<br />

0–2 AD 1991±0.1<br />

10–12 AD 1958±0.1<br />

18–20 AD 1926±0.4<br />

27–29 AD 1918±0.5<br />

39–41 AD 1907±0.4<br />

47–49 AD 1905±2.9<br />

50–51 AD 1845±8.3<br />

mentological records from the sequence, in order to establish<br />

a palaeoenvironmental framework against which human responses<br />

can be assessed.<br />

2 Site selection<br />

The best continuous, multi-proxy palaeoenvironmental<br />

archive from the Cuzco region is the 4200-year record derived<br />

from the lake site of Marcacocha (Chepstow-Lusty et<br />

al., 1996, 1998, 2003). Located within the Patacancha Valley,<br />

some 12 km north of both the major Inca settlement of Ollantaytambo<br />

and the Urubamba River, Marcacocha (13 ◦ 13 ′ S,<br />

72 ◦ 12 ′ W, 3355 m above sea-level [a.s.l.]) is a small (ca.<br />

40 m diameter) circular, in-filled lake set within a larger<br />

basin (Fig. 1). The surrounding slopes constitute an ancient,<br />

anthropogenically-modified landscape, being extensively<br />

covered with Inca and pre-Inca agricultural terraces<br />

(Fig. 2). The in-filled lake site is separated from the Patacancha<br />

River by the promontory of Huchuy Aya Orqo, which<br />

contains stratified archaeological deposits that range from the<br />

Early Horizon period (ca. 800 BC) to Inca times (Kendall<br />

and Chepstow-Lusty, 2006). Today, the flat area adjacent to<br />

the in-filled lake provides space for minor potato cultivation<br />

and is used for grazing cattle, sheep, goats and horses. In<br />

the past, the site was also important for pasturing camelids,<br />

owing to its location along a former trans-<strong>Andean</strong> caravan<br />

route that linked the Amazonian selva in the east to the main<br />

western highland trading centres. Because the Patacancha<br />

River is fed partially by melt-water, it flows throughout the<br />

year, which may explain why the basin always remains wet<br />

despite a markedly seasonal climate. The constancy of the<br />

river is likely to have been a significant factor in ensuring<br />

both the long-term occupation of the site and also the exceptional<br />

preservation of the highly organic sediments and the<br />

faunal/floral remains they contain.<br />

In 1993, an 8.25 m stratigraphically continuous sediment<br />

core that spans the last 4200 years was recovered from the<br />

centre of the basin for high-resolution palaeoecological analysis<br />

(Chepstow-Lusty et al., 1996, 1998, 2003). The chronology<br />

for the sequence was derived from seven 210 Pb dates<br />

Clim. Past, 5, 375–388, 2009 www.clim-past.net/5/375/2009/


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 377<br />

Table 2. Bulk radiocarbon dates from Marcacocha (adapted from Chepstow-Lusty et al., 2007). Calibration was carried out using the<br />

SHCal04 dataset (McCormac et al., 2004) in conjunction with version 5.0 of the CALIB calibration program (Stuiver and Reimer, 1993) and<br />

are cited as years before present (AD 1950).<br />

Depth<br />

(cm)<br />

Laboratory<br />

Reference<br />

101–102 Beta<br />

190 482<br />

14 C age<br />

(yr BP)<br />

Median<br />

calendar date<br />

400±40 AD 1540 1σ<br />

2σ<br />

115–123 Q-2917 620±50 AD 1360 1σ<br />

2σ<br />

210–218 Q-2918 1460±50 AD 630 1σ<br />

2σ<br />

310–318 Q-2919 1805±50 AD 280 1σ<br />

2σ<br />

478–486 Q-2920 2245±50 BC 240 1σ<br />

2σ<br />

610–618 Q-2921 3650±60 BC 1960 1σ<br />

2σ<br />

(Table 1) and six radiocarbon dates (Table 2). Confirmation<br />

of the sensitivity of the site to palaeoclimatic change comes<br />

from sub-centennial pollen analysis of the sequence, which<br />

documents the lake-level response to regional precipitation<br />

changes and records a series of aridity episodes broadly corresponding<br />

with regional cultural transitions indicated by<br />

an independently-derived archaeological chronology (Fig. 3)<br />

(Chepstow-Lusty et al., 2003; Bauer, 2004). Other previous<br />

work carried out on the sequence includes an assessment<br />

of basic sedimentological characteristics (such as carbonate,<br />

organic and non-organic content) and, from the top 1.9 m, an<br />

analysis of oribatid mite remains, thought to be an indirect<br />

indicator of large domestic herbivore densities in the catchment<br />

over the past ca. 1200 years (Chepstow-Lusty et al.,<br />

2007).<br />

2.1 New proxies<br />

In order to help understand potential links between past environmental<br />

and cultural change in the region, the existing<br />

palaeoenvironmental datasets from Marcacocha were assessed<br />

alongside new, complementary proxy analyses.<br />

Calibrated date<br />

AD 1458–AD 1510 (54%)<br />

AD 1554–AD 1556 (1%)<br />

AD 1574–AD 1621 (45%)<br />

AD 1454–AD 1626 (100%)<br />

AD 1315–AD 1357 (52%)<br />

AD 1381–AD 1419 (48%)<br />

AD 1295–AD 1437 (100%)<br />

AD 568–AD 678 (100%)<br />

AD 440–AD 485 (3%)<br />

AD 537–AD 775 (97%)<br />

AD 138–AD 199 (20%)<br />

AD 206–AD 403 (80%)<br />

AD 30–AD 37 (1%)<br />

AD 51–AD 543 (99%)<br />

BC 384–BC 160 (95%)<br />

BC 133–BC 117 (5%)<br />

BC 479–BC 470 (


378 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

E C U A D O R<br />

P A C I F I C O C E A N<br />

3380<br />

terraced<br />

field<br />

to<br />

Machu Picchu<br />

C O L O M B I A<br />

P E R U<br />

Lima<br />

contours<br />

(heights in metres)<br />

0 metres 50<br />

terraced field<br />

with archaeological<br />

excavation<br />

Cuzco<br />

3390<br />

3375<br />

3370<br />

3385<br />

HUCHUY AYA ORQO<br />

Patacancha River<br />

Ollantaytambo<br />

B R A Z I L<br />

Lake Titicaca<br />

Fig. 1. Location maps.<br />

3365<br />

3380<br />

3360<br />

B O L I V I A<br />

3390<br />

3385<br />

3380<br />

3375<br />

3355<br />

Marcacocha<br />

3375<br />

Urubamba<br />

grazing, potatoes<br />

and area of<br />

sedge growth<br />

72˚W<br />

CUZCO<br />

3395<br />

3390<br />

3355<br />

3385<br />

72˚W<br />

3380<br />

3375<br />

‘infilled’<br />

Marcacocha<br />

Calca<br />

pollen<br />

core<br />

3360<br />

0 km 20<br />

3365<br />

Pisac<br />

N<br />

Rio Urubamba<br />

either been washed in by surface run-off or blown in from the<br />

immediate catchment.<br />

The value of palynology in helping to determine environmental<br />

and anthropogenic change across a landscape is well<br />

documented. Pollen from the entire Marcacocha sequence<br />

has been analysed previously (Chepstow-Lusty et al., 1996,<br />

1998, 2003), although the Myriophyllum data have not hitherto<br />

been published. Myriophyllum pollen represents a group<br />

of hardy aquatic plants (the watermilfoils) that favour shallow,<br />

nutrient-rich conditions and are often associated with<br />

human impact and disturbance (e.g. Smith and Barko, 1990).<br />

Organic matter is an important constituent of lake sediments,<br />

the primary source of which is the residue of former<br />

biota (mostly plants) located in and around the lake itself<br />

(Meyers and Teranes, 2001). Although the various processes<br />

of sedimentation and diagenesis (especially degradation)<br />

will both alter the geochemical signature and reduce the<br />

concentration of organic matter ultimately preserved in the<br />

lake deposits, analyses can often still distinguish between allochthonous<br />

and autochthonous sources and determine the<br />

nature of that original material (e.g. Hodell and Schelske,<br />

1998). Carbon/nitrogen ratios (C/N) and δ 13 C data obtained<br />

from organic matter are especially useful in providing an insight<br />

into the amount and source of organic material entering<br />

Lucre<br />

3370<br />

3375<br />

3380<br />

14.5˚S<br />

N<br />

3385<br />

Fig. 2. General view north-eastwards up the Patacancha Valley over<br />

the infilled lake of Marcacocha (defined by the circle of Juncaceae<br />

vegetation), surrounded by Inca and pre-Inca terraces. The promontory<br />

on which the archaeological site of Huchuy Aya Orqo is found<br />

lies behind Marcacocha and projects from the eastern side of the<br />

valley towards the Patacancha River, which flows throughout the<br />

year. Notice the concentration of pasture adjacent to Marcacocha<br />

and the proximity of the Inca road


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 379<br />

AD<br />

BC<br />

1500<br />

1000<br />

500<br />

0<br />

500<br />

1000<br />

1500<br />

2000<br />

⊗<br />

⊗<br />

⊗<br />

⊗<br />

⊗<br />

⊗<br />

Inorganics<br />

0 20 40 60 80<br />

%<br />

Cyperaceae<br />

0 2 4 6 8 1012<br />

x10 3<br />

Alnus<br />

0 20 40<br />

x10 2<br />

Chenopodiaceae<br />

0 10 20<br />

Concentration (number of pollen grains per cm 3 )<br />

0 20 40 60 80<br />

x10 2<br />

0 10 20<br />

x10<br />

Post-<br />

Inca<br />

Killke<br />

Qotakalli<br />

F<br />

O<br />

R<br />

M<br />

A<br />

T<br />

I<br />

V<br />

E<br />

Inca<br />

Wari<br />

Cultural stratigraphy<br />

Fig. 3. Inorganic content (percentage) and pollen concentration diagram of selected taxa at Marcacocha plotted against age (adapted from<br />

Chepstow-Lusty et al., 2003). Filled rectangle indicates range of 210 Pb dates (see Table 1); shaded circles indicate calibrated radiocarbon<br />

dates (see Table 2). The independently-derived archaeological chronology for the Cuzco region is also indicated (Bauer, 2004). Shaded<br />

horizontal bands denote periods of aridity as inferred from the Cyperaceae record (Chepstow-Lusty et al., 2003).<br />

The range of δ 13 C values of phytoplankton is generally indistinguishable<br />

from the organic matter produced by terrestrial<br />

C3 plants (−20 to −30‰). However, these values are usually<br />

markedly different from those obtained from material<br />

derived from both aquatic and terrestrial C4 plants (the latter<br />

including maize, some grasses and sedges), which are generally<br />

around −8 to −13‰ (O’Leary, 1981, 1988). Under<br />

certain circumstances, however (such as a limitation of dissolved<br />

CO2 or conditions of high pH), this apparently clear<br />

distinction between C4 and other plant sources breaks down<br />

and δ 13 C values derived from C3 algal material can become<br />

as high as −9‰, comparable with values obtained from C4<br />

plant material (Meyers and Teranes, 2001). In addition, δ 13 C<br />

values derived from algal organic matter can also be unduly<br />

influenced by environmental factors such as soil erosion, the<br />

use of fertilisers or the presence of livestock sewage, all of<br />

which can potentially alter nutrient (principally nitrate and<br />

phosphate) delivery rates to the lake (Meyers and Teranes,<br />

x10 2<br />

Ambrosia<br />

Maize<br />

Early Middle Late<br />

Archaic<br />

1500<br />

1000<br />

500<br />

0<br />

500<br />

1000<br />

1500<br />

2000<br />

2001). Given these potential uncertainties in interpreting the<br />

δ 13 C signal, therefore, it is important to consider these data<br />

alongside other indicators of organic matter origin, such as<br />

C/N ratios and plant macrofossil evidence.<br />

3 Methods<br />

Three sets of 1 cm 3 volumetric sediment sub-samples were<br />

taken every centimetre throughout the top 1.9 m of the Marcacocha<br />

sequence, providing a ca. 6-year temporal resolution<br />

that spans the last 1200 years (assuming constant sedimentation<br />

rates).<br />

3.1 Plant macrofossils<br />

One set of sediment sub-samples was disaggregated in<br />

deionised water and macrofossils, including seeds and<br />

www.clim-past.net/5/375/2009/ Clim. Past, 5, 375–388, 2009<br />

AD<br />

BC


380 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

Age (years AD)<br />

1900<br />

1800<br />

1700<br />

1600<br />

1500<br />

1400<br />

1300<br />

1200<br />

1100<br />

1000<br />

900<br />

800<br />

Marcacocha<br />

u<br />

u<br />

⊗<br />

⊗<br />

Cyperaceae<br />

(×103 cm-3 0 2 4 6 8 10 12<br />

)<br />

M<br />

M<br />

M<br />

M<br />

M<br />

M<br />

Chenopodiaceae<br />

(×10 cm-3 0 10 20 30 40<br />

)<br />

Alnus (×102 cm-3 )<br />

0 20 40<br />

Myriophyllum<br />

(×10 cm-3 0 20 40 60<br />

)<br />

Macrocharcoal<br />

(×10 cm-3 0 10<br />

)<br />

C/N<br />

10 20 30<br />

TOC%<br />

0 20 40<br />

0 20 40 60 80<br />

Juncaceae seeds<br />

(cm -3 )<br />

δ<br />

-28<br />

13 C (‰)<br />

Charophyte<br />

gyrogonites<br />

(×10 cm-3 0 10 20<br />

)<br />

-26<br />

0 1 2 3<br />

0 20 40 60 80<br />

Mites (cm -3 )<br />

Fig. 4. Selected proxies and pollen taxa plotted against depth. Filled diamonds indicate 210 Pb dates (see Table 1); shaded circles indicate<br />

calibrated radiocarbon dates (see Table 2). ‘M’ denotes occurrence of maize pollen; horizontal dashed lines delineate zones discussed in text.<br />

Absence of data in certain proxies between ca. AD 1320 and 1400 reflects where material sampled for radiocarbon dating is now missing<br />

(between 115 and 123 cm depth).<br />

charophyte gyrogonites, were then hand-picked under a lowpower<br />

dissecting microscope (Fig. 4).<br />

3.2 Charcoal analysis<br />

A second set of sediment sub-samples was oxidized with<br />

30% hydrogen peroxide in test tubes placed in a hot water<br />

bath at 60 ◦ C for 2–4 h. The residue was then sieved and<br />

charcoal particles >125 µm counted using a low-power dissecting<br />

microscope. For analytical purposes, the raw charcoal<br />

accumulation rate data (C) were first interpolated to 7year<br />

time-steps, a value that corresponds approximately to<br />

the median temporal resolution of the entire charcoal record<br />

(Higuera et al., 2007, 2008) (Fig. 5a). The resulting interpolated<br />

data (Ci) were then separated into background (Cb)<br />

and peak (Cp) components (Fig. 5b). Low-frequency background<br />

variations in a charcoal record (Cb) represent changes<br />

in charcoal production, sedimentation, mixing and sampling,<br />

and are removed to obtain a residual series of “peak” events<br />

(Cp), i.e. Cp = Ci−Cb (Fig. 5c). It is assumed that Cp itself<br />

comprises two sub-components: Cnoise, representing variability<br />

in sediment mixing, sampling and analytical and naturally<br />

occurring noise; and Cfire, representing charcoal input<br />

from local fire events (likely to be related to the occurrence<br />

of one or more local fires occurring within ca. 1 km of the<br />

site; Higuera et al., 2007). For each sample, a Gaussian mixture<br />

model was used to identify the Cnoise distribution, the<br />

Post-<br />

Inca<br />

99th percentile of which was then used to define a threshold<br />

separating samples into “fire” and “non-fire” events (Higuera<br />

et al., 2008). The Cb component was estimated by means<br />

of a locally-weighted regression using a 500-yr window; in<br />

all cases, the window width used was that which maximized<br />

the signal-to-noise index and the goodness-of-fit between<br />

the empirical and modelled Cp distributions. All numerical<br />

treatments were carried out using the CharAnalysis program<br />

(© Philip Higuera). Fire history was described by quantifying<br />

the variation of fire return intervals (FRI, years between<br />

two consecutive fire events) over time, and smoothed using a<br />

locally-weighted regression with a 1000-yr window (Fig. 5d<br />

and e).<br />

3.3 Organic matter geochemistry<br />

13 C/ 12 C ratios were measured by combustion of the final set<br />

of sub-samples in a Carlo Erba 1500 elemental analyser online<br />

to a VG TripleTrap and Optima dual-inlet mass spectrometer,<br />

with δ 13 C values calculated to the VPDB scale<br />

using a within-run laboratory standard (BROC1) calibrated<br />

against NBS-19 and NBS-22. Replicate analysis of wellmixed<br />

samples indicated an absolute precision of


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 381<br />

a.<br />

CHAR<br />

(pieces cm -2 yr -1 )<br />

b.<br />

CHAR<br />

(pieces cm -2 yr -1 )<br />

c.<br />

Peak magnitude<br />

(pieces cm -2 yr -1 )<br />

d.<br />

FRI<br />

(yr fire -1 )<br />

e.<br />

Fire density (x20)<br />

30<br />

20<br />

10<br />

0<br />

20<br />

10<br />

0<br />

600<br />

400<br />

200<br />

0<br />

300<br />

200<br />

100<br />

0<br />

0.3<br />

0.2<br />

0.1<br />

1000<br />

1200<br />

1400<br />

1600<br />

Age (years AD)<br />

mean FRI = 82 yr [41-135]<br />

N FRI = 13<br />

1800<br />

0<br />

0 200 400 600 800<br />

FRI (yr fire -1 )<br />

0<br />

2000<br />

Fire events<br />

Fig. 5. Fire event reconstruction based on the raw charcoal accumulation<br />

rate (CHAR); see text for details. (a) represents the<br />

raw CHAR data interpolated to constant time intervals (Ci) using<br />

a 7-year running mean (black columns), with the background signal<br />

(Cb) also shown (grey line); (b) de-trended CHAR data (red lines<br />

denote threshold window, see text); (c) residual peaks and magnitude<br />

of detected fire events (+); (d) fire return interval (black line),<br />

inferred from time since last34 fire (grey squares), and simulated fire<br />

frequency (red line); (e) distribution of fire density.<br />

15<br />

10<br />

5<br />

Fire frequency<br />

(fires 1000-yr -1 )<br />

levels of inorganic sources from erosion and nutrients arising<br />

from livestock entering the lake. Replicate analysis of wellmixed<br />

samples indicated an absolute precision of


382 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

of the shallow water plant taxon Myriophyllum remain initially<br />

low, before undergoing expansion in the later part of<br />

this interval. C/N ratios fall abruptly from their initial peak<br />

of ca. 35, to return to values oscillating around 20 for the<br />

rest of the zone. δ 13 C increases notably in a step-wise fashion<br />

at the beginning of the interval to reach pre-AD 880 levels,<br />

but then decreases once again (to ca. −28‰) just before<br />

the end of the zone, when values once again recover to ca.<br />

−27‰. Macrocharcoal and charophyte gyrogonite levels remain<br />

suppressed for most of the interval, whereas oribatid<br />

mite abundances show a gradual increase.<br />

4.3 AD 1100–1400<br />

High concentrations of Alnus pollen are noted for the first<br />

time from the very beginning of this interval. In addition,<br />

Juncaceae seeds become prominent, reflecting the importance<br />

of this family as a major component of the lakeside<br />

vegetation. Synchronously, charophyte oospores and<br />

macrocharcoal concentrations also increase notably, the latter<br />

providing evidence for three significant fire events during<br />

this interval (Fig. 5c). C/N ratios oscillate at values >20,<br />

whereas δ 13 C experiences a progressively positive trend, increasing<br />

by around 1‰ from initial values of ca. −27‰.<br />

There is a distinct but short-lived decline in %TOC (of<br />

around 30%) at ca. AD 1170 that is also mirrored by a similar<br />

event in the %TN record.<br />

4.4 AD 1400–1540<br />

This interval is notable for the continued rise in Alnus pollen<br />

concentrations to the highest values in the sequence (peaking<br />

ca. AD 1450). This period also contains a significant<br />

decline in δ 13 C which, despite a variable signal, reaches a<br />

value of almost −28‰ before recovering at the very end of<br />

the zone. C/N ratios also decline over this interval, but much<br />

more gradually. At the end of the zone, C/N ratios also undergo<br />

a rapid increase from around 15 to almost 28, (seemingly<br />

leading the increase in δ 13 C slightly), before dropping<br />

back to ca. 20. This pattern is almost exactly the reverse<br />

of the %TN signal. Oribatid mite concentrations undergo a<br />

marked rise in the second half of the interval, only then to<br />

decline suddenly. Macrocharcoal levels are, once again, relatively<br />

suppressed over this period.<br />

4.5 Post AD 1540<br />

Concentrations of Alnus pollen generally decline throughout<br />

this zone, which is generally the converse of Cyperaceae<br />

pollen concentrations, which increase notably from<br />

around AD 1800. Although Juncaceae seed and macrocharcoal<br />

concentrations remain relatively steady throughout, they<br />

both experience marked, short-lived peaks around AD 1780<br />

(Juncaceae seed concentrations also briefly peaking around<br />

AD 1560). C/N ratios almost never dip below 15 throughout<br />

the entire zone, and they also experience a marked peak at<br />

ca. AD 1780. Interestingly, concentrations of charophytes<br />

and oribatid mite remains appear to co-vary inversely, with<br />

the principal mite peak (and minimum charophyte values)<br />

occurring ca. AD 1750. This interval is also characterised<br />

by the highest fire frequency of the entire record, with eight<br />

major fire events being identified (Fig. 5c).<br />

5 Interpretation and discussion<br />

5.1 Prior to AD 880<br />

During much of the first millennium AD, the pollen record<br />

indicates little evidence for sustained agriculture at Marcacocha<br />

(see also Fig. 3). Indeed, Chenopodiaceae pollen,<br />

which probably includes a number of high altitude-favouring<br />

cultivars such as Chenopodium quinoa, is rare (Chepstow-<br />

Lusty et al., 2003), suggesting particularly suppressed, cool<br />

temperatures over this period. Furthermore, the sediments in<br />

this interval are characterised by a lower %TOC than subsequent<br />

periods (colder climate, lower organic productivity),<br />

suggesting greater erosion in the immediate catchment,<br />

bringing in more inorganic material. Although human population<br />

levels would have been meagre, some anthropogenic<br />

disturbance is possible (as suggested by the minor peaks in<br />

charcoal at the top of this interval). More importantly, however,<br />

since plant growth would have been slow and the vegetation<br />

sparse, any natural or human agents would have enhanced<br />

erosion as the topsoil would have been less stable.<br />

An examination of the radiocarbon dates shows that the sedimentation<br />

rate is lower during this interval than the earlier<br />

part of the first millennium AD, so it is not necessarily as<br />

simple as equating higher erosion and greater inorganic input<br />

with a higher sedimentation rate. Values of C/N >15 indicate<br />

that erosion was also bringing in terrestrial organic matter,<br />

which, alongside δ 13 C values of around −27‰, shows the<br />

dominance of C3 plants.<br />

Independent evidence from ice core, archaeological, archaeobotanical<br />

and geomorphological data indicate that<br />

much of this period was characterized by relatively low temperatures,<br />

a factor that would have encouraged concentration<br />

of human populations at lower altitudes (Johannessen<br />

and Hastorf, 1990; Seltzer and Hastorf, 1990; Thompson et<br />

al., 1995; Kendall and Chepstow-Lusty, 2006). A period of<br />

burning close to the upper limit of this interval suggests that<br />

there may latterly have been limited anthropogenic activity<br />

in the basin (Figs. 4 and 5c).<br />

5.2 AD 880–1100<br />

An increase in Cyperaceae pollen concentrations at the beginning<br />

of this period probably reflects sedge colonization<br />

at the lake margin as the lake-level reduced in response to<br />

increasingly arid conditions (Chepstow-Lusty et al., 2003).<br />

This interpretation is supported by a major peak in C/N and a<br />

corresponding minimum in δ 13 C (a combination indicative of<br />

Clim. Past, 5, 375–388, 2009 www.clim-past.net/5/375/2009/


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 383<br />

abundant terrestrial plant material entering the lake from the<br />

catchment) and a short-lived peak in Chenopodiaceae pollen<br />

concentrations. This latter event (associated with the presence<br />

of maize pollen) indicates that increasing temperatures<br />

allowed limited agriculture (see Fig. 3 to compare the relative<br />

magnitude of this “peak” with that seen earlier in the record<br />

to emphasise the former importance of quinoa and related<br />

crops cultivated in this basin).<br />

From around AD 1000, however, chenopods and sedges<br />

decline in importance at the expense of the shallow-water<br />

taxon Myriophyllum, which suggests low lake-levels and increasing<br />

levels of nutrients, possibly associated with greater<br />

livestock using the pasture. This interpretation is supported<br />

by several proxies, including the progressive increase of both<br />

TN% and mite frequencies, and a decrease in δ 13 C values<br />

that, coupled with C/N ratios of ca. 20, suggest a significant<br />

C3 terrestrial input to the lake. Macrocharcoal levels<br />

remain suppressed throughout the zone (Figs. 4 and 5a–d),<br />

further implying that the basin was used for mostly pastoral<br />

purposes.<br />

5.3 AD 1100–1400<br />

The marked increase in Alnus pollen at the beginning of<br />

this interval reflects the rapid expansion into the immediate<br />

catchment (probably from lower altitudes) of Alnus acuminata.<br />

This is a tree species closely associated with the colonization<br />

of degraded soils and, subsequently, with agroforestry,<br />

because of its nitrogen-fixing properties, as well<br />

as its ability to grow fast and straight (Chepstow-Lusty and<br />

Winfield, 2000). This is the first sustained appearance of Alnus<br />

in the entire 4200-year record (Fig. 3), suggesting that<br />

the occasional pollen grains recorded prior to this are likely<br />

to represent long-distance transport. The notable increase in<br />

Juncaceae seeds indicate that lake-levels continued to drop,<br />

an interpretation supported by the rise in charophyte gyrogonites,<br />

a sustained increase in δ 13 C of around +1‰ across<br />

the zone (possibly due to an increase in C4 plants in the<br />

catchment) and C/N values generally in excess of 20. Combined,<br />

these data all point to a significant rise in temperature,<br />

though without any concurrent increase in precipitation<br />

(the hydrological requirements of A. acuminata probably being<br />

met year-round by the adjacent, glacially-fed Patacancha<br />

River).<br />

It is likely that a sustained warmer climate enabled human<br />

populations to return to traditional agricultural use of<br />

the Marcacocha Basin and the surrounding landscape, with<br />

pastoralism being pushed up to higher altitudes. Evidence of<br />

this is provided by the recovery of chenopod pollen concentrations,<br />

the presence of (rare) maize pollen and the decline<br />

in mite frequencies. Furthermore, heightened macrocharcoal<br />

values (Fig. 4) and the recognition of three significant fire<br />

events over this period (Fig. 5c) are likely to reflect increased<br />

burning and agricultural clearance. There is also evidence<br />

indicating the instigation of landscape transformation in the<br />

catchment; a brief yet marked reduction in %TOC at around<br />

AD 1170 suggests destabilization of the landscape (possibly<br />

due to terrace building), causing an increase in inorganic<br />

material being deposited in the lake. Only after this construction<br />

phase was complete, would erosion have been reduced<br />

and the landscape stabilized, after which the lake sediments<br />

became highly organic and experienced very little inorganic<br />

input over the next 650 years or so (Donkin, 1979).<br />

5.4 AD 1400–1540<br />

This interval, which coincides with a period of rapid Inca<br />

expansion outside of the Cuzco heartland, appears to have<br />

been relatively stable from a climatic point of view. Temperatures<br />

seemed to have remained warm, with precipitation<br />

(and corresponding lake-levels) being low. Despite this climatic<br />

stability, significant land-use changes occurred in the<br />

basin. The increasing commercial importance of the Patacancha<br />

Valley caravan route is reflected in the sharp rise<br />

in mite abundances from the mid-1400s, suggesting an increased<br />

density of large herbivores (particularly llamas) in<br />

the catchment (Chepstow-Lusty et al., 2007). This interpretation<br />

is supported by an increase in %TN, possibly related<br />

to higher levels of animal excreta originating from the lake<br />

margin. The influx of nutrients to the lake, coupled with<br />

likely disturbance generated by livestock, appears to have<br />

suppressed charophyte levels, which decline to almost zero<br />

by the end of the period. Low values of both C/N (ca. 15)<br />

and δ 13 C (ca. −28‰) are suggestive of increased aquatic<br />

vegetation; when coupled with the reduced charophyte levels<br />

and higher Juncaceae seed concentrations, boggy marginal<br />

conditions may be inferred. Even though there is limited evidence<br />

for maize (a C4 plant) being grown in the basin, this<br />

crop requires good drainage and so was likely to have been<br />

grown away from the immediate lake catchment (as is the<br />

case today); only crops such as potatoes were/are cultivated<br />

in the areas liable to inundation. Furthermore, there is only<br />

muted evidence for sustained burning of the landscape over<br />

this interval (Figs. 4 and 5), suggesting either that there was<br />

only limited crop production in the basin at that time, and/or<br />

that soil fertility (often maintained by periodic firing after<br />

harvests) was instead sustained by the use of animal fertilizer.<br />

It is also noticeable that Alnus pollen concentrations<br />

reach their highest levels in the entire sequence, suggesting<br />

accelerated agroforestry around the basin; this species was<br />

one of the most economic and symbolically important trees<br />

for the Inca (Chepstow-Lusty and Winfield, 2000).<br />

After about AD 1500, proxy signals at the very end of the<br />

zone witness some sharp shifts, including a marked decline<br />

in %TN and mite frequencies, a recovery in δ 13 C values, a<br />

continuing decline in Alnus pollen and a notable peak in Juncaceae<br />

seeds. These changes are likely to be linked to the<br />

collapse of the Inca Empire and the corresponding abandonment<br />

of the basin and cessation of regular llama caravan activity.<br />

www.clim-past.net/5/375/2009/ Clim. Past, 5, 375–388, 2009


384 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

5.5 Post AD 1540<br />

The marked recovery in charophyte gyrogonite concentrations<br />

at the beginning of this interval suggests a continuation<br />

of low lake-levels but a major decline in pasture usage. These<br />

inferences are supported by initially low mite abundances,<br />

declining C/N values and a recovery in δ 13 C (to ca. −26‰).<br />

A different world was imposed with the arrival of the Spanish.<br />

Against a backdrop of disease and falling population<br />

numbers, communities were often forced to migrate and/or<br />

compelled to work under the encomienda system (Chepstow-<br />

Lusty et al., 2007). Much of the previously cultivated landscape<br />

became rapidly overgrown and irrigation canals and<br />

terraces were no longer maintained, falling into neglect and<br />

disuse. A major part of the charcoal signal following the arrival<br />

of the Spanish (Fig. 4) probably represents the continual<br />

effort of the remaining population to clear agricultural land<br />

from rapidly colonizing shrubs and herbs.<br />

Mite abundances undergo one final increase from around<br />

ca. AD 1600, which probably reflects the introduction of<br />

large domesticated herbivores from the Old World, such as<br />

sheep, cattle, goats and horses, into this part of the Patacancha<br />

Valley. This increased use of the basin for pasturing<br />

animals is supported by macrocharcoal concentrations<br />

(Fig. 5a–d), which show sustained and significant burning<br />

events (presumably for clearance purposes) just prior to, and<br />

throughout, the 17th century. A hiatus in burning activity<br />

during much of the following century coincides with the collapse<br />

and subsequent recovery of mite frequencies and, by inference,<br />

livestock populations. Historical documents report<br />

the arrival in about AD 1719–1720 of an epidemic (probably<br />

smallpox) which wiped out nearly all the indigenous people<br />

in the Patacancha Valley (Glave and Remy, 1983) and,<br />

at one point, is known to have killed 600 people in a single<br />

day in Cuzco (Esquivel y Navia, 1980). This event not<br />

only overlapped with the occurrence of four intense El Niño<br />

events (manifesting as strong droughts) between 1715 and<br />

1736 (Carcelén Reluz, 2007), but also with the coldest conditions<br />

of the Little Ice Age (Thompson et al., 1988).<br />

A resumption of landscape burning accompanies the final<br />

decline in mite populations throughout the 19th and 20th centuries<br />

(Figs. 5a–d). It is interesting to note that several independent<br />

regional reconstructions of biomass burning activity<br />

in South America show a decreasing fire frequency for the<br />

past 500 years and attribute this pattern to a range of climatic<br />

controls (e.g. Bush et al., 2008; Marlon et al., 2008; Nevle<br />

and Bird, 2008). The record at Marcacocha, however, shows<br />

the opposite trend (i.e. that fire frequency increases over this<br />

period; Fig. 5d), suggesting that it is local human activity<br />

influencing the fire history of the basin, and not climatic factors.<br />

In the top part of the record, the decline of mites after<br />

ca. AD 1800 not only reflects a major drought (Tandeter,<br />

1991; Gioda and Prieto, 1999), but also the subsequent infilling<br />

of the lake. The uppermost sediments (younger than ca.<br />

AD 1830) consist of peats composed of wetland vegetation,<br />

dominated by Juncaceae; these no longer provide an environmental<br />

record comparable with the lake sediments below.<br />

6 A broader context<br />

The pattern of change witnessed at Marcacocha in the centuries<br />

that pre-dated the imperial expansion of the Inca can<br />

be summarized as: (1) a period of relative aridity developing<br />

from ca. AD 880, characterised by low lake-levels and a<br />

limited arable-based economy in the valley that gave way to<br />

more mixed agro-pastoralism at the beginning of the second<br />

millennium AD; and (2) an interval of elevated temperatures<br />

from ca. AD 1100 that probably lasted for four centuries<br />

or more (the upper limit is difficult to define). This latter,<br />

warmer period is of particular interest, since it witnessed significant<br />

human presence in the basin in terms of agriculture,<br />

early landscape modification and later trading activity. In<br />

order to properly understand the broader-scale cultural and<br />

climatic conditions over this interval, it is necessary to assess<br />

the extent to which events happening in one <strong>Andean</strong> valley<br />

may have reflected what was happening regionally. In part,<br />

this can be achieved by comparison of the Marcacocha record<br />

with other, independently derived datasets.<br />

The Quelccaya ice cap, located some 200 km to the southeast<br />

of Marcacocha at 5670 m asl in southern Peru, arguably<br />

provides the best-resolved climatic archive from the region<br />

(Thompson et al., 1984, 1986, 1988). Two overlapping<br />

cores, drilled though the ice to bedrock, provide a composite,<br />

annually-resolved record of the last 1500 years. While the<br />

use of proxies from this record as a yardstick for central <strong>Andean</strong><br />

climatic variability is not without controversy (in particular,<br />

the complex relationships between δ 18 O variability<br />

and temperature, and ice accumulation and regional precipitation;<br />

Thompson et al., 1988; Grootes et al., 1989; Ortlieb<br />

and Macharé 1993; Gartner, 1996; Thompson et al., 2000;<br />

Calaway 2005), one unambiguous dataset from the Quelccaya<br />

record is that of dust content. At present, dust layers<br />

form annually on the surface of the glacier during the dry<br />

season, when dominantly southerly to north-westerly winds<br />

pick up material from the bare, recently-harvested agricultural<br />

areas of the Altiplano (Thompson et al., 1984, 1988). It<br />

has been postulated that the two most prominent dust events<br />

from the 1500-year Quelccaya record (centring on AD 600<br />

and AD 920) may have their origin in anthropogenic activity,<br />

since they do not correspond with any marked climatic shifts.<br />

The latter of these two prominent dust events lasts around<br />

130 years and is characterized by a steady accumulation from<br />

AD 830 to a peak at AD 920, followed by a more rapid decline<br />

to about AD 960. It has been suggested that this pattern<br />

may be linked with intensified raised field agriculture<br />

around Lake Titicaca (Thompson et al., 1988) and/or may<br />

be representative of a wider spectrum of human activity in<br />

the region, including terrace construction (Erickson, 2000).<br />

Clim. Past, 5, 375–388, 2009 www.clim-past.net/5/375/2009/


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 385<br />

At Marcacocha, this interval is represented by high concentrations<br />

of Cyperaceae pollen, supported by high C/N and<br />

low δ 13 C values, suggesting reduced lake-levels and a significant<br />

proportion of terrestrial plant matter entering the lake.<br />

Although no significant agricultural palynological indicators<br />

were found in the Quelccaya record to support the notion of<br />

anthropological activity within the landscape over this interval<br />

(Thompson et al., 1988), both chenopod (probably incorporating<br />

quinoa) and maize pollen are notable at Marcacocha,<br />

strongly suggesting agricultural use of the basin at this<br />

time.<br />

It is also worth noting that some authorities have used the<br />

Quelccaya ice core data, as well as sediment records from<br />

Lake Titicaca that are indicative of low lake-levels, to infer a<br />

serious drought at ca. AD 1000–1100 (e.g. Thompson et al.,<br />

1988; Abbott et al., 1997; Binford et al., 1997). Furthermore,<br />

it is argued that this protracted period of aridity may (or may<br />

not) have been influential in hastening the demise of both the<br />

Tiwanaku and Wari cultures, centred on the southern shores<br />

of Titicaca and around Ayacucho, respectively (e.g. Binford<br />

et al., 1997; Erickson, 1999, 2000).<br />

From the perspective of Marcacocha, the record there indicates<br />

that, although dry conditions began at least a century<br />

prior to ca. AD 1000 in the Cuzco region, temperatures<br />

remained suppressed at this time. These cooler conditions<br />

would have restricted the ability of cultures such as the Wari<br />

to exploit melt-water sustained higher altitudes for agricultural<br />

purposes. This contrasts markedly with conditions just<br />

a few hundred years later when, faced with similarly arid<br />

conditions, the Inca were able to take advantage of a warmer<br />

climate to exploit new, higher altitude agro-ecological zones<br />

and adapt their irrigation technologies to harness the yearround<br />

melt-water resources.<br />

The rich archaeological record around Cuzco can also help<br />

in understanding the Marcacocha data over this crucial interval.<br />

It is from ca. AD 1200 (although perhaps as early<br />

as ca. AD 1000) that the Inca people in the Cuzco region<br />

are recognized archaeologically by the occurrence of Killke<br />

pottery (Rowe, 1944; Bauer, 2004). While there is a long<br />

history of human occupation in the Marcacocha area, there<br />

is a marked increase in the frequency of archaeological finds<br />

dated to after ca. AD 1000, such as those from the promontory<br />

of Juchuy Aya Orqo, adjacent to the lake (Kendall and<br />

Chepstow-Lusty, 2006; Kosiba 2006). This is corroborated<br />

by the Marcacocha charcoal record, which also suggests a<br />

notable increase in anthropogenic activity in the valley from<br />

ca. AD 1150. These data all point to a dramatic expansion<br />

and/or migratory influx of populations from lower altitudes<br />

over this interval, possibly accompanied by growth in economic<br />

activities. Such local trends reflect more widespread<br />

movements recorded across the <strong>Andean</strong> highlands at this<br />

time. Starting around AD 1000, regional settlement patterns<br />

shifted from valley floor and lower valley slopes to higher<br />

altitudinal locations (Arkush, 2006; Covey, 2008; Bauer and<br />

Kellett, 2009). By around AD 1300, however, the Marca-<br />

cocha area was directly incorporated into the growing Inca<br />

state and, by AD 1400, major imperial institutions had been<br />

established in the nearby town of Ollantaytambo. The incorporation<br />

of the area into the Inca state is reflected by the<br />

high oribatid mite concentrations, which suggest an increasing<br />

intensity of llama caravans utilizing the pasture adjacent<br />

to Marcacocha.<br />

From an even broader perspective, the notion that temperatures<br />

were consistently higher than modern values during<br />

the 9th–14th centuries has received increasing attention in<br />

the Northern Hemisphere (Lamb, 1965; Hughes and Diaz,<br />

1994). The prevailing view of this interval, known commonly<br />

as the “Medieval Warm Period” (MWP), is that elevated<br />

temperatures were often only intermittently experienced<br />

and, in some regions, was apparently characterized<br />

instead by climatic anomalies such as prolonged drought,<br />

increased rainfall or a stronger monsoon system (Hughes<br />

and Diaz, 1994; Stine, 1994; Bradley et al., 2003; Zhang<br />

et al., 2008). However, evidence for the MWP being a<br />

global phenomenon is contentious, especially in the Southern<br />

Hemisphere, where there are few continuous, detailed<br />

palaeoclimatic records spanning this interval (Bradley, 2000;<br />

Broecker, 2001). Nevertheless, from the Marcacocha dataset<br />

we can infer that temperatures increased from ca. AD 1100<br />

(after a period of relative aridity in comparison to much of<br />

the first millennium AD) and that conditions remained warm<br />

and stable for several centuries thereafter.<br />

7 Conclusions<br />

This study highlights some of the environmental and cultural<br />

changes that took place in the Cuzco region during the late<br />

period of prehistory. The scale of anthropological manipulation<br />

and transformation of the landscape in the south-central<br />

Andes appears to have increased after ca. AD 1100, probably<br />

in response to a climatic backdrop that was relatively<br />

warm, dry and essentially stable. The development of major<br />

irrigated terracing technology may have been increasingly<br />

necessary in these regions to obviate conditions of seasonal<br />

water stress, thereby allowing efficient agricultural production<br />

at higher altitudes. The outcome of these strategies was<br />

greater long-term food security and the ability to feed large<br />

populations. Such developments were exploited by the Inca<br />

of the Cuzco Valley, who were emerging as the dominant ethnic<br />

group of the region as early as ca. AD 1200. A healthy<br />

agricultural surplus supported their economic and political<br />

potential, enabling them to subjugate other local independent<br />

states and to effectively centralize power in the Cuzco region<br />

by ca. AD 1400 (Bauer, 2004).<br />

Fully understanding the adaptive capacity-building strategies<br />

of the Inca in the face of demanding climatic conditions<br />

is still at an early stage. However, their success has clear<br />

implications for both rural and urban <strong>Andean</strong> communities<br />

www.clim-past.net/5/375/2009/ Clim. Past, 5, 375–388, 2009


386 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

facing the environmental challenges currently being posed<br />

by global warming.<br />

Acknowledgements. We would like to thank the Palaeoenvironments<br />

Laboratory and the Centre of Bioarchaeology and Ecology<br />

at Montpellier University 2 for supporting this investigation,<br />

and gratefully acknowledge additional financial support from the<br />

CNRS and the French Ministry of Foreign Affairs. The Associación<br />

Ecosistemas Andinos (ECOAN) in Cuzco also provided invaluable<br />

facilities. We also thank K. Bennett for initiating the work at<br />

Marcacocha, A. Cundy (University of Brighton) for carrying out<br />

the 210 Pb dating, A. Kendall and the Cusichaca Trust for providing<br />

logistical support, and A. Tupayachi Herrera, C. A. Chutas and<br />

S. Echegaray for additional field assistance. Two anonymous<br />

reviewers invested much time and effort to help us improve and<br />

widen the scope of this paper, for which we are grateful.<br />

Edited by: D.-D. Rousseau<br />

References<br />

Abbott, M. B., Binford, M. W., Brenner, M., and Kelts, K. R.: A<br />

3500 14 C yr high resolution record of water-level changes in<br />

Lake Titicaca, Quat. Res., 47, 169–180, 1997.<br />

Arkush, E. N.: Collapse, conflict, conquest: the transformation of<br />

warfare in the Late Prehispanic <strong>Andean</strong> highlands, in: The Archaeology<br />

of Warfare: Prehistories of Raiding and Conquest,<br />

edited by: Arkush, E. N. and Allen, M. W., University Press of<br />

Florida, Gainesville, USA, 286–335, 2006.<br />

Arkush, E. N.: War, chronology and causality in the Titicaca basin,<br />

Lat. Am. Antiq., 19, 339–373, 2008.<br />

Bates, C. D., Coxon, P., and Gibbard, P. L.: A new method for<br />

the preparation of clay-rich sediment samples for palynological<br />

investigations, New Phytol., 81, 459–463, 1978.<br />

Bauer, B.: Ancient Cuzco: Heartland of the Inca, University of<br />

Texas Press, Austin, Texas, USA, 255 pp., 2004.<br />

Bauer, B. and Kellett, L.: Cultural transformations of the Chanka<br />

heartland (Andahuaylas, Peru) during the Late Intermediate period<br />

(AD 1000–1400), Lat. Am. Antiq., in press, 2009.<br />

Bawden, G.: The Moche, Blackwell Publishers, Cambridge, 392<br />

pp., 1996.<br />

Berglund, B. E. and Ralska-Jasiewiczowa, M.: Pollen analysis and<br />

pollen diagrams, in: Handbook of Holocene Palaeoecology and<br />

Palaeohydrology, Berglund, B. E. (Ed), John Wiley and Sons,<br />

Chichester, UK, 455–484, 1986.<br />

Binford, M. W., Kolata, A. L., Brenner, M. Janusek, J. W., Seddon,<br />

M. T., Abbott, M. and Curtis, J. H.: Climate variation and the<br />

rise and fall of an <strong>Andean</strong> civilization, Quat. Res., 47, 171–186,<br />

1997.<br />

Birks, H.: Plant macrofossils, in: Tracking Environmental Change<br />

Using Lake Sediments: Volume 3, Smol, J. P., Birks, H. J. B.<br />

and Last, W. M. (Eds.), Kluwer Academic Publishers, Dordrecht,<br />

The Netherlands, 49–74, 2001.<br />

Birks, H. J. B. and Gordon, A. D.: Numerical Methods in Quaternary<br />

Pollen Analysis, Academic Press, San Diego, California,<br />

USA, 317 pp., 1985.<br />

Bradley, R.: Paleoclimate – 1000 years of climatic change, Science,<br />

288, 1353–1354, 2000.<br />

Bradley, R. S., Hughes, M. K., and Diaz, H. F.: Climate in Medieval<br />

time, Science, 302, 404–405, 2003.<br />

Brenner, M., Whitmore, T. J., Curtis, J. H., Hodell, D. A., and<br />

Schelske, C. L.: Stable isotope (δ 13 C and δ 14 N) signatures of<br />

sedimented organic matter as indicators of historic lake trophic<br />

state, J. Paleolimnol., 22, 205–221, 1999.<br />

Broecker, W.: Was the Medieval Warm Period global?, Science,<br />

291, 1497–1499, 2001.<br />

Bush, M. B, Silman, M. R., McMichael, C., and Saatchi, S.: Fire,<br />

climate change and biodiversity in Amazonia: a Late-Holocene<br />

perspective, Philos. T. Roy. Soc. B, 363, 1795–1802, 2008.<br />

Calaway, M. J.: Ice cores, sediments and civilization collapse:<br />

a cautionary tale from Lake Titicaca, Antiquity, 79, 778–790,<br />

2005.<br />

Carcelén Reluz, C. G.:. Idolatría indígena y devoción criolla como<br />

respuestas a la variabilidad climática en Lima y Huarochirí durante<br />

el siglo XVIII, Investigaciones Sociales, 11, 171–186,<br />

2007.<br />

Chepstow-Lusty, A. and Winfield, M.: Agroforestry by the Inca:<br />

lessons from the past, Ambio, 29, 322–328, 2000.<br />

Chepstow-Lusty, A., Bennett, K. D., Fjelds˚a, J., Kendall, A.,<br />

Galiano, W. and Tupayachi Herrera, A.: Tracing 4,000 years of<br />

environmental history in the Cuzco area, Peru, from the pollen<br />

record, Mt. Res. Dev., 18, 159–172, 1998.<br />

Chepstow-Lusty, A., Bennett, K. D., Switsur, V. R., and Kendall,<br />

A.: 4000 years of human impact and vegetation change in<br />

the central Peruvian Andes – with events paralleling the Maya<br />

record?, Antiquity, 70, 823–833, 1996.<br />

Chepstow-Lusty, A., Frogley, M. R., Bauer, B. S., Bush, M. B., and<br />

Tupayachi Herrera, A.: A late Holocene record of arid events<br />

from the Cuzco region, Peru, J. Quat. Sci., 18, 491–502, 2003.<br />

Chepstow-Lusty, A., Frogley, M. R., Bauer, B. S., Leng, M.,<br />

Cundy, A., Boessenkool, K. P., and Gioda, A.: Evaluating socioeconomic<br />

change in the Andes using oribatid mite abundances<br />

as indicators of domestic animal densities, J. Archaeol. Sci., 34,<br />

1178–1186. 2007.<br />

Covey, R. A.: How the Incas Built Their Heartland: <strong>State</strong> Formation<br />

and the Innovation of Imperial Strategies in the Sacred Valley,<br />

Peru, University of Michigan Press, Ann Arbor, USA, 333 pp.,<br />

2006.<br />

Covey, R. A.: Multiregional perspectives on the archaeology of the<br />

Andes during the Late Intermediate period (c. A.D. 1000–1400),<br />

J. Archaeol. Res., 16, 287–338, 2008.<br />

Donkin, R.: Agricultural Terracing in the Aboriginal New World,<br />

Viking Fund Publications in Anthropology 56, Wenner-Gren<br />

Foundation for Anthropological Research, New York, USA, 196<br />

pp., 1979.<br />

Erickson, C.: Neo-environmental determinism and agrarian ‘collapse’<br />

in <strong>Andean</strong> prehistory, Antiquity, 73, 634–642, 1999.<br />

Erickson, C.: The Lake Titicaca basin: a Precolumbian built landscape,<br />

in: Imperfect Balance: Landscape Transformations in the<br />

Precolumbian Americas, edited by: Lentz, D. L., Columbia University<br />

Press, New York, USA, 311–356, 2000.<br />

de Esquivel y Navia, D.: Noticias cronológicas de la Gran Ciudad<br />

del Cuzco [1749]. Edición, prólogo y notas de F. Denegri Luna<br />

con la colaboración de H. Villanueva Urteaga y C. Gutiérrez<br />

Muñoz, Tomos 1 y 2, Fundación Augusto N., Lima, 1980.<br />

Flynn, W. W.: Determination of low levels of polonium-210 in environmental<br />

materials, Anal. Chim. Acta, 43, 221–226, 1968.<br />

Clim. Past, 5, 375–388, 2009 www.clim-past.net/5/375/2009/


A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context 387<br />

Gartner W. G.: Book review of ‘The Tiwanaku: Portrait of an <strong>Andean</strong><br />

Civilization’ by A. L. Kolata, Ann. Ass. Am. Geog., 86,<br />

153–156, 1996.<br />

Gioda, A. and Prieto, M. R.: Histoire des sécheresses andines. La<br />

Météorologie, 8, 33–42, 1999.<br />

Glave, L. M. and Remy, M. I.:. Estructura agraria y vida rural en<br />

una region andina: Ollantay’Tambo entre los siglos XVI y XIX,<br />

Archivos de Historia Andina, 3, Centro Bartolomé de las Casas,<br />

Lima, 584 pp., 1983.<br />

Grootes, P. M., Stuiver, M., Thompson, L. G. and Mosley-<br />

Thompson, E.: Oxygen isotope changes in tropical ice, Quelccaya,<br />

Peru, J. Geophys. Res., 91, 1187–1194, 1989.<br />

Hemming, J.: The Conquest of the Incas, Harcourt Brace Jovanovich,<br />

New York, USA, 641 pp., 1970.<br />

Heusser, L. E. and Stock, C. E.: Preparation techniques for concentrating<br />

pollen from marine sediments and other sediments with<br />

low pollen density, Palynology, 8, 225–227, 1984.<br />

Higuera, P. E., Brubaker, L. B., Anderson, P. M., Brown,<br />

T. A., Kennedy, A. T., and Hu, F. S.: Frequent fire<br />

in ancient shrub tundra: implications of paleorecords for<br />

Arctic environmental change, PLoS ONE, 3, e0001744,<br />

doi:10.1371/journal.pone.0001744, 2008.<br />

Higuera, P. E., Peters, M. E., Brubaker, L. B., and Gavin, D. G.: Understanding<br />

the origin and analysis of sediment-charcoal records<br />

with a simulation model, Quat. Sci. Rev., 26, 1790–1809, 2007.<br />

Hodell, D. A. and Schelske, C. L.: Production, sedimentation and<br />

isotopic composition of organic matter in Lake Ontario, Limnol.<br />

Oceanogr., 43, 200–214, 1998.<br />

Hughes, M. K. and Diaz, H. F.: Was there a Medieval Warm Period,<br />

and if so, where and when?, Clim. Change, 26, 109–142, 1994.<br />

Johannessen, S. and Hastorf, C.: A history of fuel management<br />

(AD 500 to the present) in the Mantaro Valley, Peru, J. Ethnobiology,<br />

10, 61–90, 1990.<br />

Kendall, A. and Chepstow-Lusty, A.: Cultural and environmental<br />

change in the Cuzco region of Peru: the rural development implications<br />

of combined archaeological and palaeoecological evidence,<br />

in: Kay Pacha: Cultivating Earth and Water in the Andes,<br />

edited by: Dransart, P., British Archaeological Reports International<br />

Series, 1478, 185–197, 2006.<br />

Kolata, A. (Ed.): Tiwanaku and its Hinterland: Archaeology and<br />

Paleocology of an <strong>Andean</strong> Civilization, Vol. 1, Agroecology,<br />

Smithsonian Press, Washington DC, USA, 322 pp., 1996.<br />

Kosiba, S. B.: Proyecto Arqueologico de la Subregion de Wat’a<br />

(PASW), Primera Temporada: Reconocimiento Intensivo del<br />

area Huarocondo – Ollantaytambo, Technical Report (Informe),<br />

Instituto Nacional de Cultura (INC), Cusco, 66 pp., 2006.<br />

Krause, W.: Characeen als Bioindikatoren für den<br />

Gewässerzustand, Limnologica, 13, 399–418, 1981.<br />

Kufel, L. and Kufel, I.: Chara beds acting as nutrient sinks in shallow<br />

lakes – a review, Aquat. Bot. 72, 249–260, 2001.<br />

Lamb, H. H.: The early medieval warm epoch and its sequel,<br />

Palaeogeogr. Palaeocl., 1, 13–37, 1965.<br />

Leng, M. J. and Marshall, J. D.: Palaeoclimate interpretation of<br />

stable isotope data from lake sediment archives, Quat. Sci. Rev.,<br />

23, 811–831, 2004.<br />

Marlon, J. R., Bartlein, P. J., Carcaillet, C., Gavin, D. G., Harrison,<br />

S. P., Higuera, P. E., Joos, F., Power, M. J., and Prentice, I. C.:<br />

Climate and human influences on global biomass burning over<br />

the past two millennia, Nat. Geosci., 1, 697–702, 2008.<br />

McCormac, F. G., Hogg, A. G., Blackwell, P. G., Buck, C. E.,<br />

Higham, T. F. G., and Reimer, P. J.: SHCal04 southern hemisphere<br />

calibration 0–11.0 cal kyr BP, Radiocarbon, 46, 1087–<br />

1092, 2004.<br />

Meyers, P. A.: Preservation of elemental and isotopic source identification<br />

of sedimentary organic matter, Chem. Geol. (Isotope<br />

Geosci. Section), 114, 289–302, 1994.<br />

Meyers, P. A. and Teranes, J. L.: Sediment organic matter, in:<br />

Tracking Environmental Change Using Lake Sediments: Volume<br />

2, edited by: Last, W. M. and Smol, J. P., Kluwer Academic Publishers,<br />

Dordrecht, The Netherlands, 239–269, 2001.<br />

Meyers, P. A., Leenheer M. J., and Bourbonniere, R. A.: Diagenesis<br />

of vascular plant organic matter components during burial in lake<br />

sediments, Aquat. Geochem. 1, 35–52, 1995.<br />

Nevle, R. J. and Bird, D. K.: Effects of syn-pandemic fire reduction<br />

and reforestation in the tropical Americas on atmospheric CO2<br />

during the European conquest, Palaeogeogr. Palaeocl., 264, 25–<br />

38, 2008.<br />

O’Leary, M. H.: Carbon isotope fractionation in plants, Phytochemistry,<br />

20, 553–567, 1981.<br />

O’Leary, M. H.: Carbon isotopes in photosynthesis, Bioscience, 38,<br />

328–335, 1988.<br />

Ortlieb, L. and Macharé, J.: Former El Niño events: records from<br />

western South America, Global Planet. Change, 7, 181–202,<br />

1993.<br />

Rowe, J. H.: An introduction to the archaeology of Cuzco, Papers of<br />

the Peabody Museum of American Archaeology and Ethnology,<br />

27, Harvard University Press, Cambridge, USA, 70 pp., 1944.<br />

Sarazin, G., Michard, G., Al Gharib, I., and Bernat, M.: Sedimentation<br />

rate and early diagenesis of particulate organic nitrogen and<br />

carbon in Aydat lake (Puy de Dome, France), Chem. Geol., 98,<br />

307–316, 1992.<br />

Seltzer, G. and Hastorf, C.: Climatic change and its effect on Prehispanic<br />

agriculture in the central Peruvian Andes, J. Field Archaeol.,<br />

17, 397–414, 1990.<br />

Smith, C. S. and Barko, J. W.: Ecology of Eurasian watermilfoil, J.<br />

Aquat. Plant Manage., 28, 55–64, 1990.<br />

Stanish, C.: Ancient Titicaca. University of California Press, Los<br />

Angeles, California, USA, 354 pp., 2003.<br />

Stine, S.: Extreme and persistent drought in California and Patagonia<br />

during Medieval time, Nature, 269, 546–549, 1994.<br />

Stuiver, M. and Reimer, P. J.: Extended 14C database and revised<br />

CALIB radiocarbon calibration program, Radiocarbon, 35, 215–<br />

230, 1993.<br />

Talbot, M. and Lærdal, T.: The Late Pleistocene – Holocene palaeolimnology<br />

of Lake Victoria, East Africa, based upon elemental<br />

and isotopic analyses of sedimentary organic matter, Journal of<br />

Paleolimnology, 23, 141–164, 2000.<br />

Tandeter, E.: La crisis de 1800-05 en el Alto Perú, Data, 1, 9–49,<br />

1991.<br />

Thompson, L. G., Davis, M. E., Mosley-Thompson, E., and Liu, K.<br />

-B.: Pre-Incan agricultural activity recorded in dust layers in two<br />

tropical ice cores, Nature, 307, 763–765, 1988.<br />

Thompson, L. G., Mosley-Thompson, E., and Arnao, B. M.: El<br />

Niño-Southern Oscillation events recorded in the stratigraphy of<br />

the tropical Quelccaya ice cap, Peru, Science, 226, 50–53, 1984.<br />

Thompson, L. G., Mosley-Thompson, E., Dansgaard, W., and<br />

Grootes, P. M.: The Little Ice Age as recorded in the stratigraphy<br />

of the tropical Quelccaya ice cap, Science, 234, 361–364,<br />

www.clim-past.net/5/375/2009/ Clim. Past, 5, 375–388, 2009


388 A. J. Chepstow-Lusty et al.: Putting the rise of the Inca Empire within a climatic and land management context<br />

1986.<br />

Thompson, L. G., Mosley-Thompson, E., Davis, M. E., Lin, P. -<br />

N., Henderson, K. A., Cole-Dai, J., Bolzan, J. F., and Liu, K.-B:<br />

Late glacial stage and Holocene tropical ice core records from<br />

Huascaran, Peru, Science, 269, 46–50, 1995.<br />

Thompson, L. G., Mosley-Thompson, E., and Henderson, K. A.:<br />

Ice-core palaeoclimate records in tropical South America since<br />

the Last Glacial maximum, J. Quat. Sci., 15, 377–394, 2000.<br />

Williams, P. R.: Rethinking disaster-induced collapse in the demise<br />

of the <strong>Andean</strong> highland states: Wari and Tiwanaku, World Archaeol.,<br />

33, 361–374, 2002.<br />

Williams, P. R.: Agricultural innovation, intensification and sociopolitical<br />

development: the case of highland irrigation agriculture<br />

on the Pacific <strong>Andean</strong> watersheds, in: Agricultural Strategies,<br />

edited by: Marcus, J. and Stanish, C., Cotsen Institute of<br />

Archaeology, UCLA, 309–333, 2006.<br />

Whitlock, C. and Larsen, C. P. S.: Charcoal as a fire proxy, in:<br />

Tracking Environmental Change Using Lake Sediments: Volume<br />

3, edited by: Smol, J. P., Birks, H. J. B. and Last W. M., Kluwer<br />

Academic Publishers, Dordrecht, The Netherlands, 75–97, 2001.<br />

Zhang, P., Cheng, H., Edwards, R. L., Chen, F., Wang, Y., Yang,<br />

X., Liu, J., Tan, M., Wang, X., Liu, J., An, C., Dai, Z., Zhou,<br />

J., Zhang, D., Jia, L., and Johnson, K. R.: A test of climate, sun<br />

and culture relationships from an 1810-year Chinese cave record,<br />

Science, 322, 940–942, 2008.<br />

Clim. Past, 5, 375–388, 2009 www.clim-past.net/5/375/2009/


Abstract<br />

Human impacts on headwater fluvial systems in the<br />

northern and central Andes<br />

Carol P. Harden<br />

Department of <strong>Geography</strong> 304 Burchfiel <strong>Geography</strong> Building, University of Tennessee, Knoxville, TN, 37996-0925, USA<br />

Received 25 August 2005; received in revised form 6 June 2006; accepted 6 June 2006<br />

Available online 17 August 2006<br />

South America delivers more freshwater runoff to the ocean per km 2 land area than any other continent, and much of that<br />

water enters the fluvial system from headwaters in the Andes <strong>Mountain</strong>s. This paper reviews ways in which human occupation of<br />

high mountain landscapes in the Andes have affected the delivery of water and sediment to headwater river channels at local to<br />

regional scales for millennia, and provides special focus on the vulnerability of páramo soils to human impact. People have<br />

intentionally altered the fluvial system by damming rivers at a few strategic locations, and more widely by withdrawing surface<br />

water, primarily for irrigation. Unintended changes brought about by human activities are even more widespread and include<br />

forest clearance, agriculture, grazing, road construction, and urbanization, which increase rates of rainfall runoff and accelerate<br />

processes of water erosion. Some excavations deliver more sediment to river channels by destabilizing slopes and triggering<br />

processes of mass-movement.<br />

The northern and central Andes are more affected by human activity than most high mountain regions. The wetter northern<br />

Andes are also unusual for the very high water retention characteristics of páramo (high elevation grass and shrub) soils, which<br />

cover most of the land above 3000 m. Páramo soils are important regulators of headwater hydrology, but human activities that<br />

promote vegetation loss and drying cause them to lose water storage capacity. New data from a case study in southern Ecuador<br />

show very low bulk densities (median 0.26 g cm − 3 ), high organic matter contents (median 43%), and high water-holding capacities<br />

(12% to 86% volumetrically). These data document wetter soils under grass than under tree cover. Effects of human activity on the<br />

fluvial system are evident at local scales, but difficult to discern at broader scales in the regional context of geomorphic adjustment<br />

to tectonic and volcanic processes.<br />

© 2006 Elsevier B.V. All rights reserved.<br />

Keywords: Human impact; Soil erosion; Fluvial geomorphology; Soil moisture; Andes<br />

1. Introduction<br />

The Andes <strong>Mountain</strong>s are the primary headwater<br />

region of the continent of South America, which<br />

delivers more freshwater runoff to the ocean per km 2<br />

land area than any other continent. Assessments of all<br />

natural freshwater resources indicate that the natural<br />

E-mail address: charden@utk.edu.<br />

Geomorphology 79 (2006) 249–263<br />

0169-555X/$ - see front matter © 2006 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.geomorph.2006.06.021<br />

www.elsevier.com/locate/geomorph<br />

internal freshwater resources of South America are<br />

second only to those of Asia (Table 1, FAO, 2003a).<br />

Rivers in the Andes, like rivers in mountain regions<br />

across the globe, respond to gradients, material properties,<br />

and inputs of water and sediment. Unlike most<br />

mountain drainage basins, however, those of the tropical<br />

Andes have been population centers for millennia and<br />

are locations of major cities. The rich history of human<br />

habitation, especially in the northern Andes, and the


250 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Table 1<br />

World water resources by continent (FAO, 2003a)<br />

Continent Total water resources<br />

km 3 − 1<br />

yr<br />

Asia 12,461.0 28.5<br />

South America 12,380.0 28.3<br />

North and Central<br />

America<br />

7443.1 17.0<br />

Europe 6619.4 15.2<br />

Africa 3950.2 9.0<br />

Oceania 910.7 2.1<br />

World 43,764.3 100<br />

% of world freshwater<br />

resources<br />

rapid economic development of <strong>Andean</strong> countries<br />

during the past half-century make this an especially<br />

interesting region in which to examine effects of human<br />

activities on fluvial geomorphology. Although physical<br />

processes by which humans affect runoff, erosion, and<br />

sedimentation are not unique to the Andes, examination<br />

of the intensity and variety of human uses of <strong>Andean</strong><br />

landscapes offers a different perspective on human im-<br />

Fig. 1. Map of the northern and central Andes.<br />

pacts on fluvial systems from that obtained in the northern<br />

hemisphere.<br />

Most (85%) of the continent of South America drains to<br />

the Atlantic Ocean. Brazil, the largest country, contains the<br />

most fresh water, but mean annual internal freshwater<br />

resources per area are greater in Colombia (1.85 m/yr),<br />

Ecuador (1.52 m/yr), Peru (1.26 m/yr), and Venezuela<br />

(0.79 m/yr) than in Brazil (0.63 m/yr)(FAO, 2003a).<br />

Focusing on the five tropical countries of Venezuela,<br />

Colombia, Ecuador, Peru, and Bolivia that form the<br />

<strong>Andean</strong> headwaters of the Amazon (Fig. 1), this paper<br />

reviews deliberate and unintentional impacts of human<br />

activities on headwater rivers in the northern and central<br />

Andes, and highlights human activities that increase<br />

overland rainfall runoff and cause more sediment to<br />

enter river channels. Zooming in from the spatial scales of<br />

a region and countries to the scales of a watershed and<br />

small plots, it investigates the unusual characteristics and<br />

hydrologic importance of páramo soils of the northern<br />

Andes and comments on the relative importance of human<br />

impacts at different spatial scales.


2. Regional setting<br />

2.1. The physical setting<br />

The cordilleras of the Andes divide the Pacific from<br />

Atlantic drainages and capture and regulate the flow of<br />

water for most of the continent of South America. On the<br />

Atlantic side, the Amazon River basin comprises 34% of<br />

the land area of South America. Other major drainage<br />

basins are the Orinoco (Atlantic), Magdelena (Caribbean),<br />

and Paraná (Atlantic). The highest peaks reach to nearly<br />

7000 m above sea level. On the Atlantic and Pacific sides,<br />

relief is dramatic and hillslopes and river gradients are<br />

steep. Westward-flowing rivers from southern Ecuador to<br />

central Chile deliver water to arid lands as they flow from<br />

mountain to ocean. Most of the region is tectonically<br />

active, and sections of it are volcanically active.<br />

Rainfall tends to increase with elevation. Moisture<br />

reaches the mountains primarily from the east, but rainfall<br />

in some Inter<strong>Andean</strong> valleys and on the western flank of<br />

the cordillera is also influenced by weather originating in<br />

the Pacific. ENSO effects are evident, although small<br />

differences in the intensity or penetration of ENSO events<br />

change ENSO influences in the higher mountains (Tarras-<br />

Wahlberg and Lane, 2003). The Magdelena River basin of<br />

Colombia has generally received less rain in El Niño years<br />

and more in the La Niña years (Restrepo and Kjerfve,<br />

2000). Similarly, pulses of deposition from <strong>Andean</strong> headwaters,<br />

recorded in floodplain sediment cores in the Beni<br />

and Mamore river basins, two Bolivian tributaries of the<br />

Amazon, were associated with La Niña events (Aalto<br />

et al., 2003). On a longer timescale, sediment cores<br />

extracted from one high, Amazon-draining lake on the<br />

western cordillera in Ecuador showed periodic episodes<br />

of high sedimentation rates that correlated well with<br />

Holocene ENSO fluctuations (Rodbelletal.,1999).<br />

The highest peaks are snow-capped. Below the snowline<br />

(ca. 4800–5000 m in Ecuador and Colombia) and<br />

above the upper limits of trees and most cultivated land<br />

(ca. 3000–3500 m), the northern Andes are characterized<br />

by the páramo environment. The páramo ecosystem, and,<br />

in particular, páramo soil, is considered to be the principal<br />

regulator of the terrestrial hydrological system of the<br />

northern Andes (Podwojewski and Poulenard, 2000;<br />

Hofstede, 2001). These soils, in Ecuador, Colombia, and<br />

Venezuela, consist of a very black, highly organic<br />

epipedon (A, Ah, and/or O horizons) discontinuously<br />

overlying an unrelated, inorganic surface (Fig. 2).<br />

Because mineral particles in páramo soils are eolian in<br />

origin, páramo soils near to and downwind from active<br />

volcanoes may be 1–2 m thick, while soils farther from<br />

ash sources or on glaciated surfaces may be only 20–<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

251<br />

30 cm deep. Páramo vegetation varies, but is most commonly<br />

grass (Calamagrostis sp., Stipa sp., and Agrostis<br />

sp.), with some shrubs in less disturbed sites. Organic<br />

matter decomposes very slowly in the moist, cool<br />

conditions that result from high elevation, frequent<br />

cloudiness, fog interception, and plentiful rainfall (ca.<br />

1000–2000 mm yr −1 ). Typical mean annual temperatures<br />

are 10–12 °C (Medina and Turcotte, 1999). Páramo soils<br />

are classified as Andosols or Histosols, depending on the<br />

organic matter content, which is typically around 30%.<br />

Low bulk-densities make them very sensitive to disturbance<br />

from humans and livestock. As in the páramo, the<br />

scarcity of trees in the drier puna grasslands of the central<br />

Andes is thought to result from anthropogenic burning<br />

and forest removal (Gade, 1999).<br />

Because the Andes are still tectonically active (Norabuena<br />

et al., 1998), the physical setting includes active<br />

volcanism, ongoing uplift, earthquakes, and high magnitude<br />

mass movements. Uplift has caused rivers to incise<br />

(Safran et al., 2005) and denudation rates to be high (Aalto<br />

et al., 2006). From Colombia to Bolivia, ten volcanoes<br />

have erupted since 1964 (Table 2), and six others erupted<br />

earlier in the 20th century. Volcanoes contribute sediment<br />

to fluvial systems by direct input, by producing bare slopes<br />

vulnerable to erosion processes, by blanketing the surrounding<br />

landscape with ash, and by oversteepening<br />

Fig. 2. Exposure of black páramo soil, under tussock grasses and<br />

perched on a fine-textured deposit, much lighter in color, Cajas<br />

National Park, Ecuador.


252 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Table 2<br />

Volcanoes active since 1964<br />

Country Volcano name Recent eruption<br />

Bolivia/Chile Irruputuncu 1995<br />

Colombia Galeras 2002<br />

Colombia Nevado del Ruiz 1991<br />

Colombia Purace 1977<br />

Ecuador Guagua Pichincha 2004<br />

Ecuador Reventador 2003<br />

Ecuador Sangay 2004<br />

Ecuador Tungurahua 2004<br />

Peru Sabancaya 2003<br />

Peru Ubinas 1969<br />

Source: Smithsonian Institution (2005).<br />

slopes, thus, making them vulnerable to landsliding. Historical<br />

reports indicate that a 1773 eruption of Tungurahua<br />

Volcano (Ecuador) dammed the Pastaza River, and that the<br />

same river had been dammed three times by eruptions in<br />

1918 (Zevallos, 1996). Ash, ice, and water in combination<br />

create lahars, which fill and alter river channels for long<br />

distances. Lahars produced by the 1985 eruption of<br />

Nevado del Ruiz volcano in Colombia moved a total of<br />

about 9×10 7 m 3 of lahar slurry to areas up to 104 km from<br />

the source (Pierson et al., 1990). In Ecuador, a lahar from<br />

an 1877 eruption of Cotopaxi volcano followed a river<br />

valley to the Pacific Ocean, over 250 km away (Hall,<br />

1977).<br />

Mass movements are important in delivering sediment<br />

to the river channels of the Andes (Table 3). On the<br />

east flanks of the Andes in Bolivia, where anthropogenic<br />

effects on erosion processes appear to have been<br />

minimal (Aalto et al., 2006), landsliding is widespread<br />

(Safran et al., 2005). In this region of steep slopes, mass<br />

movements are readily triggered by wet conditions and<br />

by earthquakes. The rugged topography favors landsliding<br />

and the formation of landslide dams. In an<br />

assessment of mass movements (rock and earth slides,<br />

debris avalanches, debris and mud flows) triggered by<br />

two earthquakes about 25 km north of Reventador Volcano<br />

in northeastern Ecuador, Schuster et al. (1996)<br />

found the greatest amount of property destruction to<br />

have been caused by flood surges of the main rivers,<br />

which had been near flood stage before large volumes of<br />

landslide debris were added to them. The largest flood<br />

surges were caused by the breaching of temporary<br />

landslide debris dams. In the case of the large 1993 slope<br />

failure at La Josefina (southern Ecuador), 30×10 6 m 3 of<br />

debris filled the channel of the Paute River for 1 km of<br />

its length (Plaza-Nieto and Zevallos, 1994). The<br />

engineered but catastrophic release of the landslide<br />

dam 33 days later re-formed the channel downstream,<br />

locally increasing the bed slope and causing rapid incision<br />

through the deposit.<br />

2.2. Population trends<br />

In the northern and central Andes, the cordillera of the<br />

Andes is not a single spine, but a region of multiple ranges<br />

with elevated valleys and plateaus. In the north, changing<br />

patterns of subduction of the Nazca plate under the South<br />

American plate have built parallel ranges, with high<br />

Inter<strong>Andean</strong> valleys between. This pattern is seen in<br />

Colombia, where three parallel north–south-trending<br />

ranges define the landscape, and in much of Ecuador,<br />

where a relatively well-defined Inter<strong>Andean</strong> valley is<br />

flanked by two parallel ranges. Farther south, the highest<br />

peaks lie east of the Altiplano. In this tropical region,<br />

higher elevation valleys and plains have long been favored<br />

for human settlement—daytime temperatures are<br />

pleasant, health-threatening insects and other disease<br />

vectors are rare, and soils support productive agriculture<br />

where moisture is adequate. Today, the Inter<strong>Andean</strong><br />

valleys contain major towns and cities, including Bogota,<br />

Colombia (2640 m), and Quito, Ecuador (2850 m).<br />

Cusco, Peru is at 3250 m, and La Paz, Bolivia between<br />

3300 and 3600 m. Bogotá had 6.8 million inhabitants in<br />

Table 3<br />

Recent, major mass movements<br />

Year Country Location Comments<br />

1962 Peru Nevados 13×10<br />

Huascaran<br />

6 m 3 , 400–500 persons<br />

killed a<br />

1970 Peru Nevados 30–50×10<br />

Huascaran<br />

6 m 3 , 18,000 killed;<br />

triggered by M 7.7 earthquake a<br />

1974 Peru Mayanmarca<br />

rockslide<br />

1985 Colombia Nevado del<br />

Ruiz lahars<br />

1987 Ecuador Reventador<br />

mass<br />

movements<br />

1993 Ecuador La Josefina<br />

rockslide<br />

1994 Colombia Paez<br />

landslides<br />

a USGS, 2005.<br />

b Martinez et al., 1995.<br />

c Pierson et al., 1990.<br />

d Schuster et al., 1996.<br />

e Plaza-Nieto and Zevallos, 1994.<br />

1.6×10 6 m 3 , created 150-m high<br />

dam on the Mantaro River, 450<br />

persons killed. b<br />

90×10 6 m 3 , killed over 23,000<br />

in Armero;<br />

triggered by volcanic eruption. c<br />

75–110 × 10 6 m 3 , ∼ 1000<br />

persons killed,<br />

earthquake-triggered rock and earth<br />

slides, debris avalanches, and<br />

debris and mud flows; most deaths<br />

from floods. d<br />

30×10 6 m 3 , N35 killed, debris dam<br />

on Paute River. e<br />

Thousands of slides in 250 km 2 area,<br />

earthquake triggered, 270 dead,<br />

1700 missing. a


2003, Quito had 1.3 million (in 2001), Cusco had over<br />

300,000 (in 2002), and La Paz had 789,585 (in 2001)<br />

(Citypopulation, 2006). Many contemporary settlements<br />

occupy sites with long histories of human use.<br />

In spite of significant rates of international emigration,<br />

populations of <strong>Andean</strong> countries continue to grow<br />

and to become more urbanized (Table 4). Fertility rates<br />

remain high by North American standards, ranging from<br />

2.6 (Colombia and Peru) to 3.9 (Bolivia), compared to<br />

1.9 (children per woman) in the United <strong>State</strong>s (Earthtrends,<br />

2003). Population increased dramatically between<br />

1950 and 2005 in the five countries of Bolivia,<br />

Columbia, Ecuador, Peru, and Venezuela, although not<br />

all of the increase was in the <strong>Andean</strong> highlands.<br />

3. Deliberate human impacts on the <strong>Andean</strong> fluvial<br />

system<br />

As populations have grown and economies developed,<br />

countries have deliberately altered fluvial systems<br />

to take advantage of water resources. The two most<br />

significant intentional interventions are dams and water<br />

withdrawls. Major dams have been built for hydroelectric<br />

power generation and to store water for irrigation<br />

(Table 5). Abundant water and dramatic drops in elevation<br />

have allowed the northern <strong>Andean</strong> countries to<br />

become dependent on hydroelectric power. In Venezuela,<br />

60% of the power is hydroelectrically generated<br />

(Earthtrends, 2003). At least 60% of the electric power<br />

of Ecuador is generated at a single large dam, the Daniel<br />

Palacios Dam on the Paute River (Hofstede, 2005). By<br />

trapping sediment and altering flow, dams profoundly<br />

influence downstream reaches.<br />

Water withdrawls for irrigation are more widespread<br />

than dams. In Bolivia, Ecuador, and Peru, more than 80%<br />

of surface water withdrawls are for agricultural irrigation<br />

Table 4<br />

Population growth and urbanization by country<br />

Country Population 1 Annual<br />

population<br />

growth 1<br />

%<br />

growth<br />

2002 Rural<br />

Areas<br />

Urban<br />

Areas<br />

Population<br />

% increase 1<br />

Percent<br />

urban 2<br />

1950–2002 1965 1989<br />

Bolivia 8,705,000 0.3 3.8 221 40 51<br />

Colombia 43,495,000 0.0 2.7 246 54 69<br />

Ecuador 13,112,000 0.2 4.0 425 37 55<br />

Peru 26,523,000 0.7 2.6 248 52 70<br />

Venezuela 25,093,000 2.0 2.1 392 70 84<br />

Source: 1 Earthtrends, 2003, 2 Valladares and Prates Coelho, 1993.<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Table 5<br />

Dams by country, as reported by FAO (1998–2002)<br />

Country Number of dams Reservoir storage<br />

Bolivia 5 large nd<br />

Colombia 26 large (N25×10 6 m 3 ) 9.1 km 3<br />

90 medium and small 3.4 km 3<br />

Ecuador 12 7.5 km 3<br />

Peru N3 2.7 km 3<br />

Venezuela 96 157 km 3<br />

(Table 6). Some irrigation systems, such as those used in<br />

Peru by the Incas and in northern Ecuador by pre-<br />

Hispanic populations (Knapp, 1991), pre-date the<br />

Spanish conquest by at least hundreds of years. The<br />

amount of water withdrawn for irrigation increased<br />

dramatically in the second half of the 20th century as a<br />

result of population growth, land reform, and government<br />

efforts to intensify agriculture. In Colombia, modern<br />

projects for public irrigation were initiated in 1936,<br />

and, in Bolivia, a commission began planning for public<br />

irrigation projects in 1938 (FAO, 1998–2002). Irrigation<br />

withdrawls in the <strong>Andean</strong> countries expanded following<br />

agrarian reform, which occurred from the 1950s (Bolivia)<br />

through the late 1960s (Peru). Irrigation has<br />

typically been small-scale (micro-irrigation), as landholdings<br />

are small (b5–10 ha) (e.g., White and Maldonado,<br />

1991), and hundreds of irrigation districts<br />

manage the withdrawl and distribution of irrigation<br />

water. Internationally available data on irrigation withdrawls,<br />

which aggregate data by country, reveal major<br />

increases in irrigation over recent decades, but do not<br />

separate <strong>Andean</strong> data from that of lowland drainages.<br />

Abstraction of water for irrigation reduces in-stream<br />

flows, especially in drier seasons. The extent to which<br />

abstraction affects geomorphically important flow levels<br />

in the Andes remains unexamined. Water added to the<br />

landscape by irrigation can reduce soil erosion by<br />

Table 6<br />

Water withdrawls by country and sector (FAO, 2003a)<br />

Country Internal<br />

renewable<br />

surface<br />

water km 3<br />

Total<br />

annual<br />

withdrawl<br />

as%of<br />

renewable<br />

water<br />

Withdrawls<br />

Agriculture Domestic Industry<br />

Bolivia 277 0.3% 87% 10% 3%<br />

Colombia 2112 0.5% 37% 59% 4%<br />

Ecuador 432 4.3% 82% 12% 6%<br />

Peru 1616 1.2% 86% 7% 7%<br />

Venezuela 700 0.8% 46% 44% 10%<br />

253


254 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

improving protective vegetative cover, but too much<br />

water added can destabilize slopes and water added<br />

rapidly on steep slopes causes soil erosion.<br />

4. Human activities that increase soil erosion and<br />

sediment in rivers<br />

Sources of sediment in rivers include hillslopes and<br />

river channel erosion. Sediment is added to the fluvial<br />

system by overland flow, mass movement, deliberate<br />

dumping, or anthropogenic reworking of sediment stored<br />

in channel beds. All of these occur in the Andes. Studies<br />

across the globe have linked increases in sediment yields<br />

of rivers with changes in land use (e.g., Ostry, 1982;<br />

Walling, 1983). Forest clearance for cultivation increases<br />

sediment yields by two to three orders of magnitude<br />

(Ongley, 1996). Agriculture can also be a primary source<br />

of eroded sediment (Bennett, 1939).<br />

Irrigation increases river sediment loads, by reducing<br />

the flow and corresponding dilution effect, and also by<br />

conveying water across steep slopes with high erosion<br />

potential. Analysis of air photos from 1976 and 1989 and a<br />

field survey in 1999 of a 900-ha catchment in the southern<br />

Ecuadorian Andes found new gullies, which were attributed<br />

to poor construction and management of irrigation<br />

infrastructure (Vanacker et al., 2003b). Gully formation<br />

was observed to be a consequence of the spillover of water<br />

from open canals and irrigation reservoirs and of mismanagement<br />

of extra irrigation water. In a similar study of<br />

a different Ecuadorian watershed, the proximity of gullies<br />

to the river appeared to control differences in suspended<br />

sediment concentrations between two subcatchments<br />

(Vanacker et al., 2003a).<br />

Mining activities contribute suspended sediment as well<br />

as metal pollutants. Since the Spanish conquest, the Andes<br />

has been known for deposits of gold, silver, and other<br />

valuable metals. Gold mining remains active, and gold<br />

extraction more than doubled in Bolivia, Colombia, and<br />

Venezuela between 1950 and 1985 (United Nations, 1990).<br />

Gold production in the Portovelo–Zaruma mining district<br />

of western <strong>Andean</strong> Ecuador increased 80% between 1994<br />

and 1999 (Tarras-Wahlberg and Lane, 2003). Copper, iron,<br />

lead, manganese, and zinc are also extracted in large<br />

quantities (United Nations, 1990). In 1991, a database of<br />

geological deposits in the Andes that have already been<br />

mined, are currently being mined, or are under evaluation<br />

for mining contained over 3300 records (BRGM, 2001).<br />

Although effects of mining on fluvial systems in the Andes<br />

have been little studied, mines are typically located in<br />

remote areas and mining regulations not well enforced<br />

(Tarras-Wahlberg and Lane, 2003). Excavations on steep<br />

slopes send debris cascading into streams and also increase<br />

the frequency of landslides. Failures of tailings impoundment<br />

also send debris into rivers. In one study of the effects<br />

of gold mining in the Puyango River basin on the western<br />

flank of the Andes in Ecuador (Tarras-Wahlberg and Lane,<br />

2003), concentrations of suspended sediment derived from<br />

mining were apparent in drier years, but represented only a<br />

small proportion of the estimated sediment yield in wet<br />

years. Restrepo and Kjerfve (2000) cited gold mining in the<br />

Cauca basin of Colombia as an important contributor to<br />

high sediment concentrations in the Magdelena River,<br />

which has one of the highest sediment yields on the<br />

continent.<br />

Another form of mining that has affected fluvial systems<br />

is the in-stream removal of sand, gravel, and river<br />

Fig. 3. Homemade weir catches marketable sand at high flow in a headwater stream.


ocks for construction purposes (Fig. 3). In-stream extraction<br />

was widely practiced in <strong>Andean</strong> streams in Ecuador<br />

in the early 1990s (Harden, 1993) but has<br />

diminished greatly because of recognition of the environmental<br />

consequences and enactment and enforcement<br />

of new regulations. In-stream mining suspends sediment<br />

at all flow levels rather than only during runoff events. In<br />

mined river systems, then, rainfall is not a good predictor<br />

of suspended sediment loads.<br />

Steep gradients promote fluvial transport of sediments.<br />

Dams and reservoirs trap sediment, calling attention<br />

to the magnitude of sediment loads when<br />

reservoir storage capacity is lost. The Daniel Palacios<br />

dam on the Paute River in Ecuador originally, in 1983,<br />

contained 120 ×10 6 m 3 of storage capacity, but that<br />

capacity had been reduced to 95×10 6 m 3 by 1993<br />

because of sedimentation. The mean annual rate of<br />

sedimentation in the reservoir increased from 2.47×10 6 m 3<br />

to 2.76×10 6 m 3 after May, 1993, when the release of the<br />

large landslide dam on the Paute River at La Josefina<br />

mobilized additional sediment (Zevallos and Jerves, 1996).<br />

An expensive dredging program has been used to maintain<br />

storage because of the important hydroelectric power plant<br />

associated with this dam.<br />

Mass movement hazards are increasing globally as<br />

population pressure and economic development push<br />

people onto steeper lands and natural vegetation is removed<br />

for cultivation and other purposes (Vanacker<br />

et al., 2003c). Although landsliding is a natural adjustment<br />

to the crustal shortening occurring in the Andes, not<br />

all of the major <strong>Andean</strong> landslides have been completely<br />

natural events. Two of 11 causal factors listed in an<br />

analysis of the 1993 La Josefina rockslide were excavation<br />

at the toe and diversion of drainage for mining<br />

purposes (Plaza-Nieto, 1996). Higher in the Paute River<br />

basin, locals attribute the 2003 reactivation of a former<br />

landslide (Soroche) to alterations in agricultural drainage<br />

above the failure (Fig. 4).<br />

5. Land use effects on rainfall runoff<br />

One of the most widespread human impacts to the<br />

fluvial system in the Andes, and in other mountain regions,<br />

is increasing the proportion of rainfall that reaches rivers as<br />

surface runoff. Changes in the rates of runoff-generation<br />

occur as unintended consequences of nearly all human<br />

activities. Forest cover, agricultural practices, grazing,<br />

urbanization, and road construction have important and<br />

spatially extensive effects on runoff, which then controls<br />

rates of erosion and sediment movement. These effects are<br />

not unique to the Andes, but are especially interesting in<br />

the Andes because of the intensity of human occupation<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

255<br />

Fig. 4. Soroche landslide in the upper Machángara River tributary of<br />

the Paute River in southern Ecuador.<br />

and the special natural and cultural characteristics of <strong>Andean</strong><br />

highland environments.<br />

Forest clearance is part of the legacy of human occupation<br />

of the <strong>Andean</strong> region. Contemporary deforestation<br />

is occurring primarily in areas, generally on the<br />

external margins of the Andes, where forests remain, or in<br />

dry climates, where tree removal promotes desertification<br />

(Ministerio de Desarrollo Sostenible y Medio Ambiente,<br />

1996). Inter<strong>Andean</strong> valleys appear to have been cleared<br />

even before the Spanish Conquest, although experts disagree<br />

about whether certain areas, which today are<br />

marginally dry or have thin soils, ever supported trees<br />

(Acosta-Solis, 1977; Ellenburg, 1979; White, 1985;<br />

Gade, 1999; Sarmiento and Frolich, 2002). Photos taken<br />

in the late 19th and early 20th centuries in the province of<br />

Tungurahua, Ecuador show far less tree cover than exists<br />

there today (Banco Central, 1984). The introduction of<br />

Eucalyptus trees in the 1860s led to the reforestation of<br />

cleared areas and the present-day dominance of Eucalyptus<br />

throughout the Inter<strong>Andean</strong> region (Dickinson, 1969).<br />

Forest cover continues to be lost, even while reforestation<br />

programs are adding trees. Between 1990 and 2000,<br />

Bolivia lost forest cover at an average rate of 0.3% per<br />

year, Colombia at 0.4%, Ecuador at 1.2%, Peru at 0.4%,<br />

and Venezuela at 0.4% (FAO, 2003b). Much of this


256 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

change occurred on external flanks of the Andes, and<br />

some occurred in lowlands rather than in <strong>Andean</strong> regions.<br />

Studies of land-cover change have documented transitions<br />

from native trees to plantations of exotic species,<br />

principally Eucalyptus and Pinus, without changing the<br />

overall extent of tree cover in the study area. Vanacker<br />

et al. (2003a) detected profound changes in land cover<br />

between 1962 and 1995 in the Machángara watershed in<br />

southern <strong>Andean</strong> Ecuador, which included an overall gain<br />

in total wooded area even though 41% of the secondary<br />

native woodland was cleared during that period. They<br />

attributed the changes to land reform programs and population<br />

growth, and the net gain to reforestation with<br />

Eucalyptus on formerly degraded lands. Gentry and<br />

Lopez-Parodi (1980) linked deforestation in the upper<br />

Amazon watershed in Peru and Ecuador to increased high<br />

water levels in the Amazon River at Iquitos and in<br />

Peruvian tributaries upstream of Iquitos.<br />

The illicit drug trade, specifically of coca in Peru,<br />

Bolivia, and Colombia, has been identified as a contributor<br />

to deforestation in the tropical Andes (U.S. Department of<br />

<strong>State</strong>, 2001). Illicit coca production has typically involved<br />

a slash-and-burn approach in which fields are abandoned<br />

after two to three growing seasons and new fields cleared<br />

to gain fertility and evade authorities. If growers have<br />

moved onto the land from other environments and are<br />

unfamiliar with local conditions, coca cultivation is more<br />

likely to promote runoff and soil loss. The U.S. <strong>State</strong><br />

Department (2001) estimated that a minimum of 2.4 Mha<br />

of forest were cleared for coca production in the <strong>Andean</strong><br />

region over the previous 20 yrs.<br />

Globally, research has demonstrated that forest clearance<br />

leads to less rainfall interception, less infiltration, less<br />

evapotranspiration, and more surface runoff (e.g., Bosch<br />

and Hewlett, 1982). Loss of forest cover, thus, increases<br />

storm hydrograph volumes and shortens lag times to peak<br />

discharges. Although forest removal typically increases<br />

runoff, reforestation does not necessarily reverse the trend<br />

and increase rainfall infiltration (Harden and Mathews,<br />

Fig. 5. Increase of cattle in northern and central <strong>Andean</strong> countries.<br />

2000). Soils are vulnerable to erosion following forest<br />

clearance, so cleared mountain slopes may readily lose soil,<br />

and, thus, the ability to absorb rainwater, between the times<br />

of clearing and reforestation. Many examples of degraded<br />

locations, at which Eucalyptus trees planted for reforestation<br />

out-competed the understory, exist in the region. Such<br />

scenarios result in trunks of trees rising from bare surfaces,<br />

which continue to erode and degrade, and to which little or<br />

no new organic matter is added. Inbar and Llenera (2000)<br />

suggested that the massive reforestation of Eucalyptus in<br />

highland Peru in 1976 was less effective than ancient<br />

terraces in preventing soil erosion. Where efforts have been<br />

made to replace páramo vegetation with pine trees to<br />

increase carbon sequestration, pines have been observed to<br />

reduce water yields and dry the soils (Hofstede, 2001).<br />

Other human activities that have altered rates and<br />

patterns of rainfall infiltration and runoff are those that<br />

cause soil compaction. Among these are the effects of<br />

livestock grazing, increased tractor use in <strong>Andean</strong> agriculture,<br />

growth of road networks, and urbanization. Grazing<br />

and trampling pressures increased greatly after the<br />

Spanish brought cattle, sheep, and horses to the Andes in<br />

the early 1500s. Compared to the native camelids (llamas,<br />

alpacas, vicuñas), hoofed animals of European origin are<br />

heavier and exert more force per foot (White and<br />

Maldonado, 1991; Gade, 1999). When heavy animals,<br />

e.g., cattle, are confined to a limited area, their weight<br />

compacts the underlying soil, and reduces infiltration<br />

capacity (Hofstede et al., 2002). The absence of a freezing<br />

winter season means that trampled soils do not have an<br />

annual period of recuperation, so infiltration capacities<br />

remain low from year to year. Grazing has been locally<br />

intense during the five centuries since Europeans arrived,<br />

and, where conditions have allowed, cattle numbers have<br />

increased in recent years (Fig. 5, FAO, 2006).<br />

Contemporary agriculture can increase or decrease<br />

rates of soil infiltration and rainfall runoff. Although<br />

clearing forested land for cultivation usually has the<br />

effect of increasing runoff, tilling cleared land increases<br />

infiltration capacities (Harden, 1991). The weight of<br />

tractors compacts the soil, so increased tractor use over<br />

the past half-century (Fig. 6, FAO, 2006) can be expected<br />

to have increased rates of runoff in <strong>Andean</strong> farmlands.<br />

The preferential generation of runoff on roads and<br />

footpaths accelerates erosion on hillslopes (Harden,<br />

1992). The location of roads near streams allows<br />

increased runoff to discharge quickly from the land to<br />

the fluvial system. Land reform programs in the 1960s<br />

and 1970s in Ecuador led to more land ownership of<br />

steeper hillsides and higher elevation fields. At the same<br />

time, land became more parcelized (Vanacker et al.,<br />

2003a). Because boundaries between parcels interrupt


Fig. 6. Increase in tractor use since 1955, by country.<br />

overland flow, parcelization can reduce soil and water<br />

loss from agricultural plots. The interrelated factors of<br />

plot steepness and length, tillage practices, and the<br />

distribution of croplands on hillsides make the hydrologic<br />

effects of cultivation difficult to generalize.<br />

Cessation of a land use that exacerbates runoff or<br />

sediment production does not necessarily mean return to<br />

a previous state of infiltration and sediment stability. In<br />

the dry central Andes of Peru, where annual rainfall is<br />

only 350 mm, Inbar and Llenera (2000) found that<br />

abandonment of agricultural terraces increases rates of<br />

soil erosion and sediment yields. More than 2 Mha of<br />

highland Peru are estimated to have been terraced, but<br />

91% of terraces in high areas have been abandoned<br />

(Inbar and Llenera, 2000). Plants do not grow naturally<br />

on the terraced slopes in this dry environment. In more<br />

humid Ecuador, where slopes were not terraced,<br />

previously cultivated sites that had been abandoned or<br />

left in long-term fallow generated significantly more<br />

runoff than sites under cultivation (Harden, 1996). Degraded<br />

soil, lack of moisture, and ongoing grazing<br />

stresses cause abandoned <strong>Andean</strong> farmlands to continue<br />

to be sources of runoff and sediment for many years<br />

(Harden, 2001). An additional factor that has reduced<br />

the ability of the soil to absorb and hold rainwater has<br />

been the wholesale removal of peaty soil. In Bolivian,<br />

turf has been sold as fuel (Godoy, 1990); in the puna of<br />

Peru, peat soil has been mined for fuel and for horticultural<br />

purposes (Llerena, 1987); and, in Ecuador, this<br />

author has observed peat mining for greenhouses that<br />

produce flowers for export.<br />

6. Human impacts on the moisture storage capacity<br />

of páramo soil<br />

Soils play a key role in determining whether rainfall is<br />

absorbed by or shed from a site; they also store and release<br />

water. Human activities that cause soil compaction reduce<br />

the available pore space in the soil and thus reduce the<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

ability of the soil to store water. Soil moisture storage<br />

capacity is also reduced when soil is lost or the organic<br />

matter content of a soil decreases. The páramo soils,<br />

which cover 35,000 km 2 of the northern Andes (Hofstede,<br />

2005), are especially vulnerable to changes that reduce<br />

pore volume. With low bulk density and high organic<br />

matter content, páramo soils are viewed as enormous<br />

sponges, which feed and regulate flow to the fluvial<br />

system (Luteyn, 2005). Water retention capacity of<br />

undisturbed páramo soils is extremely high, reaching<br />

values of more than 100% at the wilting point, and<br />

strongly correlated with the organic matter content<br />

(Buytaert et al., 2005; Luteyn, 2005). Poulenard et al.<br />

(2003) reported water contents in epipedons of páramo<br />

Hydric Melanudands at 1500 kPa≥1000 g kg −1 and<br />

attributed high porosities to the abundance of organic<br />

colloids.<br />

Loss of vegetative cover promotes drying, which<br />

irreversibly reduces pore space and the water-holding<br />

capacity of the páramo soil (Poulenard et al., 2003).<br />

Studies have shown páramo soils to become crusted and<br />

even hydrophobic following disturbance (Poulenard<br />

et al., 2001). Fire is the principal human disturbance<br />

affecting the hydrology of páramo environments. Today,<br />

páramos are primarily burned to remove old grass and<br />

promote the growth of tender new shoots as food for<br />

cattle. Gade (1999) reported that any grassy site in the<br />

high Andes is probably burned at least once each 5 yrs.<br />

Similarly, Hofstede (2005) suggested that only the most<br />

remote or most protected páramo sites are not affected<br />

by livestock. Páramos are also burned to clear land for<br />

cultivation, improve hunting, or implement local belief<br />

systems (e.g., bring rain, deter evil spirits) (Hofstede,<br />

2001). Grass páramos have been used as grazing lands at<br />

least since the arrival of cattle, sheep and horses with the<br />

Spanish in the 1500s, so widespread burning, coupled<br />

with trampling and vegetation removal by grazing,<br />

would have reduced moisture storage and increased<br />

rates of runoff in páramo environments over the last five<br />

centuries. These practices have been so widespread that<br />

the region lacks control sites for comparative studies.<br />

7. Case study of soil moisture<br />

257<br />

To more closely examine the soil moisture conditions<br />

in the headwater region and to investigate differences in<br />

soil moisture between grass- and tree-covered highland<br />

sites, a study was conducted of surface soil moisture<br />

characteristics in and near the 50 km 2 Llaviucu watershed,<br />

in the western cordillera of the Andes near Cuenca,<br />

Ecuador. The Llaviucu watershed is of special interest<br />

because it yields 20–30% of the water used by the city of


258 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Cuenca and because most of it lies within Cajas National<br />

Park (Fig. 7). Elevations in the watershed range from<br />

3000 to 4300 m. Vegetation is predominantly grass<br />

páramo, with occasional small stands of Polylepis trees at<br />

high elevations (Fig. 8), tropical montane forest vegetation<br />

below 3400 m on the steep flanks of the glacial trough<br />

of the lower watershed, and imported pasture grasses on<br />

the trough floor. The national park is managed by the<br />

Municipality of Cuenca through a corporation with major<br />

leadership by ETAPA (Empresa Pública Municipal de<br />

Telecomunicaciones, Agua Potable y Alcantarillado y<br />

Saneamiento de Cuenca), the local utility company. In<br />

addition to the usual goals of protecting the natural environment<br />

and fostering tourism, recreation, and environmental<br />

education, this park is also managed to maximize<br />

water storage and dry season river flow. Cattle grazing in<br />

the park has successfully been reduced from tens of<br />

thousands to a very small number, and the corporation<br />

must now decide whether to continue to burn the páramo<br />

regularly or allow vegetative succession to occur. Evidence<br />

in the region indicates that trees, which presently<br />

occur in isolated islands up to 4300 m (Gade, 1999), may<br />

become dominant in the absence of burning (White and<br />

Maldonado, 1991), but the water resources effects of such<br />

a change are not known.<br />

Fig. 7. Location of Cajas National Park, Ecuador.<br />

For this study, pairs of soil moisture study plots were<br />

established at elevations between 3163 m and 3527 m to<br />

compare the effects of grass and tree cover where other<br />

factors–elevation, location, slope, parent material, soil<br />

development–were the same. A HydroSense (Campbell<br />

Scientific) Time Domain Reflectometer (TDR) was<br />

deployed in 4 m 2 plots at 23 sites to obtain 5–15 replications<br />

of in situ measurements of volumetric moisture<br />

content (VMC) at each site. The 12-cm TDR probes integrated<br />

VMC along their length. In each plot, a 12-cm<br />

deep soil sample was collected between one set of TDR<br />

probe holes for laboratory determination of gravimetric<br />

moisture content (GMC), and a second sample (0–12 cm<br />

deep), taken within 1 m of the first, was extracted for bulk<br />

density determination. The sand replacement method was<br />

used in the field to determine in situ volume for bulk<br />

density calculations. All soil samples were air-dried and<br />

then oven-dried (24 h at 105–110 °C) and weighed. GMC<br />

samples were weighed before drying, so that GMC could<br />

be calculated as the ratio of the mass of water (mwater=wet<br />

soil mass−oven-dried soil mass) to the mass of oven-dried<br />

soil (GMC=mwater/msoil).<br />

Loss-on-Ignition (LOI), interpreted as an approximate<br />

measure of the organic material (% by mass), was<br />

determined as the percentage of the sample mass


emaining after burning the oven-dried sample at 550 °C<br />

for 2 h. All sampling was done in June and July of 2004, a<br />

relatively moist year.<br />

Volumetric moisture content is the ratio of the volume<br />

of water to the volume of soil. As volume is the ratio of<br />

mass to density, VMC may also be expressed as<br />

VMC ¼ðmwater=q waterÞ=ðmsoil=q soilÞ: ð1Þ<br />

Using soil bulk density as ρsoil, and setting the density<br />

of water (ρwater) at1gml − 1 , the three values for VMC,<br />

GMC and bulk density at each site can be related as<br />

GMC ¼ðVMC⁎1 gml 1 Þ=ð bulk densityÞ: ð2Þ<br />

The resulting data show extremely low bulk densities<br />

(median 0.26 g cm − 3 ), high organic matter contents of<br />

surface horizons (median 43%), and high water-holding<br />

capacities (median GMC 1.52 g g − 1 ; VMC ranges from<br />

12% to 86%). In other words, the mass of water in these<br />

soils is typically 1.5 times the dry mass of solid soil<br />

material. Measurements from paired plots showed that<br />

soils under grass cover consistently had higher VMC<br />

than soils under trees (Table 7), and field observations<br />

showed that soils under trees contained more macropores<br />

compared to soils under grass cover. In all pairs,<br />

soil under tree cover was less dense and drier than soil<br />

under grass. A comparison of soil properties at four<br />

páramo soils sites, three that had burned recently and<br />

one that had not, showed little or no difference in bulk<br />

density or VMC between the four (Table 8).<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Fig. 8. Páramo grassland in Cajas National Park, Ecuador.<br />

259<br />

Although bulk densities in the Llaviucu watershed<br />

study were very low, generally good agreement (same<br />

order of magnitude, most within 50%) between measured<br />

and calculated values of soil moisture and field<br />

observation provided confidence in the results. This<br />

comparison between grass- and tree-covered soils was<br />

limited by the small size of the database and by the<br />

difficulty of finding sites at which all other factors were<br />

equal, as trees were observed to typically occupy steeper<br />

and rockier sites. These initial results suggest that treecovered<br />

sites store less moisture than grass-covered sites<br />

and raise the question of whether páramo soils only<br />

behave as “sponges” under grass cover, which may, in<br />

turn, be an artifact of anthropogenic fire management.<br />

Lightning-caused fires have not been reported in<br />

páramos, although lightning strikes have occasionally<br />

been observed (Luteyn, 2005).<br />

The lack of difference in soil moisture and soil bulk<br />

density between recently burned páramo sites and an<br />

unburned, but similarly grassy site was not unexpected,<br />

as the longer-term history of all grass páramo sites<br />

appears to include relatively frequent fires. Results from<br />

this study suggest that, not the fire, but the establishment<br />

of woody plants in the absence of fire would increase<br />

evapotranspiration rates and soil drainage (through<br />

macropore formation), causing páramo soil to lose its<br />

capacity as sponge. A small sample can only be illustrative,<br />

however.<br />

A further confounding factor in the moisture storage<br />

capacity of páramo soils is the depth of soil at a given<br />

location. Soil in much of the Llaviucu watershed is thin


260 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Table 7<br />

Data from paired plots in Llaviucu watershed, Ecuador<br />

Vegetation Altitude<br />

(m)<br />

Moisture GMC<br />

− 1<br />

0–12 cm g g<br />

Bulk density 0–<br />

− 1<br />

12 cm g cc<br />

Moisture VMC<br />

average 0–12 cm<br />

%<br />

(20–30 cm deep), compared to soils at other high<br />

elevation páramo sites less affected by glaciers.<br />

Although it is presently below the lower limit of<br />

permanent ice, the upper Llaviucu watershed was<br />

under part of a larger ice cap (Rodbell et al., 2002),<br />

and the lower Llaviucu watershed is the trough-shaped<br />

valley of an outlet glacier. Field observations made in the<br />

Llaviucu catchment documented the flashiness of runoff<br />

on thin, quickly saturated soils in this glacially scoured<br />

valley. The flashiness of runoff in the watershed and<br />

relative dryness of its wooded soils, even in a wet time,<br />

can be readily explained, but are contrary to conventional<br />

expectations. This underscores the importance of<br />

localized differences and the need for more such case<br />

studies, which should help management make informed<br />

decisions for this national park in its effort to sustain and<br />

VMC range of<br />

obs values %<br />

VMC ratio<br />

grass:<br />

wooded<br />

improve downstream water resources as well as better<br />

understand the páramo in other areas.<br />

8. Summary and conclusions<br />

− 1<br />

GMC g g<br />

calculated as<br />

VMC/BD<br />

PAIR 1 2.8<br />

Grass 1 3170 1.52 0.20 68 48–76 3.4 0.4<br />

Grass 2 3170 1.80 0.34 65 53–79 1.9 0.9<br />

Wood 1 3175 1.06 0.04 20 8–34 5.0 0.2<br />

Wood 2 3175 nd nd 27 13–34<br />

PAIR 2 2.8<br />

Grass 3 3248 0.49 0.54 35 18–57 0.6 0.8<br />

Grass 4 3248 0.70 0.48 0.7 1.0<br />

Wood 3 3260 0.73 0.09 12 8–24 1.3 0.5<br />

Wood 4 3260 0.89 0.11 1.1 0.8<br />

PAIR 3 4.0<br />

Grass 5 3170 1.45 0.32 71 64–76 2.2 0.7<br />

Wood 5 3176 1.31 0.06 20 6–28 3.3 0.4<br />

Wood 6 3176 2.50 0.05 15 10–27 3.0 0.8<br />

PAIR 4 1.3<br />

Grass 6 3163 1.45 1.08 60 53–65 0.6 2.6<br />

Wood 7 3273 1.10 1.06 47 38–58 0.4 2.5<br />

Table 8<br />

Bulk density, Loss-On-Ignition (LOI), volumetric soil moisture<br />

(VMC) and gravimetric soil moisture (GMC) of páramo soils at four<br />

sites in Cajas National Park, southern Ecuador<br />

Vegetation Elev.<br />

m<br />

Bulk<br />

Density g<br />

− 1<br />

cc<br />

LOI<br />

%<br />

VMC<br />

avg.<br />

%<br />

VMC<br />

min.<br />

%<br />

VMC<br />

max.<br />

%<br />

GMC<br />

− 1<br />

gg<br />

Páramo 3465 0.47 29.3 74 72 76 153<br />

burned 1 32.5<br />

Páramo 3467 0.37 42.2 80 78 82 211<br />

burned 2 48.5<br />

Páramo 3469 0.28 54.5 84 83 85 238<br />

burned 3 58.5<br />

Páramo 3527 0.27 57.6 86 84 88 252<br />

unburned<br />

GMC ratio<br />

of obs/calc<br />

Ways in which <strong>Andean</strong> people and their activities<br />

affect the flow and sediment loads of mountain rivers<br />

are essentially the same as on other continents. The<br />

tropical location of the northern Andes, belief systems<br />

that motivate certain land-management strategies, and<br />

the dominance of grass páramo distinguish the Andes<br />

from other steep, volcanically active mountain regions<br />

of the world. Beyond deliberately engineered changes<br />

of river impoundments and water withdrawls, human<br />

activities have many more subtle and unintentional<br />

geomorphic effects on the fluvial system. Forest<br />

clearance, grazing, agriculture, roadways, and urbanization<br />

increase the proportion of rainfall that flows to<br />

the channel network during storms, and steep slopes<br />

give overland flow the energy to erode and move fine<br />

sediments. Human interventions that destabilize slopes,<br />

from mining, to tree removal, to problems with<br />

irrigation canals, contribute to triggering mass movements,<br />

the largest of which dam rivers and subsequently<br />

serve as fluvial sediment sources. Páramo soils, which<br />

are hydrologically important in the northern Andes, are<br />

particularly sensitive to drying, which occurs when<br />

vegetation is removed by burning, grazing, or tilling.<br />

The Llaviucu case study verified the low bulk density<br />

and high moisture-holding capacities of páramo and<br />

highland soils and showed that soils under grass cover


hold more moisture in a moist season than soils under<br />

trees.<br />

At the regional scale, natural processes of uplift and<br />

denudation, volcanic activity, steep slopes, and mass<br />

movement create a geomorphically active environment<br />

in the northern and central Andes. Where human settlement<br />

is sparse and relief is high, such as on the external<br />

flanks of the Andes, the impacts of human activities are<br />

negligible compared to the magnitudes of natural processes<br />

and adjustments in the fluvial system. A recent<br />

study in the Upper Beni River basin in the Bolivian<br />

Andes found no significant change in rates of erosion,<br />

determined from 10 Be analysis of quartz grains, over the<br />

last millions of years (Safran et al., 2005). Likewise, a<br />

broader study of rates of erosion from 47 rapidly eroding<br />

drainage basins in the Bolivian Andes concluded that<br />

anthropogenic disturbance was minimal (Aalto et al.,<br />

2006). Given the great distance and the number of<br />

sediment sinks in the Amazon basin between the Andes<br />

and the Atlantic Ocean, it is unlikely that human impacts<br />

in the <strong>Andean</strong> headwaters are noticed at the mouth of the<br />

Amazon.<br />

On the scale of 10 3 km 2 in human-dominated landscapes<br />

of the Inter<strong>Andean</strong> valleys and on a temporal scale<br />

of years to centuries, unintended human impacts on soil<br />

erosion and runoff are evident, if not well documented.<br />

These are visible as truncated soils, reservoir sedimentation,<br />

stream incision, increased duration of stream turbidity,<br />

and accelerated rates of mass movement where<br />

people have steepened slopes through construction and<br />

mining. At the scale of plots (10 m 2 ) definite differences in<br />

soil properties are associated with differences in land cover<br />

and land use.<br />

Many <strong>Andean</strong> river valleys are narrow and steep, so<br />

changes in slope stability or changes that cause soil to<br />

erode or rainfall to run off are rapidly transmitted to the<br />

river system. The land uses today follow a legacy, begun<br />

long before the Colonial era, of forest clearance, agriculture,<br />

and urbanization in the <strong>Andean</strong> region. Land uses<br />

have changed as populations have increased, local and<br />

global economies and technologies have changed, and<br />

land reforms have been implemented. Human impacts in<br />

these mountains may now be more intensive and<br />

extensive, but they are not new.<br />

Acknowledgements<br />

The author completed the case study in Cajas<br />

National Park with the support from the Fulbright<br />

Commission, U.S. Department of <strong>State</strong>, with collaboration<br />

of faculty and students from the University of<br />

Cuenca, Ecuador, and with permission from Cajas<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

National Park, ETAPA, and the Municipalidad de<br />

Cuenca. Additional thanks go to Alan Moore and<br />

Marisa Ernst for assistance in the field and to Will<br />

Fontanez and David Moore for help with databases and<br />

graphics.<br />

References<br />

261<br />

Aalto, R., Dunne, T., Guyot, J.-L., 2006. Geomorphic controls on<br />

<strong>Andean</strong> denudation rates. The Journal of Geology 114, 85–99.<br />

Aalto, R., Maurice-Bourgoin, L., Dunne, T., Montgomery, D.,<br />

Nittrouer, C., Guyot, J.-L., 2003. Episodic sediment accumulation<br />

on Amazonian flood plains influenced by El Niño/Southern<br />

Oscillation. Nature 425 (2), 493–496.<br />

Acosta-Solis, M., 1977. Conferencias Fitogeográficas. I.P.G.H.,<br />

Sección Nacional del Ecuador, Quito.<br />

Banco Central, 1984. Paisajes del Ecuador. Colección Imágenes, vol. 5.<br />

Banco Central del Ecuador, Centro de Investigación y Cultura,<br />

Archivo Histórico, Quito, Ecuador.<br />

Bennett, H.H., 1939. Soil Conservation. McGraw-Hill Book Company,<br />

Inc., New York.<br />

Bosch, J.M., Hewlett, J.D., 1982. A review of catchment experiments<br />

to determine the effect of vegetation changes on water yield and<br />

evapotranspiration. Journal of Hydrology 55, 3–23.<br />

BRGM (Bureau de Recherches Géologiques et Minières), 2001. The<br />

GIS Andes Mineral Deposit Database. http://gisandes.brgm.fr/<br />

index.htm (updated 2001, accessed 2005).<br />

Buytaert, W., Sevink, J., De Leeuw, B., Deckers, J., 2005. Clay<br />

mineralogy of the soils in the south Ecuadorian páramo region.<br />

Geoderma 127, 114–129.<br />

Citypopulation, 2006. The principal cities and agglomerations in<br />

America. http://www.citypopulation.de/America.html (accessed<br />

in March 2006).<br />

Dickinson III, J., 1969. The eucalypt in the Sierra of southern Peru.<br />

Annals of the Association of American Geographers 59 (2), 294–307.<br />

Earthtrends, 2003. Country profiles. Water Resources Institute. http://<br />

earthtrends.wri.org (accessed May 2005).<br />

Ellenburg, H., 1979. Man's influence on tropical mountain ecosystems<br />

in South America. Journal of Ecology 67, 401–416.<br />

FAO (Food and Agriculture Organization of the United Nations),<br />

1998–2002. Perfiles Hídricos por Pais. http://www.fao.org/ag/agl/<br />

aglw/aquastat/countries/indexesp.stm.<br />

FAO (Food and Agriculture Organization of the United Nations),<br />

2003a. Review of World Water Resources by Country. FAO<br />

Technical Paper, vol. 23. Rome. http://www.fao.org/DOCREP/<br />

005/Y4473E/Y4473E00.pdf.<br />

FAO (Food and Agriculture Organization of the United Nations),<br />

2003b. Situación Forestal en la Región de América Latina y el<br />

Caribe 2002. FAO Oficina Regional de la FAR para America<br />

Latina y el Caribe, Santiago, Chile. http://www.fao.org/Regional/<br />

LAmerica/pubs/document.htm (accessed January 2005).<br />

FAO (Food and Agriculture Organization of the United Nations),<br />

2006. FAOSTAT, FAO Statistical Databases. http://faostat.fao.<br />

org/ (accessed March 2006).<br />

Gade, D., 1999. Nature and culture in the Andes. The University of<br />

Wisconsin Press, Madison, WI.<br />

Gentry, A., Lopez-Parodi, J., 1980. Deforestation and increasing<br />

flooding of the upper Amazon. Science 210, 1354–1356.<br />

Godoy, R.A., 1990. Mining and agriculture in highland Bolivia.<br />

University of Arizona Press, Tucson, AZ.


262 C.P. Harden / Geomorphology 79 (2006) 249–263<br />

Hall, M., 1977. El Volcanismo en el Ecuador. I.P.G.H., Sección<br />

Nacional del Ecuador, Quito.<br />

Harden, C.P., 1991. <strong>Andean</strong> soil erosion: a comparison of soil erosion<br />

conditions in two <strong>Andean</strong> watersheds. National Geographic<br />

Research and Exploration 7 (2), 216–231.<br />

Harden, C.P., 1992. Incorporating roads and footpaths in watershedscale<br />

hydrologic and soil erosion models. Physical <strong>Geography</strong> 13<br />

(4), 368–385.<br />

Harden, C.P., 1993. Upland erosion and sediment yield in a large<br />

<strong>Andean</strong> drainage basin. Physical <strong>Geography</strong> 14 (3), 254–271.<br />

Harden, C.P., 1996. Interrelationships between land abandonment and<br />

land degradation: a case from the Ecuadorian Andes. <strong>Mountain</strong><br />

Research and Development 16 (3), 274–280.<br />

Harden, C.P., 2001. Soil erosion and sustainable mountain development:<br />

experiment, observations, and recommendations from the<br />

Ecuadorian Andes. <strong>Mountain</strong> Research and Development 21 (1),<br />

77–83.<br />

Harden, C.P., Mathews, L.E., 2000. Rainfall response of degraded soil<br />

following reforestation in the Copper Basin, Tennessee, USA.<br />

Environmental Management 26 (2), 163–174.<br />

Hofstede, R., 2001. El impacto de las actividades humanas sobre el<br />

páramo. In: Mena, P., Medina, G., Hofstede, R. (Eds.), Los<br />

Páramos del Ecuador: Paticularidades, Problemas y Perspectives.<br />

Abya Yala/Proyecto Páramo, Quito, Ecuador, pp. 161–185.<br />

Hofstede, R., 2005. Distribución, impacto humano y conservatión de<br />

los páramos neotropicales. In: Kappelle, M., Horn, S. (Eds.),<br />

Páramos de Costa Rica. Editorial INBio, Santo Domingo de<br />

Heredia, Costa Rica, pp. 701–724.<br />

Hofstede, R., Groenendijk, J., Coppus, R., Fehse, J., Sevink, J., 2002.<br />

The impact of pine plantations on soils and vegetation in the<br />

Ecuadorian high Andes. <strong>Mountain</strong> Research and Development 22<br />

(2), 159–167.<br />

Inbar, M., Llenera, C., 2000. Erosion processes in high mountain<br />

agricultural terraces in Peru. <strong>Mountain</strong> Research and Development<br />

20 (1), 72–79.<br />

Knapp, G., 1991. <strong>Andean</strong> Ecology. Westview Press, Boulder, CO.<br />

Luteyn, J., 2005. Introducción al ecosistema de páramo. In: Kappelle,<br />

M., Horn, S. (Eds.), Páramos de Costa Rica. Editorial INBio, Santo<br />

Domingo de Heredia, Costa Rica, pp. 37–99.<br />

Llerena, C., 1987. Erosion and sedimentation issues in Peru. Erosion<br />

and Sedimentation in the Pacific Rim (Proceedings of the Corvallis<br />

Symposium, August, 1987)IAHS Publ., vol. 165, pp. 3–14.<br />

Martinez, J., Avila, G., Agudelo, A., Schuster, R., Casadevall, T.,<br />

Scott, K., 1995. Landslides and debris flows triggered by the 6<br />

June 1994 Paez earthquakes, southwestern Colombia. Landslide<br />

News 9, 13–15.<br />

Medina, L., Turcotte, P., 1999. Calidad de las aguas de los<br />

páramos. In: Josse, C., Mena, P., Medina, G. (Eds.), El Páramo<br />

Como Fuente de Recursos Hídricos. Serie Páramo, vol. 3.<br />

Grupo de Trabajo en Páramos del Ecuador/Abya Yala, Quito,<br />

Ecuador, pp. 15–24.<br />

Ministerio de Desarrollo Sostenible y Medio Ambiente, 1996. Mapa<br />

preliminar de erosión de suelos: región árida, semiárida y subhumeda<br />

seca de Bolivia: Memoria explicativa. Secretaria Nacional de<br />

Recursos Naturales y Medio Ambiente, La Paz, Bolivia.<br />

Norabuena, E., Leffler-Griffin, L., Mao, A., Dixon, T., Stein, S., Sacks,<br />

I., Ocola, L., Ellis, M., 1998. Space geodetic observations of<br />

Nazca–South America convergence across the central Andes.<br />

Science 279, 358–362.<br />

Ongley, E.D., 1996. Control of water pollution from agriculture. FAO<br />

Irrigation and Drainage Paper, vol. 55. Food and Agriculture<br />

Organization of the United Nations, Rome.<br />

Ostry, R.C., 1982. Relationship of water quality and pollutant loads to<br />

land uses in adjoining watersheds. Water Resources Bulletin 18,<br />

99–104.<br />

Pierson, T., Janda, R., Thouret, J.-C., Borrero, C., 1990. Perturbation<br />

and melting of snow and ice by the 13 November eruption of<br />

Nevado del Ruiz, Colombia, and consequent mobilization, flow<br />

and deposition of lahars. Journal of Volcanology and Geothermal<br />

Research 41, 17–66.<br />

Plaza-Nieto, G., 1996. El deslizamiento-represamiento. In: Zevallos,<br />

O., Fernández, M.A., Plaza-Nieto, G., Klinkicht-Sojos, S. (Eds.),<br />

Sin Plazo Para la Esperanza. Escuela Politécnica Nacional, Quito,<br />

Ecuador, pp. 57–76.<br />

Plaza-Nieto, G., Zevallos, O., 1994. The 1993 La Josefina Rockslide<br />

and Rio Paute landslide dam, Ecuador. Landslide News 8, 4–6.<br />

Podwojewski, P., Poulenard, J., 2000. Los suelos de los páramos del<br />

Ecuador. In: Mena, P., Josse, C., Medina, G. (Eds.), Los Suelos del<br />

Páramo. Serie Páramo, vol. 5. Grupo de Trajabo en Páramos del<br />

Ecuador/Abya Yala, Quito, Ecuador, pp. 1–26.<br />

Poulenard, J., Podwojewski, P., Janeau, J.L., Collinet, J., 2001. Runoff<br />

and soil erosion under rainfall simulation of Andisols from the<br />

Ecuadorian Páramo: effects of tillage and burning. Catena 45,<br />

185–207.<br />

Poulenard, J., Podwojewski, P., Herbillon, J., 2003. Characteristics of<br />

non-allophanic Andisols with hydric properties from the Ecuadorian<br />

paramos. Geoderma 117 (3–4), 267–281.<br />

Restrepo, J.D., Kjerfve, B., 2000. Magdelena River: interannual<br />

variability (1975–1995) and revised water discharge and sediment<br />

load estimates. Journal of Hydrology 235, 137–149.<br />

Rodbell, D., Seltzer, G., Anderson, D., Abbott, M., Enfield, D.,<br />

Newman, J., 1999. An ∼15,000-year record of El Niño-driven<br />

alluviation in Southwestern Ecuador. Science 283, 516–520.<br />

Rodbell, D., Bagnato, S., Nebolini, J., 2002. A Late-Glacial–Holocene<br />

tephrochronology for glacial lakes in southern Ecuador. Quaternary<br />

Research 57, 343–354.<br />

Safran, E., Bierman, P., Aalto, R., Dunne, T., Whipple, K., Caffee, M.,<br />

2005. Erosion rates driven by channel network incision in the<br />

Bolivian Andes. Earth Surface Processes and Landforms 30,<br />

1007–1024.<br />

Sarmiento, F., Frolich, L., 2002. <strong>Andean</strong> cloud forest tree lines.<br />

<strong>Mountain</strong> Research and Development 22 (3), 278–287.<br />

Schuster, R., Nieto, A., O'Rourke, T., Crespo, E., Plaza-Nieto, G.,<br />

1996. Mass wasting triggered by the 5 March 1987 Ecuador<br />

earthquakes. Engineering Geology 42, 1–23.<br />

Smithsonian Institution, 2005. Global Volcanism Program. http://<br />

www.volcano.si.edu/world/ (accessed in March 2005).<br />

Tarras-Wahlberg, N.H., Lane, S.N., 2003. Suspended sediment yield and<br />

metal contamination in a river catchment affected by El Niño events<br />

and gold mining activities: The Puyango river basin, southern Ecuador.<br />

Hydrological Processes 17, 3101–3123.<br />

United Nations, 1990. The water resources of Latin America and the<br />

Caribbean-Planning, hazards and pollution. United Nations<br />

Economic Commission for Latin America and the Caribbean,<br />

Santiago, Chile.<br />

U.S. Department of <strong>State</strong>, 2001. The Andes Under Siege: Environmental<br />

consequences of the drug trade. http://usinfo.state.gov/<br />

products/pubs/archive/ (accessed May 2005).<br />

USGS (United <strong>State</strong>s Geological Survey), 2005. 25 Major Landslides<br />

of the 20th Century. http://landslides.usgs.gov/html.files/majorls.<br />

html (accessed April 2005).<br />

Valladares, L., Prates Coelho, M., 1993. Urban research in Latin<br />

America: towards a research agenda. UNESCO. http://www.unesco.<br />

org/most/vallens.htm#urbanresearch (accessed in March 2006).


Vanacker, V., Govers, G., Barros, S., Poesen, J., Deckers, J., 2003a.<br />

The effect of short-term socio-economic and demographic change<br />

on landuse dynamics and its corresponding geomorphic response<br />

with relation to water erosion in a tropical mountainous catchment,<br />

Ecuador. Landscape Ecology 18, 1–15.<br />

Vanacker, V., Govers, G., Poesen, J., Deckers, J., Dercon, G., Loaiza,<br />

G., 2003b. The impact of environmental change on the intensity<br />

and spatial pattern of water erosion in a semi-arid mountainous<br />

<strong>Andean</strong> environment. Catena 51, 329–347.<br />

Vanacker, V., Vanderscaeghe, M., Govers, G., Willems, E., Poesen, J.,<br />

Deckers, J., DeBievre, B., 2003c. Linking hydrological, infinite slope<br />

stability and land-use change models through GIS for assessing the<br />

impact of deforestation on slope stability in high <strong>Andean</strong> watersheds.<br />

Geomorphology52(3–4), 299–315.<br />

Walling, D., 1983. The sediment delivery problem. Journal of<br />

Hydrology 65, 209–237.<br />

C.P. Harden / Geomorphology 79 (2006) 249–263<br />

263<br />

White, S., 1985. Relations of subsistence to the vegetation mosaic of<br />

Vilcabamba, southern Peruvian Andes. Yearbook, Conference of<br />

Latin Americanist Geographers 11, 5–12.<br />

White, S., Maldonado, F., 1991. The use and conservation of natural<br />

resources in the Andes of southern Ecuador. <strong>Mountain</strong> Research<br />

and Development 11 (1), 37–55.<br />

Zevallos, O., 1996. Introducción. In: Zevallos Moreno, O., Fernández,<br />

M.A., Plaza Nieto, G., Klinkicht Sojos, S. (Eds.), Sin Plazo Para la<br />

Esperanza: Reporte sobre el desastre de la Josefina-Ecuador, 1993.<br />

Escuela Politécnica Nacional, Quito, Ecuador, pp. 9–17.<br />

Zevallos, O., Jerves, L., 1996. Efectos de la crecida en el embalse<br />

Amaluza y en el río Paute. In: Zevallos Moreno, O., Fernández,<br />

M.A., Plaza Nieto, G., Klinkicht Sojos, S. (Eds.), Sin Plazo<br />

Para la Esperanza: Reporte sobre el desastre de la Josefina-<br />

Ecuador, 1993. Escuela Politécnica Nacional, Quito, Ecuador,<br />

pp. 199–211.


This article was downloaded by: [<strong>Appalachian</strong> <strong>State</strong> University]<br />

On: 30 August 2012, At: 09:41<br />

Publisher: Routledge<br />

Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered<br />

office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK<br />

Journal of Sustainable Tourism<br />

Publication details, including instructions for authors and<br />

subscription information:<br />

http://www.tandfonline.com/loi/rsus20<br />

Developing sustainable tourism through<br />

adaptive resource management: a case<br />

study of Machu Picchu, Peru<br />

Lincoln R. Larson a & Neelam C. Poudyal a<br />

a Warnell School of Forestry and Natural Resources, University of<br />

Georgia, 180 E. Green Street, Athens, GA, USA<br />

Version of record first published: 19 Mar 2012<br />

To cite this article: Lincoln R. Larson & Neelam C. Poudyal (2012): Developing sustainable tourism<br />

through adaptive resource management: a case study of Machu Picchu, Peru, Journal of Sustainable<br />

Tourism, 20:7, 917-938<br />

To link to this article: http://dx.doi.org/10.1080/09669582.2012.667217<br />

PLEASE SCROLL DOWN FOR ARTICLE<br />

Full terms and conditions of use: http://www.tandfonline.com/page/terms-andconditions<br />

This article may be used for research, teaching, and private study purposes. Any<br />

substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,<br />

systematic supply, or distribution in any form to anyone is expressly forbidden.<br />

The publisher does not give any warranty express or implied or make any representation<br />

that the contents will be complete or accurate or up to date. The accuracy of any<br />

instructions, formulae, and drug doses should be independently verified with primary<br />

sources. The publisher shall not be liable for any loss, actions, claims, proceedings,<br />

demand, or costs or damages whatsoever or howsoever caused arising directly or<br />

indirectly in connection with or arising out of the use of this material.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism<br />

Vol. 20, No. 7, September 2012, 917–938<br />

Developing sustainable tourism through adaptive resource<br />

management: a case study of Machu Picchu, Peru<br />

Lincoln R. Larson ∗ and Neelam C. Poudyal<br />

Warnell School of Forestry and Natural Resources, University of Georgia, 180 E. Green Street,<br />

Athens, GA, USA<br />

(Received 5 August 2011; final version received 6 February 2012)<br />

Machu Picchu, Peru, is recognized as a top international travel destination. Pressure<br />

from the approximately 900,000 tourists who annually visit the ancient Inca city threatens<br />

the ecological integrity, physical substance and cultural authenticity of the World<br />

Heritage Site and surrounding area, including the Inca Trail. Multiple organizations and<br />

agencies currently involved in the management of Machu Picchu have distinct agendas<br />

for the conservation and development of the city, and conflicts regarding public access,<br />

economic growth and cultural preservation are rampant. Attempts to establish carrying<br />

capacities have failed, with proposed daily visitor levels ranging from 800 to 4000. This<br />

paper explores the complex issues surrounding tourism at Machu Picchu and presents a<br />

potential solution: an adaptive management approach based on the UN World Tourism<br />

Organization’s (UNWTO) sustainable tourism framework. This integrative strategy accounts<br />

for multiple perspectives and synthesizes disparate goals embraced by diverse<br />

stakeholders, including the Peruvian government, international conservation organizations,<br />

foreign tourists, private tour operators, regional authorities and indigenous<br />

communities. The focus on Machu Picchu as an adaptive management case study site<br />

outlines key steps leading to implementation, offering planning and policy implications<br />

for sustainability initiatives at numerous developing-world tourism destinations facing<br />

similar political and socio-economic challenges.<br />

Keywords: adaptive management; community development; indicators; Machu Picchu;<br />

sustainability; world heritage site<br />

Introduction<br />

Few places in the world can match the natural beauty and historical significance of Machu<br />

Picchu, Peru. The ecological and cultural allure of the ancient Inca city has earned Machu<br />

Picchu a place on the United Nations Educational, Scientific & Cultural Organization’s<br />

(UNESCO) World Heritage List (UNEP, 2008). Machu Picchu is also formally recognized<br />

as one of the “New Wonders of the World” (World of New7Wonders, 2011), and the image<br />

of Peru’s “Lost City” remains a powerful symbol of Peruvian culture and heritage. With its<br />

dramatic setting and mysterious past, the former Inca citadel has become a popular tourism<br />

destination and the centerpiece of a booming tourism industry (Desforges, 2000).<br />

About 2500 tourists visit Machu Picchu each day, and the remote site is under increasing<br />

pressure from developers and government officials, who want to expand tourism operations<br />

in the area (Leffel, 2005). Threats posed by unregulated use, inadequate planning, deficient<br />

∗ Corresponding author. Email: llarson@uga.edu.<br />

ISSN 0966-9582 print / ISSN 1747-7646 online<br />

C○ 2012 Taylor & Francis<br />

http://dx.doi.org/10.1080/09669582.2012.667217<br />

http://www.tandfonline.com


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

918 L.R. Larson and N.C. Poudyal<br />

monitoring mechanisms and weak policy enforcement have caused Machu Picchu to be<br />

ranked as one of the most rapidly deteriorating World Heritage Sites (Hawkins, Chang, &<br />

Warnes, 2009). Because of these concerns, UNESCO has urged the government of Peru<br />

to revise its Master Plan for managing the Historic Sanctuary to emphasize sustainable<br />

development and prevent Machu Picchu’s possible inscription on the list of World Heritage<br />

Sites in danger (UNESCO, 2009). The state is currently working with UNESCO to construct<br />

a new Master Plan (Vecchio, 2011), but – with many diverse stakeholders and interests to<br />

balance – reaching consensus regarding Machu Picchu’s future has proven to be extremely<br />

difficult.<br />

Machu Picchu’s governing body consists of multiple organizations and agencies – from<br />

local to international – that have very different interests, ranging from preservation to<br />

utilization. Advocates of the mass tourism strategy want to increase access to the site,<br />

generate revenue for regional governments, private operators and local communities, and<br />

promote Inca culture as a marketable commodity. Opponents of mass tourism want to limit<br />

access, preserve ecological, archeological and spiritual assets, and protect existing cultures<br />

and livelihoods in Peru’s <strong>Andean</strong> highlands. David Ugarte, a former regional director of<br />

Cusco’s National Cultural Institute, summed up the core issue: “The (tour) companies are<br />

thinking of profit. Our task is to give to the next generation the opportunity to continue<br />

seeing this wonder for centuries to come ... In ten years’ time there will no longer be a<br />

Machu Picchu. It’s not only part of our heritage. It’s a part of humanity’s” (Collyns, 2007).<br />

Although research suggests that resource protection and development are not mutually<br />

exclusive in the tourism sector (Weaver, 2011), the successful integration of these principles<br />

will likely require a shift from reactive to proactive management paradigms (Allen,<br />

Fontaine, Pope, & Garmenstani, 2011). In this respect, Machu Picchu presents a compelling<br />

opportunity for case study. Tourism management at the site is a classic example of what<br />

McCool and Moisey (2008) call a “messy situation” – a context where goals conflict,<br />

uncertainty abounds and relationships between stakeholders can be polarizing or hostile.<br />

However, if a sustainable solution were to emerge under these challenging circumstances,<br />

it could be used to resolve tourism problems in similar settings around the world. This<br />

paper explores the general concept of sustainable tourism, analyzes the complex issues surrounding<br />

tourism in Machu Picchu and presents a proactive, systematic, objective-driven,<br />

indicator-based adaptive management framework that may facilitate a progression from the<br />

sustainable rhetoric prevalent in management plans to sustainable solutions and action (Zan<br />

& Lusiani, 2011).<br />

What is sustainable tourism?<br />

The issue of sustainable development is at the core of the debate over Machu Picchu’s<br />

use. To ecologists, sustainable development is concerned with preserving the status and<br />

function of ecosystems (Rees, 1990). From the economic standpoint espoused by the World<br />

Commission on Environment and Development, sustainable development is “development<br />

that meets the needs of the present without compromising the ability of future generations<br />

to meet their own needs” (Toman, 1992, p. 3). Within tourism, sustainable development<br />

typically refers to tourism that satisfies the needs of present tourists and host regions while<br />

protecting and enhancing opportunities for the future (Vaughan, 2000).<br />

Multiple meanings have been attached to the term “sustainable” in the tourism context<br />

(Bramwell & Lane, 1993). McCool and Moisey (2008) suggest that sustainable tourism can<br />

refer to a business that perseveres and flourishes over a long period of time or an industry that<br />

acknowledges biophysical and social limits and intentionally remains small in scope. Hunter


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 919<br />

(1995) observes that, in its purest form, sustainable tourism is a vital tool that augments<br />

large-scale economic and social development programs. Butler succinctly summarizes the<br />

adaptive paradigm by stating, “sustainable tourism is that which is developed and maintained<br />

in an area in such manner and at such a scale that ...it remains viable over an infinite period<br />

of time and does not degrade or alter the environment (human and physical) in which it<br />

exists” (1999, p. 12). An effective sustainable tourism approach should maximize benefits<br />

and minimize impacts, thereby increasing the likelihood of long-term persistence. Saarinen<br />

(2006) argues that sustainability in tourism accounts for resource-based (e.g. impacts on<br />

natural and cultural capital), activity-based (e.g. growth and development of industry) and<br />

community-based (e.g. involvement of social capital in a local context) traditions. Each<br />

of these perspectives is relevant at Machu Picchu, where tourism threatens the ecological<br />

integrity and cultural authenticity of a cherished resource while bringing the promise of<br />

economic enhancement and community development. To understand how the sustainable<br />

tourism model may function in this context, the specific situational factors and challenges<br />

that make Machu Picchu unique must be considered.<br />

Machu Picchu: an overview<br />

The ancient citadel of Machu Picchu was built as a royal estate for the Inca ruler Pachacuti<br />

between 1460 and 1470 AD. The citadel remained hidden for centuries until 1911, when<br />

American explorer Hiram Bingham became the first non-native Peruvian to discover the<br />

ruins of the mythical “Lost City”. Bingham immediately understood the magnitude of his<br />

find, noting that “Machu Picchu might prove to be the largest and most important ruin<br />

discovered in South America since the days of the Spanish conquest” (Bingham, 1913).<br />

Bingham’s discovery brought international attention to the ancient city and the awe-inspiring<br />

landscape around it. In 1983, UNESCO officially recognized the cultural and natural value<br />

of the area by designating it a World Heritage Site and making the preservation of Machu<br />

Picchu a global priority (ICOMOS, 1983).<br />

To further protect its national treasure, the Peruvian government created a National Historic<br />

Sanctuary in 1981. Today, Machu Picchu continues to provide evidence and artifacts<br />

that help archeologists reconstruct elements of Inca civilization (Gordon & Knopf, 2007).<br />

The spirit of Machu Picchu has also pervaded the public imagination, fueling a nationalist<br />

movement among Peru’s indigenous people (Flores Ochoa, 2004; van den Berghe & Flores<br />

Ochoa, 2000). Throughout Peru’s turbulent past, Machu Picchu has also remained a pillar<br />

of stability and an emblem of cultural fortitude. Overall, this enduring heritage highlights<br />

a powerful cultural landscape that warrants protection (Alberts & Hazen, 2010).<br />

The Machu Picchu ecosystem contains a variety of habitats and incredible biodiversity.<br />

Its elevation ranges from 1850 to 4600 m and includes dry subtropical forest along the<br />

river valleys, humid montane cloud forests on the steep mountain slopes and high-elevation<br />

paramo grassland (Galiano Sanchez, 2000; UNEP, 2008). Machu Picchu provides refuge<br />

for many wildlife species, including the <strong>Andean</strong> cock-of-the-rock, the ocelot and the endangered<br />

spectacled bear (Young & Leon, 2000), and the discovery of new species in the Machu<br />

Picchu area is not uncommon. These unique ecological features convinced the International<br />

Council on Monuments and Sites (ICOMOS) that Machu Picchu should be expanded into<br />

a biological protected zone for the surrounding areas. Machu Picchu is now recognized as a<br />

“managed resource protection area” by the World Conservation Union (International Union<br />

for Conservation of Nature, IUCN), set aside for the sustainable use of natural ecosystems<br />

and the associated cultural resources (IUCN, 1994). The protected zone currently covers<br />

80,535 acres and reaches far beyond the ruins (Flores Ochoa, 2004).


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

920 L.R. Larson and N.C. Poudyal<br />

Tourism at Machu Picchu<br />

Conservation designations by international organizations underscored Machu Picchu’s universal<br />

appeal, and the tourism industry was quick to respond (Vaughan, 2000). Before<br />

1990, Peru accounted for less than 5% of all tourism in South America (Aguilar, Hinojosa,<br />

Milla, & Nordt, 1992). In the 1990s, with the introduction of improved infrastructure and<br />

reduction in violence and unrest in Peru, its exceptional natural and cultural features became<br />

more appealing to tourists (Casado, 1998; Desforges, 2000). A publicity campaign<br />

with the slogan “El turista es su amigo” (The tourist is your friend) was used to encourage<br />

positive public attitudes toward tourists and promote tourism as the panacea for Peru’s<br />

stagnant economy (Desforges, 2000). These political changes, economic reforms and aggressive<br />

marketing strategies created an international tourist boom in the mid-1990s that<br />

continues today. From 2000 to 2010, international tourist arrivals and expenditures in Peru<br />

doubled (World Economic Forum, 2011) and tourism became the fastest-growing sector of<br />

the Peruvian economy (Mitchell & Eagles, 2001).<br />

Based on the per capita number of tourist bed nights, the Cusco region assumes a dominant<br />

position in Peru’s international tourism hierarchy (O’Hare & Barrett, 1999); Cusco<br />

is the nearest major city to Machu Picchu, with airline and rail connections (Mendoza<br />

Quintana, 1997). In fact, over 90% of all international visits to Peru feature a stop in the<br />

Cusco Department, and nearly half of these trips include a visit to Machu Picchu (Desforges,<br />

2000; Solano, 2005). Cusco hotels and hostels receive an average of 1.5 domestic tourists<br />

to every foreign traveler, much less than the 7:1 domestic to international ratio observed<br />

across Peru (O’Hare & Barrett, 1999). The spatially uneven nature of tourist promotion<br />

in Peru explains a large portion of this discrepancy. Government departments that oversee<br />

the tourism industry are hesitant to endorse travel to outlying areas, preferring instead to<br />

concentrate investments in established hotspots. Although there are 36,000 known archeological<br />

sites in the Cusco region, the potential for tourism in most of these underdeveloped<br />

locations has not been explored (Del-Arroyo, 2005; McGrath, 2004).<br />

Machu Picchu, however, is threatened by its extreme global visibility and a growing<br />

influx of visitors that threatens the limits of sustainability. Between 400 and 3000 people<br />

visit the ancient Inca city every day, and visitor numbers are increasing 6–10% every year<br />

(Emmott, 2003; Leffel, 2005; UNEP, 2008). The record annual high of almost 900,000<br />

visitors was recorded in 2008 and – despite a slight downturn in 2009 and 2010, primarily<br />

due to site closures related to flooding and mudslides (<strong>Andean</strong> Tour Operator, personal<br />

communication, March 21, 2011) – that number is likely to increase (Vecchio, 2011).<br />

Although projections indicate that international tourist numbers across Peru may begin<br />

to stabilize in the near future (Divino & McAleer, 2010), the intensifying pressure on<br />

Machu Picchu itself will not subside (UNEP, 2008). The UNESCO Chief Irina Boklova<br />

acknowledged this alarming pattern in early 2011, remarking that “Machu Picchu is a victim<br />

of its own success” (Vecchio, 2011). If Peru is going to protect one of its most valuable<br />

resources, then management plans must address multiple challenges.<br />

Management challenges at Machu Picchu<br />

Tourism management within and around Machu Picchu is affected by a variety of environmental,<br />

economic and social factors often associated with World Heritage Sites in<br />

developing countries (Regalado-Pezua & Arias-Valencia, 2006; UNEP, 2008). Mitigating<br />

these factors can be a daunting task, but a successful management framework may be able<br />

to address each of the following concerns.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 921<br />

Ecosystem fragility<br />

The Machu Picchu ecosystem is extremely fragile. Nearly 90% of South America’s <strong>Andean</strong><br />

cloud forests have already been lost, and anthropogenic changes have already damaged<br />

cloud forests on mountain slopes within the Historic Sanctuary (Hamilton, 1995). Scientists<br />

with Peru’s Institute of Natural Resources (INRENA) believe that noise pollution<br />

from helicopters and other vehicular traffic led to the disappearance of <strong>Andean</strong> condors<br />

in the region (Collyns, 2006). Current expansion of civilization and tourism infrastructure<br />

also threatens the migration corridors and montane habitats of several endangered<br />

species on the World Conservation Union’s Red List (IUCN, 2011; Peyton, 1980). The<br />

ecological viability and resiliency of Machu Picchu is a global concern. With tourist activity<br />

rising, managers need to act swiftly and decisively to protect the area’s biological<br />

diversity.<br />

The unique topography and geological instability of Machu Picchu add another element<br />

to the management equation. Landslides are common along the valley’s steep slopes and<br />

additional construction of visitor facilities atop Machu Picchu could precipitate a disaster<br />

(Hadfield, 2001; Sassa, Fukuoka, Wang, & Wang, 2005). Scientists report that the eastern<br />

portion of the ancient city is sliding downhill at a rate of 0.4 inches per month, and this<br />

movement could be the precursor stage of a rockslide (Sassa et al., 2005). Prominent<br />

Peruvian archeologist Federico Kauffman believes that in Inca times, no more than 500<br />

small, barefooted people occupied Machu Picchu (LaFranchi, 2001), but modern tourists<br />

– whose behavior is generally much more destructive – often exceed 2000 on a single day.<br />

The UNESCO-supported proposed management plan incorporates a satellite monitoring<br />

system that will track earth movements and visitor activity patterns around the historic<br />

ruins, and UNESCO has also urged managers to develop a thorough risk preparedness plan<br />

for the site (UNESCO, 2006). As tourist numbers increase, managers must devise consistent<br />

and reliable strategies for preventing major site degradation.<br />

Site accessibility<br />

Although tourism in Machu Picchu is increasing, growth in the region has been hindered by<br />

the ruins’ remote location. With relatively limited access, Machu Picchu is a prime example<br />

of how inadequate transportation systems have constrained spatial expansion of the tourist<br />

industry in Peru (O’Hare & Barrett, 1999; World Economic Forum, 2011). About 150 km<br />

of rugged <strong>Andean</strong> highlands separate the Inca citadel from the urban center of Cusco, but<br />

a visit to Machu Picchu is much more than a casual day trip. Most tourists arrive via a<br />

four-hour train ride that carries passengers from Cusco to the town of Aguas Calientes at<br />

the foot of Machu Picchu. From there, buses transport visitors on a 20-minute ride along a<br />

dirt road up to the ruins 350 m above the Urubamba River. The ancient city itself sits on<br />

the saddle (elevation 2430 m), surrounded by the jagged peak of Huayna Picchu (2667 m)<br />

and Machu Picchu <strong>Mountain</strong> (2795 m).<br />

Efforts have been made to increase the accessibility of Peru’s premier tourist destination.<br />

The Peruvian company Inkaterra was recently cleared to open a helicopter service from<br />

Cusco to Aguas Calientes, but the Peruvian Ministry of Transport and Communications<br />

(MTC) reversed the decision after complaints from archeologists and environmentalists<br />

(Collyns, 2006; Higgins, 2006). Another government plan to increase tourism capacity<br />

– a cable car system that would transport visitors from Aguas Calientes to a proposed<br />

tourist village atop the ridge – was indefinitely suspended following public animosity<br />

regarding potential destruction of primary forests and important archeological remains<br />

(Burger & Salazar, 2004). The newly constructed Carilluchayoc Bridge, inaugurated despite


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

922 L.R. Larson and N.C. Poudyal<br />

a prohibitory court order, opened a road corridor between Cusco and the small town of<br />

Santa Teresa. This bridge provides alternative road access to Machu Picchu for visitors who<br />

wish to bypass the expensive tourist train (however, road travelers must still be willing to<br />

endure several treacherous days of travel on perilous unpaved roads).<br />

Limited vehicular access encourages travelers and tourist companies to explore alternative<br />

access to the ancient Inca city. A popular option is the Inca Trail, a stone path built by the<br />

Incas to connect important sites throughout their Sacred Urubamba Valley. Tourists hike the<br />

trail from near Cusco, spending 3–4 days traversing scenic and historic mountainous terrain<br />

before descending to Machu Picchu. They experience the physical demands and the spiritual<br />

nature of the Inca’s ancient lifestyle in an authentic, ecologically friendly way (Arellano,<br />

2004). However, many of these “eco-travelers” are contributing to degradation of a historic<br />

path not built to accommodate such frequent use. In 1984, 6000 tourists hiked the Inca Trail;<br />

by 1998, 66,000 hiked the trail, with over 1500 travelers on the Trail any given day (Roach,<br />

2002). Because of severe damage, and based on UNESCO’s recommendations, a daily limit<br />

of 500 travelers (200 tourists, 300 porters) was implemented in 2001. The entry fee for a full<br />

trail hike was also raised from US $17 to $50. Today, only licensed tour operators are allowed<br />

to sell Inca Trail packages (Barcelona Field Studies Centre, 2007). Stringent regulations<br />

have created a three- to six-month waiting list for tourists hoping to hike the Trail, and many<br />

hikers are now seeking cheaper alternative routes (Healy, 2006). To reduce pressure on the<br />

ruins, UNESCO has discouraged the development of new access routes to Machu Picchu<br />

(Flores Ochoa, 2004). Increased access, however, could boost tourist numbers and increase<br />

foreign expenditure in the region. Hence, the debate over access to Machu Picchu continues.<br />

Local development<br />

The promise of huge profits from increased tourist activity around Machu Picchu could<br />

outweigh the problems of increased access and the related costs of potential site degradation<br />

for many local communities. The Peruvian government formally recognizes the value of<br />

tourism in employment creation and endorses the industry as an important development<br />

strategy (Brohman, 1996; Desforges, 2000). The Sierra Highlands of Peru contain half of<br />

the country’s people but produce just 12.5% of its GNP (O’Hare & Barrett, 1999). Many<br />

small Amerindian villages are scattered throughout <strong>Andean</strong> Peru, and very few of these<br />

communities derive any direct benefits from the tourist boom occurring in the Cusco area.<br />

Aguas Calientes (also known as Machu Picchu Pueblo), the small town at the base of the<br />

ruins, is an exception. The population of Aguas Calientes has grown from 400 to almost<br />

4000 in less than a decade, the fastest rate of population growth in Peru (Emmott, 2003;<br />

UNEP, 2008). This rapid growth is directly related to the international appeal of Machu<br />

Picchu, and Aguas Calientes is earning a reputation as a tourist trap for visitors to the ancient<br />

ruins (Leffel, 2005). Unplanned commercial growth and an overflow of transient settlers<br />

have created other unanticipated problems for the town and its inhabitants. For example,<br />

a lack of adequate water treatment facilities forced the town to pump untreated human<br />

waste into local rivers, polluting the local ecosystem. Seasonal fluctuations in employment,<br />

restricted livelihood choices and a lack of collective identity have also been cited as factors<br />

contributing to social inequities (McGowan, 2010; UNEP, 2008).<br />

The unequal distribution of Machu Picchu tourism profits has done little to help the<br />

plight of most Aguas Calientes residents (<strong>Andean</strong> Tour Operator, personal communication,<br />

March 21, 2011). PeruRail, owned by the British company Orient Express Hotels, has held<br />

a monopoly on transportation in the Sacred Valley for nearly a decade. The company also<br />

owns the only hotel adjacent to the ruins, the Machu Picchu Sanctuary Lodge. Locals argue


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 923<br />

that the foreign company, which carries 92% of the tourists from Cuzco to Machu Picchu,<br />

takes all the money out of the region (Collyns, 2007). Studies of other tourism destinations<br />

in the Peruvian Andes highlight the devastating effect of leakages; in many cases, over 90%<br />

of gross tourism revenues never reach the local community (Bury, 2008; Mitchell & Eagles,<br />

2001). Although tour operators often promote tourism as a sustainable activity, improving<br />

the welfare of local people, in reality, this is rarely a primary goal (Blamey, 1997). To<br />

combat this, UNESCO has advised management bodies to give 10% of ticket receipts from<br />

Machu Picchu to the town of Aguas Calientes (Barcelona Field Studies Centre, 2007).<br />

This new source of money could help transform local infrastructure, but research shows<br />

that high levels of social integration, communication and trust-building between actors in<br />

the tourism industry are necessary to guarantee enduring socio-economic benefits to host<br />

communities (Cole, 2006). For example, tourism’s social ties to the host community are<br />

increased when the industry creates employment opportunities for local residents to serve<br />

as educators or interpretive guides (Jensen, 2010; McGrath, 2004). Increased ownership<br />

and control has been a vital component of sustainable tourism projects in rural <strong>Andean</strong><br />

settings (Mitchell, 2008; Mitchell & Eagles, 2001); comparable approaches could produce<br />

positive results in tourist-busy villages throughout the Machu Picchu area.<br />

Persistence of Peruvian culture<br />

Even if an equitable distribution of tourism revenue in struggling communities can be<br />

achieved, local cultures may still suffer. Many of Peru’s indigenous people, the descendants<br />

of Machu Picchu’s Inca builders, resent the government’s push for more tourist facilities and<br />

greater access to the Cusco region (van den Berghe & Flores Ochoa, 2000). The <strong>Andean</strong><br />

people, proud of their cultural heritage, are concerned that international tourism threatens<br />

the sanctity of their sacred sites. A Peruvian movement known as incanismo has responded<br />

to these concerns, sparking new controversy over Machu Picchu’s use. Incanismo extols<br />

the virtues of Inca civilization and vilifies Europeans as the scourge of the Americas (van<br />

den Berghe & Flores Ochoa, 2000). On a local level, incansimo principles advocate the<br />

preservation of historical treasures, such as Machu Picchu, for traditional purposes. For<br />

example, the residents of Aguas Calientes, hoping to regain ownership of Machu Picchu and<br />

repossess treasures hailed as part of their “cultural identity”, recently asked Yale University<br />

to return Inca artifacts taken by explorer Hiram Bingham almost a century ago (Cornwell,<br />

2006). Yale began returning the artifacts in 2011, and they will eventually be displayed at<br />

a museum in Cusco (Regalado, 2011).<br />

On a larger scale, the exploitation of incanismo ideology contributes to a novel form<br />

of cultural degradation – ethnic tourism. Modern tourism packages in the Cusco region are<br />

all linked in some way to the Inca theme, and foreigners embracing the heritage tourism<br />

experience are eager to accept embellishments of Inca culture propagated by their guides<br />

(McGrath, 2004; van den Berghe & Flores Ochoa, 2000). Tour operators, particularly<br />

members of the Cusco urban elite, have therefore been able to capitalize on incanismo as<br />

a marketable tourist commodity. The phenomenon of “staged authenticity”, where native<br />

people adopt a contrived culture to appeal to tourist interests, is slowly pervading the<br />

Peruvian highlands (MacCannell, 1973). The subsequent re-invention of tradition and<br />

cultural change precipitated by the tourism-mediated commercialization of Inca culture has<br />

transformed many aspects of life in the Machu Picchu area (Cohen, 1988).<br />

Rising entrance fees at Machu Picchu represent another obstacle threatening to diminish<br />

the site’s importance to local residents. Entrance fees have been raised several times in the<br />

past 10 years, from US $10 to $45 for international tourists and approximately half that


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

924 L.R. Larson and N.C. Poudyal<br />

price for Peruvian residents – with occasional free days for locals (<strong>Andean</strong> Travel Web,<br />

2011; Barcelona Field Studies Centre, 2007). Though the price seems to not discourage<br />

international travelers, Amerindians who wish to visit the site for spiritual or cultural<br />

purposes can rarely afford access and are often displaced by large-scale tourism activities.<br />

Such conflicts between local interests and tourism-associated demands are not unique to<br />

Machu Picchu (e.g. Rugendyke & Son, 2005). The ICOMOS highlights the difficulties in<br />

integrating cultural resource management and tourism in its International Cultural Tourism<br />

Charter, asserting that physical, intellectual and emotive access to cultural heritage sites is<br />

a right that should not be denied (ICOMOS, 1999).<br />

Many locals believe the cultural “existence value” of a place like Machu Picchu cannot<br />

be expressed in economic terms or exchanges (Navrud & Ready, 2002), and some Peruvian<br />

residents that oppose foreign tourism are fighting to preserve their authentic cultural<br />

heritage. In 1999 and 2000, Cusqueños conducted a “March to Machu Picchu” denouncing<br />

government management policies in the Historic Sanctuary (Flores Ochoa, 2004). Yachay<br />

Wasi, a Cusco-based nongovernmental organization that works on cultural issues and sustainable<br />

development to benefit indigenous people, urges the world to critically analyze<br />

the consequences of mass tourism. At a 2006 United Nations forum, Yachay Wasi issued<br />

its Inka Challenge: “Will world governments, scientists, nonprofit sponsors and tourists<br />

respect indigenous people’s spiritual heritage, religion, burial sites, and human remains,<br />

and will the international community respect and allow them to protect their sacred sites”<br />

(Yachay Wasi, 2006). Concerns of indigenous people are now an integral part of UN agendas,<br />

and efforts to support local pride and regional identity must become an important part<br />

of management plans in places like Machu Picchu.<br />

Institutional complexity<br />

To compound the problems already facing the site, the Management Unit of Machu Picchu<br />

– the entity charged with carrying out Machu Picchu’s Master Plan – is composed of<br />

many different agencies. A recent restructuring of the Peruvian government has further<br />

complicated matters. Each of the disparate bodies that govern Machu Picchu has a distinct<br />

agenda, and each considers management options that balance public access, economic<br />

opportunity, and cultural and biological preservation in the region to different degrees.<br />

Thus, Peru’s National Institute of Culture (INC, now the Ministry of Culture) is charged<br />

with preserving the country’s national heritage and manages the cultural and historic aspects<br />

of the site, including the actual ruins. The INRENA (now the National Service of Protected<br />

Areas – or SERNANP - in the Ministry of the Environment) is responsible for the flora<br />

and fauna in the Historic Sanctuary. The Ministry of Industry, Tourism, Integration and<br />

International Trade Negotiations (MITINCI, now the Ministry of Foreign Commerce and<br />

Tourism – or MINCETUR) regulates tourism and development in the area, and the state’s<br />

Tourism Promotion Commission (PromPeru) actively markets the site to potential visitors.<br />

All activity in and around Machu Picchu is also overseen and monitored by two international<br />

organizations, UNESCO and IUCN. A heavy emphasis on centralized decision-making<br />

orchestrated by powerful government elites, not governance that involves a full range<br />

of invested individuals and organizations, has been a major constraint for developing<br />

countries trying to promote community participation in the tourism industry (Plummer<br />

& Fennell, 2009; Tosun, 2000). At Machu Picchu, regional and local authorities have a<br />

small but growing voice in management decisions, and efforts are underway to expand<br />

a management committee that incorporates public and private stakeholders not currently<br />

represented (PromPeru, personal communication, July 22, 2011). Although Machu Picchu’s


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 925<br />

managers agree on the general goal of sustainable development, the current Master Plan<br />

does not adequately describe how this philosophy should dictate management strategies<br />

or who should be responsible for implementing them (UNESCO, 2011). The absence<br />

of an effective autonomous collaborative committee to create and implement directives<br />

and dictate the future course of Machu Picchu represents a major management problem<br />

(Regalado-Pezua & Arias-Valencia, 2006).<br />

Sustainable tourism priorities at Machu Picchu<br />

To maintain the value of the site and to mollify some of the challenges described above,<br />

most management-oriented documents at Machu Picchu have focused on a central question<br />

– the establishment and regulation of an appropriate carrying capacity. The concept of<br />

carrying capacity has been applied in a variety of settings and generally describes the level<br />

of visitor use that can be appropriately accommodated in a site without altering the physical<br />

environment or the overall visitor experience (Manning, 1999; McCool & Lime, 2001). At<br />

Machu Picchu, tourism is pushing the limits on both fronts. Excessive use has destroyed<br />

important archeological remains and overcrowding has negatively affected the aesthetic<br />

enjoyment, historical immersion, imagination and solitude that appeal to many visitors<br />

(Emmott, 2003). Larger crowds may impact the international public’s perception of Machu<br />

Picchu as a must-see tourist attraction and detract from its value as a World Heritage Site.<br />

Although Machu Picchu’s managers and operators acknowledge the importance of<br />

carrying capacity studies for the preservation of their site, they cannot agree on specific<br />

numbers. UNESCO, a conservation-minded organization, suggests that tourist numbers<br />

should be cut to 800 per day and visitors should wear soft shoes to reduce pressure on the<br />

ruins (Barcelona Field Studies Centre, 2007). Agencies hoping to increase tourism revenue<br />

in the region believe that the estimate is far too conservative. Peru’s INC, which oversees<br />

day-to-day running of Machu Picchu, claims that the site can cope with 3000 tourists per day<br />

(Emmott, 2003). Orient Express Hotels, the private British company that runs the Machu<br />

Picchu Sanctuary Lodge and the PeruRail tourist train from Cuzco, believes that the site<br />

can easily sustain more than 4000 daily visitors (Collyns, 2007). Seeking a compromise,<br />

Peru’s Regional Office of Culture tentatively accepted a management plan (effective 15 July<br />

2011) created with input from UNESCO that capped daily entries to Machu Picchu at 2500<br />

visitors, with further restrictions for specific locations within the Sanctuary (PromPeru,<br />

personal communication, July 22, 2011). However, enduring consensus in Machu Picchu’s<br />

carrying capacity debate is unlikely given the conflicting priorities of the site’s complex<br />

management conglomerate. In fact, research in multiple tourism settings has shown that<br />

attempts to define optimal carrying capacity in complex real-world situations are generally<br />

futile (McCool & Lime, 2001). In many cases, the failure of management plans is due<br />

in part to an unbalanced emphasis on overall carrying capacity and a lack of specificity<br />

regarding management goals and objectives (UNEP, 2008).<br />

The first step toward a sustainable future for tourism at Machu Picchu therefore involves<br />

the identification of appropriate management priorities. Regional authorities generally place<br />

a premium on tourism promotion and economic expansion, while international agencies<br />

often focus on conservation and preservation. Integration of these concepts is central to<br />

successful management in parks where multiple issues influence management decisions<br />

(McCool & Moisey, 2008). Therefore, all parties with a vested interest in tourism need<br />

to come together and engage in participatory planning focused on unified goals (Mitchell<br />

& Eagles, 2001; Regalado-Pezua & Arias-Valencia, 2006). At Machu Picchu, these goals<br />

can be expressed in a pyramid of priorities (Figure 1). Site management should begin with


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

926 L.R. Larson and N.C. Poudyal<br />

Sustainable Tourism:<br />

A Pyramid of Priorities<br />

Sustainable Use<br />

(Maintain resource value)<br />

Social Viability<br />

(Stakeholder satisfaction, collaborative planning)<br />

Resource Utility<br />

(Support local development, increase economic benefits)<br />

Resource Integrity<br />

(Protect biological diversity, prevent site degradation, preserve cultural heritage)<br />

Figure 1. Pyramid of priorities to guide sustainable tourism management.<br />

foundational efforts to preserve resource integrity, capturing the fundamental essence of<br />

the site and its unique “spirit of place” (Shackley, 2006). Once the protection of basic assets<br />

is secured, resource utility leading to social viability becomes the primary focus. This<br />

synergistic, hierarchical network of factors results in sustainable resource use, generating<br />

a positive feedback loop that theoretically persists in perpetuity. Although the pyramid<br />

of priorities concept may serve as a valuable point of origin for planning efforts, actual<br />

management requires far more explicit goals and actions. Successful growth of tourism in<br />

Machu Picchu and other World Heritage Sites may ultimately depend on an objective-driven,<br />

indicator-based adaptive management framework.<br />

Sustainable tourism: an adaptive resource management framework<br />

Defining management objectives and actions<br />

A major criticism of Machu Picchu’s existing Master Plan, as well as management guidelines<br />

for international tourist destinations in other developing countries, has been the ambiguity<br />

of goals and strategies and a conspicuous absence of detail for possible actions (Schianetz<br />

& Kavanagh, 2008; UNEP, 2004; Zan & Lusiani, 2011). This criticism could be resolved<br />

using an adaptive resource management (ARM) framework, which applies knowledge from<br />

related disciplines to address contemporary tourism issues (Farrell & Twining-Ward, 2004).<br />

Variations of the ARM approach to informed decision-making have been applied in<br />

a variety of settings. Early iterations still in use today include the Limits of Acceptable<br />

Change (LAC; Stankey, Cole, Lucas, Peterson, & Frissell, 1985) and the Visitor Impact<br />

Management Model (VIM; Graefe, Kuss, & Vaske, 1990), both of which aim to set limits<br />

and minimize negative impacts from recreation and tourism on public lands in the United<br />

<strong>State</strong>s. Newer strategies such as the Tourism Optimization Management Model (TOMM;<br />

Miller & Twining-Ward, 2005; Twining-Ward & Butler, 2002) and the Integrated Monitoring<br />

and Adaptive Management System (iMAMS; QStation, 2009) have been used by<br />

Australian tourism operators to assess, monitor and successfully achieve sustainable outcomes.<br />

Although research has identified several distinct decision-making methods used<br />

in the adaptive management process, many similarities exist (McFadden, Hiller, & Tyre,<br />

2011). Each framework relies on specific objectives or standards, associated indicators


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 927<br />

that facilitate monitoring of these objectives and corresponding actions that help to reduce<br />

uncertainty associated with decision outcomes and ensure that desired objectives are being<br />

met (Allen et al., 2011; McCool & Lime, 2001).<br />

An effective ARM approach consists of several basic steps (Knutson et al. 2011; Miller<br />

& Twining-Ward, 2005). First, managers must define the problem and a corresponding key<br />

objective that represents a desired outcome. At Machu Picchu, the fundamental objective<br />

might be to maintain the site’s value as a unique natural, cultural and economic resource.<br />

Next, managers must identify a set of means objectives that support the fundamental<br />

objective. At Machu Picchu, these objectives might include simultaneously minimizing<br />

the impacts and maximizing the benefits of tourism. Other targets or means subobjectives<br />

could then be constructed, creating a transparent network structure to help guide the<br />

decision-making process. For example, means subobjectives under the broader category of<br />

“minimizing the impacts of tourism” might address the challenge of ecosystem fragility,<br />

whereas means subobjectives under the “maximizing the benefits of tourism” could focus<br />

on issues such as site accessibility and community development. Each means subobjective<br />

is associated with a corresponding set of management actions necessary to achieve the<br />

desired goal (Figure 2). In this proposed framework, objectives are designed to integrate<br />

stakeholder concerns and specifically address the major challenges facing Machu Picchu<br />

from a social–ecological perspective that incorporates human, natural and support systems<br />

(Schianetz & Kavanagh, 2008). The traditionally narrow focus of previous efforts<br />

to define tourism goals has restricted progress in other protected areas. By incorporating<br />

noneconomic factors and simultaneously balancing costs and benefits, the ARM objectives<br />

outlined here provide a solid foundation for long-term success (Moscardo, 2011).<br />

The input of all groups, including local residents, tourists, tour operators and site managers,<br />

is a critical component of objective and action specification (Plummer & Fennell,<br />

2009; Stronza, 2001). A meta-analysis of international adaptive management supported this<br />

assertion, revealing that decision-theoretic approaches emphasizing stakeholder communication<br />

early in the process typically resulted in less complex models with greater efficacy<br />

addressing specific decision problems (McFadden et al., 2011). Adaptive management attempts<br />

that experience only limited success are often generic top-down systems focused<br />

around expert opinion, not place-based frameworks guided by local knowledge and concerns<br />

(Miller & Twining-Ward, 2006; Schianetz & Kavanagh, 2008). For example, a study<br />

of a rural Bolivian ecotourism destination demonstrated the importance of context-specific<br />

outcomes and the invaluable role of local contributions in the accomplishment of long-term<br />

goals (Jamal & Stronza, 2009). For ARM to function properly at Machu Picchu, local<br />

stakeholders should be involved throughout the constantly evolving planning process. This<br />

intentional inclusion of local input could help to resolve two of the major challenges facing<br />

Machu Picchu: local development and the persistence of Peruvian culture.<br />

Selecting appropriate indicators<br />

Although the identification of explicit management outcomes is important, goals and objectives<br />

alone are insufficient. A set of measured attributes must also be included to monitor<br />

progress and ensure that action goals are met. Recognizing the value of this approach,<br />

the UN World Tourism Organization (UNWTO) has created a guide for the selection of<br />

performance indicators that measure the effects of tourism on the environment and poverty<br />

alleviation in the developing world (Miller & Twining-Ward, 2005; WTO, 2004). The nature<br />

of these sustainability indicators varies from site to site, but core indicators in the<br />

UNWTO framework generally focus on aspects including critical ecosystems, maintenance


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

928 L.R. Larson and N.C. Poudyal<br />

Figure 2. Proposed adaptive management framework for Historic Sanctuary of Machu Picchu.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 929<br />

of natural capital stock, long-term use intensity, local involvement, well-developed plans,<br />

resident/customer satisfaction and tourism’s contribution to the economy (Miller, 2001;<br />

WTO, 1996). Research indicates that tourism experts around the world believe both objective<br />

(quantitative) and subjective (qualitative) metrics can provide important information<br />

(Choi & Sirakaya, 2006; Miller, 2001). Quantitative indicators often include raw data, ratios<br />

and percentages; qualitative indicators might incorporate categorical indices, and normative<br />

or nominal information associated with a resource (WTO, 2004). For indicators to function<br />

properly, they should be condensed into a concise set carefully selected by integrated, multidisciplinary<br />

advisory panels composed of expert and nonexpert stakeholders (Schianetz &<br />

Kavanagh, 2008). The indicators should also be subjected to a systematic screening process<br />

to identify limitations and methodological challenges (Miller & Twining-Ward, 2005).<br />

Overall, effective indicators are relevant, reliable, feasible and stable over an extended<br />

period of time (QStation, 2009; Twining-Ward & Butler, 2002).<br />

At Machu Picchu, a diverse suite of quantitative and qualitative indicators based on<br />

the UNWTO’s framework could measure and monitor progress and direct actions related<br />

to target objectives (Table 1). Some of these indicators include information that is already<br />

available through standard surveillance monitoring; others require additional research and<br />

data collection. For example, the role of tourism in community development could be<br />

tracked through quantitative evaluations of tourism integration (e.g. percentage of guides<br />

that are locals or percentage of hotels operated by locals) or qualitative assessments of<br />

community involvement (e.g. stakeholder ratings of perceived collaboration in the tourism<br />

planning process). Similarly, the economic benefits of tourism could be assessed by quantitative<br />

methods (e.g. daily tourism revenues) or qualitative approaches (e.g. stakeholder<br />

ratings of the distributional equity of tourism profits). Machu Picchu’s Management<br />

Committee has already identified many potential indicators in the site’s comprehensive<br />

234-page Master Plan (INC, 2005). However, the Committee has yet to devise a consistent<br />

strategy for implementing and monitoring these performance indicators (UNESCO, 2011;<br />

Zan & Lusiani, 2011). This stage of adaptive management – the monitoring phase – is<br />

where many sustainable tourism projects break down (Twining-Ward & Butler, 2002).<br />

Monitoring progress<br />

Monitoring is more than a stand-alone activity used to detect trends. In the ARM framework,<br />

monitoring can help managers determine which management practices meet specified<br />

objectives and develop flexible strategies for resolving recurring problems (Knutson et<br />

al., 2011; Plummer & Fennell, 2009). In practice, the major limitation of the UNWTO’s<br />

approach to indicators has been a heavy focus on the development of indicators with little<br />

emphasis on their actual implementation (Miller & Twining-Ward, 2006).<br />

For monitoring in ARM to be beneficial, it must be an iterative component of<br />

management-based science (Nichols & Williams, 2006). Managers should initiate a particular<br />

management option, evaluate its impact and learn from the results – adjusting management<br />

strategies as understanding improves (Williams, 2011a; Figure 3). This learning can<br />

occur “actively” through the deliberate reduction of uncertainty associated with particular<br />

outcomes, often using an experimental approach involving different actions on multiple<br />

units simultaneously. Alternatively, the learning can occur “passively” as a byproduct of<br />

systems that focus on changes in resource conditions with respect to desired objectives<br />

using one model at a time (Williams, 2011b). The passive approach is probably more feasible<br />

in the nascent stages of ARM at Machu Picchu, where managers are under pressure<br />

to implement actions that generate an immediate response (e.g. temporary site closures,


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

930 L.R. Larson and N.C. Poudyal<br />

Table 1. Suggested indicators for specified objectives in adaptive management framework at Machu<br />

Picchu a .<br />

Objectives (with corresponding<br />

actions) Potential objective indicators b Potential subjective indicators b<br />

Protect biological diversity<br />

Maximize amount of protected<br />

habitat<br />

Increase populations of T&E<br />

species<br />

Minimize land use change and<br />

fragmentation in Urubamba<br />

Valley<br />

Prevent physical degradation<br />

Minimize erosion and<br />

landslide potential<br />

Minimize and properly<br />

dispose off waste<br />

Preserve cultural heritage<br />

Minimize damage to historic<br />

Inca structures<br />

Increase site accessibility for<br />

spiritual purposes<br />

Minimize exploitation of<br />

ethnic tourism<br />

Minimize socio-cultural<br />

impacts of tourism-related<br />

activities<br />

Support local development<br />

Increase local involvement in<br />

tourism industry<br />

Improve access and local<br />

infrastructure<br />

Increase educational<br />

opportunities<br />

No. of local species<br />

downgraded on IUCN Red<br />

List/yr<br />

Amount of habitat restored<br />

for T&E species<br />

Size/viability of threatened<br />

populations<br />

Percent change in interior<br />

forest area and edge/yr.<br />

Population density within 10<br />

km of site<br />

No. of people and visitor<br />

density across<br />

spatial/temporal scales<br />

Soil loss in valley and in<br />

ruins/yr<br />

Total weight of waste<br />

generated/month<br />

Percent wastewater treated<br />

before disposal<br />

No. of historic structures<br />

damaged/yr<br />

Percent revenue spent on<br />

structural renovation/yr<br />

No. of people attending<br />

cultural events at site/yr<br />

No. of interpretive signs<br />

highlighting Inca heritage<br />

No. of locals visiting site/yr<br />

No. of crimes committed in<br />

region/yr<br />

Site protection priority ratings<br />

(IUCN conventions)<br />

Site resiliency and habitat<br />

integrity ratings (experts)<br />

Site susceptibility to climate<br />

change ratings (experts)<br />

Landslide potential ratings<br />

(experts)<br />

Ratings of waste management<br />

procedures<br />

Perceived impacts of waste in<br />

local communities<br />

General appearance of site<br />

ratings<br />

Perceived authenticity of<br />

interpretive efforts<br />

Tour guide knowledge and<br />

performance ratings<br />

Ratings of perceived cultural<br />

degradation among locals<br />

Opinions toward current tourism<br />

practices<br />

Percentage of guides at site Perceived collaboration in<br />

that are locals<br />

planning process<br />

Percentage of hotels operated Rankings for levels of planned<br />

by locals<br />

development control<br />

No. of locals employed in Regulatory framework efficacy<br />

tourism industry<br />

ratings<br />

Ratio of foreign tourists to Business environment and<br />

local residents<br />

infrastructure ratings (locals<br />

No. of visitors/week reaching and managers)<br />

site from various access Perceivedcontributionof<br />

points (road, train, etc.) tourism to local development<br />

No. of social services<br />

(including education<br />

programs) available to<br />

locals<br />

projects<br />

(Continued on next page)


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 931<br />

Table 1. Suggested indicators for specified objectives in adaptive management framework at Machu<br />

Picchu a . (Continued)<br />

Objectives (with corresponding<br />

actions) Potential objective indicators b Potential subjective indicators b<br />

Increase economic benefits<br />

Increase amount of foreign No. of tourists visiting site per Prioritization of tourism<br />

expenditures<br />

day, week, etc.<br />

rankings on multiple scales<br />

Increase flow of tourism Tourism revenue per day, Tourism marketing materials<br />

profits to local communities week,etc.(+ leakage) efficacy ratings<br />

(reduce leakage)<br />

No. of agencies/operators Distributional efficiency of<br />

Enhance marketing/promotion using site<br />

tourism profit ratings<br />

of tourism<br />

No. of jobs added by tourism Perceived impact of imported<br />

Ensure stakeholder satisfaction<br />

sector<br />

Percent economy based on<br />

tourism (local, regional,<br />

national)<br />

Indirect/direct economic<br />

impacts of tourism in the<br />

area<br />

tourism goods and services<br />

Maximize visitors’ satisfaction No. of conflicts/confrontations Satisfaction level ratings for<br />

Maximize locals’ satisfaction between tourists and locals, visitors and managers<br />

Maximize managers’<br />

locals/yr.<br />

Visitor perceptions of trip value<br />

satisfaction<br />

No. of protests/complaints given investment<br />

filed against<br />

Perceived crowding<br />

management/yr.<br />

Sustainability ratings for<br />

No. of repeat visitors/yr. operations<br />

a Full adaptive management framework is depicted in Figure 2.<br />

b Potential indicators are based on previous research and guidelines specified by the World Tourism Organization<br />

(WTO, 2004). The Machu Picchu Management Committee could adapt this framework to provide more specific,<br />

measurable and verifiable indicators.<br />

regulating visitor numbers, altered pricing schemes, restorations, education campaigns). As<br />

information is gathered, comparisons of desired outcomes and observed responses of the<br />

selected indicators facilitate a movement toward more promising management alternatives<br />

(Miller & Twining-Ward, 2005). Tourism managers at North Head Quarantine Station in<br />

New South Wales have already put these principles into practice (QStation, 2009). Using<br />

their iMAMS integrated monitoring approach, the QStation team developed a sustainability<br />

index to determine the percentage of environmental, cultural, social and economic indicators<br />

performing within their accepted range. Adaptive management responses are initiated<br />

when index scores are unsatisfactory.<br />

Although proposed solutions to complex management problems can be difficult to<br />

identify, ARM is specifically designed to deal with complicated circumstances. The ARM<br />

approach functions best in situations where controllability is high (i.e. management has the<br />

ability to affect resources) and uncertainty abounds (i.e. responses to management actions<br />

may vary), making it an ideal fit for remote World Heritage Sites like Machu Picchu (Allen<br />

et al., 2011). In summary, the constantly evolving implementation of ARM involves goals<br />

and objectives that are used to specify desired outcomes, indicators that serve as metrics<br />

for measuring the success of these outcomes, and monitoring that provides a mechanism to<br />

determine if these outcomes are being met. These interrelationships highlight the iterative,


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

932 L.R. Larson and N.C. Poudyal<br />

Continue<br />

monitoring<br />

Set objectives/<br />

Define community values<br />

Develop appropriate indicators<br />

Develop a monitoring program/<br />

Inventory conditions<br />

Collect information/<br />

Are objectives being met?<br />

Yes No<br />

Initiate a management<br />

response<br />

Monitor<br />

response<br />

Figure 3. Schematic management diagram demonstrating adaptive management monitoring principles<br />

(adapted from Hammitt & Cole, 1998; Miller & Twining-Ward, 2005). Adapted version reprinted<br />

with permission of the authors and John Wiley & Sons, Inc.<br />

cyclical nature of an effectively flexible implementation of ARM and help explain why<br />

adoption of ARM may be critical to promoting sustainable tourism at Machu Picchu.<br />

Implementing ARM at Machu Picchu<br />

The Historic Sanctuary of Machu Picchu already has many of the essential ingredients for<br />

ARM in place. The site’s Master Plan, backed by UNESCO, encompasses a comprehensive<br />

resource assessment and outlines management recommendations (INC, 2005). The Plan<br />

would provide a valuable starting point for conversations among stakeholders focused<br />

on objectives, potential actions and corresponding indicators (Miller & Twining-Ward,<br />

2005). The Machu Picchu Management Committee, composed of local to international<br />

agencies, has already expressed a desire to expand and incorporate a broader range of<br />

actors in the public and private sectors (PromPeru, personal communication, July 22,<br />

2011). If open lines of communication are established and the ARM process is initiated,<br />

Machu Picchu’s managers may finally be able to successfully overcome the obstacle of<br />

institutional complexity with a long-term management framework that is flexible and open<br />

to participatory decision-making (Williams, 2011a). Presumably, this cooperative approach<br />

would generate greater social and financial support across multiple scales.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 933<br />

Other challenges remain, however. The Management Committee must determine which<br />

agencies and entities are responsible for implementing specified actions and monitoring<br />

progress. These agencies must be individually accountable for certain aspects of the site<br />

(e.g. protected habitat, historic structures, tourism profits, visitor satisfaction), and they<br />

must be collectively devoted to the fundamental objective of resource protection. Managers<br />

must simultaneously recognize the local value and global importance of Machu Picchu,<br />

balancing conservation-oriented edicts from organizations such as UNESCO with regional<br />

economic growth (Saarinen, 2006). A stakeholder meeting would set the ARM process in<br />

motion (Miller & Twining-Ward, 2005), extending the framework described here to create<br />

a more explicit action plan guided by site-specific knowledge of management objectives<br />

and measurable indicators of immediate, mid-term and long-term utility. Ultimately, a<br />

collaborative ARM approach should help to systematically resolve some of the challenges<br />

that impede the development of sustainable tourism at Machu Picchu, helping Peru to<br />

ensure that its “Lost City” is preserved in perpetuity.<br />

Conclusion<br />

Sustainable tourism resource management is an elusive goal in most developing countries,<br />

where ecological and cultural heritage is often sacrificed in pursuit of economic wellbeing<br />

(Keatinge, 1982). Peru’s economic growth and aggressive promotion of tourism have placed<br />

severe demographic pressure on its most valuable tourism asset, Machu Picchu, creating<br />

divergent opinions over management priorities. This conflict of interest is magnified by the<br />

site’s unparalleled combination of biological and archeological resources, fragility, remote<br />

location, extreme local poverty, emerging cultural tensions and institutional complexity. Existing<br />

management plans have provided few answers, generally exacerbating disagreement<br />

among stakeholders and pushing the groups charged with protecting Machu Picchu to the<br />

breaking point (Regalado-Pezua & Arias-Valencia, 2006; Zan & Lusiani, 2011). The future<br />

of the ancient Inca city depends on a delicate balance between preservation, utilization and<br />

sustainable growth.<br />

This paper suggests that an adaptive resource management approach may help planners<br />

and managers guide Machu Picchu’s growth. The ARM tourism framework would<br />

help to identify management priorities, facilitating the creation of cohesive goals and objectives<br />

among the various agencies responsible for conservation and development in the<br />

Historic Sanctuary. By individually monitoring specific indicators of quality across various<br />

spatial and temporal scales, managers could potentially address multiple management considerations<br />

that affect local residents, foreign tourists, private tour operators and regional<br />

governments. Furthermore, managers could foster resilience and flexibility by learning<br />

from inevitable mistakes and surprising responses, adjusting actions to better meet specified<br />

goals (Allen et al., 2011; QStation, 2009). Implementation of ARM would likely<br />

require substantial international investment, but a global commitment may be necessary<br />

for long-term conservation and appreciation of Machu Picchu and other premier World<br />

Heritage Sites (Saarinen, 2006). As more unique places around the world begin to feel<br />

the impending pressure of increased visitation (Weaver, 2011), the ARM approach may<br />

be necessary to support sustainable development that addresses environmental, economic<br />

and social challenges. If ARM is adopted successfully at Machu Picchu, then lessons<br />

learned from this sustainability framework could inform tourism practices at other heritage<br />

destinations surrounded by complex political circumstances and socioeconomic contexts.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

934 L.R. Larson and N.C. Poudyal<br />

Acknowledgements<br />

The authors wish to thank the anonymous <strong>Andean</strong> tour operators who supplied first-hand accounts of<br />

the current tourism situation at Machu Picchu. The authors also wish to thank PromPeru (La Comisión<br />

de Promoción del Perú para la Exportación y el Turismo) for providing updated information about<br />

the current state of management efforts within the Historic Sanctuary.<br />

Notes on contributors<br />

Lincoln R. Larson is a graduate student in the Natural Resource Recreation and Tourism program at the<br />

University of Georgia’s Warnell School of Forestry and Natural Resources, USA. His interdisciplinary<br />

research focuses on a range of topics, including outdoor recreation, environmental education and<br />

human dimensions of conservation. Lincoln’s interest in sustainable tourism stems from his work on<br />

ecotourism projects in Peru.<br />

Neelam C. Poudyal is an Assistant Professor of Natural Resource Recreation and Tourism at the<br />

Warnell School of Forestry and Natural Resources, University of Georgia, USA. He teaches courses<br />

on Ecotourism and Sustainable Development and Recreation Resource Management, and the main<br />

themes of his research program include the human dimensions and economic analysis of natural<br />

resource recreation and tourism in the United <strong>State</strong>s and beyond.<br />

References<br />

Aguilar, V., Hinojosa, L., Milla, C., & Nordt, W. (1992). Turismo y desarrollo: Posibilidades en la<br />

region Inka. Cuzco, Peru: CARTUC y CERA.<br />

Alberts, H.C., & Hazen, H.D. (2010). Maintaining authenticity and integrity at cultural world heritage<br />

sites. The Geographical Review, 100(1), 56–73.<br />

Allen, C.R., Fontaine, J.J., Pope, K.L., & Garmenstani, A.S. (2011). Adaptive management for a<br />

turbulent future. Journal of Environmental Management, 92, 1339–1345.<br />

<strong>Andean</strong> Travel Web. (2011). Machu Picchu tourist information. Retrieved April 13, 2011, from<br />

http://www.andeantravelweb.com/peru/destinations/machupicchu/index.html<br />

Arellano, A. (2004, January). Mobility systems and performance at Machu Picchu. Paper presented<br />

at the Alternative Mobility Futures Conference, Lancaster University, Lancaster, UK.<br />

Barcelona Field Studies Centre. (2007). Machu Picchu: Attitudes to management. Retrieved March<br />

8, 2007, from http://geographyfieldwork.com/MachuTourismAttitudes.htm<br />

Bingham, H. (1913). In the wonderland of Peru. National Geographic, 24(4), 387–573. Retrieved<br />

from http://ngm.nationalgeographic.com/1913/04/machu-picchu/bingham-text<br />

Blamey, R.K. (1997). Ecotourism: The search for an operational definition. Journal of Sustainable<br />

Tourism, 5(2), 109–130.<br />

Bramwell, B., & Lane, B. (1993). Sustainable tourism: An evolving global approach. Journal of<br />

Sustainable Tourism, 1(1), 1–5.<br />

Brohman, J. (1996). New directions in tourism for Third World development. Annals of Tourism<br />

Research, 23, 48–70.<br />

Burger, R.L., & Salazar, L.C. (Eds.). (2004). Machu Picchu: Unveiling the mystery of the Incas.New<br />

Haven, CT: Yale University Press.<br />

Bury, J. (2008). New geographies of tourism in Peru: Nature-based tourism and conservation in the<br />

Cordillera Huayhuash. Tourism Geographies, 10(3), 312–333.<br />

Butler, R.W. (1999). Sustainable tourism: A state-of-the-art review. Tourism Geographies, 1(1), 7–25.<br />

Casado, M.A. (1998). Peru’s tourism industry: Growth, setbacks, threats. Cornell Quarterly: Hotel<br />

Restaurant and Administration, 39(1), 68–73.<br />

Choi, H.C., & Sirakaya, E. (2006). Sustainability indicators for managing community tourism.<br />

Tourism Management, 27, 1274–1289.<br />

Cohen, E. (1988). Authenticity and commoditization in tourism. Annals of Tourism Research, 15,<br />

371–386.<br />

Cole, S. (2006). Information and empowerment: The keys to achieving sustainable tourism. Journal<br />

of Sustainable Tourism, 14(6), 629–644.<br />

Collyns, D. (2006, September 8). Peru bans flight over Inca ruins. BBC News. Retrieved from<br />

http://news.bbc.co.uk/2/hi/americas/5326042.stm<br />

Collyns, D. (2007, February 1). Bridge stirs the waters in Machu Picchu. BBC News. Retrieved from<br />

http://news.bbc.co.uk/2/hi/americas/6292327.stm


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 935<br />

Cornwell, R. (2006, February 3). Peru tells Yale it wants its Machu Picchu treasures back (after 100<br />

years). The Independent. Retrieved from http://www.independent.co.uk/news/world/americas/<br />

peru-tells-yale-it-wants-its-machu-picchu-treasures-back-after-100-years-465452.html.<br />

Del-Arroyo, T. (2005). Choquequirao: Crib of gold. Retrieved from http://www.american.edu/<br />

ted/peru-culture.htm<br />

Desforges, L. (2000). <strong>State</strong> tourism institutions and neo-liberal development: A case study of Peru.<br />

Tourism Geographies, 2(2), 177–192.<br />

Divino, J.A., & McAleer, M. (2010). Modelling and forecasting daily international mass tourism to<br />

Peru. Tourism Management, 31(6), 846–854.<br />

Emmott, R. (2003, November 26). Tourist boom threatens Peru’s Machu Picchu. Reuters Foundation<br />

Alert Net. Retrieved http://www.alertnet.org<br />

Farrell, B.H., & Twining-Ward, L. (2004). Reconceptualizing tourism. Annals of Tourism Research,<br />

31(2), 274–295.<br />

Flores Ochoa, J.A. (2004). Contemporary significance of Machu Picchu. In R.L. Burger & L.C.<br />

Salazar (Eds.), Machu Picchu: Unveiling the mystery of the Incas (pp. 109–123). New Haven,<br />

CT: Yale University Press.<br />

Galiano Sanchez, W. (2000). Situacion ecologico-ambiental del Santuario Historico de Machupiqcchu:<br />

Una aproximacion. Cuzco, Peru: Programa Machu Picchu.<br />

Gordon, R., & Knopf, R. (2007). Late horizon silver, copper, and tin from Machu Picchu, Peru.<br />

Journal of Archaeological Science, 34, 38–47.<br />

Graefe, A.R., Kuss, F.R., & Vaske, J.J. (1990). Visitor Impact Management: The planning framework.<br />

Washington, DC: National Parks and Conservation Association.<br />

Hadfield, P. (2001, March 7). Slip siding away. New Scientist. Retrieved from http://www.<br />

newscientist.com/article/dn491-slip-sliding-away.html.<br />

Hamilton, L.S. (1995). <strong>Mountain</strong> cloud forest conservation and research: A synopsis. <strong>Mountain</strong><br />

Research and Development, 15(3), 259–266.<br />

Hammitt, W.E., & Cole, D.N. (1998). Wildland recreation: Ecology and management. NewYork:<br />

John Wiley & Sons.<br />

Hawkins, D.E., Chang, B., & Warnes, K. (2009). A comparison of the National Geographic Stewardship<br />

Scorecard Ratings by experts and stakeholders for selected World Heritage destinations.<br />

Journal of Sustainable Tourism, 17(1), 71–90.<br />

Healy, P.O. (2006, November 12). Taking the back roads to Machu Picchu. New York Times: Travel.<br />

Retrieved from http://travel2.nytimes.com/2006/11/12/travel/12machu.html<br />

Higgins, M. (2006, July 9). A faster way: New helicopter service to Machu Picchu. New York Times:<br />

Travel. Retrieved from http://travel.nytimes.com/2006/07/09/travel/09transmachu.html<br />

Hunter, C.J. (1995). On the need to reconceptualize sustainable tourism development. Journal of<br />

Sustainable Tourism, 3, 155–165.<br />

Instituto Nacional de Cultura (INC). (2005). Plan maestro del santuario historic de Machupicchu.<br />

Cusco, Peru: Instituo Nacional de Cultura, Instituo Nacional de Recursos Naturales y Dirección<br />

Regional de Cusco.<br />

International Council on Monuments and Sites (ICOMOS). (1983). Historic sanctuary<br />

of Machu Picchu. Retrieved from http://whc.unesco.org/archive/advisory_body_<br />

evaluation/274.pdf<br />

International Council on Monuments and Sites (ICOMOS). (1999). International cultural<br />

tourism charter. Retrieved from http://www.international.icomos.org/charters/<br />

charters.pdf<br />

Jamal, T., & Stronza, A. (2009). Collaboration theory and tourism practice in protected areas:<br />

Stakeholders, structuring and sustainability. Journal of Sustainable Tourism, 17(2),<br />

169–189.<br />

Jensen, O. (2010). Social mediation in remote developing world tourism locations – the significance<br />

of social ties between local guides and host communities in sustainable tourism development.<br />

Journal of Sustainable Tourism, 18(5), 615–633.<br />

Keatinge, R.W. (1982). Economic development and cultural resources management in the Third<br />

World: An example from Peru. Journal of Anthropological Research, 38(2), 211–226.<br />

Knutson, M., Laskowski, H., Moore, C., Lonsdorf, E., Lor, S., & Stevenson, L. (2011). Defensible<br />

decision making: Harnessing the power of adaptive resource management. The Wildlife<br />

Professional, 4(4), 58–62.<br />

LaFranchi, H. (2001). Machu Picchu’s slide. Christian Science Monitor, 93(112), 7.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

936 L.R. Larson and N.C. Poudyal<br />

Leffel, T. (2005, November). Saving Machu Picchu: Responsible tourism is a 3-way<br />

deal. Transitions Abroad. Retrieved from http://www.transitionsabroad.com/publications/<br />

magazine/0511/saving_machu_picchu.shtml<br />

MacCannell, D. (1973). Staged authenticity: Arrangements of social space in tourist settings.<br />

American Journal of Sociology, 79(3), 589–603.<br />

Manning, R.E. (1999). Studies in outdoor recreation: Search and research for satisfaction (2nd ed.).<br />

Corvallis, OR: Oregon <strong>State</strong> University Press.<br />

McCool, S.F., & Lime, D.W. (2001). Tourism carrying capacity: Tempting fantasy or useful reality?<br />

Journal of Sustainable Tourism, 9(5), 372–388.<br />

McCool, S.F., & Moisey, R.N. (2008). Introduction: Pathways and pitfalls in the search for sustainable<br />

tourism. In S.F. McCool & R.N. Moisey (Eds.), Tourism, recreation and sustainability (2nd ed.,<br />

pp. 1–16). Wallingford: CAB International.<br />

McFadden, J.E., Hiller, T.L., & Tyre, A.J. (2011). Evaluating the efficacy of adaptive management approaches:<br />

Is there a formula for success? Journal of Environmental Management, 92, 1354–1359.<br />

McGowan, J.L. (2010). The sustainable livelihoods and tourism intersect: Gender, power structures<br />

and local market vendors in Aguas Calientes, Peru. Halifax, Nova Scotia: Dalhousie University.<br />

McGrath, G. (2004). Including the outsiders: The contribution of guides to integrated heritage of<br />

tourism management in Cusco, Southern Peru. Current Issues in Tourism, 7(4), 426–432.<br />

Mendoza Quintana, A. (1997). Evaluacion del patrimonio turistico del Peru. Investigaciones en<br />

Turismo, 1, 149–177.<br />

Miller, G. (2001). The development of indicators for sustainable tourism: Results of a Delphi survey<br />

of tourism researchers. Tourism Management, 22, 351–362.<br />

Miller, G., & Twining-Ward, L. (2005). Monitoring for a sustainable tourism transition: The challenge<br />

of developing and using indicators. Cambridge, MA: CABI Publishing.<br />

Miller, G., & Twining-Ward, L. (2006). Monitoring as an approach to sustainable tourism. In D.<br />

Buhalis & C. Costa (Eds.), Tourism management dynamics: Trends, management, and tools (pp.<br />

51–57). Oxford: Elsevier Butterworth-Heinemann.<br />

Mitchell, R.E. (2008). Community perspectives in sustainable tourism: Lessons from Peru. In S.F.<br />

McCool & R.N. Moisey (Eds.), Tourism, recreation and sustainability (2nd ed., pp. 158–182).<br />

Wallingford: CAB International.<br />

Mitchell, R.E., & Eagles, P.F.J. (2001). An integrative approach to tourism: Lessons from the Andes<br />

of Peru. Journal of Sustainable Tourism, 9(1), 4–28.<br />

Moscardo, G. (2011). Exploring social representations of tourism planning: Issues for governance.<br />

Journal of Sustainable Tourism, 19(4), 423–436.<br />

Navrud, S., & Ready, R.C. (Eds.). (2002). Valuing cultural heritage: Applying environmental valuation<br />

techniques to historic buildings, monuments and artefacts. Cheltenham: Edward Elgar.<br />

Nichols, J.D., & Williams, B.K. (2006). Monitoring for conservation. Trends in Ecology & Evolution,<br />

21(12), 668–673.<br />

O’Hare, G., & Barrett, H. (1999). Regional inequalities in the Peruvian tourist industry. The Geographical<br />

Journal, 165(1), 47–61.<br />

Peyton, B. (1980). Ecology, distribution, and food habits of spectacled bears, Tremarctus ornatus, in<br />

Peru. Journal of Mammalogy, 61(4), 639–652.<br />

Plummer, R., & Fennell, D.A. (2009). Managing protected areas for sustainable tourism: Prospects<br />

for adaptive co-management. Journal of Sustainable Tourism, 17(2), 149–168.<br />

QStation. (2009). Bi-annual monitoring report for the north head quarantine station. Retrieved from<br />

http://www.qstation.com.au/pdfs/conservation/Q%20Station%20Bi-Annual%20Report%203.pdf<br />

Rees, W.E. (1990). The ecology of sustainable development. Ecologist, 20(1), 18–23.<br />

Regalado, A. (2011, February 14). Yale agrees to return Machu Picchu artifacts to Peru. ScienceInsider.<br />

Retrieved from http://news.sciencemag.org/scienceinsider/2011/02/yale-agrees-to-returnmachu-picchu.html<br />

Regalado-Pezua, O., & Arias-Valencia, J. (2006). Sustainable development in tourism: A proposition<br />

for Machu Picchu, Peru. In A. Leask & A. Fyall (Eds.), Managing world heritage sites (pp.<br />

196–204). Burlington, MA: Butterworth-Heinemann.<br />

Roach, J. (2002, April 15). Machu Picchu under threat from pressures of tourism.<br />

National Geographic News. Retrieved from http://news.nationalgeographic.com/news/<br />

2002/04/0415_020415_machu.html<br />

Rugendyke, B., & Son, N.T. (2005). Conservation costs: Nature-based tourism as development at<br />

Cuc Phuong National Park, Vietnam. Asia Pacific Viewpoint, 46(2), 185–200.


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

Journal of Sustainable Tourism 937<br />

Saarinen, J. (2006). Traditions of sustainability in tourism studies. Annals of Tourism Research, 33(4),<br />

1121–1140.<br />

Sassa, K., Fukuoka, H., Wang, F., & Wang, G. (Eds.). (2005). Landslides: Risk analysis and sustainable<br />

disaster management. Berlin: Springer.<br />

Schianetz, K., & Kavanagh, L. (2008). Sustainability indicators for tourism destinations: A complex<br />

adaptive systems approach using systemic indicator systems. Journal of Sustainable Tourism,<br />

16(6), 601–628.<br />

Shackley, M. (2006). Visitor management at World Heritage Sites. In A. Leask & A. Fyall (Eds.),<br />

Managing world heritage sites (pp. 83–94). Burlington, MA: Butterworth-Heinemann.<br />

Solano, P. (2005). La esperanza es verde: Areas naturales protegidas en el Peru. Lima: Sociedad<br />

Peruana de Derecho Ambiental.<br />

Stankey, G.H., Cole, D.N., Lucas, R.C., Peterson, M.E., & Frissell, S.S. (1985). The limits of acceptable<br />

change (LAC) system for wilderness planning (Gen. Tech. Rep. INT-176). Ogden, UT:<br />

Intermountain Research Station, USDA Forest Service.<br />

Stronza, A. (2001). Anthropology of tourism: Forging new ground for ecotourism and other alternatives.<br />

Annual Review of Anthropology, 30, 261–283.<br />

Toman, M.A. (1992). The difficulty in defining sustainability. Resources, 106, 3–6.<br />

Tosun, C. (2000). Limits to community participation in the tourism development process in developing<br />

countries. Tourism Management, 21, 613–633.<br />

Twining-Ward, L., & Butler, R. (2002). Implementing STD on a small island: Development and use<br />

of sustainable tourism development indicators in Samoa. Journal of Sustainable Tourism, 10(5),<br />

363–387.<br />

United Nations Educational, Scientific & Cultural Organization (UNESCO). (2006). Decision:<br />

30COM 7B.35. Retrieved from http://whc.unesco.org/en/decisions/1121<br />

United Nations Educational, Scientific & Cultural Organization (UNESCO). (2009). Decision:<br />

30COM 7B.42. Retrieved from http://whc.unesco.org/en/decisions/4150<br />

United Nations Educational, Scientific & Cultural Organization (UNESCO). (2011). Decision:<br />

35COM 7B.38. Retrieved from http://whc.unesco.org/en/decisions/4446<br />

United Nations Environment Programme (UNEP). (2004). Tourism and local agenda 21: The<br />

role of local authorities in sustainable tourism. Paris: UNEP Division Techonology, Industry<br />

and Economics & International Council for Local Environmental Initiatives. Retrieved from<br />

http://www.unep.fr/shared/publications/pdf/3207-TourismAgenda.pdf<br />

United Nations Environment Programme (UNEP). (2008). World Conservation Monitoring<br />

Centre, Protected Areas and World Heritage Sites: Historic sanctuary of Machu<br />

Picchu, Peru. Retrieved from http://www.unep-wcmc.org/medialibrary/2011/06/29/91fbcd4e/<br />

Machu%20Picchu.pdf<br />

van den Berghe, P.L., & Flores Ochoa, J. (2000). Tourism and nativistic ideology in Cuzco, Peru.<br />

Annals of Tourism Research, 27(1), 7–26.<br />

Vaughan, D. (2000). Tourism & biodiversity: A convergence of interests? International Affairs (Special<br />

Biodiversity Issue), 76(2), 283–297.<br />

Vecchio, R. (2011, February 22). UNESCO Chief: “Machu Picchu is a victim of its own success”.<br />

Fertur Peru Travel Blog. Retrieved from http://www.ferturperu.info/2011/unesco-chief-machupicchu-is-a-victim-of-its-own-success/2297#more-2297<br />

Weaver, D.B. (2011). Organic, incremental and induced paths to sustainable mass tourism convergence.<br />

Tourism Management. doi: 10.1016/j.tourman.2011.08.011<br />

Williams, B.K. (2011a). Adaptive management of natural resources – framework and issues. Journal<br />

of Environmental Management, 92, 1346–1353.<br />

Williams, B.K. (2011b). Passive and active adaptive management: Approaches and an example.<br />

Journal of Environmental Management, 92, 1371–1378.<br />

World Conservation Union (IUCN). (1994). Guidelines for protected area management categories.<br />

Retrieved from http://app.iucn.org/dbtw-wpd/edocs/1994-007-En.pdf<br />

World Conservation Union (IUCN). (2011). The IUCN Red List of threatened species. Retrieved from<br />

IUCN: http://www.iucnredlist.org/<br />

World Economic Forum. (2011). Peru: Travel & tourism indicators. Retrieved from<br />

http://www.weforum.org/reports/travel-tourism-competitiveness-report-2011<br />

World of New Seven Wonders. (2011). The official New7Wonders of the World. Retrieved from<br />

http://world.n7w.com/the-official-new7wonders-of-the-world/


Downloaded by [<strong>Appalachian</strong> <strong>State</strong> University] at 09:41 30 August 2012<br />

938 L.R. Larson and N.C. Poudyal<br />

World Tourism Organization (WTO). (1996). What tourism managers need to know: A practical guide<br />

to the development and use of indicators of sustainable tourism. Madrid: WTO.<br />

World Tourism Organization (WTO). (2004). Indicators of sustainable development for<br />

tourism destinations. Madrid: WTO. Retrieved from http://sdt.unwto.org/en/content/<br />

indicators-sustainability-tourism-destinations<br />

Yachay Wasi. (2006, May). Sacred sites and the environment from an indigenous perspective. Paper<br />

presented at the Fifth UN Permanent Forum on Indigenous Issues, New York. Retrieved from<br />

http://www.yachaywasi-ngo.org/InkaChallenge06<br />

Young, K.R., & Leon, B. (2000). Biodiversity conservation in Peru’s eastern montane forests. <strong>Mountain</strong><br />

Research and Development, 20(3), 208–211.<br />

Zan, L., & Lusiani, M. (2011). Managing change and master plans: Machu Picchu between conservation<br />

and exploitation. Archaeologies: Journal of the World Archaeological Congress. doi:<br />

10.1007/s11759-011-9167-7


ADAPTIVE GOVERNANCE AND CLIMATE CHANGE IN THE<br />

TROPICAL HIGHLANDS OF WESTERN SOUTH AMERICA<br />

KENNETH R. YOUNG and JENNIFER K. LIPTON<br />

Department of <strong>Geography</strong> and the Environment, University of Texas at Austin,<br />

Austin, TX 78712, U.S.A<br />

E-mail: kryoung@mail.utexas.edu<br />

Abstract. Climate changes occurring during the past several decades in the high elevations of the<br />

tropical Andes <strong>Mountain</strong>s have implications for the native plant and animal species, for the ecological<br />

integrity of the affected land cover, and for the human-biophysical systems involved. Consequences are<br />

also probable for rural inhabitants and their livelihoods, especially for farmers and pastoralists. Biophysical<br />

factors have always changed in these mountainous zones; the extent and degree of alteration<br />

acting on native and agricultural biodiversity is the concern. Addressing these climate changes is probably<br />

within the adaptive capacity of many local land-use systems, unless external socioeconomic or<br />

political forces are unsupportive or antagonistic. Suitable programs to provide information, subsidies,<br />

or alternatives could be designed. We highlight some of the inherent resiliencies of natural and cultural<br />

systems in the Andes and suggest that these systems contain lessons that could be useful elsewhere, in<br />

terms of the traits that allow for the sustainable utilization of dynamic and heterogeneous landscapes.<br />

1. Introduction<br />

Indications of dramatic and accelerating changes in the high elevations of the Andes<br />

<strong>Mountain</strong>s come from scientific observers who commented on retreating valley<br />

glaciers and ice cap margins above 5000 m elevation (Hastenrath and Ames, 1995;<br />

Kaser and Georges, 1999; Francou et al., 2003). One example is from the Quelccaya<br />

ice cap in southern Peru where 37 years of observations document an increased<br />

rate of retreat during the 1990s (Thompson et al., 2003). While loss of glacial ice<br />

signals climatic changes at higher elevations, at lower elevations the signs are more<br />

subtle, influencing landscapes and their human-biophysical systems. A human imprint<br />

in the Andes is near ubiquitous: agricultural fields reach to above 4000 m<br />

and livestock, especially cattle and sheep, graze to snowline. Local people interact<br />

with landscapes as they grow crops, extract natural resources such as firewood, and<br />

graze livestock (Gade, 1999; Mayer, 2002). Additionally, the rural inhabitants of the<br />

Andes are immersed in formal and informal institutions concerned with governing<br />

resources. To some extent, their decisions in relation to changing environmental<br />

controls must themselves be regulated by economic, cultural, political, and legal<br />

concerns. Their livelihoods and local landscape perspectives are important to consider<br />

when studying climate change and integrating its effects into large-scale policy<br />

considerations.<br />

Climatic Change (2006) 78: 63–102<br />

DOI: 10.1007/s10584-006-9091-9 c○ Springer 2006


64 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

The consequences of changing climatic processes on these areas are more subtle<br />

than solely the loss of glacial ice upslope. They will be complicated for both natural<br />

and cultural systems: shifts in plant species dominances, alterations in primary<br />

productivity, and changes in the kind and tempo of disturbance regimes. Human<br />

responses to the effects of climate change will likely be complex and variable.<br />

Individuals living within kilometers of mountain glaciers are directly affected by<br />

environmental perturbations. Implications for them might be alterations in the cover<br />

of spontaneously occurring and cultivated plants, in water resources, in exposure<br />

to natural hazards, and in household or community decisions on when and where<br />

to farm. Those shifts will affect the native biological diversity of the Andes, which<br />

is notable in its richness and uniqueness. In particular, the tropical latitudes of the<br />

Andes were recently ranked as one of the world’s most important and threatened<br />

biodiversity hotspots by Manne et al. (1999), Myers et al. (2000), and Brooks<br />

et al. (2002). This biodiversity is often adjacent to landscapes with high cultural<br />

diversity, as measured by ethnic or linguistic differences (Fjeldsä et al., 1999). There<br />

are numerous species with limited ranges (Young, 1995), leading to high placeto-place<br />

differences in species composition, for example, for birds (Poulsen and<br />

Krabbe, 1998; Ruggiero, 2001), plants (Young et al., 1997; Luteyn, 1999; Young and<br />

León, 1999; Keating et al., 2002; Young et al., 2002; Leimbeck et al., 2004), insects<br />

(Brehm et al., 2003), and aquatic organisms (González and Watling, 2003; Koínek<br />

and Villalobos, 2003). Biodiversity concerns are mediated through institutions,<br />

including community and household decisions on extraction and management of<br />

species considered useful.<br />

In this paper, we evaluate the likely consequences of ongoing and future climate<br />

changes for land-use systems of rural parts of the tropical Andes, for the<br />

various institutions that make land management decisions for the respective countries<br />

and landscapes, for the native plants and animals, and for the institutions<br />

involved in their conservation or utilization. Most of our experiences and examples<br />

are drawn from fieldwork in the Peruvian Andes, but many of our conclusions<br />

appear to be valid regionally. We obtained additional insights from participation<br />

in formal and informal projects examining the state of conservation and rural development<br />

programs in Peru. These included national-level reviews (Rodríguez<br />

and Young, 2000), and regional assessments (León and Young, 1996; Young and<br />

León, 1999, 2001). Our eventual goal is to provide sufficient information so that<br />

the climate-related changes in the tropical Andes can be compared and contrasted<br />

with fluxes in other places worldwide. We suspect that some of the inherent resiliencies<br />

of natural and cultural systems in the Andes contain lessons for use<br />

elsewhere.<br />

We begin with a case study from an area in and around Huascaran National Park,<br />

containing the world’s largest protected tropical mountain landscape. We then place<br />

those findings in the context of the general features of <strong>Andean</strong> rural livelihoods,<br />

stressing the aspects that seem most relevant to climate change. Finally, we evaluate<br />

the likely resilience of those coupled human-biophysical systems, especially as


KENNETH R. YOUNG AND JENNIFER K. LIPTON 65<br />

viewed through the lens of governance and decision making by the people and<br />

institutions involved.<br />

2. Case Study: Huascaran National Park and Surroundings<br />

in North-Central Peru<br />

Huascaran National Park (HNP), located at approximately 9 ◦ 30 ′ S and 77 ◦ 49 ′ Win<br />

the Department of Ancash, north-central Peru, protects tropical alpine vegetation<br />

communities along the highest mountain chain in the tropical Andes, the Cordillera<br />

Blanca (Figures 1 and 2). The information in this case study serves as an example of<br />

the interlinkages that exist among different institutional frameworks all operating<br />

within one defined study area. Our goal is to reveal some of the challenges and<br />

benefits facing decision-makers concerned with biodiversity and climate change in<br />

tropical mountain environments.<br />

The Cordillera Blanca is one of the world’s youngest exposed granitic batholithic<br />

intrusions (Atherton and Petford, 1996; McNulty and Farber, 2002). Extending approximately<br />

180 km parallel to the Peruvian coast with over 200 peaks that surpass<br />

5000 m elevation, HNP encompasses most of the mountain range. It covers 340,000<br />

ha and includes Mt. Huascarán (6768 m) as the highest point (Figure 2). Given the<br />

tropical location, annual seasonal temperature variability is minimal staying at<br />

approximately 15 ◦ C at mid-elevations. However, in the upper elevations diurnal<br />

temperatures may fluctuate dramatically, resulting in daily freeze-thaw events. The<br />

Cordillera Blanca glaciers rarely extend below about 4700 m and as evidence suggests,<br />

this limit has been moving upwards as glaciers retreat, increasing the size and<br />

number of proglacial lakes, which can be dangerous if they become destabilized<br />

(Ames et al., 1989; Lliboutry et al., 1977; Kaser and Osmoston, 2002). Glacial<br />

runoff, 296 glacial lakes, and seasonal rainfall (October to March) provide not only<br />

water for drinking, agriculture, and industry for the park’s peripheral communities<br />

and cities, but also hydroelectricity for distant urban centers on the Peruvian<br />

coast. Amidst the numerous glacial valleys that intersect the eastern and western<br />

flanks of the Cordillera Blanca, with steep elevation gradients ranging from 1900<br />

m to upwards of 6000 m, is an heterogeneous mosaic of vegetation and land-cover<br />

types. High <strong>Andean</strong> woodlands of Polylepis, Buddleia, and Gynoxis and stands<br />

of Puya raimondii, form isolated patches throughout the national park (Smith,<br />

1988). Cushion-plant communities, wetlands, and natural grasslands are also considered<br />

a top priority by the park managers for flora and fauna habitat conservation.<br />

Vicuña (Vicugna vicugna), the <strong>Andean</strong> bear (Tremarctos ornatus), <strong>Andean</strong><br />

huemul deer (Hippocamelus antisensis), and the <strong>Andean</strong> condor (Vultur gryphus)<br />

are species under protection and conservation priorities within the national park<br />

(INRENA, 2003).<br />

Approximately 337,520 inhabitants, mainly bilingual Quechua-Spanish speakers,<br />

live on both sides of the mountain range and are engaged in livelihood strategies


66 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Figure 1. National protected area system in Peru, with the general location of the case study indicated<br />

by arrows.<br />

of subsistence and market-based agro-pastoralism, mining, and tourism. Population<br />

in urban centers, such as the departmental capital of Huaraz, has increased<br />

over the past few decades due to migration from rural areas and has been driven<br />

by job opportunities in commercial centers. Situated just to the west and parallel to<br />

the Cordillera Blanca is the intensively cultivated Santa River Valley, where urban<br />

development, infrastructure, and services are more pronounced than on the eastern<br />

side of the mountains (Figure 2). On the eastern flank of the mountain range<br />

are smaller populated centers and more isolated valleys, including the Mosna and<br />

Pomabamba River Valleys, which flow to the Marañon River system and eventually<br />

drain to the Amazon basin. Private property owners and peasant communities


KENNETH R. YOUNG AND JENNIFER K. LIPTON 67<br />

Figure 2. Location of the case study in north-central Peru of Huascaran National Park (which includes<br />

much of the Cordillera Blanca), the Cordillera Huayhuash Reserved Zone, and major urban centers.<br />

(comunidades campesinas) have lands or perceived property rights that extend into<br />

the national park. Agricultural production of alfalfa, corn, wheat, barley, and beans<br />

occurs predominantly at elevations between 2000–3200 m. Within intermontane<br />

basins and valleys, potatoes, ulluco, oca, broad beans, and quinoa are frequently<br />

cultivated from 2500–4000 m. Communities in the study area also often utilize


68 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

upper elevations (3900 to more than 5000 m), dominated by grasslands, woodland<br />

patches, shrublands, and wetland plants, for seasonal rotational grazing of mostly<br />

cattle, horses, and sheep and, less frequently, alpacas and llamas.<br />

The delimiting boundary for the national park was established at roughly the<br />

4000 m elevation line in 1975. Prior to the creation of the park, the highland<br />

area was under the jurisdiction of large farms (haciendas), private ownership, or<br />

mining concessions. Peruvian agrarian reform took place between the years of<br />

1969–1975 and was coincident with regional rebuilding and restructuring following<br />

a destructive 1970 earthquake and subsequent ice and rock avalanche that took the<br />

lives of over 18,000 people in the immediate area (Lliboutry et al., 1977; Kaser<br />

and Osmoston, 2002). The park was established for the conservation of the unique<br />

flora and fauna and as a tourist destination for Peruvians and foreigners alike.<br />

Local people, who often self-identify as campesinos (peasants), are involved in<br />

seasonal adventure- and ecotourism-based economies as guides or porters and in<br />

commercial services, in addition to traditional agropastoral and mining livelihoods<br />

(Bartle, 1993). In many zones of the park, conflicts exist between local interests<br />

and park directives regarding access to and use of what is legally considered park<br />

territory.<br />

Located approximately 120 km to the southeast of HNP is another range, the<br />

Cordillera Huayhuash (Figures 1 and 2). This mountain range extends for 26 km<br />

between 10 ◦ 12 ′ and 10 ◦ 27 ′ S and about 76 ◦ 55 ′ N and has over 117 glaciers with<br />

a majority draining to the Pacific (Ames et al., 1989). Long-term agro-pastoral use<br />

has facilitated culturally created biological diversity while areas of Polylepis forest<br />

patches harbor distinct avifauna habitat (Cerrate, 1979; Oxford University, 1996).<br />

The area surrounding the Cordillera Huayhuash was officially integrated in 2002<br />

into the Peruvian national protected-area system and is categorized as a Reserved<br />

Zone (Resolucion Ministerial, 2002), while it awaits formal categorization and<br />

delimitation.<br />

In-residence during 2002 and 2003, the second author conducted fieldwork in<br />

core and periphery communities of HNP and the Cordillera Huayhuash. As part<br />

of a study on landscape change and national park management, the research included<br />

collecting ground truth data for different land cover types. Semi-structured<br />

interviews were carried out with 117 individuals (89 men, 28 women). All interviewees<br />

lived in the park periphery and had a vested interest in the resources of<br />

the region; they were, therefore, defined as stakeholders. Key informants interviewed<br />

were engaged in park management, glacial research, peasant community<br />

groups, and/or individuals with private interests. Interviews were conducted in<br />

homes, fields, and offices in Spanish and Quechua with the help of field assistants.<br />

Interviews and discussions with household, community, and regional-level<br />

officials provided qualitative data for the analysis of perspectives on landscape<br />

change and potential responses to climate change. During the semi-structured interviews,<br />

the informants were asked open-ended questions covering the following<br />

areas: household and community land use and land-cover change, perceptions of


KENNETH R. YOUNG AND JENNIFER K. LIPTON 69<br />

climatic changes, glacial ice limit differences, agro-pastoral production rates and<br />

changes, institutional participation, and interactions with the park. Fieldwork also<br />

included community mapping to obtain Quechua toponyms, land-cover categories,<br />

and historical land-use and land-cover descriptions. To obtain the interviewees’<br />

perspectives and discourse on park management and landscape change, data were<br />

sought at various levels of institutional hierarchy. The institutional interactions and<br />

forms of adaptive governance that occurred within this one region help illustrate<br />

the complex levels and issues involved in resiliency of land-use systems and biodiversity<br />

conservation.<br />

We became sensitized to the perceptions of climate change on future landscape<br />

scenarios by conversations with local residents. A typical example is Felix Valverde,<br />

who lives in the park periphery. We visited his lands in July 2002 and 2003. Born and<br />

raised in the same locale, he lives with his family at 4290 m. They are members of the<br />

peasant community of Aquia, located 18 km to the southwest of his dwellings. Their<br />

livelihood is dependent upon livestock, subsistence agriculture, and occasional<br />

wage-work. The Valverdes personally own over 100 head of livestock, mostly sheep,<br />

horses and burros. In addition, in exchange for goods, they serve as guardians for<br />

other animals that are owned by Aquia community members who live closer to<br />

town. The Valverde family has agricultural lands at lower elevations closer to the<br />

town where they grow crops such as wheat, beans and maize, plus <strong>Andean</strong> tubers:<br />

mashua, oca, potatoes, and ulluco. With a newly paved road, they are able to travel<br />

to their fields and to town more frequently than in the past.<br />

Señor Valverde’s concerns about glacial retreat and the impact that it has made<br />

in his life—and would make in the future—were striking. Over the years he has<br />

watched as the glaciers have receded, and he knows that fairly soon they might be<br />

gone; some small glacial caps in the valley where he lives have already disappeared.<br />

As he identified different (de)glaciated mountain peaks, with elevations ranging<br />

from approximately 5095–5250 meters around his homestead, his recall of where<br />

the glacial margins used to be was often associated with personal events in his life.<br />

For example, in the 1960s he worked for a Peruvian-owned silver mine and on his<br />

way up to work over the pass, he used to cut off pieces of the glacial ice. Today<br />

there is a layer of talus to demarcate this past glacial lobe boundary. Thinking<br />

about this kind of change provoked him to comment upon how the mountains will<br />

soon be “desnudo” (naked) and only rock will remain. He is certain that this will<br />

happen within his lifetime and that it will affect his livelihood. He further noted that<br />

the immediate environmental changes were not only related to the disappearance<br />

of the glaciers, but also to a difference in water drainage patterns and a drop in<br />

precipitation. As glaciers have receded, some drainage and runoff channels that<br />

were accessed by the Valverde family diminished in flow and were now either<br />

intermittent or dry. As a result of this change, he needed to obtain water from<br />

sources further away where runoff and glacial melt had increased. Also, he related<br />

how, in some valley bottoms, lakes and wetland areas increased in size, so they had<br />

changed the sites where they took their sheep for pasturing.


70 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

According to more than half (56%) of informants, glacial land cover is undergoing<br />

the most evident rate of change throughout the area. This is corroborated<br />

by glacial inventories and long term studies of glacial recession completed in<br />

the study area (Ames et al., 1989; Hastenrath and Ames, 1995a, 1995b; Kaser et<br />

al., 1996; Kaser and Georges, 1997; Morales, 2000; Thompson, 2002; Kaser and<br />

Osmoston, 2002; Thompson et al., 2003; Francou et al., 2003; Silverio and Jaquet,<br />

2005). Georges (2004) reported that glacial cover has decreased from an estimated<br />

extent of 800–850 km 2 in the 1930s to only 619 km 2 by the beginning of the 1990s.<br />

We observed that people from different communities in our study area have, with<br />

a sense of humor, renamed the recently deglaciated mountains. Local toponyms<br />

that once described a shape of an ice cap, now refer to the loss of that glacier. For<br />

example, a mountain that was called “sleeping lion” for the shape of the glacier is<br />

now called “lion has left” (that is, león dormido has become león se ha ido). Or in<br />

the local dialect a mountain that once had a glacier is now called simply “mountain<br />

without glacier.”<br />

Quechua households and communities within the study area had environmental<br />

perceptions that were informal indicators of seasonal climate variability and<br />

change. Approximately two thirds (68%) of informants, including Sr. Valverde,<br />

when questioned commented on how precipitation has decreased compared to the<br />

past. Many local residents noted a decline in hail and sleet, which seasonally blanketed<br />

landscapes with overnight accumulations of more than a few centimeters.<br />

Sleet seemed to serve as an important environmental predictor for the likelihood of<br />

damaging frosts, affecting seasonal agricultural patterns. These traditional signals<br />

provided farmers with an ecological agricultural decision-making toolset that in<br />

many cases might be relied upon more firmly than technological or scientificallyvalidated<br />

signals. As another example, in a couple of communities, the presence of<br />

an <strong>Andean</strong> swift circling and flying in low to mid-elevation zones indicated that the<br />

rains were going to arrive early. Conversely, a relatively low population of firefly<br />

beetles indicated that it would be a dry season.<br />

Woodland and grassland cover were two other land cover categories that informants<br />

claimed to have changed over time. Data from Byers (2000), who used<br />

repeat photography of specific valleys within HNP from the years of 1936 and 1998,<br />

clarifies that native forest cover of Polylepis stayed the same or increased, while<br />

nonnative forest cover of Eucalyptus and Pinus increased. Contrarily, our informants<br />

claimed that native forest cover had reduced. Reasons stated for the change<br />

in forest cover were that tourists utilized wood for campfires, mining camps used<br />

it for fire and construction, people from communities removed wood for cooking<br />

and construction, and that natural fires destroyed woodland patches. According to<br />

34% of informants interviewed, grassland cover had reduced in terms of quality of<br />

production for livestock. Grasslands were described by informants as overgrazed,<br />

insufficient for cattle or with a low quality of native grasses. An indicator of reduced<br />

forage value is said to be the presence of a high-elevation cactus, Opuntia flocossa,<br />

which becomes more common with overgrazing. Both of these land cover changes


KENNETH R. YOUNG AND JENNIFER K. LIPTON 71<br />

were identified by local rural residents as a change in spatial extent and quality<br />

of the land cover itself. When discussing changes in climate, people often directly<br />

correlated the change in climate to a change or potential change in agro-pastoral<br />

productivity.<br />

For highland families, livelihood modifications and adaptations are imminent in<br />

the ongoing landscape change associated with glacial retreat and climate change.<br />

As in the case of the Valverdes, decisions on grazing, herd size, dealing with<br />

disease, and agricultural patterns are ongoing. An example of the complexity of<br />

these interactions is in regards to grazing conditions. Cattle are considered an<br />

investment for many rural families; a source of livelihood and wealth. Informants<br />

commented that reduced fodder has led to inadequate milk production. One possible<br />

household-level response to these perceptions would be to adjust herd size or, more<br />

commonly, to diversify livelihoods and seek supplementary income sources, such<br />

as wage labor, to compensate for declining production. Another response might be<br />

to acquire usage rights for grazing at higher elevations.<br />

Informants indicated that livelihoods based solely on agriculture were currently<br />

insufficient given the amount and rising costs of inputs, labor, resulting yield, and<br />

market prices. In fact, of 67 men interviewed in Ancash in 2002 and 2003, 52<br />

(78%) participated in part-time wage labor in other farms, jobs, or regions of Peru.<br />

The majority of jobs sought within Ancash included tourism, mining, and labor on<br />

other farmsteads, as well as short-term employment in governmental programs for<br />

forestry, roads, housing, and other construction infrastructure projects. In addition,<br />

young men from agricultural communities frequently out-migrated to urban areas<br />

seeking wage-labor and opportunities. In one sector of a community in Ancash,<br />

87% of the young men between the ages of 27 and 33 had left the community. This<br />

type of demographic change will have an impact on the future of community level<br />

decisions and will reduce the experiential traditional knowledge that is required to<br />

sustain culturally maintained agricultural biodiversity (e.g., Stobart and Howard,<br />

2002). Women are involved in almost all of the agro-pastoral functions and are in<br />

charge while husbands and sons migrate seasonally. However, they are frequently<br />

marginalized from direct positions in community social organizations, assemblies,<br />

or training classes.<br />

Informants indicated that current market returns for subsistence crops do not<br />

compensate for their cost of agricultural inputs. This economic stress was emphasized<br />

on a regional level: in July 2002, campesinos participated in three days of<br />

marches and strikes to voice these and other concerns, effectively shutting down<br />

urban areas and principal transportation routes. In addition to these acts of resistance,<br />

farmers responding to market-driven forces adapt to new management<br />

practices and build institutional connections with technical agro-pastoral programs.<br />

National and international nongovernmental (NGO) networks that promote agricultural<br />

production along economic corridors spread information about new techniques<br />

and varieties. For example, on the eastern periphery of HNP, in response to a<br />

regional development program, farmers are integrating different varieties of maize,


72 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

primarily from southern Peru, which are known to be in demand and can be cultivated<br />

at slightly higher elevations. Community institutions may seek out respected<br />

NGOs that offer technical assistance, because of a keen interest in enhancing commercial<br />

production and bettering local skill-sets and knowledge. NGO and other<br />

institutional technical programs will need to update and modify their advice, skills,<br />

and techniques to respond to potential changes in pests or other management and<br />

production issues associated with climate change. These transnational networks<br />

simultaneously link producers to potential international export markets.<br />

Many individuals in the study area had serious concerns about running water<br />

and were aware that, although now they might have water, as the glaciers recede<br />

irrigation canals would possibly have less flow and some highland crops would be<br />

in jeopardy. A fear generally expressed was that the glaciated mountains someday<br />

would resemble the semiarid, rain-shadow mountains visible on the horizon towards<br />

the dry Pacific slopes; thus, this fear raised their concerns about possible loss<br />

of water-availability, increased fire-prone areas, and shifts in agricultural zonation.<br />

In one northern valley where water was being diverted for hydroelectric purposes,<br />

informants discussed their concerns about the lack of sufficient water for household<br />

use and for irrigation of mid-elevation crops. In other areas, freshwater primarily<br />

from glacial melt was channeled for mining and industrial needs that would otherwise<br />

be used by the rural populations for agriculture and consumption. A subsidiary<br />

concern is that with glacial retreat, increased destabilization of slopes might occur,<br />

causing natural hazards, such as landslides and mudflows, as has occurred in<br />

the past (Lliboutry et al., 1977; Kaser and Osmoston, 2002). Local communities<br />

are vulnerable to these conditions. Regional and national institutions will need to<br />

respond to damage when it occurs and to plan for increased risks in the future.<br />

Regional institutions stabilize and drain glacial lakes, as well as reinforce dams.<br />

However, what are lacking are appropriate social mechanisms, such as regional and<br />

community-based warning systems, logistical emergency plans, infrastructure, and<br />

financial support.<br />

Although current research efforts can monitor runoff from the glaciers of the<br />

Cordillera Blanca for local needs (Kaser et al., 2001; Mark and Seltzer, 2003), a<br />

scale-related disjuncture exists between the scientific-technical studies that examine<br />

hydrologic resources of the Cordillera Blanca, the national institutions involved<br />

in water use and planning, and the demands and needs of local populations. Private,<br />

transnational corporate institutions with political and economic clout have<br />

access to water resources and are assured of compliance from national and regional<br />

institutions that manage HNP’s natural resources and protect its biodiversity. Communities<br />

have protested and petitioned to regional officials for better safeguards of<br />

community water rights. Informants commented in interviews and in public meetings<br />

that, given the private, regional, and national interests associated with water<br />

usage, the local communities will be the first affected by water scarcity and imposed<br />

restrictions, while potentially the last to have their needs addressed. Due to previous<br />

political situations, there is a legacy of mistrust and apprehension that many


KENNETH R. YOUNG AND JENNIFER K. LIPTON 73<br />

of the campesinos have toward the national and corporate interests involved with<br />

water resources. What remains to be seen is how these multiple institutions will<br />

collectively negotiate the potential impacts of hydrologic resource scarcity with<br />

impending (and continuing) climate change.<br />

Differences in perspectives, by individual and community institutions, for tenure<br />

and rights to highland resources of the “protected” mountain landscape are critical<br />

factors in the conservation of biodiversity in the Cordillera Blanca. Although the<br />

boundary of the national park has been established for 30 years, from a local livelihood<br />

perspective, individual and community needs take precedence when it comes<br />

to resource use. Individuals and representative community institutions surrounding<br />

HNP consider a number of the regional and national institutions affiliated with<br />

conservation as “hypocritical,” “corrupt,” and “paternalistic,” as striving to satisfy<br />

the centralized Lima economy and political needs before they address local conditions.<br />

Despite a norm of some internal disharmony amongst peasant community<br />

members, they share an overriding bond of collective action, trust, and community<br />

cohesion that supercedes the external authority imposed by regional and national<br />

conservation institutions. Elements of trust and collective action are what facilitate<br />

adaptive capacity and governance in social ecological systems. In one of the sites<br />

where interviews were carried out, informants were asked to state the organizations<br />

and institutions that they were directly involved with: 82% participated in at least<br />

seven different committees or groups. The predominant committees that had membership<br />

were concerned with agricultural lands, pastures, forests, irrigation, sports,<br />

health, and the community kitchen. Community institutions are viewed as having<br />

served as reliable social scaffolding to provide support, information, and expertise<br />

to members and affiliates when other external institutions have been unreliable,<br />

self-interested, or unavailable. Based on these and other observations, our conclusion<br />

is that the community level of adaptability and resiliency, aided by clever<br />

land-use decisions by households, will allow many rural <strong>Andean</strong> residents to cope<br />

with climate change.<br />

Regional, national, and international institutions interested in preserving biodiversity<br />

conservation in the face of climate change, for example by establishing conservation<br />

corridors and additional protected areas, will need to address and surmount<br />

embedded negative perspectives to achieve long term goals and plans. Piecemeal<br />

approaches, historical legacies, and improper management of conservation areas in<br />

<strong>Andean</strong> areas have been the result of centralization and lack of funding for projects,<br />

despite ample local technical expertise and knowledge. For example, local community<br />

actors are in the position of deciding the status of the Cordillera Huayhuash<br />

as a potential protected area and are involved in asserting local level agendas into<br />

future national conservation planning and institutional structures. However, there<br />

is a divergence of interests and agendas among different sociopolitical actors in the<br />

Cordillera Huayhuash. Fujimori-era neoliberal economic reforms attracted foreign<br />

mining investors to the region, and while the mines offered jobs and opened up new<br />

roads, communities also perceived a state-mandated loss of access to the land from


74 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

national or external interests. Local institutions organized by charismatic community<br />

leaders are in the process of facilitating demands to maintain autonomy over<br />

the landscape and natural resources of the Cordillera Huayhuash, whether it is for<br />

conservation or tourism purposes or for the extraction of mineral resources. Vocal,<br />

self-motivated familial networks organize to actively declare their positions while<br />

other networks seek to develop bridges and coordinate activities between multiple<br />

communities and exogenous institutions. The role of trusted external sympathetic<br />

actors to assist in the adjudication, translation of policy, and technological expertise<br />

is important to the adaptive capacity of the institutional structures involved in<br />

conservation and resource governance in the Cordillera Huayhuash.<br />

With local, regional, national, and international institutions and actors establishing<br />

coordination and information networks, the future of the Cordillera Huayhuash<br />

might be developed as part of both a horizontal and vertical conservation corridor,<br />

connecting to HNP (and potentially to other protected areas), and providing an<br />

altitudinal range within the protected area that would allow for internal ecological<br />

shifts. Local communities however are positioned as the ultimate decision-makers<br />

for the governance of this area as a conservation zone. This highland area facilitates<br />

an interconnection of the high <strong>Andean</strong> environment while directing management<br />

towards culturally selected agro-biodiversity and native endemic flora and fauna<br />

diversity along the altitudinal and ecological gradients. However, more research on<br />

land cover needs to be completed to obtain a better perspective of how variables of<br />

climate and biodiversity are changing in the tropical alpine environment. Highland<br />

areas that will be more available to tourism, road development, and disturbances<br />

are also susceptible to dynamic changes pending climate change. Therefore, better<br />

inventories and studies on the feedbacks between community decisions and<br />

land cover should be undertaken. Conservation mechanisms, such as offered by<br />

the designation of the Cordillera Huayhuash as a protected area, and the promotion<br />

of transparent, innovative formal and informal institutional linkages provide a<br />

supportive framework that could attend to climate change implications.<br />

3. <strong>Andean</strong> Land Use and Climate Change<br />

As noted in the case study, spatial heterogeneity characterizes the biophysical realm<br />

of mountainous environments (Troll, 1968; Gerrard, 1990; Körner, 1999). The tropical<br />

Andes have particularly long, complex environmental gradients associated with<br />

elevation because mountaintops can exceed 6000 m with adjacent valley bottoms<br />

3000–4000 m below. The outer flanks of the Andes connect contiguously to Pacificand<br />

Atlantic-drainage basins and descend to or near sea level. Intermontane valleys<br />

often have drier environments due to rain shadow effects, thus juxtapositioning varied<br />

humidity zones within steep elevational gradients. Underlying this topographic<br />

variation are substantial intraregional differences in bedrock, from limestones or<br />

andesites to a great variety of metamorphic rocks. The resulting soil differences can


KENNETH R. YOUNG AND JENNIFER K. LIPTON 75<br />

at times create local-to-regional variations in the color and physical and chemical<br />

characteristics of surface soils.<br />

The steep elevational gradients of the Andes create zonal changes in ecological<br />

conditions related to altitude. It is possible to distinguish plant formations (ie. grasslands,<br />

shrublands, forests) and vegetation types along these gradients. Grasslands<br />

and other low herbaceous plant communities are found above about 3600 m. A<br />

broad ecotonal area from that elevation down to as low as 1000 m alternates between<br />

forests and shrublands as predominant land cover types. The shrublands are<br />

found in semi-arid to seasonally dry areas, while forests are found in wetter areas<br />

or are restricted to protected ravines in drier places (Young and León, 2001). The<br />

distribution of biological diversity follows these trends (Gentry, 1995).<br />

Young (1998) evaluated some of the biogeographical consequences of the inherently<br />

complex environmental mosaics of <strong>Andean</strong> landscapes that are further<br />

subdivided and pushed toward coverage by nonforest vegetation due to long-term<br />

human influences. <strong>Andean</strong> forests are often small, surrounded by nonforest vegetation,<br />

exposed to altered physical conditions along forest margins (ie. edge effects),<br />

and isolated from similar forest tracts. This habitat fragmentation would exclude<br />

most forest interior species, for example certain kinds of raptors (Thiollay, 1996)<br />

and frogs (Toral et al., 2002). The forests and their biota may be exposed to chronic<br />

anthropogenic disturbances from fires, firewood harvesting, grazing, and cutting to<br />

establish agricultural plots. The plants and animals in the most heavily altered landscapes<br />

would seemingly need to be a robust subset of the original flora and fauna.<br />

Young (1998; also Young and Keating, 2001) listed some of the characteristics<br />

expected for the plants, including shade intolerance, ability to resprout following<br />

mechanical damage, and seed dispersal by small birds or wind. The animal types<br />

expected to be successful would be small-bodied birds that are granivores or frugivores,<br />

and inconspicuous, small mammals. Larger native animals either are more<br />

common in forest edge or shrub habitats, such as the white-tailed deer (Odocoileus<br />

virginianus), or are currently restricted to high mountain environments (<strong>Andean</strong><br />

huemul deer, vicuña), isolated shrublands (guanaco, Lama guanicoe), or distant<br />

forests (<strong>Andean</strong> bear).<br />

Given these difficulties, what can currently be said about likely climate change<br />

consequences? Warmer temperatures (Vuille and Bradley, 2000) would tend to<br />

shift plants, animals, and ecosystems upslope (e.g., Inouye et al., 2000; Enquist,<br />

2002), except when other environmental controls, such as local edaphic conditions<br />

(soil depth or drainage) or aspect-related differences (e.g., south-facing slopes,<br />

windward-facing slopes, etc.) prevent those shifts from occurring. Increased carbon<br />

dioxide levels may promote primary productivity of some ecosystems (Woodward,<br />

2002), but the effect appears likely to be dampened due to other limiting factors<br />

such as the availability of soil nutrients and the downward adjustment of number<br />

of stomata per unit area in developing leaves exposed to higher ambient carbon<br />

dioxide levels (Körner, 2000; Brownlee, 2001). In some cases, aquatic and wetland<br />

habitats may be expected to increase in size, particularly in the lakes and high


76 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

elevations of retreating glaciers. However, with a potential for increased sediment<br />

load, these environments might undergo other permutations having a direct and<br />

generally negative impact on aquatic life.<br />

A valuable component of <strong>Andean</strong> biodiversity was actually created by the domestication<br />

of native plants and animals during the evolution of land-use systems<br />

(Piperno and Pearsall, 1998). Most of this additional diversity is genetic in nature,<br />

with new gene complexes maintained in domesticated varieties that are distinct<br />

from their wild relatives growing nearby in habitats adjacent to the agricultural<br />

fields. Farmers have developed different varieties with particular desired properties.<br />

This process appears to have reached its zenith in the potato cultivation of the<br />

central Andes. One field may have 40–60 potato variants—each with a common<br />

name and often each with a unique use, because of taste or texture, or because of<br />

the growth conditions required (Brush et al., 1995; Zimmerer, 1996; Ochoa, 1999;<br />

Hijmans and Spooner, 2001). This type of biodiversity can be kept intact through<br />

interventions such as cold storage gene banks, although at great expense. In situ<br />

approaches to maintaining genetic diversity are dependent on farmers who keep<br />

planting and utilizing their full range of crop varieties (e.g., Perales et al., 2003).<br />

This practice allows for further evolution and cultural selection of the genotypes. It<br />

requires either intact traditional systems of seed/tuber interchanges or some kind of<br />

subsidy. Some genetic biodiversity programs are beginning to contemplate climate<br />

change implications. All such efforts are useful because a wider base of genetic<br />

diversity is kept available for future generations.<br />

Yet another aspect involving climate change is the adaptive nature of many of<br />

the <strong>Andean</strong> land-use systems. As we reveal in the case study above, farmers in<br />

the Andes are attuned to biophysical patterns, processes, and the resulting spatial<br />

heterogeneity. Agro-pastoral livelihoods are realized in patchy, complex environments<br />

that in some land-poor communities may be sub-optimal for various kinds<br />

of production (Denevan, 2001; Zimmerer, 2003). <strong>Andean</strong> farmers lower risks and<br />

accommodate ecological variability by having diversified household economies,<br />

utilizing multiple fields, and selecting numerous seed and crop types. One household<br />

might graze livestock in upper-elevation communal land, grow potatoes in their<br />

fields at the highest elevations, maize and wheat at intermediate levels, and cassava<br />

in valley bottoms (Gade, 1975; Brush, 1976; Young, 1993). Still other fields are<br />

created out of wetlands that are modified with drainage canals and raised seedbeds<br />

(Zimmerer, 1991).<br />

Knowledge of how to manage useful plants and animals is embedded within the<br />

agricultural practices and traditions that are passed on to children in households<br />

(Young, 2002). The social networks established in youth are often the same that are<br />

implemented later in life to organize collective harvesting or tillage, to maintain<br />

irrigation systems, and to allocate land-use rights. Often these systems nestle hierarchically<br />

within municipal, district, and provincial governing frameworks. But<br />

this need not be the case if, for example, the people living in towns do not consider<br />

land-use decisions in outlying rural areas to be worthy of attention. Or there may


KENNETH R. YOUNG AND JENNIFER K. LIPTON 77<br />

be ethnic or age differences between rural inhabitants and those wielding political<br />

power that negate tendencies to collaborate. Also, as was noted in and around HNP,<br />

increased out-migration of young people seeking job opportunities and experience<br />

in urban centers can lead to younger generations not participating in household or<br />

community decision-making institutions and local political networks. This might<br />

be one of the inevitable consequences of urban areas having better opportunities<br />

for education, health care, and jobs.<br />

As was the situation in the case study, probably the biggest ongoing concern in<br />

much of the rural Andes is access to irrigation water. Often elaborate local systems<br />

are in place to assure some sort of equitable distribution of water in relation to need<br />

(Mitchell and Guillet, 1993; Gelles, 1999). For example, time allocation of waterflow<br />

to individual farm plots within an irrigation network, known as mitas de agua,<br />

can be controlled through community-based institutions responding to household<br />

demands and personal interests. This kind of social capital is also critical for extending<br />

planting seasons beyond the two to eight months a year when the rains fall and<br />

for good development of particular water-demanding crops. Farmers avoiding risk<br />

might determine which fields to cultivate depending upon household water rights,<br />

thereby providing a type of insurance in dry years. The retreat of <strong>Andean</strong> glaciers<br />

signals a negative glacial mass balance, which may be waning due to less precipitation,<br />

less snow accumulation in the form of glacial firn, more melt due to higher<br />

temperatures, more ablation, or a combination of these factors. Vuille et al. (2003)<br />

concluded, from an admittedly sparse regional data set, that glacial retreat in the<br />

tropical Andes is most likely due to warmer temperatures at high elevations and possibly<br />

higher relative humidity. This regional conclusion, however, does not exclude<br />

other changes from being important in particular areas: Kaser and colleagues (Kaser<br />

et al., 1990; Kaser, 1999; Kaser and Osmaston, 2002) have stressed the likelihood<br />

of slight decreases in precipitation being important for tropical snowline retreat.<br />

Over the next few decades, increased demand for water from growing population<br />

centers and industry will affect farmers in rural areas, adding to the complexity<br />

of making predictions. It is not yet clear how urbanization and globalization will<br />

affect land-use strategies and rural natural resource management. While larger and<br />

more accessible markets provide potential incentives to intensify agriculture and<br />

increase <strong>Andean</strong> production of commodities, those same national and transnational<br />

socioeconomic forces connect to other producers, perhaps in other countries, who<br />

may be able to produce the same product more economically or efficiently (e.g.,<br />

Bebbington, 2001).<br />

The community-based systems that regulate or control access to water also function<br />

in regards to other natural resources, including the use of lands for grazing or<br />

planting (Flores Ochoa, 1977; Guillet, 1981). Community formal and informal institutions<br />

are built upon pre-European and Spanish-derived land-use practices. The<br />

governance of community lands entails making decisions about the type of agricultural<br />

production, for example, whether it is for local consumption or for larger<br />

markets. Roads, particularly if paved, dramatically change options by transforming


78 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

accessibility and cost-effort assessments by rural inhabitants selling excess production<br />

or even growing items specifically for urban markets, as seen in the example<br />

of Sr. Valverde’s land use decisions. In some cases, this transformation has<br />

shifted tenure to more privately owned lands. This shift was especially pervasive<br />

in the 1990s when neoliberal reforms became an international mode and expectation<br />

(Inter-American Development Bank, 1997; Thorp, 1998; Bury, 2004), with<br />

governments allowing for direct purchase of lands in places where once such transactions<br />

would not have occurred. Another new socioeconomic phenomenon is the<br />

increased importance of remittances from family members living in cities or other<br />

countries (e.g., Jokisch, 2002). This change can alter the intensity and nature of<br />

land use, especially if valuable lands are controlled by absentee owners.<br />

<strong>Andean</strong> lands have supported millions of people and vigorous societies throughout<br />

millennia. Local observations and perceptions of climate phenomenon are embedded<br />

in familial and community traditions and are an important aspect of technical<br />

knowledge. The farmers near Lake Titicaca are able to forecast ENSO (El Niño<br />

Southern Oscillation) events based on the visibility in the night sky of certain constellations<br />

(Orlove et al., 2000). Those events affect the high elevation lake basin<br />

with prolonged droughts, so several months warning helped local people make decisions<br />

on when and where to plant. Understanding these and other traditional and<br />

folk-based indicators of these climate variations and landscape-change perceptions,<br />

such as documented in the case study, will be important for studies of both <strong>Andean</strong><br />

land-use management and climate. Responses of greater exploitation would<br />

be expected from people who are observant of the rapidity of recovery of vegetation<br />

following a disturbance or the swiftness of repopulation of an area by some<br />

animal species affected by harvesting or other population alteration. This feedback<br />

could result in the masking of biotic responses to climate change. It might require<br />

field experiments or studies in natural areas kept free of land use to distinguish<br />

climate-caused responses from those same responses as mediated, lessened, or increased<br />

by concurrent human exploitation or manipulation. An important question<br />

is, Are the progressive or more extreme changes to be expected in the next several<br />

decades in the Andes within the adaptive capacity of these people and their<br />

land-use systems?<br />

4. <strong>Andean</strong> Rural Decision Making and Multi-Scale Governance<br />

A critical aspect for the analysis of adaptive capacity for environmental decision<br />

making is the assessment of the institutions that people are engaged in, with<br />

consideration to the historical, familial, and community contexts from which they<br />

operate (Blaikie and Brookfield, 1987; Robbins, 1998; Ostrom et al., 1999). As<br />

we showed in the case study and as reported from throughout the tropical Andes,<br />

many of the agro-pastoral activities and decisions are made and implemented on a<br />

household, extended family, and community level (Brush and Guillet, 1985; Mayer,


KENNETH R. YOUNG AND JENNIFER K. LIPTON 79<br />

2002). Transnational linkages and market forces are increasingly integrating once<br />

geographically and socially marginalized <strong>Andean</strong> communities into larger networks<br />

of formal and informal institutions (Price, 1994; Bebbington, 2001; Miles, 2004),<br />

thereby having an impact on household decisions regarding resource management.<br />

In an <strong>Andean</strong> household, distinct activities and livelihoods are dependent upon<br />

labor and priorities, evaluated in relation to season, access to different ecological<br />

zones, and the ease or cost of connections between rural and urban centers. This<br />

level of decision making in turn influences the types of participation in various<br />

institutions operating on other scales. Thus, all these scales must be considered<br />

when assessing the potential consequences of climate change to livelihoods and<br />

biodiversity.<br />

4.1. HOUSEHOLDS<br />

Many highland families engage in multiple livelihood strategies in order to satisfy<br />

and sustain basic needs. A household might be involved in field agriculture,<br />

raising livestock, or wage labor all at the same time. Kin associations, reciprocity,<br />

community duties and responsibilities all factor into one household’s daily and<br />

seasonal functions (e.g., Allen, 1988; Mayer, 2002). Scheduling tasks, measuring<br />

inputs, trade-offs, assessing risks and other opportunities are a few of the agricultural<br />

factors that concern households (Browman, 1987; Netting, 1993; Zimmerer,<br />

1996). Concurrently, agricultural decisions within a household can be based<br />

upon internal domestic factors such as age and gender, labor availability, priorities,<br />

and acceptance or avoidance of risk (Valdivia et al., 1996). Likewise external<br />

factors, including political stability, can be influential. Also a large degree of individuality<br />

exists, with some households specializing in certain tasks or products<br />

due to the experiences of household members. Individualism in communities can<br />

lead to increased pressure on agricultural land as people shorten fallow periods<br />

for maximum production and gradually depart from communal controls (Mayer,<br />

2002; Zimmerer, 2002). Economic choices required for agricultural intensification<br />

and the production of off-season crops can also compete with the labor demands<br />

necessary to maintain crop diversity and selection meant for subsistence.<br />

Under the influence of changing biophysical and environmental conditions, the<br />

information available to households about type and quantity of inputs will be a<br />

limiting factor.<br />

Pastoral work often involves multi-tasking and is frequently carried out by either<br />

gender depending upon labor availability (Brush, 1987). In a number of situations,<br />

women and children combine pastoralism with other responsibilities of child rearing,<br />

plant collection, knitting, and food preparation for home or market. For the<br />

agro-pastoralists living within the study area multi-tasking and livelihood diversification<br />

are important aspects of adaptive capacity because of the risk of dependency<br />

on any one resource. Livelihood diversification contributes to the resiliency of


80 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

<strong>Andean</strong> households by creating access to potential subsidiary channels of income.<br />

Combining market oriented and subsistence agriculture, the export of meat and<br />

wool, renting of communal lands, economic assistance through familial bonds, or<br />

engaging in wage labor can be essential for household self-sufficiency. Diversifying<br />

can buffer financial shortfalls when unpredictable events or disturbances occur<br />

(Valdivia et al., 1996). Frequently, men will seek supplemental wages, working for<br />

larger farmsteads, on construction activities, or on corporate farms, and tend to their<br />

own family plots with available time.<br />

The ongoing transnational influences that are progressively integrating rural people<br />

into market-based economies will have a direct effect upon household resource<br />

management and technologies for agricultural production (Bebbington, 2001). Culturally<br />

created biological diversity, such as the numerous potato varieties, will likely<br />

only be maintained through households and other local social institutions that mediate<br />

and implement agricultural and pastoral land-use decisions. Farmers consciously<br />

select traditional crops, such as tubers and maize, to ensure a diverse and vigorous<br />

seed-pool that does not require the high inputs that commercial production<br />

of wheat, barley, oats, and potato varieties need (Zimmerer, 1996; Mayer, 2002).<br />

Thus, the question in reference to climate change is, Does the capacity for invention<br />

and adaptation still exist among these informal and local institutions? The stressors<br />

on household resource management certainly include educational and other attractions<br />

that pull youth away from pursuing rural livelihoods, plus the availability of<br />

imported food products, which compete with the sales of extra produce from <strong>Andean</strong><br />

households. We have talked with farmers who consciously maintain high crop<br />

diversity, partly out of tradition, partly from a sense of cultural or local pride. They<br />

mentioned neighbors that had switched to cash crops, such as alfalfa to be sold for<br />

fodder and barley for beer-making companies. Farmers also contrasted the dozens<br />

of potato varieties they may maintain on their lands with the actions of others who<br />

have converted to the much more limited set of improved varieties promoted by<br />

governmental offices.<br />

4.2. COMMUNITIES<br />

Community and inter-community relationships create and then retain agricultural<br />

biodiversity. However, these institutions are experiencing stresses and undergoing<br />

transformations. Although, in general, urbanization and globalization lessen agricultural<br />

biodiversity, several countervailing processes do occur. A resurgence of<br />

ethnic or regional pride in the Andes often includes assertions that native plants<br />

and animals should receive special attention. In areas of Peru and elsewhere, rural<br />

communities show great interest in participating in community seed exchanges<br />

and agricultural fairs, some of which is promoted by regional and international<br />

NGOs concerned with agro-biodiversity. Increased access in the difficult terrain<br />

of the Andes by an improved transportation network means that there is more


KENNETH R. YOUNG AND JENNIFER K. LIPTON 81<br />

potential for peoples of similar environments, separated by high mountain barriers,<br />

to come into personal contact. Improved roads and transportation networks allow<br />

for capital-intensive agricultural production in more competitive and interlinked<br />

markets (e.g., Sierra et al., 2003). Even extremely remote areas receive radio programs<br />

or satellite television broadcasts. Frequently, early morning programming<br />

includes information meant for agriculturalists.<br />

In an <strong>Andean</strong> community, social networks built upon familiarity, close contact,<br />

and trust, are indispensable and create social capital (Ostrom, 1990; Bebbington,<br />

1996, 1997). Networks of reciprocity and shared goals regarding resources are an<br />

integral aspect of <strong>Andean</strong> society (Mayer, 2002). Households that participate in<br />

informal or formal resource-based institutions may express their obligations to the<br />

entire community by helping to look after community resources. Of course, participation<br />

might not always be the case. Individuals and households may have personal<br />

and competing interests, such that personal or familial agendas can circumvent<br />

community directives (Orlove, 1977). The obligations intrinsic in informal institutions<br />

and peer groups may demand commitments away from the habitual tasks<br />

of the household. Disharmony or competition amongst members of the same community<br />

can occur and may affect communal land-use management or result in the<br />

disrespect of communal norms. Or other changes may erode traditional claims on<br />

individual actions.<br />

Participation in communal institutions governing land use, social activities, or<br />

labor needs is invaluable not only for the community, but also for the individual<br />

household (Mayer, 2002). In the tropical Andes, many informal institutions operating<br />

within a community consist of the personal arrangements and bonds that<br />

are formed through reciprocity, where unspoken agreements and contracts between<br />

people are fulfilled through mutual assistance. This was demonstrated in the case<br />

study by the high number of community organizations that individuals participated<br />

in. Collective action as mediated through community institutions may have<br />

in place mutual coercion mechanisms to assist with compliance (Robbins, 1998).<br />

In the Andes, commitment to communal activities, often in the form of labor exchanges<br />

known as minkas, collectively guarantees access to resources, participation<br />

in fiestas, and can be considered a form of social insurance (Mayer, 2002). Spending<br />

a day or two working in a communal labor effort known as a faena can ensure that<br />

access to water rights, land, or help with a harvest will be available to a household<br />

or extended family.<br />

Throughout the world, common property resources are frequently governed<br />

and managed by institutions, which increasingly are subject to extra-local<br />

dynamics (McCay and Acheson, 1987; Guillet, 1992; Robbins, 1998; Mayer, 2002;<br />

Zimmerer, 2002; Adams et al., 2003; Dietz et al., 2003; Giordano, 2003). Important<br />

<strong>Andean</strong> pastoral production zones, consisting of tropical alpine grasslands and<br />

grass-shrub communities, are often governed as communal grazing areas through<br />

multi- or intra-community-based management (Orlove and Godoy, 1986; Knapp,<br />

1991; Zimmerer, 2002). Many upper elevation communities that are land poor or


82 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

do not have access to other ecological zones rely upon income-generating resource<br />

management strategies of essential products, such as meat, wool, and the sale of<br />

milk for cheese production. Formalized community institutions, such as resourceuser<br />

groups, act together in order to manage and coordinate resources within one<br />

community. Conducting inventories; making assessments; compiling complaints or<br />

concerns from other households; handling allotment, disbursement, and management<br />

of resources are some of the responsibilities involved in communal resource<br />

management.<br />

Participation by at least one member from each sector of a community in the<br />

resource-user group is important for facilitating the transfer of information and for<br />

maintaining access by that sector to resources in different ecological zones, provided<br />

the community has such access. For example, a pasture committee might include<br />

at least one individual from each community sector to assist with management of<br />

communal lands located in upper elevations. As many community institutions function<br />

out of a commitment to shared goals and values, they also are shaped by the<br />

cultural norms, information exchanges, and perceptions of the participating members<br />

(de Janvry et al., 1993; Adger, 2000). Through these community institutions,<br />

the knowledge of landscape change and placement of new rules for governance of<br />

resources could be encouraged and adapted if it is deemed important by community<br />

members. As we found in our case study, adaptive governance is strengthened by<br />

the ability to maintain a checks-and-balance system or transparency among local<br />

networks.<br />

Community social networks often operate to transmit news. They are usually<br />

the first informal institutions to disseminate information about market prices, opportunities<br />

or district-wide agendas (Rodríguez and Pascual, 2004). They are also<br />

the first to be contacted by district and regional governmental institutions for proposed<br />

work or resource-based projects. Increasingly, community organizations are<br />

actively seeking and negotiating relationships with other institutions at other scales<br />

for economic and technical advantages (Bebbington, 1997). This aspect of community<br />

governance is especially important in projects that involve a regionally<br />

oriented approach for addressing climate change issues. However, there may be a<br />

great deal of skepticism directed toward extra-local institutions, particularly if the<br />

extra-local interests are going to impinge upon current resources or livelihoods, as<br />

was exemplified for the lands within Huascaran National Park. In some situations<br />

where community institutions are opposed to regional or national projects, organized<br />

activism occurs, instigating strikes and general resistance (e.g. Healy, 1991).<br />

4.3. REGIONAL, NATIONAL, AND INTERNATIONAL<br />

Regional formalized institutions, both governmental and nongovernmental, that are<br />

focused on resources in the Andes must typically have community contacts and access<br />

to networks in order to successfully carry out projects and policies. Establishing


KENNETH R. YOUNG AND JENNIFER K. LIPTON 83<br />

linkages with community institutions and making strategic alliances are political<br />

processes that can strengthen the goals of a broad-based project (Agrawal, 1997;<br />

Healy, 2001). Depending upon their particular agendas, regional formalized institutions<br />

may sponsor educational and technical training for the betterment of<br />

production or for alternative forms of production. This training is often sought<br />

out by communities, as was the case for farmers seeking access to improved crop<br />

varieties in northern Peru. Some community institutions have intentionally positioned<br />

themselves to access national or international technical or financial resources<br />

(Bebbington and Perreault, 1999).<br />

An important part of the relationship between extra-local institutions and communities<br />

is the ability to maintain and follow through on commitments made. As<br />

regional programs seek to work on a larger scale, they tend to overextend resources,<br />

personnel, and time commitments. Unfortunately, our informants have commented<br />

on how this propensity for over-extension can lead to a lack of follow-through on<br />

proposals and a resultant waning of trust and interest in projects at the community<br />

level. These efforts can also be hampered when neighboring communities compete<br />

for limited project resources. In particular, during the 1990s, many governmental<br />

institutions, typically under funded and short-staffed, were besieged by neoliberal<br />

reforms that sought policies of privatization and natural resource extraction<br />

(Escobar, 1995; Stiglitz, 2002; Zimmerer, 2002). These reforms resulted in offices<br />

and personnel with inadequate resources and technologies for maintaining or<br />

carrying out projects.<br />

With the global phenomenon of climate change forcing local and regional<br />

changes, institutions of the tropical Andes would have advantages in becoming<br />

globally networked. Nested institutional relationships at the local, regional, national<br />

and international level will be necessary for creating, implementing, and monitoring<br />

climate change policies and projects. For example, regional institutions have<br />

benefited from connections with international institutions. Recently, committed individuals<br />

working in the upper and middle tiers of governmental programs have<br />

taken a more proactive approach to solidify international and national networks<br />

to achieve goals for both research and policy-making regarding climate change.<br />

In some cases in Peru, administrators have sought out linkages with foreign researchers<br />

and NGOs to bring foreign funding, skills, and global media attention<br />

to the condition of <strong>Andean</strong> glaciers with the intent of reinforcing and enhancing<br />

existing national institutional structures and infrastructure. One goal for these connections<br />

is to promote better coordination and directives regarding research and<br />

planning for global climate change in an international context. These new linkages<br />

could position local, regional, and national institutions to adequately prepare<br />

for water conservation, and hydroelectric demands and respond to natural hazard<br />

situations. Combined with grassroots support, flexibility will add to the decision<br />

making process.<br />

With democracy, national governmental institutions receive their mandates from<br />

the voting patterns of large urban populations, with elected and career administrators


84 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

serving as mediators. The rights and needs of minorities, consisting of ethnic and<br />

often marginalized, dispersed rural inhabitants, have been historically neglected in<br />

Latin America (Maybury-Lewis, 2002). Policies and programs for climate change<br />

will need to rely upon the nested institutional information networks. However,<br />

these are often bureaucratically dominated by corporate or political interests. For<br />

example, in the study area and many other locales, resources such as minerals,<br />

water, and/or timber are often extracted by exogenous transnational institutions<br />

with powerful political connections. As a result, a spatial and socioeconomic disconnect<br />

from landscape changes occurring in the rural highlands may take place.<br />

The institutional governance that accompanies climate change phenomenon necessitates<br />

ongoing flows of information, institutional transparency, and cooperation<br />

between the many entities involved. Possibly, rural development banks could provide<br />

short-term loans and agricultural extension agents could facilitate suggestions<br />

for adaptations; neither system however, is well developed in the tropical <strong>Andean</strong><br />

countries.<br />

Less clear is how regional, national, and international institutions interface with<br />

processes affecting sustainability concerns. This lack of clarity is not only due to<br />

a paucity of relevant research, but is also a result of significant region-to-region<br />

differences in predominant biophysical and land-use systems. The people near<br />

Sajama National Park in northwest Bolivia utilize lands above 4200 m, so their<br />

needs must be met from livestock raising and trading for other staples. The farmers<br />

near Otavalo, Ecuador not only raise a wide variety of crops for sale in Quito,<br />

but are tied to international economic forces through commodity production of<br />

handicrafts and remittances from other countries (Korovkin, 1998; Bebbington,<br />

2000). The national institutions involved with farming systems embed regional<br />

disparities in their structures, and cohesive national priorities seldom endure as<br />

they are prone to reversals when new political parties take office. Even wellmeant<br />

international efforts may be carried out in ways that act to reduce local<br />

adaptive capacity.<br />

In addition to rural livelihoods, the institutions involved with the conservation<br />

of biological diversity need to be evaluated. These institutions include the protected<br />

areas systems of the <strong>Andean</strong> countries. They are especially weak in implementing<br />

policy and programs, not at the national and international levels, but at the local and<br />

regional levels (Young and Rodríguez, in press). There may be compelling global<br />

and national reasons to protect a particular place or species, but convincing local<br />

people to accept the risks or costs to do so is fraught with contradictions (Young,<br />

1997; Young and Zimmerer, 1998; Zerner, 2000; Terborgh et al., 2002). Sometimes<br />

the success or failure of a particular national park or nature reserve to conserve<br />

native ecosystems is due to the managerial, technical, and social capacity of the<br />

on-site park administration. If personnel are charismatic, local communities can be<br />

convinced to merge their goals with park objectives for mutual benefit. But past bad<br />

experiences, incompetent personnel, or some other historical or personal legacy can<br />

make this impossible.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 85<br />

The dilemma for national and international institutions involved in biodiversity<br />

conservation in the <strong>Andean</strong> highlands is complicated by the fact that no<br />

protected area has been evaluated rigorously for climate change consequences.<br />

Thus, little scientific basis exists for taking a particular action, except for extrapolations<br />

of approaches from more general concerns (e.g., conservation corridors)<br />

or other geographical regions (Western and Wright, 1994; Soulé and Terborgh,<br />

1999). For example, most organized park-people programs in the tropical Andes<br />

are premised on possibilities of sustainable extraction of natural resources.<br />

Difficulties are numerous, mainly because sustainable extraction rates may be<br />

fundamentally unknowable if carrying capacities are not set by equilibrial ecological<br />

processes and if there are stochastic and/or progressive changes in physical<br />

environmental conditions (Reynolds et al., 2001). Philosophically, these conservation<br />

efforts do not include protection and management goals that cover the<br />

other non-economic values of biodiversity (Perlman and Adelson, 1997). One<br />

such value connects directly to climate change and is part of the founding legislation<br />

of all the nationally protected area systems of the Andes: those areas<br />

are to be used for scientific research to better understand the functioning of<br />

natural environments.<br />

Although conservation corridors that can connect together the parks and reserves<br />

of national and regional protected-area systems have been amply discussed from<br />

both theoretical and practical viewpoints (Soulé and Terborgh, 1999), their applications<br />

for climate change are less well developed. Many of the conservation corridor<br />

studies concentrate on the species level, focusing on an indicator species that would<br />

benefit from habitat conservation (Gutzwiller, 2002). We suggest that conservation<br />

corridors could potentially serve two different goals: connecting together similar<br />

environments across relatively large distances and providing conservation protection<br />

along steep environmental gradients, such as those associated with altitudinal<br />

change.<br />

The former are herein called horizontal conservation corridors and would potentially<br />

lessen the isolation of organisms and permit interchange of individuals and<br />

genes, both key conservation concerns (Frankham et al., 2002). The effective area of<br />

protected habitat is increased, although there are many real-world concerns of how<br />

wide these corridors could or should be, how restrictive their land use or tenure can or<br />

must be, if the terrain between protected areas is such that environmental conditions<br />

do not change enough to create pre-existing biogeographical barriers, and if species<br />

of concern will even move through them. Vertical conservation corridors, on the<br />

other hand, are already contained within many of the large national parks of the <strong>Andean</strong><br />

countries. As an example, Table 1 shows that all but two of Peru’s national parks<br />

include elevational gradients of more than 3500 m, effectively connecting together<br />

in one conservation unit the lowlands to the highlands. This is an ideal way to buffer<br />

for unpredictable climate change consequences. In mountainous environments,<br />

such as in the Andes, it is likely that environmental conditions will shift upslope,<br />

for example with warmer environments replacing those now found near snowline.


86 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

High elevation ecosystems and species restricted to them may be lost from mountain<br />

ranges (Peters and Lovejoy, 1992). This would especially be the case for the<br />

nonglaciated cordilleras.<br />

Our suggested solution is to combine vertical and horizontal conservation corridors<br />

in national, regional, and continental planning. Currently, the national protected<br />

area system in Peru consists mostly of unconnected parks and reserves (Figure 1).<br />

The vertical corridors would connect disparate elevations and the horizontal corridors<br />

would provide additional habitat and connectivity among protected areas.<br />

As seen in the case of Peru (Table I), some of the other types of protected areas,<br />

for example the national reserves, are designed to feature one major environmental<br />

zone and so do not include much elevational or other variation (Rodríguez and<br />

Young, 2000). These areas would require both horizontal and vertical corridors to<br />

function effectively in the future.<br />

There are a number of international NGOs that are working to promote conservation<br />

corridors with the involvement of regional and national institutions. In one<br />

instance, Conservation International is organizing a series of activities in lands adjacent<br />

to formally protected areas in a plan meant to create the “Vilcabamba-Amboró<br />

Conservation Corridor” from southern Peru to central Bolivia on the eastern flanks<br />

of the Andes. For our case study, the Cordillera Huayhuash reserve would in effect<br />

form a conservation corridor with Huascaran National Park. In addition, corridortype<br />

areas might be implemented by enforcing pre-existing wildlife laws within<br />

their length and extent. For example, there are laws on the books in Peru that<br />

restrict hunting of deer and fishing of trout to certain seasons of the year, or that<br />

completely outlaw the killing of rare species. However, such laws are not effectively<br />

enforced. If they were to be applied in the long, but narrow areas that connect one<br />

conservation area to another, some conservation-corridor benefits would be gained<br />

without the need to create new legal mechanisms.<br />

We believe that the criteria suggested above for conservation corridors and<br />

for protected areas would build in enough connectivity and redundancy for most<br />

species. That said, we reiterate that the necessary research to classify the thousands<br />

of restricted-species of the Andes for their conservation risks is just beginning<br />

(Stattersfield et al., 1998; Valencia et al., 2000). In none of the recent studies was<br />

exposure to future climate change used as an evaluation criterion. We also feel that<br />

local perspectives and multi-scale institutional adaptations need to be incorporated<br />

into long term planning for the human and biophysical systems that will be affected<br />

by environmental shifts.<br />

5. An <strong>Andean</strong> Climate Research Agenda<br />

Much of the scientific research needed to address these concerns has yet to be done<br />

for the tropical Andes, although inspiration could be drawn from advances elsewhere<br />

(Mount, 1994; Jackson and Weng, 1999; Kappelle et al., 1999; Foster, 2001;


KENNETH R. YOUNG AND JENNIFER K. LIPTON 87<br />

TABLE I<br />

Nationally protected conservation areas of Peru, including national parks, reserves, and sanctuaries.<br />

Those areas with relatively large sizes (>100,000 ha) and large elevational ranges (>3000 m) already<br />

act as vertical conservation corridors, in the sense proposed here.<br />

Name Category Year Established Area (has) Elevational Range (m)<br />

Cutervo National Park 1961 2,500 2,350–3,350<br />

Tingo Maria National Park 1965 18,000 680<br />

Huascaran National Park 1975 340,000 2,400–6,768<br />

Cerros de Amotape National Park 1975 91,300 75–1550<br />

Rio Abiseo National Park 1983 274,520 350–4,350<br />

Yanachaga-Chemillen National Park 1986 122,000 250–3,700<br />

Bahuaja-Sonene National Park 2000 1,091,416 200–2450<br />

Cordillera Azul National Park 2001 1,353,190 150–2,320<br />

Manu National Park 1973 1,692,137 250–4,050<br />

Otishi National Park 2003 305,973 700–4,150<br />

Pampa Galeras National Reserve 1967 6,500 3,850–4,150<br />

Junin National Reserve 1974 53,000 4,008–4,125<br />

Paracas National Reserve 1975 335,000 9–786<br />

Lachay National Reserve 1977 5,070 100–750<br />

Titicaca National Reserve 1978 36,180 3810<br />

Salinas y Aguada Blanca National Reserve 1979 366,963 2,800–6,050<br />

Calipuy National Reserve 1981 64,000 400–4,000<br />

Pacaya Samiria National Reserve 1982 2,080,000 83–160<br />

Tambopata National Reserve 2000 274,690 200–400<br />

Allpahuayo-Mishana National Reserve 2004 58,070 104–185<br />

Huayllay National Sanctuary 1974 6,815 4,100–4,600<br />

Calipuy National Sanctuary 1981 4,500 400–4,000<br />

Lagunas de Mejia National Sanctuary 1984 690 0–50<br />

Ampay National Sanctuary 1987 3,635 2,780–5,235<br />

Manglares de Tumbes National Sanctuary 1988 2,972 0<br />

Tabaconas-Namballe National Sanctuary 1988 29,500 1,500–3,200<br />

Chacamarca Historical Sanctuary 1974 2500 4,112–4,438<br />

Pampa de Ayacucho Historical Sanctuary 1980 300 0<br />

Machupicchu Historical Sanctuary 1981 32592 1,800–6,250<br />

Bosque de Pomac Historical Sanctuary 2001 5887 100–150<br />

Laquipampa Reserved Zone 1982 11,347 200–2,550<br />

Pantanos de Villa Reserved Zone 1989 263 0–5<br />

Tumbes Reserved Zone 1994 75,102 0–900<br />

Algarrobal el Moro Reserved Zone 1995 321 20–50<br />

Chancaybanos Reserved Zone 1996 2,628 1,300–2,400<br />

Aymara Lupaca Reserved Zone 1996 300,000 3,825–4,500<br />

Gueppi Reserved Zone 1997 625,971 200–250<br />

Rio Rimac Reserved Zone 1998 28 km 220–1,000<br />

Santiago-Comaina Reserved Zone 2000 1,642,567 200–2,700<br />

Alto Purus Reserved Zone 2002 2,724,264 200–650<br />

Codillera de Colan Reserved Zone 2002 64,115 750–3,600<br />

Huayhuash Reserved Zone 2002 67,590 3,650–6,600


88 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Guisan and Theurillat, 2001; Walther et al., 2002; Malhi and Phillips, 2004; Thomas<br />

et al., 2004; Dockerty et al., 2005). For example, Walker and colleagues (Walker,<br />

2003; Walker et al., 2004) recently provided insights from household-level research<br />

in the Amazon that could also be adapted towards the <strong>Andean</strong> examples explored in<br />

this paper. They found that a behavioral model of the consequences of numerous individual<br />

household decisions in regard to land use could successfully reproduce the<br />

landscape transformations occurring during colonization and frontier expansion.<br />

<strong>Andean</strong> rural residents are making decisions in relation to shifting environmental<br />

conditions, to the lands and social capital available through community-level decisions,<br />

and to constraints imposed by market forces and governmental decisions<br />

(e.g., Swinton et al., 2003; Rodríguez and Pascual, 2004). Each household evaluates<br />

their vulnerability with respect to their goals and needs. While it is important to<br />

develop computerized models that can relate these factors to shifting land use and<br />

land cover, it is just as important for more empirical field work to be carried out to<br />

better characterize the range of adaptive practices and responses found along and<br />

across the <strong>Andean</strong> highlands (e.g., Vanacker et al., 2003).<br />

Murra (1975) and others (e.g., Stanish, 1992) pointed out that long-distance<br />

interchanges of products and information have anciently characterized how the<br />

Andes <strong>Mountain</strong>s were used by people, thus permitting existence in areas otherwise<br />

inhospitable for the production of certain necessities. Globalization now<br />

adds to those spatial linkages while accelerating rates of socioeconomic change,<br />

with implications for vulnerability to climate change (Adger, 1999; Metz, 2000;<br />

Griffin, 2003; Tanner, 2003; O’Brien et al., 2004; Young and Rodríguez, in press).<br />

Although the basics of how households are nested within increasingly complex<br />

networks from community to global levels are understood, the practical benefits<br />

of nurturing linkages that could add to resiliency have been little explored either<br />

conceptually or in terms of the place-to-place heterogeneity of <strong>Andean</strong> landscapes<br />

and livelihood modes. Comparisons with how change occurs in other mountainous<br />

regions would also provide insights (e.g., Bartlein et al., 1997; Boggs and<br />

Murphy, 1997; Still et al., 1999; Beniston, 2003; Dirnböck et al., 2003; Fagre<br />

et al., 2003; Halloy and Mark, 2003; Konvicka et al., 2003; Storch et al., 2003;<br />

Whitlock et al., 2003). In addition, the higher levels of social capital, such as<br />

traditionally found in the <strong>Andean</strong> land use systems and as still exists in rural<br />

Ancash, Peru, are known to promote climate resiliency elsewhere (Adger, 2003;<br />

Fraser et al., 2003).<br />

The Andes have experienced much climate change in the past (e.g., Clapperton,<br />

1993; Marchant et al., 2002; Abbott et al., 2003; Hansen et al., 2003), with current<br />

climatic envelopes only established in the last 3000 years, a long time for human<br />

societies, but very brief compared to the evolutionary trajectories that establish<br />

species adaptations (Davis and Shaw, 2001). Thus, many native species will have<br />

experienced past climate fluxes similar or greater than what will occur in the coming<br />

decades. It is human alteration of land-cover types and population densities of<br />

biota that will differ into the foreseeable future, creating novel stresses and land


KENNETH R. YOUNG AND JENNIFER K. LIPTON 89<br />

cover configurations never seen before (e.g., Pyke, 2004). Combining land-use and<br />

land-cover change studies with insights from rural livelihood research will help<br />

in further evaluations, including more integrated assessments (e.g., Hannah et al.,<br />

2002; Pielke et al., 2002). In fact, we predict that the land-use systems of the Andes<br />

will be found to contain redundancy and an ability to predict and accommodate<br />

changes in seasonality and climatic variables.<br />

Miles et al. (2004) recently used biogeographic simulations of 69 Amazonian<br />

plant species to predict dramatic future declines in population viability of almost<br />

half of the species. The tropical Andes cover a smaller geographical region than the<br />

Amazon, but have a similar amount of plant (Henderson et al., 1991) and mammal<br />

diversity (Mares, 1992). The modeling of potential future species distributions for<br />

the Andes would need to accommodate the much finer ecological mosaic found in<br />

the mountains, with biocomplexity augmented by the barriers to dispersal formed<br />

by high elevations and deep river valleys. Lindberg and Olesen (2001) showed<br />

that habitat loss would affect both a wild passion plant species and its specialized<br />

pollinating hummingbird. Davis et al. (1998) demonstrated in general the need to<br />

include the effect of climate change on species interactions and there are numerous<br />

ongoing attempts to model the likely distributions of species in relation to climatic<br />

envelopes (e.g., Peterson et al., 2001; Erasmus et al., 2002; Kadmon et al., 2003;<br />

Oberhauser and Peterson, 2003). Some aspects of ecosystem function may also be<br />

predicted based not on the occurrence of individual species, but on the relationship<br />

of physical environmental factors to vegetation height, biomass, and trophic organization<br />

(Chapin et al., 2000, 2001). The predicting of shifts in distributions will<br />

need to take these considerations into account (e.g., Bush, 2002; Téllez-Valdés and<br />

Dávila-Aranda, 2003; Cabeza et al., 2004; Mayle et al., 2004; Opdam and Wascher,<br />

2004), as will the assessments of biodiversity conservation areas (e.g., Harrison and<br />

Bruna, 1999; Scott et al., 2002; Aremteras et al., 2003; Cabeza and Moilanen, 2003;<br />

Araújo et al., 2004; Meir et al., 2004; Saxon et al., 2005). In the meanwhile, the<br />

establishment of vertical and horizontal conservation corridors provides a means<br />

to proceed that could yield other practical benefits, for example for the livelihoods<br />

of rural <strong>Andean</strong> people, if done appropriately. This land-use planning might permit<br />

climate change researchers to build in “degree of human usage” of landscapes<br />

as one of the variables they examine or control for, to compare more isolated<br />

sites with others that are more utilized. Field experiments of the sort that manipulate<br />

atmospheric composition and soil temperature also would be useful (e.g., de<br />

Valpine and Harte, 2001; Morgan et al., 2004), but do not appear to be ongoing in<br />

the study area.<br />

Research finding correlations between climate change and the fates of <strong>Andean</strong><br />

societies (e.g., Binford et al., 1997) sends a cautionary message ameliorated by<br />

long-term evidence of persistence despite climatic stresses (Dillehay and Kolata,<br />

2004; Janusek and Kolata, 2004). What are called “traditions” are in fact ways that<br />

information is culturally passed to future generations, sometimes with insights that<br />

can be applied elsewhere (e.g., Cash et al., 2003; Altieri, 2004; Stoorvogel et al.,


90 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

2004). The means by which households utilize <strong>Andean</strong> landscapes, through environmental<br />

knowledge and social networks of hierarchical linkages and obligations,<br />

may embody transferable lessons.<br />

6. Conclusion<br />

Glaciers store water that fell as rain and snow years earlier. As such, they serve as<br />

buffers for this natural resource’s availability: melt water can be used for domestic,<br />

agricultural, or industrial purposes, even in times when rainfall is low or missing<br />

(Barry and Seimon, 2000). The people of the rural Andes who live below glaciated<br />

peaks are aware of the dangers and risks posed by steep slopes, tectonic movements,<br />

and possible catastrophic mass movements. If questioned, they will also likely<br />

indicate that they are aware that there are compensations, because the water can be<br />

used during dry periods for irrigation. Burger (1992) noted that the archaeological<br />

record from three to four thousand years ago indicates that the need for control and<br />

access to water was already apparent in settlement and architectural designs. Those<br />

highland slopes that are not overtopped by glaciers or which are too far for canals<br />

to be made are limited to rain-fed agriculture.<br />

Many of our field observations reported in the case study come from the<br />

Cordillera Blanca of Ancash, Peru, where the climate change phenomena of concern<br />

include reduced permanent ice and hence more erratic water supplies and possibly<br />

more hazardous conditions downhill and downstream. However, the Cordillera<br />

Negra lies just to the west (Figure 2) and has probably not had permanent ice since<br />

the last glacial period. The agricultural landscapes there are dry for more than half<br />

the year, rural populations are less dense, and it seems clear that decreased water<br />

availability will impact local people directly. This contrast is likely to be a crucial<br />

planning difference to include in institutional analyses: the relative buffering of the<br />

physical environment utilized by the social group of concern.<br />

Traditional ways of carrying out land use presumably include wisdom acquired<br />

from trial and error in the past. In practice, <strong>Andean</strong> land use is an agglomeration<br />

of traditions begun thousands of years ago, leavened by adaptations to new<br />

plants, animals, and agricultural products introduced from overseas, and then implemented<br />

by individuals drawing upon knowledge and influenced by needs from<br />

their households and communities. For example, total land-use options diversified<br />

with the introduction of additional plant and animal species from the Old World,<br />

with wheat, oats, and alfalfa integrated into systems based on tuber crops, maize,<br />

beans, and other <strong>Andean</strong> species (Gade, 1992). Most traditional <strong>Andean</strong> households<br />

raise native domesticated animals such as guinea pigs (Morales, 1999), and<br />

the camelids, llama and alpaca, are used especially in Peru and Bolivia for wool,<br />

meat, and cargo transport (Wheeler, 1995). The Old World contribution of useful<br />

animals has been extensive: most <strong>Andean</strong> landscapes are utilized in conjunction<br />

with pigs and poultry near households, and goats, sheep, cattle, and horses out


KENNETH R. YOUNG AND JENNIFER K. LIPTON 91<br />

across the landscape, using the shrublands, wetlands, and grasslands for grazing.<br />

We have been struck by the multiplicity of land-use strategies in our study area, as<br />

expressed by spatial and compositional differences within and among agricultural<br />

fields, and by the complex integration that goes on with the raising of a variety<br />

of domesticated animals. In many parts of the tropical Andes, adaptation to fairly<br />

dramatic shifts in the location of ecological zones for planting and grazing appears<br />

to work by switching the species used or at least their relative proportions.<br />

Programs to alleviate stresses originating from climate change should start from<br />

this premise and look for ways to maintain, encourage, or even spread through<br />

example local adaptations to other locales with similar biophysical parameters. It<br />

is not clear that this strategy is within the present-day capacity of most national<br />

agricultural and rural development institutions. Perhaps as important, there appear<br />

to be fundamental limitations on how responsive those national entities can be to<br />

local concerns. For that matter, there are aspects of <strong>Andean</strong> belief systems that<br />

are quite different in origin and orientation from those of many western cultures<br />

(Greenway, 1998).<br />

Native biodiversity in the spatially and temporally dynamic <strong>Andean</strong> landscapes<br />

also represent an object lesson for evaluating climate change adaptiveness. These<br />

plants and animals have been through past climate fluctuations and have survived.<br />

Current limitations on their distributions often appear to originate only in part from<br />

climatic or other biophysical controls; the other limitations originate with past and<br />

present human-caused and land-use-related alterations. Many <strong>Andean</strong> landscapes<br />

appear to be the result of ancient human settlements and land uses (Lumbreras,<br />

1974; Bruhns, 1994; Gade, 1999; Young and Keating, 2001), making them similar<br />

to those described in continental Europe (Ellenberg, 1979; Birks et al., 1988;<br />

Young, 1997). The biodiversity conservation institutions have not yet assimilated<br />

this distinction into practice in the Andes: most efforts have gone into naming<br />

large areas as national parks or reserves, while ignoring opportunities for ecological<br />

restoration in the remainder of the countries involved. Even more intractable<br />

has been the conservation of biodiversity that originates and is maintained by cultural<br />

practices associated with land use. Programs focused on species of concern,<br />

on landscapes with significant native vegetation, and on agricultural land races<br />

and breeds are all needed; climate change concerns should be part of the decision<br />

matrices used for setting goals and priorities (e.g., Balmford et al., 2005). Institutional<br />

support needs to be strengthened at local, national, and international levels<br />

(e.g., Strigl, 2003). The largest difficulty will be the need to proceed without clear<br />

scientific precedents or guidelines. However, this uncertainty is a part of all governance<br />

faced by climate change (e.g., O’Hare, 2000; Reilly et al., 2001; Lynch<br />

et al., 2003; Pearson and Dawson, 2003; Johnson and Gillingham, 2004; Mastrandrea<br />

and Schneider, 2004) and some solutions likely already exist among the<br />

responses of native plants and animals and within the dynamism of local land-use<br />

systems.


92 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Acknowledgments<br />

We are grateful to Paul Robbins and the Adaptive Research and Governance in<br />

Climate Change program of the Ohio <strong>State</strong> University for the invitation to participate<br />

in the conference that led to the presentation of this paper. Blanca León<br />

and Carol Squires cheerfully helped with the manuscript. K. Young most recently<br />

received travel and research assistance from a Mellon Foundation faculty grant for<br />

research through the University of Texas at Austin’s Population Research Center. J.<br />

Lipton’s fieldwork in Peru was made possible through a University of Texas at<br />

Austin Thematic Fellowship, a David Boren fellowship, a National Science Foundation<br />

Dissertation Improvement Grant (NSF #0117806), and a Mellon Foundation<br />

Summer Research Award. Sincere gratitude is extended to people of the communities<br />

of Ancash who participated in this study. In addition, a profound thank you<br />

to the many individuals who assisted, encouraged, and informed throughout the<br />

duration of the project, including: Hector Abignal, Roberto Arevalo, Alton Byers,<br />

Pedro Camacho, Asunción Cano, Juan Carlos Castro, Jose Carrasco, Lorenzo<br />

Champa, Maria Victoria Chavez, Doña Coti, Hildegardo Cordova, Brandie Fariss,<br />

Jesus Gomez, Oswaldo Gonzáles de Paz, Blanca León, Carlos Meza, Nelson Santillain,<br />

Karen Price, Jorge Recharte, Hugo Romero, Vidal Rondan, Pablo Tadeo,<br />

Willy Tamayo, Mirriam Torres, Felix Valverde, Selwyn Valverde, Mario Villanueva,<br />

Marco Zapata.<br />

References<br />

Abbott, M. B., Wolfe, B. B., Wolfe, A. P., Seltzer, G. O., Aravena, R., Mark, B. G., Polissar, P.<br />

J., Rodbell, D. T., Rowe, H. D., and Vuille, M.: 2003, ‘Holocene Paleohydrology and Glacial<br />

History of the Central Andes using Multiproxy Lake Sediment Studies’, Palaeogeogr. Palaeoclim.<br />

Palaeoecol 194, 123–138.<br />

Adams, W. M.: 2003, ‘Managing Tragedies: Understanding Conflict over Common Pool Resources’,<br />

Science 302, 1915–1916.<br />

Adger, W. N.: 1999, ‘Social Vulnerability to Climate Change and Extremes in Coastal Vietnam’,<br />

World Develop. 27, 249–269.<br />

Adger, W. N.: 2000, ‘Institutional Adaptation to Environmental Risk Under the Transition in Vietnam’,<br />

Annals Assoc. Amer. Geogr 90, 738–758.<br />

Adger, W. N.: 2003, ‘Social Capital, Collective Action, and Adaptation to Climate Change’, Economic<br />

Geogr 79, 387–404.<br />

Allen, C. J.: 1988, The Hold Life has: Coca and Cultural Identity in an <strong>Andean</strong> Community, Washington<br />

D.C., Smithsonian Press.<br />

Ames, A., Muñoz, G., Verastegui, J., Vigil, R., Zamora, M., and Zapata, M.: 1989, Glacier Inventory<br />

of Peru, Volume 1. Hidrandina S. A. Unidad de Glaciologia e Hidrologia.<br />

Araújo, M. B., Cabeza, M., Thuiller, W., Hannah, L., and Williams, P. H.: 2004, ‘Would Climate<br />

Change Drive Species out of Reserves? An Assessment of Existing Reserve-Selection Methods’,<br />

Global Change Biol. 10, 1618–1626.<br />

Armenteras, D., Gast, F., and Villareal, H.: 2003, ‘<strong>Andean</strong> Forest Fragmentation and the Representativeness<br />

of Protected Natural Areas in the Eastern Andes, Colombia’, Biol. Conservation 113,<br />

245–256.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 93<br />

Balmford, A., Bennun, L., ten Brink, B., Cooper, D., Côté, I. M., Crane, P. Dobson, A., Dudley,<br />

N., Dutton, I., Green, R. E., Gregory, R. D., Harrison, J., Kennedy, E. T., Kremen, C., Leader-<br />

Williams, N., Lovejoy, T. E., Mace, G., May, R., Mayaux, P., Morling, P., Phillips, Joanna,<br />

Redford, K., Ricketts, T. H., Rodríguez, J. P., Sanjayna, M., Schei, P. J., van Jaarsveld, A. S.,<br />

and Walther, B. A.: 2005, ‘The Convention on Biological Diversity’s 2010 Target’, Science 307,<br />

212–213.<br />

Barry, R. G. and Seimon, A.: 2000, ‘Research for <strong>Mountain</strong> Area Development: Climatic Fluctuations<br />

in the <strong>Mountain</strong>s of the Americas and Their Significance’, Ambio, 29, 364–370.<br />

Bartle, J.: 1993, Parque Nacional Huascarán. Nuevas Imágenes S., Lima.<br />

Bartlein, P. J., Whitlock, C., and Shafer, S. L.: 1997, ‘Future Climate Change in the Yellowstone<br />

National Park Region and its Potential Impact on Vegetation’, Conservation Biol. 11, 782–<br />

792.<br />

Bebbington, A.: 1996, ‘Movements, Modernizations, and Markets’, in Peet, R. and Watts, M. (eds.),<br />

Liberation Ecologies: Environment, Development, and Social Movements, Routledge, London,<br />

pp. 86–109.<br />

Bebbington, A.: 1997, ‘Social Capital and Rural Intensification: Local Organizations and Lands of<br />

Sustainability in the Rural Andes’, Geographical J. 163(2), 189–199.<br />

Bebbington, A.: 2000, ‘Reencountering Development: Livelihood Transitions and Place Transformations<br />

in the Andes’, Annals Assoc. Amer. Geogr. 90, 495–520.<br />

Bebbington, A.: 2001, ‘Globalized Andes? Livelihoods, Landscapes and Development’, Ecumene<br />

8(4), 414–436.<br />

Bebbington, A. and Perreault, T.: 1999, ‘Social Capital, Development, and Access to Resources in<br />

Highland Ecuador’, Economic <strong>Geography</strong> 75(4), 395.<br />

Beniston, M.: 2003, ‘Climatic Change in <strong>Mountain</strong> Regions: A Review of Possible Impacts’, Climatic<br />

Change 59, 5–31.<br />

Blaikie, P. and Brookfield, H. (Eds): 1987, Land Degradation and Society, Methuen, New York.<br />

Binford, M. W., Kolata, A. L., Brenner, M., Janusek, J. W., Seddon, M. T., Abbottt, M., and Curtis,<br />

J. H.: 1997, ‘Climate Variation and the Rise and Fall of an <strong>Andean</strong> Civilization.’ Quat. Research<br />

47, 235–248.<br />

Birks, H. H., Birks, H. J. B., Kaland, P. E., and Moe, D. (eds.): 1988, The Cultural Landscape: Past,<br />

Present and Future. Cambridge University Press, Cambridge.<br />

Boggs, C. L. and Murphy, D. D.: 1997, ‘Community Composition in <strong>Mountain</strong> Ecosystems: Climatic<br />

Determinants of Montane Butterfly Distributions’, Global Ecol. Biogeogr. Letters 6, 39–48.<br />

Brehm, G., Süssenbach, D., and Fiedler, K.: 2003, ‘Unique Elevational Diversity Patterns of Geometrid<br />

Moths in an <strong>Andean</strong> Montane Rainforest’, Ecography 26, 456–466.<br />

Brooks, T. M., Mittermeier, R. A., Mittermeier, C. G., Da Fonseca, G. A. B, Rylands, A. B., Konstant,<br />

W. R., Flick, P., Pilgrim, J., Oldfield, S., Magin, G., and Hilton-Taylor, C.: 2002, ‘Habitat Loss<br />

and Extinction in the Hotspots of Biodiversity’, Conservation Biol. 16, 909–923.<br />

Brownlee, C.: 2001, ‘The Long and the Short of Stomatal Density Signals’, Trends in Plant Science<br />

6, 441–442.<br />

Browman, D.: 1987, Arid Land Use Strategies and Risk Management in the Andes: A Regional<br />

Anthropological Perspective, Westview Press, Boulder, Colorado.<br />

Bruhns, K. O.: 1994, Ancient South America, Cambridge World Archaeology, Cambridge University<br />

Press, Cambridge, pp. 424<br />

Brush, S. B.: 1976, ‘Man’s Use of an <strong>Andean</strong> Ecosystem’, Human Ecology 4, 147–166.<br />

Brush, S. B.: 1987, ‘Diversity and Change in <strong>Andean</strong> Agriculture’, in Little, P. D. et al. (eds.), Lands<br />

at Risk in the Third World: Local Level Perspectives, Westview Press, Boulder, Colorado, pp.<br />

271–289.<br />

Brush, S. B. and Guillet, D.: 1985, ‘Small-Scale Agro-Pastoral Production in the Central Andes’,<br />

<strong>Mountain</strong> Research Develop. 5(1), 19–30.


94 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Brush, S. B., Kesseli, R., Ortega, R., Cisneros, P., Zimmerer, K., and Quiros, C.: 1995.<br />

Potato Diversity in the <strong>Andean</strong> Center of Crop Domestication’, Conservation Biol. 9,<br />

1189–1198.<br />

Burger, R. L.: 1992, Chavin and the Origins of <strong>Andean</strong> Civilization, Thames and Hudson, London.<br />

Bury, J.: 2004, ‘Livelihoods In Transition: Transnational Gold Mining Operations and Local Change<br />

in Cajamarca, Peru’, Geograph J. 170(1), 78–91.<br />

Bush, M. B.: 2002, ‘Distributional Change and Conservation on the <strong>Andean</strong> Flank: A Palaeoecological<br />

Perspective’, Global Ecol. Biogeogr 11, 463–473.<br />

Byers, A.: 2000, ‘Contemporary Landscape Change in the Huascarán National Park and Buffer Zone,<br />

Cordillera Blanca, Peru’, <strong>Mountain</strong> Research Develop. 20, 52–63.<br />

Cabeza, M., Araújo, M. B., Wilson, R. J., Thomas, C. D., Cowley, M. J. R., and Moilanen, A.:<br />

2004, ‘Combining Probabilities of Occurrence with Spatial Reserve Design’, J. Applied Ecol. 41,<br />

252–262.<br />

Cabeza, M. and Moilanen, A.: 2003, ‘Site-Selection Algorithms and Habitat Loss’, Conservation<br />

Biol. 17, 1402–1413.<br />

Cash, D. W., Clark, W. C., Alcock, F., Dickson, N. M., Eckley, N., Guston, D. H., Jäger, and Mitchell,<br />

R. B.: 2003, ‘Knowledge Systems for Sustainable Development’, Proc. Nat. Acad. Sci. 100,<br />

8086–8091.<br />

Cerrate de Ferreyra, E.: 1979: Vegetación del Valle de Chiquian (Provincia de Bolognesi, Departamento<br />

de Ancash), Editorial Los Pinos, Lima.<br />

Chapin III, F. S., Matson, P. A., and Mooney, H. A.: 2001, Principles of Terrestrial Ecosystem Ecology,<br />

Springer, New York.<br />

Chapin III, F. S., Sala, O. E., and Huber-Sannwald, E. (eds.): 2000, ‘Global Biodiversity in a Changing<br />

Environment. Scenarios for the 21st Century’, Springer, New York.<br />

Clapperton, C. M.: 1993, ‘Nature of Environmental Changes in South America at the Last Glacial<br />

Maximum’, Palaeogeogr. Palaeoclimatol. Palaeoecol 101, 189–208.<br />

Davis, A. J., Jenkinson, L. S., Lawton, J. H., Shorrocks, B., and Wood, S.: 1998, ‘Making Mistakes<br />

when Predicting Shifts in Species Range in Response to Global Warming’, Nature 391, 783–786.<br />

Davis, M. B. and Shaw, R. G.: 2001, ‘Range Shifts and Adaptive Responses to Quaternary Climate<br />

Change’, Science 292, 673–679<br />

de Janvry, A. Sadoulet, E., and Thorbecke, E.: 1993, ‘<strong>State</strong>, Market and Civil Organizations: New<br />

Theories, New Practices and their Implications for Rural Devleopment: Introduction. World<br />

Develop. 21, 565–575.<br />

Denevan, W. M.: 2001, Cultivated Landscapes of Native Amazonia and the Andes, Oxford University<br />

Press, New York.<br />

de Valpine, P. and Harte, J.: 2001, ‘Plant Responses to Experimental Warming in a Montane<br />

Meadow’, Ecology. 82, 637–648.<br />

Dietz, T., Ostrom, E., and Stern, P.: 2003, ‘The Struggle to Govern the Commons.’ Science 302,<br />

1907–1912.<br />

Dillehay, T. D. and Kolata, A. L.: 2004, ‘Long-term Human Response to Uncertain Environmental<br />

Conditions in the Andes’, Proc. Nat. Acad. Sci. 101, 4325–4330.<br />

Dirnböck, T., Dullinger, S., and Grabherr, G.: 2003, ‘A Regional Impact Assessment of Climate and<br />

Land-Use Change on Alpine Vegetation’, J. Biogeography 30, 401–417.<br />

Dockerty, T., Lovett, A., Sünnenberg, G., Appleton, K., and Parry, M.: 2005, ‘Visualizing the<br />

potential impacts of climate change on rural landscapes’, Computers, Environment and Urban<br />

Systems 29, 297–320.<br />

Ellenberg, H.: 1979, ‘Man’s Influence on Tropical <strong>Mountain</strong> Ecosystems in South America’, J. Ecol.<br />

67(2), 401–416.<br />

Enquist, C. A. F.: 2002, ‘Predicted Regional Impacts of Climate Change on the Geographical<br />

Distribution and Diversity of Tropical Forests in Costa Rica’, J. Biogeography 29, 519–534.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 95<br />

Erasmus, B. F., Van Jaarsveld, A. S., Chown, S. L., Kshatriya, M., and Wessels, K. J.: 2002, ‘Vulnerability<br />

of South African Animal Taxa to Climate Change’, Global Change Biol. 8, 679–693.<br />

Escobar, A.: 1995, Encountering Development: The Making and Unmaking of the Third World,<br />

Princeton University Press, New Jersey.<br />

Fagre, D. B., Peterson, D. L., and Hessl, A. E.: 2003, ‘Taking the Pulse of <strong>Mountain</strong>s: Ecosystem<br />

Responses to Climatic Variability’, Climatic Change 59, 263–282.<br />

Fjelds˚a, J., Lambins, E., and Mertens, B.: 1999, ‘Correlation between Endemism and Local<br />

Ecoclimatic Stability Documented by Comparing <strong>Andean</strong> Bird Distributions and Remotely<br />

Sensed Land Surface Data’, Ecography 22, 63–78.<br />

Flores Ochoa, J. (eds.): 1977, Pastores de Puna: Uywamichiq Punarunakuna, Instituto de Estudios<br />

Peruanos, Lima.<br />

Foster, P.: 2001, ‘The Potential Negative Impacts of Global Climate Change on Tropical Montane<br />

Cloud Forests’, Earth Science Rev. 55, 73–106.<br />

Francou, B., Vuille, M., Wagnon, P. Mendoza, J., and Sicart, J. E.: 2003, ‘Tropical Climate Change<br />

Recorded by a Glacier in the Central Andes during the Last Decades of the 20th Century:<br />

Chacaltaya, Bolivia, 16 ’, J. Geophysical Res. 108 D5, 4154, doi:10.1029/2002JD002959.<br />

Frankham, R., Ballou, J. D., and Briscoe, D. A.: 2002, Introduction to Conservation Genetics,<br />

Cambridge University Press, New York.<br />

Fraser, E. D. G., Mabee, W., and Slaymaker, O.: 2003, ‘Mutual Vulnerability, Mutual Dependence:<br />

The Reflexive Relation Between Human Society and Environment’, Global Environmental<br />

Change 13, 137–144.<br />

Gade, D. W.: 1975, Plants, Man and the Land in the Vilcanota Valley of Peru, W. Junk, The Hague.<br />

Gade, D. W.: 1992, ‘Landscape, System, and Identity in the Post-Conquest Andes’, Annals Assoc.<br />

Geogr. 82, 460–477.<br />

Gade, D. W.: 1999, Nature and Culture in the Andes, The University of Wisconsin Press, Madison,<br />

Wisconsin.<br />

Gelles, P. H.: 1999. Water and Power in Highland Peru. New Brunswick, Rutgers University Press.<br />

Gentry, A. H.: 1995, ‘Patterns of Diversity and Floristic Composition in Neotropical Montane Forests’,<br />

in Churchill, S. P., Balslev, H., Forero, E., and Luteyn, J. L. (eds.), Biodiversity and Conservation<br />

of Neotropical Montane Forests, New York Botanical Garden, New York, pp. 103–126.<br />

Georges, C.: 2004, ‘20th-Century Glacier Fluctuations in the Tropical Cordillera Blanca, Peru’,<br />

Arctic, Antarctic, Alpine Research 36, 100–107.<br />

Gerrard, A. J.: 1990. <strong>Mountain</strong> Environments: An Examination of the Physical <strong>Geography</strong> of<br />

<strong>Mountain</strong>s, MIT Press, Cambridge, Massachusetts.<br />

Giordano, M.: 2003, ‘The <strong>Geography</strong> of the Commons: The Role of Scale and Space’, Annals Assoc.<br />

Amer. Geogr. 93, 365–375.<br />

González, E. R. and Watling, L.: 2003, ‘Two New Species of Hyalella from Lake Titicaca, and Redescriptions<br />

of Four Others in the Genus (Crustacea: Amphipoda)’, Hydrobiologia 497, 181–204.<br />

Griffin, K.: 2003, ‘Economic Globalization and Institutions of Global Governance’, Development<br />

Change 34, 789–807.<br />

Greenway, C.: 1998, ‘Hungry Earth and vengeful stars: soul loss and identity in the Peruvian Andes’,<br />

Soc. Sci Med. 47, 993–1004.<br />

Guillet, D.: 1981, ‘Land Tenure, Ecological Zone and Agricultural Regime in the Andes’, American<br />

Ethnologist 8(1), 1339–156.<br />

Guillet, D.: 1992, Covering Ground. Communal Water Management and the <strong>State</strong> in the Peruvian<br />

Highlands, The University of Michigan Press, Ann Arbor.<br />

Guisan, A. and Theurillat, J. -P.: 2001. ‘Assessing Alpine Plant Vulnerability to Climate Change: A<br />

Modeling Perspective’, Integrated Assessment 1, 307–320.<br />

Gutzwiller, K. J. (eds): 2002, Applying Landscape Ecology in Biological Conservation, Springer-<br />

Verlag, New York.


96 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Halloy, S. R. P. and Mark, A. F.: 2003, ‘Climate-Change Effects on Alpine Plant Diversity: A New<br />

Zealand Perspective on Quantifying the Threat’, Arctic, Antarctic, Alpine Research 35, 248–254.<br />

Hannah, I., Midgley, G. G., and Millar, D.: 2002, ‘Climate Change-Integrated Conservation<br />

Strategies’, Global Ecol. Biogeography 11, 485–495.<br />

Hansen, B. C. S., Rodbell, D. T., Seltzer, G. O., León, B., Young, K. R., and Abbott, M.: 2003,<br />

‘Late-glacial and Holocene Vegetational History from two Sites in the Western Cordillera of<br />

Southwestern Ecuador’. Palaeogeogr. Palaeoclim. Palaeoecol 194, 79–108.<br />

Harrison, S. and Bruna, E.: 1999, ‘Habitat Fragmentation and Large-scale Conservation: What do<br />

we know for sure?’, Ecography 22, 225–232.<br />

Hastenrath, S. and Ames, A.: 1995a, ‘Diagnosing the Imbalance of Yanamarey Glacier in the<br />

Cordillera Blanca of Peru’, J. Geophysical Res. 100, 5105–5112.<br />

Hastenrath, S. and Ames, A.: 1995b, ‘Recession of Yanamarey Glacier in Cordillera Blanca, Peru,<br />

during the 20th Century’, J. Glaciology 41, 191–196.<br />

Healy, K.: 1991: ‘Political Ascent of Bolivia’s Peasant Coca Leaf Producers’, Journal of Interamerican<br />

Studies World Affairs 33(1), 87–121.<br />

Healy, K.: 2001: Llamas, Weavings, and Organic Chocolate: Multicultural Grassroots Development<br />

in the Andes and Amazon of Bolivia, University of Notre Dame Press, Notre Dame.<br />

Henderson, A., Churchill, S. P., and Luteyn, J. L.: 1991. ‘Neotropical plant diversity’, Nature 351,<br />

21–22.<br />

Hijmans, R. J. and Spooner, D. M.: 2001, ‘Geographic Distribution of Wild Potato Species’, Amer.<br />

J. Bot. 88, 2101–2112.<br />

Inouye, D. W., Barr, B., Armitage, K. B., and Inouye, B. D.: 2000, ‘Climate Change is Affecting<br />

Altitudinal Migrants and Hibernating Species’, Proc. Nat. Acad. Sci. 97, 1630–1633.<br />

Instituto Nacional de Recursos Naturales – INRENA: 2003, ‘Parque Nacional Huascarán Plan<br />

Maestro 2003–2007’ INRENA-Instituto de Montaña y Profonanpe.<br />

Inter-American Development Bank: 1997, Economic and Social Progress in Latin America, Johns<br />

Hopkins University Press, Baltimore, Maryland.<br />

Jackson, S. T. and Weng, C.: 1999, ‘Late Quaternary Extinction of a Tree Species in Eastern North<br />

America’, Proc. Nat. Acad. Sci. 96, 13847–13852.<br />

Janusek, J. W. and Kolata, A. L.: 2004, ‘Top-down or Bottom-up: Rural Settlement and Raised Field<br />

Agriculture in the Lake Titicaca Basin, Bolivia.’, J. Anthrop. Arachaeol 23, 404–430.<br />

Johnson, C. J. and Gillingham, M. P.: 2004, ‘Mapping Uncertainty: Sensitivity of Wildlife Habitat<br />

Ratings to Expert Opinion’, J. Applied Ecol. 41, 1032–1041.<br />

Jokisch, B. D.: 2002, ‘Migration and Agricultural Change: the Case of Smallholder Agriculture in<br />

Highland Ecuador’, Human Ecol. 30, 523–550.<br />

Kadmon, R., Farber, O., and Danin, A.: 2003, ‘A Systematic Analysis of Factors Affecting the<br />

Performance of Climatic Envelope Models’, Ecol. Applications 13, 853–867.<br />

Kappelle, M., Van Vuuren, M. M. I., and Baas, P.: 1999, ‘Effects of Climate Change on Biodiversity:<br />

A Review and Identification of Key Research Issues’, Biodiversity Conservation 8, 1383–1397.<br />

Kaser, G.: 1999, ‘A Review of the Modern Fluctuations of Tropical Glaciers’, Global Planetary<br />

Change 22, 93–103.<br />

Kaser, G., Ames, A., and Zamora, M.: 1990, ‘Glacier Fluctuations and Climate in the Cordillera<br />

Blanca, Peru.’ Ann. Glaciology 14, 136–140.<br />

Kaser, G. and Georges, C.: 1997, ‘Changes of the Equilibrium Line Altitude in the Tropical Cordillera<br />

Blanca (Perú) Between 1930 and 1950 and their spatial variations’, Ann. Glaciology 24, 344–349.<br />

Kaser, G. and Georges, C.: 1999, ‘On the Mass Balance of Low Latitude Glaciers with Particular<br />

Consideration of the Peruvian Cordillera Blanca’, Geogr. Ann. 81A, 643–651.<br />

Kaser, G., Juen, I., Georges, C., Gomez, J., and Tamayo, W.: 2001, ‘The impact of glaciers on the<br />

runoff and the reconstruction of mass balance history from hydrological data in the tropical<br />

Cordillera Blanca, Perú.’, J. Hydrology 282(1–4), 130–144.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 97<br />

Kaser, G. and Osmaston, H.: 2002, Tropical Glaciers. Cambridge: Cambridge University Press.<br />

Keating, P. L., Young, K. R., and León, B.: 2002, ‘Variation in High <strong>Andean</strong> Vegetation at a Site in<br />

Southwestern Ecuador’, Pennsylvania Geograph 40(2), 15–35.<br />

Knapp, G.: 1991, ‘<strong>Andean</strong> Ecology. Adaptive Dynamics in Ecuador’, Westview Press, Boulder,<br />

Colorado.<br />

Konvicka, M., Maradova, M., Benes, J., Fric, Z., and Kepka, P.: 2003, ‘Uphill Shifts in Distribution<br />

of Butterflies in the Czech Republic: Effects of Changing Climate Detected on a Regional Scale’,<br />

Global Ecol. Biogeography 112, 403–410.<br />

Koínek, V. and Villalobos, L.: 2003, ‘Two South American Endemic Species of Daphnia from High<br />

<strong>Andean</strong> Lakes’, Hydrobiologia 490, 107–123.<br />

Körner, C.: 1999, Alpine Plant Life: Functional Plant Ecology of High <strong>Mountain</strong> Ecosystems.<br />

Springer, Berlin.<br />

Körner, C.: 2000, ‘Biosphere Responses to CO2 Enrichment’, Ecol. Applications 10, 1590–1619.<br />

Korovkin, T: 1998, ‘Commodity Production and Ethnic Culture: Otavalo, Northern Ecuador’, Econ.<br />

Develop. Cultural Change 47(1), 125–154.<br />

Leimbeck, R. M., Valencia, R., and Balslev, H.: 2004, ‘Landscape Diversity Patterns and Endemism<br />

of Araceae in Ecuador’, Biodiversity Conservation 13, 1755–1779.<br />

León, B. and Young, K. R.: 1996, ‘Aquatic Plants of Peru: Diversity, Distribution, and Conservation’,<br />

Biodiver. Conserv. 5,1169–1190.<br />

Lindberg, A. B. and Olesen, J. M.: 2001, ‘The Fragility of Extreme Specialization: Passiflora<br />

mixta and its Pollinating Hummingbird Ensifera ensifera’, J. Tropical Ecol. 17,<br />

323–329.<br />

Lliboutry, L., Morales, B., Pautre, A., and Schneider, B.: 1977. ‘Glaciological Problems Set By the<br />

Control of Dangerous Lakes in the Cordillera Blanca, Peru. I. Historical failures of Morainic<br />

Dams, Their Causes and Prevention.’ J. Glaciology 18, 239–254.<br />

Lumbreras, L. G.: 1974, The Peoples and Cultures of Ancient Peru, Translated by B. J. Meggers,<br />

Smithsonian Institution Press, Washington, D.C.<br />

Luteyn, J. L.: 1999, ‘Páramos: A Checklist of Plant Diversity, Geographical Distribution and<br />

Botanical Literature”, Mem. New York Bot. Gard. 84, I–xv, 1–278.<br />

Lynch, A. H., Rivers, A. R., and Bartlein, P. J.: 2003, ‘An Assessment of the Influence of Land Cover<br />

Uncertainties on the Simulation of Global Climate in the Early Holocene’, Climate Dynam. 21,<br />

243–256.<br />

Malhi, Y. and Phillips, O. L.: 2004, ‘Tropical forests and global atmospheric change: a synthesis’,<br />

Phil. Trans R. Soc. Lond. B 359, 549–555.<br />

Manne, L. L, Brooks, T. M., and Pimm, S. L.: 1999, ‘Relative Risk of Extinction of Passerine Birds<br />

on Continents and Islands’, Nature 399, 258–261.<br />

Marchant, R., Behling, H., Berrio, J.-C., Cleef, A., Duivenvoorden, J., Hooghiemstra, H., Kuhry,<br />

P., Melief, B., Schreve-Brinkman, E., Van Geel, B., Van der Hammen, T., Van Reenen, G., and<br />

Wille, M.: 2002, ‘Pollen-based Biome Reconstructions for Colombia at 3000, 6000, 9000, 12000,<br />

15000 and 18000 14C yr ago: Late Quaternary Tropical Vegetation Dynamics’, J. Quaternary<br />

Sci. 17, 113–129.<br />

Mares, M. A.: 1992, ‘Neotropical Mammals and the Myth of Amazonian Biodiversity’. Science 255,<br />

976–979.<br />

Mark, B. G. and Seltzer, G. O: 2003, ‘Tropical Glacier Meltwater Contribution to Stream Discharge:<br />

A Case Study in the Cordillera Blanca, Peru’, J. Glaciology 49, 271–281.<br />

Mastrandrea, M. D. and Schneider, S. H.: 2004, ‘Probalistic Integrated Assessment of “Dangerous”<br />

Climate Change’, Science 304, 571–575.<br />

Maybury-Lewis, D. (ed.): 2002, The Politics of Ethnicity: Indigenous Peoples in Latin American<br />

<strong>State</strong>s, Harvard University Press and David Rockefeller Center for Latin American<br />

Studies.


98 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Mayer, E.: 2002, The Articulated Peasant: Household Economies in the Andes, Westview Press,<br />

Boulder, Colorado.<br />

Mayle, F. E., Beerling, D. J., Gosling, W. D., and Bush, M. B.: 2004, ‘Reponses of Amazonian<br />

Ecosystems to Climatic and Atmospheric Carbon Dioxide Changes since the Last Glacial<br />

Maximum’, Phil. Trans. R. Soc. Lond. B359, 499–514.<br />

McCay, B. and Acheson, J. M. (eds.): 1987, The Question of the Commons: The Culture and Ecology<br />

of Community Resources. University of Arizona Press, Tucson, Arizona.<br />

Meir, E., Andelman, S., and Possingham, H. P.: 2004, ‘Does Conservation Planning Matter in a<br />

Dynamic and Uncertain World?’, Ecol. Letters 7, 615–622.<br />

Metz, B.: 2000, ‘International Equity in Climate Change Policy’, Integrated Assessment 1, 111–126.<br />

Miles, A.: 2004, From Cuenca to Queens: An Anthropological Story of Transnational Migration.<br />

Austin, University of Texas Press.<br />

Miles, L., Grainger, A., and Phillips, O.: 2004, ‘The Impact of Global Climate Change on Tropical<br />

Forest Biodiversity in Amazonia.’ Global Ecol. Biogeogr 13, 553–565.<br />

Mitchell, W. P. and Guillet, D. W. (eds.): 1993, Irrigation at High Altitudes: The Social Organization<br />

of Water Control Systems in the Andes, Washington D.C., Society for Latin American<br />

Anthropology and the American Anthropological Association.<br />

Morales Arnao, B.: 2000, ‘Los eternos nevados en el Peru estan retrocediendo en forma cada vez<br />

mas acelerada’, El Medio Ambiente en el Peru, Instituto Cuanto, Lima, pp. 17–24.<br />

Morales, E.: 1999, The Guinea Pig: Healing, Food, and Ritual in the Andes. University of Arizona<br />

Press, Tucson, Arizona.<br />

Morgan, J. A., Mosier, A. R., Milchunas, D. G., LeCain, D. R., Nelson, J. A., and Parton, W. J.:<br />

2004, ‘CO2 Enhances Productivity, Alters Species Composition, and Reduces Digestibility of<br />

Shortgrass Steppe Vegetation’, Ecol. Applications 14, 208–219.<br />

Mount, T. P.: 1994, ‘Climate Change and Agriculture: A Perspective on Priorities for Economic<br />

Policy’, Climatic Change 25, 59–83.<br />

Murra, J. V.: 1975. Formaciones Económicas y Políticas del Mundo Andino. Instituto de Estudios<br />

Andionos, Lima, Peru.<br />

Myers, N., Mittermeier, R. A., da Fonseca, C. G., Gustavo, A. B., and Kent, J.: 2000, ‘Biodiversity<br />

Hotspots for Conservation Priorities’, Nature 403, 853–858.<br />

Netting, R. M.: 1993. Smallholders, Householders: Farm Families and the Ecology of Intensive,<br />

Sustainable Agriculture, Stanford University Press, Palo Alto, California.<br />

Oberhauser, K. and Peterson, A. T.: 2003, ‘Modeling Current and Future Potential Wintering Distributions<br />

of Eastern North America Monarch Butterflies’, Proc. Nat. Acad. Sci. 100, 14063–14068.<br />

O’Brien, K., Leichenko, R., Kelkar, U., Venema, H., Aandahl, G., Tompkins, H., Javed, A., Bhadwal,<br />

S., Barg, S., Nygaard, L., and West, J.: 2004, ‘Mapping Vulnerability to Multiple Stressors:<br />

Climate Change and Globalization in India’, Global Environ. Change 14, 303–313.<br />

Ochoa, C. M.: 1999, Las Papas de Sudamérica. Part I. Perú, Centro Internacional de la Papa, Allen<br />

Press, Lawrence, Kansas.<br />

O’Hare, G.: 2000, ‘Reviewing the Uncertainties in Climate Change Science’, Area 32.4, 357–368.<br />

Opdam, P. and Wascher, D.: 2004, ‘Climate Change meets Habitat Fragmentation: Linking Landscape<br />

and Biogeographical Scale Levels in Research and Conservation’, Biol. Conservation 117,<br />

285–297.<br />

Orlove, B.: 1977, ‘Inequalities Among Peasants: the Forms and uses of Reciprocal Exchange in<br />

<strong>Andean</strong> Peru’, in Halperin, R. (ed.), Peasant Livelihood: Studies in Economic and Cultural<br />

Ecology, St. Martin’s Press, New York, pp. 201–214.<br />

Orlove, B. S., Chiang, J. C. H., and Cane, M. A.: 2000, ‘Forecasting <strong>Andean</strong> Rainfall and Crop Yield<br />

from the Influence of El Niño on Pleiades Visibility’, Nature 403, 68–71.<br />

Orlove, B. S. and Godoy, R.: 1986, ‘Sectoral Fallow Systems in the Central Andes’, J. Ethnobiology<br />

6(1), 169–204.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 99<br />

Ostrom, E.: 1990, Governing the Commons: The Evolution of Institutions for Collective Action,<br />

Cambridge University Press, Cambridge.<br />

Ostrom, E., Burger, J., Field, C. B., Norgaard, R. B., and Policansky, D.: 1999, ‘Revisiting the<br />

Commons: Local Lessons, Global Challenges’, Science 284, 278–282.<br />

Oxford University: 1996, Preliminary Report of the 1996 Oxford University Expedition to Peru:<br />

Upon the Avifauna of the High Altitude Polylepis Forests of the Cordillera Huayhuash. Report.<br />

Instituto de Montaña, Huaraz.<br />

Pearson, R. G. and Dawson, T. P.: 2003, ‘Predicting the Impacts of Climate Change on the Distribution<br />

of Species: are Bioclimate Envelope Models Useful?’, Global Ecol. Biogeogr 12, 361–371.<br />

Perales R., H., Brush, S. B., and Qualset, C. O.: 2003, ‘Landraces of Maize in Central Mexico: An<br />

Altitudinal Transect’, Econ. Bot. 57, 7–20.<br />

Perlman, D. L. and Adelson, G.: 1997, Biodiversity. Exploring Values and Priorities in Conservation,<br />

Blackwell Science, Malden, Massachusetts.<br />

Peters, R. L. and Lovejoy, T. E. (eds.): 1992, Global Warming and Biological Diversity, Yale<br />

University Press, New Haven, Connecticut.<br />

Peterson, A. T., Sánchez-Cordero, V., Soberón, J., Bartley, J., Buddemeier, R. W., and Navarro-<br />

Sigüenza, A. G.: 2001, ‘Effects of Global Climate Change on Geographic Distributions of<br />

Mexican Cracidae’, Ecol. Modelling 144, 21–30.<br />

Pielke, R. A., Marland, G., Betts, R. A., Chase, T. N., Eastman, J. L., Niles, J. O.,<br />

Niyogi, D., and Running, S. W.: 2002, ‘The Influence of Land-Use Change and<br />

Landscape Dynamics on the Climate System: Relevance to Climate-Change Policy<br />

Beyond the Radiative Effect of Greenhouse Gases’, Phil. Trans. R. Soc. Lond.<br />

A 360, 1–15.<br />

Piperno, D. R. and Pearsall, D. M.: 1998, The Origins of Agriculture in the Lowland Neotropics,<br />

Academic Press, San Diego.<br />

Poulsen, B. O. and Krabbe, N.: 1998, ‘Avifaunal Diversity of Five High-Altitude Cloud Forests on the<br />

<strong>Andean</strong> Western Slope of Ecuador: Testing a Rapid Assessment Method’, J. Biogeogr 25, 83–93.<br />

Price, M. F.: 1994, ‘Should <strong>Mountain</strong> Communities be Concerned about Climate Change?’, In<br />

<strong>Mountain</strong> Environments in Changing Climates. Beniston, M. (ed.)., Routledge, London. 431–457<br />

Pyke, C. R.: 2004, ‘Habitat Loss Confounds Climate Change Impacts’, Front. Ecol. Environ 2,<br />

178–182.<br />

Reilly, J., Stone, P. H., Forest, C. E., Webster, M. D., Jacoby, H. D., and Prinn, R. G.: 2001,<br />

‘Uncertainty and Climate Change Assessments’, Science 293, 430–433.<br />

Resolucion Ministerial No. 1173–2002-AG,: 2002, Declaran superficie ubicada en los departamentos<br />

de Ancash, Huánuco y Lima. como “Zona Reservada Cordillera Huayhuash”, El Peruano Normas<br />

Legales, Tuesday, December 24, 2002, pp. 235808.<br />

Reynolds, J. D., Mace, G. M., Redford, K. H., and Robinson, J. G. (eds.): 2001, Conservation of<br />

Exploited Species, Cambridge University Press, Cambridge, United Kingdom.<br />

Robbins, P.: 1998, ‘Authority and Environment: Institutional Landscapes in Rajasthan, India’, Annals<br />

Assoc. Amer. Geogr. 88, 410–435.<br />

Rodríguez, L. C. and Pascual, U.: 2004, ‘Land Clearance and Social Capital in <strong>Mountain</strong> Agroecosystems:<br />

The Case of Opuntia Scrubland in Ayacucho, Peru’, Ecological Economics 49,<br />

243–252.<br />

Rodríguez, L. O. and Young, K. R.: 2000, ‘Biological Diversity of Peru: Determining Priority Areas<br />

for Conservation’, Ambio 29, 329–337.<br />

Ruggiero, A.: 2001, ‘Size and Shape of the Geographical Ranges of <strong>Andean</strong> Passerine Birds: Spatial<br />

Patterns in Environmental Resistance and Anisotropy’, J. Biogeography 28, 1281–1294.<br />

Saxon, E., Baker, B., Hargrove, W., Hoffman, F., and Zganjar, C.: 2005, ‘Mapping Environments<br />

at Risk under different Global Climate Change Scenarios’, Ecol. Letters 8, 53–<br />

60.


100 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Scott, D., Malcolm, J. R., and Lemieux, C.: 2002, ‘Climate Change and Modeled Biome Representation<br />

in Canada’s National Park System: Implications for System Planning and Park Mandates’,<br />

Global Ecol. Biogeography 11, 475–484.<br />

Sierra, R., Tirado, M., and Palacios, W.: 2003, ‘Forest Cover Change From Labor- and Capitalintensive<br />

Commercial Logging in the Southern Choco Rainforests., Professional Geogr. 55,<br />

477–490.<br />

Silverio, W. and Jaquet, J. -M.: 2005, ‘Glacial cover mapping (1987–1996) of the Cordillera Blanca<br />

(Peru) using satellite imagery’, Remote Sensing Environ. 95, 342–350.<br />

Smith, D. N.: 1988, Flora and Vegetation of the Huascaran National Park, Ancash, Peru, with Preliminary<br />

Taxonomic Studies for a Manual of the Flora. Ph.D. Thesis, Iowa <strong>State</strong> University, Ames.<br />

Soulé, M. E. and Terborgh, J. (ed.): 1999, Continental Conservation. Scientific Foundations of<br />

Regional Reserve Networks, Island Press, Washington, D.C.<br />

Stanish, C.: 1992, Ancient <strong>Andean</strong> Politcal Economy. University of Texas Press, Austin.<br />

Stattersfield, A. J., Crosby, M. J., Long, A. J., and Wege, D. C.: 1998, Endemic Bird Areas of the<br />

World: Priorities for Biodiversity Conservation, BirdLife Conservation Series No. 7. Cambridge,<br />

United Kingdom.<br />

Stiglitz, J. E.: 2002, Globalization and its Discontents, W. W. Norton, New York.<br />

Strigl, A. W.: 2003, ‘Science, Research, Knowledge and Capacity Building’, Environm. Develop.<br />

Sustain 5, 255–273.<br />

Still, C. J., Foster, P. N., and Schneider, S. H.: 1999, ‘Simulating the Effects of Climate Change on<br />

Tropical Montane Cloud Forests’, Nature 398, 608–610.<br />

Stobart, H. and Howard, R. (eds.): 2002, Knowledge and Learning in the Andes: Ethnographic<br />

Perspectives, Liverpool University Press, Liverpool.<br />

Stoorvogel, J. J., Antle, J. M., and Crissman, C. C.: 2004, ‘Trade-off analysis in the Northern Andes<br />

to study the Dynamics in Agricultural Land Use’, J. Environ. Mngmt. 72, 23–33.<br />

Storch, D., Konvicka, M., Benes, J., Martinková, J., and Gaston, K. J.: 2003, ‘Distribution Patterns<br />

in Butterflies and Birds of the Czech Republic: Separating Effects of Habitat and Geographical<br />

Position’, J. Biogeography 30, 1195–1205.<br />

Swinton, S. M., Escobar, G., and Reardon, T.: 2003, ‘Poverty and Environment in Latin America:<br />

Concepts, Evidence and Policy Implications’, World Develop. 31, 1865–1872.<br />

Tanner, T.: 2003, ‘Peopling <strong>Mountain</strong> Environments: Changing <strong>Andean</strong> Livelihoods in North-West<br />

Argentina’, Geographical J. 169, 205–214.<br />

Téllez-Valdés, O. and Dávila-Aranda, P.: 2003, ‘Protected Areas and Climate Change: A Case Study of<br />

the Cacti in the Tehuacán-Cuicatlán Biosphere Reserve, México’, Conservation Biol. 17, 846–853.<br />

Terborgh, J., van Schaik, C, Davenport, L., and Rao, M. (eds.): 2002, Making Parks Work: Strategies<br />

for Preserving Tropical Nature, Island Press, Washington, D.C.<br />

Thiollay, J. -M.: 1996, ‘Distributional Patterns of Raptors along Altitudinal Gradients in the Northern<br />

Andes and Effects of Forest Fragmentation’, J. Tropical Ecol. 12, 535–560.<br />

Thomas, C. D., Cameron, A., Green, R. E., Bakkenes, M. Beaumont, L. J., Collingham, Y. C.,<br />

Erasmus, B. F. N., de Siqueira, M. F., Grainger, A., Hannah, L., Hughes, L., Huntley, B., van<br />

Jaarsveld, A. S., Midgley, G. F., Miles, L., Ortega-Huerta, M. A., Peterson, A. T., Phillips, O. L.,<br />

and Williams, S. E.: 2004, ‘Extinction Risk from Climate Change’, Nature 427, 145–148.<br />

Thompson, L. G.: 2000, ‘Ice Core Evidence for climate change in the Tropics: implications for our<br />

future’, Quat. Science Reviews 19, 19–35.<br />

Thompson, L. G., Mosley-Thompson, E., Davis, M. E., Lin, P.-N., Henderson, K., and Mashiotta, T.<br />

A.: 2003, ‘Tropical Glacier and Ice Core Evidence of Climate Change on Annual to Millennial<br />

Time Scales’, Climatic Change 59, 137–155.<br />

Thorp, R.: 1998, Progress, Poverty and Exclusion. An Economic History of Latin America in the<br />

20th Century, The Johns Hopkins University Press and Inter-American Development Bank,<br />

Baltimore, Maryland.


KENNETH R. YOUNG AND JENNIFER K. LIPTON 101<br />

Toral C., E., Feinsinger, P., and Crump, M. L.: 2002, ‘Frogs and a Cloud-Forest Edge in Ecuador’,<br />

Conservation Biol. 16, 735–744.<br />

Troll, C.: 1968, ‘The Cordilleras of the Tropical Americas: Aspects of Climatic, Phytogeographical<br />

and Agrarian Ecology’, Colloquium Geographicum 9, 15–56.<br />

Valdivia, C., Dunn, E., and Jette, C.: 1996, ‘Diversification as a Risk Management Strategy in an<br />

<strong>Andean</strong> Agropastoral Community’, Amer. J. Agric. Econ. 78, 1329–1334.<br />

Valencia, R., Pitman, N., León-Yánez, S., and Jørgensen, P. M. (eds.): 2000, Libro Rojo de las Plantas<br />

Endémicas del Ecuador 2000, Herbario QCA, Pontificia Universidad Católica del Ecuador,<br />

Quito.<br />

Vanacker, V., Govers, G., Barros, S., Poesen, J., and Deckers, J.: 2003, ‘The Effect of Short-Term<br />

Socio-Economic and Demographic Change on Landuse Dynamics and its Corresponding<br />

Geomorphic Response with Relation to Water Erosion in a Tropical <strong>Mountain</strong>ous Catchment,<br />

Ecuador’, Landscape Ecol. 18, 1–15.<br />

Vuille, M. and Bradley, R. S.: 2000, ‘Mean Annual Temperature Trends and their Vertical Structure<br />

in the Tropical Andes’, Geophys. Res. Lett. 27, 3885–3888.<br />

Vuille, M., Bradley, R. S., Werner, M., and Keimig, F.: 2003, ‘20th Century Climate Change in the<br />

Tropical Andes: Observations and Model Results’, Climatic Change 59, 75–99.<br />

Walker, R.: 2003, ‘Mapping Process to Pattern in the Landscape Change of the Amazonian Frontier.’<br />

Annals Assoc. Amer. Geogr. 93, 376–398.<br />

Walker, R., Drzyzga, S. A., Li, Y. Qi, J., Caldas, M., Arima, E., and Vergara, D.: 2004, ‘A Behavioral<br />

Model of Landscape Change in the Amazon Basin: The Colonist Case’. Ecol. Applications 14<br />

Supplement, S299–S312.<br />

Walther, G. -R., Post, E., Convey, P., Menzel, A., Parmesan, C., Beebee, T. J. C., Fromentin, J.-M.,<br />

Hoegh-Guldberg, O., and Bairlein, F.: 2002, ‘Ecological Responses to Recent Climate Change’,<br />

Nature 416, 389–395.<br />

Western, D. and Wright, R. M.: 1994, Natural Connections. Perspectives in Community-based<br />

Conservation, Island Press, Washington, D.C.<br />

Wheeler, J. C.: 1995, ‘Evolution and Present Situation of the South American Camelidae’, Biol. J.<br />

Linnean Soc. 54, 271–295.<br />

Whitlock, C., Shafer, S. L., and Marlon, J.: 2003, ‘The Role of Climate and Vegetation Change<br />

in Shaping Past and Future Fire Regimes in the Northwestern US and the Implications for<br />

Ecosystem Management’, Forest Ecol. Management 178, 5–21.<br />

Woodward, F. I.: 2002, ‘Potential Impacts of Global Elevated CO2 Concentrations on Plants’, Current<br />

Opinion Plant Biol. 5, 207–211.<br />

Young, K. R.: 1993, ‘National Park Protection in Relation to the Ecological Zonation of a Neighboring<br />

Human Community: an Example from Northern Peru’, <strong>Mountain</strong> Res. Develop. 13, 267–280.<br />

Young, K. R.: 1995. ‘Biogeographical Paradigms Useful for the Study of Tropical Montane Forests<br />

and their Biota’, in Churchill, S. P., H. Balslev, E. Forero and J. L. Luteyn (eds.), Biodiversity<br />

and Conservation of Neotropical Montane Forests, New York Botanical Garden, Bronx, New<br />

York, pp. 79–87.<br />

Young, K. R.: 1997, ‘Wildlife Conservation in the Cultural Landscapes of the Central Andes’,<br />

Landscape and Urban Planning 38, 137–147.<br />

Young, K. R.: 1998, ‘Deforestation in Landscapes with Humid Forests in the Central Andes: Patterns<br />

and Processes’, in Zimmerer, K. S., and Young, K. R. (eds.). Nature’s <strong>Geography</strong>: New Lessons<br />

for Conservation in Developing Countries, University of Wisconsin Press, Madison, Wisconsin,<br />

pp. 75–99.<br />

Young, K. R.: 2002, ‘Minding the Children: Knowledge Transfer and the Future of Sustainable<br />

Agriculture’, Conserv. Biol. 16, 855–856.<br />

Young, K. R. and Keating, P. L.: 2001. ‘Remnant forests of Volcán Cotacachi, northern Ecuador’.<br />

Arctic, Antarctic, and Alpine Research 33, 165–172.


102 ADAPTIVE GOVERNANCE AND CLIMATE CHANGE<br />

Young, K. R., León, B., Cano, A., and Herrera MacBryde, O.: 1997, ‘Peruvian puna, Peru’, in Davis,<br />

S. D., V. H. Heywood, O. Herrera MacBryde, and Hamilton, A. C. (eds.). Centres of Plant<br />

Diversity: A Guide and Strategy for their Conservation. Volume 3: The Americas. WWF and<br />

IUCN, London, pp. 470–476.<br />

Young, K. R. and León, B.: 1999, Peru’s Humid Eastern Montane Forests: An Overview of their<br />

Physical Settings, Biological Diversity, Human Use and Settlement, and Conservation Needs.<br />

DIVA, Technical Report No. 5, p. 1–97.<br />

Young, K. R. and León, B.: 2001, ‘Perú’, in M. Kappelle and A. D. Brown (eds.). Bosques Nublados<br />

del Neotrópico. INBio, Heredia, Costa Rica, pp. 549–580.<br />

Young, K. R. and L. O. Rodríguez: In press, ‘Development of Peru’s national protected area system:<br />

Historical continuity in conservation goals’, in K. S. Zimmerer (ed.). Globalization and New<br />

Geographies of Conservation, University of Chicago Press, Chicago.<br />

Young, K. R., Ulloa Ulloa, C., Luteyn, J. L., and Knapp, S.: 2002, ‘Plant Evolution and Endemism<br />

in <strong>Andean</strong> South America: an Introduction’, Bot. Review 68, 4–21.<br />

Young, K. R. and Zimmerer, K. S.: 1998, ‘Conclusion: biological conservation in developing countries’,<br />

in Zimmerer, K. S. and Young, K. R. (eds.), Nature’s <strong>Geography</strong>: New Lessons for Conservation<br />

in Developing Countries, University of Wisconsin Press, Madison, Wisconsin, pp. 327–344.<br />

Zerner, C. (ed.): 2000, People, Plants and Justice. The Politics of Nature Conservation, Columbia<br />

University Press, New York.<br />

Zimmerer, K. S.: 1991, ‘Wetland Production and Smallholder Persistence: Agricultural Change in a<br />

Highland Peruvian Region’, Annals Assoc. Amer. Geogr. 81, 443–463.<br />

Zimmerer, K. S.: 1996, Changing Fortunes: Biodiversity and Peasant Livelihood in the Peruvian<br />

Andes, University of California Press, Berkeley and Los Angeles, California, 308.<br />

Zimmerer, K. S.: 2002, ‘Common Field Agriculture as a Cultural Landscape of Latin America: Development<br />

and History in the Geographical Customs of Land Use’, J. Cultural Geogr. 19(2), 37–63.<br />

Zimmerer, K. S.: 2003, ‘Just Small Potatoes (and Ulluco)? The Use of Seed-Size Variation in “Native<br />

Commercialized” Agriculture and Agrobiodiversity Conservation among Peruvian Farmers’,<br />

Agricul. Human Values 20, 107–123.<br />

(Received 25 May 2004; accepted in final form 9 November 2005)


2<br />

Adaptive Management for Biodiversity Conservation<br />

under Climate Change – a Tropical <strong>Andean</strong><br />

Perspective<br />

David G. Hole, Kenneth R. Young, Anton Seimon, Carla Gomez<br />

Wichtendahl, Dirk Hoffmann, Klaus Schutze Paez, Silvia Sanchez,<br />

Douglas Muchoney, H. Ricardo Grau, and Edson Ramirez<br />

The tropical Andes are a globally significant region for biodiversity (Mittermeier et al. 2004).<br />

Dramatic environmental heterogeneity across the region including steep gradients in elevation<br />

and humidity, and complex mosaics of bedrocks and soils (Young, Chapter 8, this volume),<br />

together with the wide range of historical variability in <strong>Andean</strong> climates (Chapter 5, this<br />

volume), have helped shape this remarkable biological diversity (e.g., Trenel et al. 2008;<br />

Antonelli et al. 2009; Guarnizo et al. 2009). These same processes have also shaped the human<br />

context and provided natural resources that now sustain the wellbeing of millions of people,<br />

including lands for farming and grazing, water for households, irrigation and industries, and<br />

space for settlements. The region is one of the key places in the world for the development of<br />

early human societies, from organized settlements, to irrigated agriculture and the domestication<br />

of plant and animal species (Denevan 2001). Hence, it now contains large expanses of humandominated<br />

landscapes, from highly urbanized centers to more rural areas, where native forests<br />

have been replaced by agriculture and non-native tree plantations. Even the dry environments<br />

and high elevation sites are often used for extensive grazing of livestock. This leaves relatively<br />

few areas without a human presence, principal exceptions being the cool, moist cloud forests and<br />

the very humid paramos on the high mountains of the northern Andes. As a result of these<br />

historical and contemporary land-use patterns, there have long been tensions among the needs for<br />

conservation and protection of natural biodiversity, versus those for economic development and<br />

the reduction of social inequalities (Terborgh 1999). These tensions substantially increase the<br />

complexities of conservation planning in the region.<br />

Historically, the rate of change in the land cover of tropical landscapes is unlikely to have<br />

occurred as rapidly as it has in the last 50 years (Young 2007). Deforestation and conversion of<br />

land has led to an unprecedented loss of natural habitats in recent decades, with profound<br />

ramifications for the continued functioning of entire ecosystems (MEA 2005). Ongoing and<br />

projected climate change adds a substantial new component to this mix. Dramatic changes to<br />

19


oth biotic and abiotic systems and processes are already being seen across the tropical Andes,<br />

with glacial ice diminishing and upward biotic range extensions occuring even into high alpine<br />

areas (Seimon et al. 2007), with local people already altering their land uses in response (Postigo<br />

et al. 2008; Young 2008). Such changes are having profound impacts on species’ phenologies,<br />

distributions and abundance (Parmesan 2006) that will increase in magnitude in the future. Of<br />

particular conservation concern are the likely changes in representation and abundance of species<br />

within existing protected areas and across networks (Araujo et al. 2004; Hole et al. 2009), as well<br />

as the high likelihood of profound changes in the location and continued functioning of many<br />

<strong>Andean</strong> ecosystems (Anderson et al. Chapter 1, this volume). Indeed, such changes are likely to<br />

result in the formation of ‘no-analog’ communities (i.e. species assemblages for which there are<br />

no present-day examples) (Williams and Jackson 2007). Given the projected pace and likely<br />

consequences of climate change, magnified as they are in regions such as the high Andes<br />

(Bradley et al. 2006; IPCC 2007), it is critical that we adapt conservation management strategies<br />

in an effort to maintain their effectiveness under climate change. Without such effort, the region<br />

risks losing substantial components of its biodiversity (Larsen et al. Chapter 3, this volume), loss<br />

of key ecological processes (Aguirre et al. Chapter 4, this volume) and disruption of its<br />

ecosystems and consequent reduction or loss of the services they provide (Anderson et al.<br />

Chapter 1, this volume).<br />

In this chapter, we will focus on the conservation spotlight on adaptive management. It is<br />

not intended to be exhaustive in its coverage - the topics included are broad and need<br />

consideration in greater detail. Instead, we begin the process of identifying the range of risks and<br />

opportunities for biodiversity conservation and adaptive management presented by climate<br />

change within the unique context of the tropical <strong>Andean</strong> region. We highlight some of the<br />

principal tools available for assessing the vulnerability of biodiversity and ecosystems, and<br />

describe a range of conservation and management options that might be selected, based on the<br />

degree of manipulation and use required in order to maintain human wellbeing. In some cases,<br />

strict protection of very fragile ecosystems and endangered wild species is likely to be needed. In<br />

other cases, a mix of conservation through protected area systems and integrated planning for<br />

sustainable land use will likely be more appropriate. We then look briefly at options for<br />

monitoring climate change impacts and the effectiveness of management actions, before<br />

highlighting the opportunities (however limited) that climate change may bring for conservation.<br />

Finally, we identify critical institutional capacity needs within the region that are urgently<br />

required in order to effectively, efficiently, and equitably enable adaptation to the profound<br />

challenges posed by climate change.<br />

Current Status of <strong>Andean</strong> Biodiversity Conservation<br />

Protected areas are the single most important tool for biodiversity conservation in the tropical<br />

Andes region and have seen a substantial increase in number and area covered over the past 15<br />

years (Hoffmann et al. Chapter 22, this volume). Currently, around 15% of the four <strong>Andean</strong><br />

nations' land area is under national protected area status. While designation of protected areas<br />

has generally been based on biodiversity targets, it has not necessarily resulted in the<br />

representation of the most biologically pristine or valuable areas within priority ecosystems.<br />

More recently, protected areas have in many cases been created simply on the basis of sociopolitical<br />

opportunity.<br />

20


Even though robust assessments of the representativeness of the different protected area systems<br />

have not been carried out in a systematic manner, around 70-80% of species are likely to be<br />

represented within national protected areas (excluding municipal and departmental level<br />

protected areas, indigenous and community conservation areas, and indigenous communal<br />

lands). However, representation is biased towards lowlands and foothills, and little attention has<br />

been paid to ecological processes, especially in western regions.<br />

Capacity and data limitations mean that potential gaps in the network under climate<br />

change are largely unknown. However, there is now a push to integrate the large number of<br />

additional conservation areas managed at local and regional levels, into the National Protected<br />

Area Systems (Hoffmann 2009; Hoffmann et al. Chapter 22, this volume), in some cases led by<br />

municipal governments, and in others by native peoples, which could play a vital role in adapting<br />

national biodiversity conservation efforts to the challenges of climate change. There are also<br />

efforts under way to define national and regional conservation corridors that serve to link<br />

protected areas (e.g., Vilcanota-Amboró Corridor in Bolivia and Peru; three altitudinal corridors<br />

on the Eastern slopes in Colombia). Only in Colombia, however, has climate change recently<br />

been incorporated as an explicit component of the conservation planning process behind these<br />

efforts (see Appendix 2.1 and Hoffmann et al. Chapter 22, this volume, for further details).<br />

Hence, substantial basic research, including gathering of baseline data, as well as modelling of<br />

potential future shifts in species distributions and consequent changes in the provision of<br />

ecosystem services, is a critical priority for the tropical Andes region.<br />

Adaptive Management and the Identification of Future Vulnerabilities and<br />

Opportunities<br />

If we are to be proactive in addressing climate change, the identification of likely vulnerabilities<br />

and evaluation of possible adaptation responses, as well as the identification of potential<br />

opportunities, is paramount. Proactive strategies are likely to prove both more cost-effective<br />

(Hannah et al. 2007) and ethically responsible - in terms of preventing or ameliorating some of<br />

the worst potential impacts of climate change (Adger et al. 2009). However, given the magnitude<br />

of uncertainty in projections of climate change and in species, ecosystems and human responses,<br />

conservation planning must be set in the context of a range of potential future scenarios. Such an<br />

approach is not an excuse for inaction, but a call for adaptive management.<br />

Adaptive management is an iterative process of optimal decision making in the face of<br />

uncertainty, that attempts to reduce that uncertainty over time by system-level monitoring (see<br />

Sutherland 2006 for more details). Broadly speaking, such an approach has five stages (Figure<br />

2.1): 1) define plausible future scenarios and set conservation targets within this plausible range;<br />

2) perform conservation actions; 3) actions will lead to new behaviour within the system; 4)<br />

monitor to detect changes in the system; 5) analyze impacts of conservation actions and adjust<br />

initial targets accordingly. The cycle is then repeated.<br />

However, the high degree of historical and contemporary variability in <strong>Andean</strong> climate<br />

presents a substantial challenge to the identification of future scenario’s (Figure 2.1, Stage 1) in<br />

relation to climate change impacts on species, processes and ecosystems. For example, for much<br />

of the high <strong>Andean</strong> region above 3000 m, climatological data indicates that inter-annual thermal<br />

variability is regionally synchronous and largely controlled by the phases of the El Niño<br />

Southern Oscillation (ENSO), whereas the precipitation pattern varies spatially. At Cusco in<br />

21


Figure 2.1. Simple schematic of the five stages of the adaptive management process: 1) Define plausible future<br />

scenarios based on modelling and/or expert opinion and set conservation targets within this plausible range; 2)<br />

Perform conservation actions; 3) Actions will lead to new behaviour within the system, which will also be subject to<br />

influence by multiple external factors; 4) Monitor to detect changes in the behaviour of the system, as well as the<br />

influence of external factors, including measurement of the direct and indirect costs of management actions, both<br />

positive (e.g. rehabilitation of ecosystem services) and negative (e.g. opportunity cost of foregone agricultural<br />

production); 5) Feed these data into an analysis of the impacts of the conservation actions, and adjust the initial<br />

models/scenario’s and targets accordingly. This cycle is then repeated at appropriate time intervals.<br />

southern Peru, the inter-annual range of variation in daily maximum temperature over the<br />

summer growing season has been as much as 3.8 0 C (from +2.5 0 C accompanying a strong El<br />

Niño event in 1983, to -1.3 0 C under La Niña conditions in 1984). Historically, such a range of<br />

variability is likely the norm, although glaciological evidence suggests downward shifts in<br />

22


aseline mean temperature followed by a rebound to present levels probably occurred during the<br />

Little Ice Age (1500-1900 AD).<br />

Such natural climatic variability creates two challenges. Firstly, it translates into greater<br />

uncertainty in projections of future climatic change in the region, particularly at the relatively<br />

fine scales required for most spatial planning needs (Vuille et al. 2008). Secondly, it has shaped<br />

the evolutionary environment of regional biodiversity, potentially enhancing the resilience of<br />

species and ecosystems to sequential climatic extremes. Moreover, the amplitude of change of<br />

such short term climatic shocks (i.e. 3.8 0 C) exceeds the expected net thermal increase projected<br />

for the end of the 21st century under all but the most extreme emissions scenarios depicted by<br />

climate models for the Andes. Hence, there is a risk that some assessments could overestimate<br />

the impacts of climate change, projecting extinction in regions where a species or component<br />

populations are, at least potentially, pre-adapted. Perhaps more pertinent, however, is the<br />

capacity of individual species and ecological systems to adapt to a shifting thermal baseline that<br />

increasingly biases the distribution towards higher temperatures likely to surpass historical<br />

experience. Hence, the <strong>Andean</strong> context highlights a key issue in the development of scenarios of<br />

potential climate change impacts: what is the level of acceptable uncertainty? Practically<br />

however, when considering biodiversity targets, such uncertainty necessitates the use of multiple,<br />

but necessarily limited data sources, and therefore a judicious assessment of vulnerability.<br />

Assessing Vulnerability<br />

Direct Impacts<br />

Broadly speaking, the vulnerability of a species to climate change is a product of its<br />

susceptibility (defined by its intrinsic biological traits), its exposure (does it occur in a region of<br />

high climatic change?) and its adaptive capacity (can it adapt to climatic change?) (Figure 2.2).<br />

The vulnerability of an ecosystem can then be defined by the likely complex interactions and<br />

synergies between the relative vulnerabilities of its component species and proximate abiotic<br />

processes (see Chapter 4 for a discussion of the complexity of the interactions).<br />

Figure 2.2. Schematic representation of a species vulnerability to climate change, where each component varies<br />

from ‘low’ to ‘high’ according to the colour gradient, such that ‘X’ represents greatest vulnerability (i.e., the<br />

intersection of the three components – high susceptibility, high exposure and low adaptive capacity)<br />

23


These three components can be estimated using a variety of data sources and<br />

methodologies requiring varying degrees of technical capacity, input data quality and quantity,<br />

ability to provide robust future projections, definition of operational scale, and – critically – the<br />

associated uncertainties (Table 2.1). The simplest approach is to identify those biological traits of<br />

individual species (or across a taxon) that, based on expert opinion, are likely to predispose a<br />

species to being susceptible to climate change (Foden et al. 2009). Such traits might include a<br />

narrow altitudinal range at a high elevation, a high degree of habitat specialization, or a<br />

dependency on only a few prey or host species. Similarly, a species’ adaptive capacity might be<br />

deemed limited if it is a poor disperser, or if its genetic diversity is low. Exposure can then be<br />

estimated based on, for example, its thermal tolerance (Jiguet et al. 2006) in comparison to<br />

projected climatic anomalies across its area of occupancy. Such an approach has many<br />

advantages – for example, it relates directly to a specie's ecology; it avoids the need to develop a<br />

more complex model of the relationship between climate and species distribution; and it is data<br />

‘light’ in that it can be applied to a wide range of species and taxa. Disadvantages are that it<br />

provides limited information on the likely future spatial distribution of biodiversity that is critical<br />

for conservation planning and interactions between components of vulnerability may be missed.<br />

Table 2.1. Examples of the methods currently available for assessing vulnerability and<br />

identifying future scenarios of biodiversity pattern.<br />

Method Description Identifies Example Reference<br />

Trait-based Expert derived identification of Susceptibility; Adaptive (Foden et al. 2009)<br />

assessments broad life-history characteristics<br />

that predispose susceptibility<br />

capacity<br />

Shifts in climatic Spatial representation of shifts Exposure; Refugia (Ohlemuller et al. 2006;<br />

parameters in climatic parameters derived<br />

from GCM or RCM outputs<br />

Williams et al. 2007)<br />

Species distribution Modelled statistical association Susceptibility via e.g. present (Pearson and Dawson<br />

models (SDMs) between a species present-day and future range overlap or 2003; Hole et al. 2009)<br />

distribution and current climate change in future range extent;<br />

– can then project relationship Exposure; Spatially explicit<br />

onto future climates (e.g.<br />

Maxent; Generalized Additive<br />

Models; Boosted-Regression<br />

Trees)<br />

projections<br />

Dynamic [global] Process based models that Susceptibility of individual (Scott et al. 2002;<br />

vegetation models simulate shifts in, and dynamics modelled plant species; Hannah et al. 2008)<br />

(D[G]VMs) of, vegetation, in response to Susceptibility of other species<br />

climate and other drivers (e.g. via spatially explicit<br />

LPJ; VECODE; BIOME) projections of future presence<br />

of suitable habitat<br />

Population viability Data-intensive, species-specific Vulnerability (Brito and Figueiredo<br />

analysis (PVA) model that determines viability<br />

of one or more populations over<br />

time, based on internal and<br />

external drivers of population<br />

dynamics<br />

2003; Vargas et al. 2007)<br />

24


Towards the other extreme of the data requirement spectrum is species distribution<br />

modelling (SDM) (Graham et al. Chapter 21, this volume; Pearson and Dawson 2003). Given<br />

adequate and robust gridded data on the presence and absence of a species across its entire range,<br />

or sufficient accurate and fine-scale presence localities across an environmentally representative<br />

sample of its range, together with technical capacity in statistical analysis and GIS, it is now<br />

relatively easy to generate present and future projections of individual species distributions at<br />

local, regional or continental scales (although interpretation of model outputs still requires<br />

ecological expertise). Advantages include quantified estimates of sensitivity (e.g.,<br />

reduction/increase in future range extent), and spatially explicit projections of species’ potential<br />

future distributions – a product that is unavailable by any other means. Disadvantages include the<br />

wide-range of assumptions that such models rely on (see Graham et al. Chapter 21, this volume)<br />

and the consequent requirement for judicious interpretation. Recent advances in the development<br />

of the next generation of SDMs seek to overcome many of these disadvantages and assumptions<br />

(e.g., Keith et al. 2008), by incorporating habitat requirements, population dynamics and<br />

dispersal into the bioclimatic modelling framework. On the downside, such models will require a<br />

vastly improved understanding of a specie's ecology. The pros and cons inherent across these<br />

methodologies should therefore be carefully considered when applying analyses to inform the<br />

adaptive management process.<br />

Indirect Impacts<br />

An understanding of the direct impacts of climate change on biodiversity is beginning to emerge<br />

in the <strong>Andean</strong> region, as is an awareness of the diversity and magnitude of the responses by<br />

people whose lives and livelihoods are altered by climate change (Young and Lipton 2006).<br />

However almost no attention has been paid to the impacts that human responses to climate<br />

change will have upon biodiversity. Yet these ‘indirect’ impacts have the potential to be of a<br />

magnitude and scope that will rival or exceed the direct impacts. Here we identify several areas<br />

in which the interactions between people, biodiversity and climate change in the tropical Andes<br />

have the potential to be critical and which urgently warrant further research and consideration in<br />

the adaptive management process.<br />

Food: Climate change will have profound impacts on our ability to grow food. Regions currently<br />

suitable for a particular crop may become unsuitable, as climate change pushes local<br />

microclimates beyond the crop's temperature or water tolerance (Lobell et al. 2008). Unless new<br />

varietals can be developed, cropping practices adapted, or alternative crops identified, farmers<br />

may be forced to migrate to new areas (Warner et al. 2009), putting increased pressure on natural<br />

habitats or urban zones. Of course the converse may also be true, with changes in climate<br />

generating increases in yield and allowing some crops to be grown in areas that were previously<br />

unsuitable. In Peru’s Cordillera Vilcanota for example, cultivation practices on highland slopes<br />

have moved progressively higher in recent decades as moderating temperatures have expanded<br />

the domain of a viable growing season upward. Studies at Nunoa performed in 1964-65<br />

identified a relatively frost-free 5-month period representing a suitable growing season at<br />

4,236m, close to the highest tilled fields at that time. A nearby climate station meanwhile, at<br />

4,543m, recorded frosts throughout the year (Winterhalder and Thomas 1978). By 2003,<br />

however, potato cultivation in the nearby Pitumarca valley had moved to 4,550m around several<br />

25


communities. The warming regional climate has thus been accompanied by a rise in the limit of<br />

cultivation by as much as 300m over the 38 year period (Halloy et al. 2005; Figure 2.3). Further<br />

upward shifts in cultivation may put still largely pristine high elevation habitats under pressure.<br />

In the Cordillera Apolobamba in Bolivia, the upward shift in agricultural activities has displaced<br />

livestock (camélidos) to even greater elevations, with consequent impacts on high altitude<br />

ecosystems in the region (Schulte 1996). Where higher elevations have limited area, the same<br />

number of animals is often packed into a smaller area resulting in increased erosion. Yields of<br />

cash-crops crucial to livelihoods in the region, such as coffee, will also respond to changes in<br />

temperature, as projected for other regions of the America’s (e.g., Mexico; Schroth et al. 2009).<br />

Increased pressures on upslope protected areas, as cost-effective yields of Coffea arabica<br />

become increasingly limited to higher elevations, are a likely consequence.<br />

Water: Deglaciation caused by climate change over the past several decades, has led to a<br />

significant reduction in dry season flows in glacier-fed rivers (Francou et al. 2005; Vergara et al.<br />

2007). This has necessitated the creation of upstream reservoir capacity to provide a buffer to<br />

ensure sufficient flow for hydropower generation and agricultural needs downstream. For<br />

example, the largest high-alpine lake in the Andes, Peru’s 32 km 2 Laguna Sibinacocha at 4,900<br />

m, which provides habitat for thousands of flamingos and other waterfowl, was artificially<br />

enlarged by a dam in the mid-1990s to augment diminishing dry season flow far downstream at<br />

the Machu Picchu hydroelectric generation plant on the Vilcanota-Urubamba river (Seimon<br />

2001). Anecdotal evidence suggests that avian diversity and abundance on the lake may have<br />

decreased as a result, due to inundation of traditional nesting areas (see Anderson et al. Chapter<br />

1, this volume for further examples).<br />

Health: A major potential impact of climate change on human health is the change in the<br />

incidence and geographic range of vector-borne diseases (Martens 1998; Patz et al. 2002).<br />

Changes in ambient temperatures and rainfall have affected the seasonality, duration of<br />

outbreaks and morbidity profiles of both malaria and dengue fever, diseases whose transmission,<br />

distribution and seasonality are linked to climatic conditions (Poveda et al. 2001; Ruiz et al.<br />

2006). Projected increases in temperature under climate change may now drive these diseases<br />

into formerly mosquito-free territories. As in the past, communities may choose to seek out<br />

malarial-free areas (Gade 1999) beyond the expanding disease front, putting further pressure on<br />

natural habitats within these regions.<br />

Energy: Biofuels have been proposed as one method for mitigating greenhouse gas emissions.<br />

Whether or not biofuels actually offer carbon savings depends on how they are produced. In<br />

many regions the production of food crop-based biofuels is leading to the conversion of<br />

rainforests, savannas, peatlands, grasslands and other natural ecosystems to agriculture,<br />

generating a substantial “carbon debt” (Fargione et al. 2008), while simultaneously causing<br />

widespread degradation of natural ecosystems. In the <strong>Andean</strong> region, it is unclear as yet whether<br />

the increase in biofuel demand will have substantial negative impacts on <strong>Andean</strong> biodiversity as<br />

a result of deforestation and other legal, illegal or politically motivated land use changes.<br />

Although laws have been passed recently in Bolivia for example, encouraging biofuel production<br />

for national consumption, the government has so far opposed producing biofuels for exports,<br />

because of the potentially negative impacts on food security and on small farmers.<br />

26


Hydroelectric power generation meanwhile is affected by changes in climate and stream<br />

flows. In Peru, the prospect of decreasing dry season flows from glacier-fed rivers has motivated<br />

an adaptive electricity generation strategy shifting away from hydropower, which formerly<br />

contributed 90% of the country’s power supply, to greenhouse gas producing gas/thermoelectric<br />

power generation facilities (Vergara et al. 2007).<br />

Cordillera Vilcanota, Peru<br />

Frost frequency vs. elevation(1964-65)<br />

monthly percentage of days < 0°C<br />

4,236m 4,543m<br />

<br />

season<br />

---------- ~4,550 m<br />

Pitumarca valley,<br />

2003<br />

Figure 2.3. Rising altitudinal limit to cultivation in the Andes associated with regional warming.<br />

Photograph © Anton Seimon 2003<br />

27


Human migration: Large-scale displacements of people resulting from climate change is already<br />

a reality, with the United Nations High Commissioner for Refugees in 2008 arguing for the first<br />

time that the increase in displaced people globally was at least partly attributable to climate<br />

change-related conflict. However, even in the absence of conflict, this trend is likely to increase<br />

substantially in synergy with many of the drivers outlined above, generating new, or modifying<br />

existing patterns. Rural-urban migration is already significant in the Andes, and will be further<br />

enhanced by the increased unpredictability of agricultural production. While this process may<br />

result in socioeconomic improvements, as well as more efficient use of natural resources, it is<br />

also likely to result in increased human pressure in urban areas, local increases in water demand<br />

and water diversion, with consequent impacts on freshwater and terrestrial ecosystems.<br />

Conversely, the reduction in the human footprint in rural areas could lead to a beneficial indirect<br />

impact on biodiversity (Grau et al. 2003; Grau and Aide, 2007), assuming the land has not been<br />

badly degraded and is allowed, or actively encouraged to revert to a natural state.<br />

Adaptive Management Responses<br />

Mitigating identified vulnerabilities to climate change in the tropical <strong>Andean</strong> region will be<br />

challenging, given the complex patterns and associated interactions of biodiversity and human<br />

land use. Here we highlight key adaptive management responses (see Heller and Zavaleta 2009<br />

for a more comprehensive review), noting that these responses require coordination and<br />

integration at local, national and regional levels (Figure 2.4) if they are to be effective. Although<br />

here we break these responses down into species-, site- and landscape-level approaches, such<br />

distinctions must become necessarily blurred when developing a holistic adaptation strategy<br />

under climate change.<br />

Regional planning process<br />

A systematic conservation planning process (Margules and Pressey 2000) for the region is a<br />

critical research and policy requirement for the near term. While the potential foundations for<br />

such a process exist (e.g., the Estrategia Regional de Biodiversidad para los Paises del Tropico<br />

Andino), climate change is at best a marginal component in such plans. Any process must also be<br />

cognizant of the specific national and local contexts in which biodiversity conservation will be<br />

carried out.<br />

Species-level approaches<br />

Removing current stresses and threats to a vulnerable species or ecosystem (e.g., hunting<br />

pressure, habitat fragmentation) requires minimal knowledge of potential climate change<br />

impacts, yet it is a practical way to increase resilience since a system facing multiple stressors is<br />

less able to cope with additional pressures from climate change. The <strong>Andean</strong> bear (Tremarctos<br />

ornatus) is one example that would likely benefit from such an approach, given its large homerange<br />

and lack therefore of site-specific conservation options. Individual species action plans<br />

such as the Biodiversity Action Plans (BAPs) in Europe, or the US Endangered Species Program<br />

that are backed up by legal enforcement, also reduce the vulnerability of a species to climate<br />

change. A limited number of such plans already exist in the <strong>Andean</strong> region, e.g., for Polylepis in<br />

the Callejon de Conchucos in Ancash, Peru; and Southern Horned Curassow Pauxi unicornis in<br />

28


Bolivia. Identification of further species for which alternative adaption options are unlikely to be<br />

sufficient to maintain viable populations should be a priority.<br />

Figure 2.4. Adaptive management options for increasing resilience to biodiversity, across spatial scales.<br />

Where a species is imminently at risk of extinction, its translocation (or assisted migration) to<br />

suitable but inaccessible habitat (perhaps due to distance from the species’ current range) and ex<br />

situ conservation have to be seriously considered, despite acknowledged costs, difficulties and<br />

risks (e.g., Hoegh-Guldberg et al. 2008). For example, following precipitous population declines<br />

caused principally by Chytridiomycosis (Pounds et al. 2006), many <strong>Andean</strong> amphibians now<br />

require an ex situ approach, namely bringing the most threatened species into captive breeding<br />

until release back into the wild becomes feasible. In other situations, this may be done in situ,<br />

translocating plants and animals to sites that will be buffered from the effects of climate change<br />

(see ‘refugia’ below) or at sites that can be managed such that the negative effects of climate<br />

change are mitigated. Crucially, translocation will also likely have to include long-lived elements<br />

of critical habitat types, such as seedlings of trees that upon maturity will provide food, shelter<br />

and other resources, for the range of species and taxa that will be shifting their distributions into<br />

these newly climatically suitable sites/regions.<br />

Site-level approaches<br />

Protected areas (PAs) and PA networks remain the optimal strategy for conserving global<br />

biodiversity (Bruner et al. 2001). Increasing the coverage of the global PA network in order to<br />

fill ‘gaps’ (i.e., species whose ranges are not included in the current network (Rodrigues et al.<br />

2004)) is an urgent conservation priority. However, under climate change simply trying to<br />

achieve adequate representation of current biodiversity pattern when choosing new sites is likely<br />

29


to be insufficient. Instead, other considerations, particularly potential shifts in both biodiversity<br />

pattern and process, must be explicitly considered.<br />

Relict forest patches above the mean closed timberline are a common feature on many steep<br />

mountain slopes up to 5,400 m, especially along the eastern Andes. Composed primarily of<br />

Polylepis and Gynoxys (Fjeldså and Kessler 1996; Coblentz and Keating 2008), these forest<br />

islands harbour assemblages of other species including endemic fauna and flora such as the<br />

critically endangered bird, the Royal Cinclodes (Cinclodes aricomae). Indeed, these patches have<br />

likely functioned as biological refugia through climatically stressful epochs in the past,<br />

including most recently the Little Ice Age. Their apparent stability therefore suggests resilience<br />

in the face of climate change. It has therefore been argued that they should be considered high<br />

priority targets for conservation, both for their current ecological function and for their potential<br />

to serve as refugia under the changed climatic conditions of the future. However, some caution is<br />

needed since these relict patches could simply represent remnants of much more extensive<br />

forests fragmented by humans over centuries or millennia (Fjeldså and Kessler 1996; Fjeldså<br />

2002; Kessler 2002), raising doubt over their ability to act as future refugia. Other putative<br />

refugia have also been identified in the <strong>Andean</strong> region, including areas with stable mist<br />

formation on the Pacific slope and the intermontane basins (Fjeldså et al. 1999). Clearly, the<br />

robust identification of functional refugia in the tropical Andes must be a research priority.<br />

Environmental gradients containing sufficient natural habitat to create land-use mosaics that<br />

will maintain functional habitat connectivity, should be incorporated into conservation units<br />

wherever possible (e.g. sites, corridors, watersheds). Both short, sharp, and broad, diffuse<br />

gradients should be considered where relevant, because of the different options they provide for<br />

rapid dispersal of species or adaptation through genetic plasticity (Killeen and Solorzano 2008).<br />

Gradients of particular relevance to the Andes include elevational, edaphic, and humidity<br />

gradients. This strategy is already being utilized in Peru, where three large national parks have<br />

been created that include entire elevational gradients (Rio Abiseo, Yanachaga-Chemillen, and<br />

Manu National Parks) (Young and Lipton 2006), in Bolivia (Madidi, Carrasco, and Amboró<br />

National Parks), in Ecuador (La Reserva Cofanes-Chingual), and in Colombia.<br />

The tracking of ecotones (the transition between two ecosystems or biomes) offers a further<br />

adaptive management option. An important example in the <strong>Andean</strong> context is the treeline, the<br />

dynamics of which are sensitive to both climatic and human impacts (Bader et al. 2007; Young<br />

and Leon 2007). Given the relatively high environmental and climatic variability at ecotones,<br />

populations in or near these areas are likely to be pre-adapted to a relatively high level of<br />

physiological stress and may possess adaptive genetic traits that are absent from core populations<br />

(Killeen and Solorzano 2008). Monitoring and tracking of ecotones is now feasible thanks to<br />

recent advances in remote sensing (see Monitoring section).<br />

Riverine forest corridors have served as pathways or refugia for many forest taxa (e.g., in<br />

Madagascar; Wilme et al. 2006) and are likely to continue to do so under climate change. By<br />

connecting higher elevation watersheds to the lowlands, they incorporate many of the<br />

environmental gradients already highlighted as key targets in the <strong>Andean</strong> region. Examples<br />

include the Topo and Palora rivers in Ecuador, whose watersheds are protected by the Llangantes<br />

and Sangay National Parks, thereby covering an elevational gradient of almost 3000 m.<br />

30


Maintaining riverine forest can reduce the build up of sediment and/or chemicals leaching into<br />

the water from the surrounding landscape, preserving water quality for aquatic species. The<br />

rivers themselves are also critical for both freshwater and biodiversity.<br />

In order to reduce or remove threats to a vulnerable species or ecosystems, strict protection of<br />

some localities within PAs, or of entire PAs, will be required. Most tropical <strong>Andean</strong> national<br />

protected area systems include strict protection as a management option, typically defining<br />

specific vulnerable zones as “no impact” or “no visit”. Use of this management option is likely to<br />

become even more crucial under climate change and sensitivity will need to be shown in<br />

deriving the correct balance between sustainable use of an area by local communities and the<br />

preservation of species and key habitats.<br />

SDM projections provide a means of identifying regions or even sites that could represent<br />

concentrations of future species richness and/or regions projected to experience large numbers<br />

of species moving through them, as species’ ranges shift. As a result, they have begun to be<br />

incorporated into adaptive management planning, e.g. for the Cape Proteaceae in South Africa<br />

(Williams et al. 2005); for the African IBA network (Hole et al. in press). If the assumptions and<br />

uncertainties in such modelling exercises are fully acknowledged, these techniques currently<br />

provide the only method of proactively identifying future regions of potentially high<br />

conservation value that may not be included in current conservation plans.<br />

Similarly, use of SDM and dynamic vegetation models (DVM) projections provide a means of<br />

informing future site-based management strategies by identifying and targeting broad<br />

management strategies across networks of conservation sites, that reflect potential changes in the<br />

species composition of those sites. Broad strategies can be characterized from SDM derived<br />

projections of inter-site differences in the number of immigrant, emigrant and persistent species<br />

across a network of sites (e.g., for the African IBA network; Hole et al. in press), or based on<br />

DVM projections of future vegetation patterns within sites (e.g., the Canadian Parks system;<br />

Scott et al. 2002).<br />

Landscape-level approaches – corridors and landscape permeability<br />

Large protected areas may include sufficient environmental gradients such that they can be<br />

managed as a single landscape unit (Figure 2.5 A). Within such sites, species range shifts will<br />

likely leave relict distributions in addition to forming new species assemblages along elevation<br />

and humidity gradients. A variety of approaches may then be viable for preservation,<br />

translocation, and restoration. Most <strong>Andean</strong> landscapes, however, include a wide variety of<br />

human land uses, creating mosaics within which conservation strategies are more constrained<br />

(Figure 2.5 B). Given the potential speed and magnitude of species range shifts in response to<br />

climate change, and the very high likelihood of substantial species turnover within conservation<br />

areas (Hole et al. 2009), most site-based conservation strategies will continue to fulfil their role<br />

only if the landscapes within which they are embedded (i.e., the matrix) allow species to move<br />

through them across ecologically relevant temporal and spatial scales (Gascon et al. 1999).<br />

Corridors are landscape scale conservation and management units (Soule and Terborgh 1999)<br />

that have already been defined across parts of the <strong>Andean</strong> region. Within such landscape units,<br />

promotion of land use efficiency may prove to be key. Encouraging the concentration of<br />

agricultural production within restricted, but high yield areas, may favour de-intensification in<br />

31


Figure 2.5. <strong>Andean</strong> landscapes have strong environmental gradients, with dramatic changes in elevation joined by<br />

shifts in humidity, mediated through rainfall amounts, degree of seasonality, and soil moisture.<br />

XA, represents a protected landscape, wherein the goals of biodiversity conservation take priority. Climate change<br />

can force species shifts along the elevational gradient (b), along the humidity gradient (c), or in some cases may not<br />

cause species shifts or may not affect a particular site (a). Strategies within this protected area may include special<br />

attention to a, given it serves as a conservation refugia, while facilitated transformation of important habitat<br />

elements (e.g., tree plantings) and translocation of species of concern to places predicted to be important for<br />

altitudinal shifts (b) and range shifts to wetter or drier sites (c).<br />

XB, represents an inhabited landscape, utilized by people and with multiple land owners and land tenures. Here<br />

there could also be refugia (a), elevational shifts (b) and shifts with changing humidity regimes (c), but the<br />

conservation strategies may be more constrained in deference to competing needs for land use. Instead management<br />

strategies or conservation incentives could be used to lessen the landscape matrix’s resistance to species movements,<br />

through formal conservation corridors or through land uses such as agroforestry or shade coffee. The refugia (a)<br />

could be particularly important conservation targets, requiring active management interventions including strict<br />

protection and habitat management.<br />

32


less productive and more vulnerable areas (e.g., on slopes and at higher elevations) and reduce<br />

encroachment on natural habitats (Green et al. 2005; Grau and Aide 2007), although the<br />

effectiveness of such ‘land sparing’ remains questionable (Ewers et al. 2009).<br />

Monitoring<br />

Monitoring is a critical component of the adaptive management process (Figure 2.1, Stage 4). It<br />

provides baseline assessments of biodiversity and environmental parameters of interest (e.g.,<br />

climate) from which to interpret current and potential future changes; a means to validate model<br />

outputs (e.g. is a species range shifting in the manner projected by its SDM?); an early-warning<br />

mechanism of unexpected climate change impacts; and a measure of the effectiveness of a<br />

conservation action, that must then be fed back into adaptive management. Choosing the correct<br />

indicators to monitor can be challenging given limited resources, and they will depend on the<br />

goal of the monitoring. Suitable indicators can be individual species (e.g., amphibians with a<br />

recognized sensitivity to climatic conditions); assemblages (repeat censusing of floral<br />

assemblages is already underway through the GLORIA initiative but is otherwise lacking in the<br />

Andes); ecosystems (e.g., paramos are a crucial ecosystem to monitor since there is nowhere<br />

‘upwards’ for constituent species to go (Anderson et al. Chapter 1, this volume)); processes or<br />

interactions (e.g., the well documented relationship between fig trees (Ficus spp.) and their figwasp<br />

pollinators (Agaoninae) represents a powerful indicator since its disruption through climate<br />

change would resonate across communities or even entire ecosystems); composites of these (e.g.,<br />

GLORIA http://www.gloria.ac.at/?a=20 and TEAM http://www.teamnetwork.org/en/ networks);<br />

or derived products from remote sensing (e.g., NDVI to assess seasonal ‘greening’). Utilizing a<br />

broad range of indicators increases our ability to detect climate change signals and rapidly adapt<br />

management responses accordingly. Long-term data series are also vital for understanding the<br />

natural range of variability of systems, yet standardized long-term monitoring records are lacking<br />

throughout the majority of the tropical Andes.<br />

The utility of remote sensing data for monitoring purposes has increased rapidly over the<br />

past decade and merits closer inspection, given its potential for relatively cheap, standardized<br />

monitoring over much of the globe, including the tropical Andes. There is now a wide array of<br />

current and/or planned Earth observing systems (satellite, aerial, and in situ) at local, national<br />

and regional scales, that collect and disseminate monitoring data. The Global Earth Observing<br />

System of Systems (GEOSS) is coordinating the collection, distribution and use of these data<br />

across multiple themes of importance to society: climate, weather, energy, health, agriculture,<br />

water, disasters, biodiversity, and ecosystems. Recent changes in data policy have made remote<br />

sensing data available at little or no cost. Table 2.2 provides an overview of current satellite and<br />

associated sensors most relevant to monitoring climate change effects on biodiversity and<br />

ecosystems in the tropical Andes. As an example of the utility of such data, satellite and aerial<br />

photography, together with radar and LIDAR data, when coupled with field-based calibration<br />

and validation, now provide an operational means of defining and monitoring ecotones,<br />

including changes in treelines, through estimation of tree height, basal area and stem volume<br />

(Holmgren 2004), treeline structure (Rees 2007), species composition (Holmgren and Persson<br />

2004), tree migration (Næsset and Nelson 2007), and changes in treeline over time (Zhang et al.<br />

2009). The number of satellite-borne sensors has increased dramatically and many hyperspectral,<br />

radar, and LIDAR missions are on-going or planned. Yet access to this high data volume, to<br />

33


usable, derived products, and to technical capacity for interpretation remain limiting factors for<br />

monitoring climate change impacts on biodiversity.<br />

Table 2.2. An overview of principal current remote sensing data, of relevance to biodiversity and<br />

ecosystem monitoring in the tropical Andes . * = commercial system; Pan = panchromatic; XS =<br />

multispectral.<br />

Satellite / Sensor Temporal<br />

Resolution<br />

(days)<br />

Moderate-resolution (250-8000m)<br />

Spatial<br />

Resolution<br />

(m)<br />

Spectral<br />

Resolution /<br />

Band<br />

Program<br />

Life<br />

Derivative<br />

Products<br />

AVHRR 1 1000-8000 XS, thermal 1982- IGBP Land Cover,<br />

GIMMS NDVI<br />

SeaWIFS 1 1000 XS 1998- Ocean Color<br />

SPOT-Vegetation (VGT) 2 1000 XS 1998- GLC 2000, GIMMS<br />

MODIS 1 250-1000 36 channels in<br />

visible - thermal<br />

MERIS 1 300 36 channels in<br />

visible -<br />

thermal<br />

Medium-resolution optical (10-250m)<br />

NDVI<br />

2000- MODIS LAI/fPAR, VI,<br />

NPP, Land Cover, Fire.<br />

Land Surface<br />

Temperature<br />

2000- GLC 2005<br />

Landsat 1-3 MSS 16 80 XS 1973-87 Global Land Survey<br />

1970<br />

Landsat 4-5 TM 16 28.5 XS, thermal 1984- Global Land Survey<br />

1970, 1990, 2000,<br />

2005<br />

Landsat 7 ETM 16 15-28.5m XS, thermal. 1999- Global Land Survey<br />

pan<br />

2000, 2005<br />

SPOT 1-5 2 10-20 Green, red, 1990-<br />

near-IR<br />

ASTER 16 15, 30, 90 XS, thermal 1999- ASTER Global Digital<br />

Elevation Model<br />

IRS Variable 2.5, 20, 30 Pan, XS 1988-<br />

CBERS 1-3 5 20 XS, Pan 2003-<br />

ALOS 46 10 XS 2006-<br />

Medium-resolution radar(10-250m)<br />

ERS 1-2 35 25 C-band SAR 1996-<br />

JERS-1 44 10 L-band 1992-98<br />

Radarsat 1-2 24 8, 20, 50 C-band 1995-<br />

PALSAR 46 10, 100 L-band 2006-<br />

SRTM short-lived 30,90, 1000 Interferometric 2002 SRTM30, SRTM90<br />

radar<br />

DEMS<br />

ENVISAT ASAR 35 C-band 2002-<br />

High-resolution (


Opportunities – Demonstrating the Value of Maintaining Functional<br />

Ecosystems<br />

Climate change is already having negative repercussions for both biodiversity and people and<br />

poses a multitude of future risks. If there is any positive side to this, it may be that climate<br />

change has helped push biodiversity, ecosystems and the services they provide, to finally being<br />

recognized and valued as critical components of the Earth system, rather than as a global<br />

commons to be used and abused without concern for the consequences. Attaching a ‘value’ to<br />

nature is seen as anathema by some and care must be taken to ensure that the value of an<br />

ecosystem reflects not just its direct commercial or marketable value (e.g., provision of<br />

hydrological services), but also those goods and services that are extremely difficult to quantify<br />

and value (e.g., cultural services such as spiritual value). Nevertheless, the lack of any sort of<br />

valuation is an underlying cause of the degradation of ecosystems and the loss of biodiversity<br />

that we see today (TEEB 2008). Such a valuation becomes even more critical given that the<br />

maintenance of functional ecosystem services is likely to represent the most cost-effective way<br />

to avoid and/or adapt to many of the projected impacts of climate change on human wellbeing.<br />

So-called ‘ecosystem based adaptation’ (an adaptation strategy now defined and endorsed by<br />

IUCN, the World Bank and many other international organizations) is likely to be of particular<br />

relevance in much of the developing world, where ‘technical’ adaptation options, such as the<br />

building of a large-scale water treatment plant in response to diminished freshwater quality, may<br />

be a far less pertinent adaptation option than, for example, maintaining or restoring forest-cover<br />

throughout a key watershed. In light of these developments, including a valuation (even a “backof-the-envelope”<br />

calculation) of the ecosystem services preserved or restored as part of any<br />

adaptive management strategy, represents a key tool to leverage biodiversity-favourable policy<br />

across stakeholders.<br />

Payment for Ecosystem Services (PES)<br />

Paying individuals or communities for the services provided by natural ecosystems on their land<br />

is still rare in the <strong>Andean</strong> region. However, successful PES schemes do exist. For example, in the<br />

Los Negros river watershed in Bolivia, Fundación Natura has initiated a project that seeks to<br />

reduce or prevent reductions in water quality and quantity for downstream users (principally<br />

farmers relying on irrigation) caused by upstream deforestation. Through the scheme, these<br />

downstream irrigators agree to compensate upstream farmers to protect certain forests and<br />

restore others, thereby ensuring the quality of their water supply. In Ecuador meanwhile, the<br />

Pimampiro initiative places a specific fee for watershed forest protection onto the water bills of<br />

nearly 1300 families; 20% of these funds are then used to pay 19 upstream farmers to conserve<br />

their forests (390 ha of forest and 163 ha of paramo) (Camacho 2008). Major cities such as<br />

Bogotá in Colombia and Santa Cruz in Bolivia are now also beginning to utilize water-based<br />

PES schemes through small additions to water or electricity fees to pay for watershed<br />

conservation. In Quito, Ecuador, which receives its water supply from the high plateaus of the<br />

surrounding <strong>Andean</strong> range, a Water Conservation Fund (FONAG) was set up in 2000 to manage<br />

and direct revenue generated by a water consumption fee, to fund conservation and restoration<br />

projects in surrounding watershed. These projects include, for example, improving sheep and<br />

cattle production practices in order to reduce negative impacts on land cover and water quality.<br />

35


On a potentially far greater scale, the introduction of REDD (Reduced Emissions from<br />

Deforestation and Degradation) as a climate change mitigation strategy, could provide a<br />

potentially massive increase in the funds available for forest conservation through their role in<br />

carbon sequestration and storage. The benefits to biodiversity could be increased further still if<br />

policymakers could agree to prioritize forests with high biodiversity for REDD funding. The<br />

potential for REDD to contribute to biodiversity conservation in the tropical <strong>Andean</strong> region is<br />

substantial and numerous pilot studies are being carried out. For example, a recent study in<br />

Peru’s Cordillera Azul National Park, has demonstrated that management activities over the past<br />

six years have decreased deforestation rates in its buffer zone and prevented further deforestation<br />

in the park. However, without a continuous source of revenue, such as REDD could provide,<br />

continued management will be impossible and the park will likely succumb to threats from<br />

logging, oil extraction and the expansion of the agricultural frontier that have impacted the park<br />

in the past.<br />

Incorporating Biodiversity Conservation into Development Planning<br />

There is a growing awareness, from local NGOs to the large multilateral organizations, that<br />

poverty and biodiversity are intimately linked. The poor, especially in rural areas, depend on<br />

biodiversity for food, fuel, shelter, medicines and a wealth of other ecosystem services.<br />

Biodiversity loss therefore exacerbates poverty, while poverty in turn is a major threat to<br />

biodiversity, through unsustainable land use. The conservation and sustainable use of<br />

biodiversity also plays a critical role in the economic survival of a variety of production sectors<br />

such as fisheries, agriculture, and tourism. Consideration of biodiversity must therefore be<br />

integrated into the development process. This ‘mainstreaming’ of biodiversity into production<br />

sectors, poverty reduction plans and national sustainable development plans has been an<br />

internationally acknowledged goal for some time. For instance, goal 3.3 of the Strategic Plan of<br />

the Convention on Biodiversity (CBD), developed in 2002, requires that “biodiversity concerns<br />

are being integrated into relevant national sectoral and cross-sectoral plans, programmes and<br />

policies”. Yet progress on achieving this goal in the <strong>Andean</strong> region, as elsewhere, has been<br />

limited. It must now become a priority, both in terms of collating the scientific and socioeconomic<br />

evidence to support it, and the development of the policy initiatives to make it happen.<br />

Obstacles – Institutional Capacity<br />

Perhaps the single greatest limitation to robustly addressing the combination of the direct and<br />

indirect effects of climate change, together with the impacts of other global change (e.g.,<br />

deforestation, species invasion) on biodiversity in the <strong>Andean</strong> Region, is the lack of institutional<br />

capacity. This lack is principally a result of competing and overlapping responsibilities,<br />

compounded by lack of interconnections between and among institutions, authorities and other<br />

stakeholders in the region. Although skilled and technically adept scientists and policy makers<br />

exist across the region, they are in profoundly short supply. In order to improve capacity<br />

therefore, there is a clear need to: 1) Develop interdisciplinary higher education programs, to<br />

train qualified local researchers in the processing and analysis of ever increasing quantities of<br />

data (e.g., by promoting international programmes for students to study in leading institutions in<br />

36


Europe, Asia and North America). 2) Promote training opportunities for local leaders to facilitate<br />

their understanding and interpretation of complex analyses. 3) Involve the social, behavioral, and<br />

economic sciences in the necessary planning, implementation, and monitoring; advances are<br />

being made in the areas of environmental economics and common pool resource theory, used for<br />

understanding decisions relating to natural resource use and valuation. 4) Establish and promote<br />

social networks, for example, through the internet, to coordinate and facilitate information<br />

dissemination to decision makers and other stakeholders.<br />

At the scale of the tropical Andes, institutions like the Comunidad Andina (CAN) could<br />

play a leading role in promoting and facilitating integration and adaptive practices among<br />

signatory countries. Ideally, this would result in a regional planning process at least every five<br />

years, timed to coincide with the availability of new assessments resulting from the IPCC<br />

process. The planning process may also need to expand to include the Venezuelan Andes, northwest<br />

Argentina and north-east Chile, given the extension of tropical <strong>Andean</strong> ecosystems across<br />

these country’s borders. At the national level, there are existing planning initiatives in both the<br />

private and public sectors that could be expanded and replicated elsewhere. Not only should this<br />

involve the scientific community, but also social communicators. At the local level,<br />

municipalities and local communities will need access to results to allow them to assess and plan<br />

for locally-led initiatives. It will therefore be important to promote the provision of training to<br />

meet identified needs, so that local concerns are accounted for. In some regions, municipalities<br />

coordinate with protected area systems and should also be partners when setting climate change<br />

goals. Throughout the regional planning process, it will be crucial to facilitate public<br />

dissemination in order to ensure that the process is transparent and accountable.<br />

Conclusions<br />

Given the global importance of the tropical Andes for biodiversity, and the considerable risks<br />

posed by climate change, it is critical that both a regional and an international response be<br />

oriented to provide the necessary information and resources at the appropriate regional, national,<br />

and local scales, in order to inform robust adaptive management responses. Given current social<br />

inequalities which will likely be further exacerbated by climate change, the implementation of<br />

strategies that incorporate the use of economic, policy, and legal instruments for biodiversity<br />

conservation across the tropical <strong>Andean</strong> region will need to consider equity, fairness and<br />

distributional issues. How are these policies impacting stakeholders? Are they imposing burdens<br />

on the poorest sectors? These and other similar questions must be considered when designing<br />

and implementing integrated conservation and development policies. We highlight nine critical<br />

needs:<br />

• Convene a region-wide systematic conservation planning process, that explicitly<br />

incorporates the impacts of climate change, and that reconvenes every five years, in order<br />

to coincide with the availability of new knowledge from the IPCC and other assessments.<br />

• Continue to develop a comprehensive understanding of <strong>Andean</strong> climatology of the<br />

present and recent past to provide a baseline for detecting change and for assessing<br />

species and ecosystem capacities for resilience to climate stresses.<br />

• Implement standardized monitoring protocols to provide baseline evaluations of species<br />

distributions, population status, and ecosystem integrity, drawing from taxonomy, field<br />

37


ecology, and remote sensing. It is critical that data transparency and sharing is widely<br />

promoted.<br />

• Continue to develop and test the next generation of SDMs for projecting species and<br />

ecosystem (or proxy) spatial responses to climate change, since these provide the only<br />

way of assessing potential synergies and conflicts with people in an uncertain future.<br />

• Improve the understanding of the indirect impacts of climate change, resulting from<br />

planned and unplanned human adaptation and mitigation responses, on biodiversity and<br />

the provision of ecosystem services.<br />

• Demonstrate the direct and indirect benefits of ecosystem based adaptation as a key tool<br />

for making lives and livelihoods more resilient to climate change.<br />

• Build institutional capacity to design and implement robust adaptive management<br />

strategies, at regional, national and local scales, including all stakeholders.<br />

• Incorporate consideration of biodiversity into local, national and regional development<br />

planning, across all economic and societal sectors. Biodiversity and the ecosystem<br />

services it underpins must be front and center with economic and other considerations.<br />

Literature Cited<br />

Adger, W. N., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D. R. Nelson, L. O. Naess, J.<br />

Wolf, and A. Wreford. 2009. Are there social limits to adaptation to climate change?<br />

Climatic Change 93:335-354.<br />

Antonelli, A., J. A. A. Nylander, C. Persson, and I. Sanmartin. 2009. Tracing the impact of the<br />

<strong>Andean</strong> uplift on Neotropical plant evolution. Proceedings of the National Academy of<br />

Sciences of the United <strong>State</strong>s of America 106:9749-9754.<br />

Araujo, M. B., M. Cabeza, W. Thuiller, L. Hannah, and P. H.Williams. 2004. Would climate<br />

change drive species out of reserves? An assessment of existing reserve-selection<br />

methods. Global Change Biology 10:1618-1626.<br />

Bader, M. Y., M. Rietkerk, and A. K. Bregt. 2007. Vegetation structure and temperature regimes<br />

of tropical alpine treelines. Arctic, Antarctic, and Alpine Research 39:353-364.<br />

Bradley, R. S., M. Vuille, H. F. Diaz, and W. Vergara. 2006. Threats to water supplies in the<br />

tropical Andes. Science 302:1755-1756.<br />

Brito, D., and M. D. L. Figueiredo. 2003. Minimum viable population and conservation status of<br />

the Atlantic Forest spiny rat Trinomys eliasi. Biological Conservation 113:153-158.<br />

Bruner, A. G., R. E. Gullison, R. E. Rice, and G. A. B. da Fonseca. 2001. Effectiveness of parks<br />

in protecting tropical biodiversity. Science 291:125-128.<br />

Camacho, D. C. 2008. Esquemas de pagos por servicios ambientales para la conservacion de<br />

cuencas hidrograficas en el Ecuador. Investigacion Agraria: Sistemas y Recursos<br />

Forestales 17:54-66.<br />

Coblentz, D., and P. L. Keating. 2008. Topographic controls on the distribution of tree islands in<br />

the high Andes of south-western Ecuador. Journal of Biogeography 35:2026-2038.<br />

Denevan, W. F. 2001. Cultivated landscapes of native Amazonia and the Andes. Oxford: Oxford<br />

University Press.<br />

Ewers, R. M., J. P. W. Scharlemann, A. Balmford, and R. E. Green. 2009. Do increases in<br />

agricultural yield spare land for nature? Global Change Biology 15:1716-1726.<br />

38


Fargione, J., J. Hill, D. Tilman, S. Polasky, and P. Hawthorne. 2008. Land clearing and the<br />

biofuel carbon debt. Science 319:1235-1238.<br />

Fjeldså, J. 2002. Polylepis forests – vestiges of a vanishing ecosystem in the Andes. Ecotropica<br />

8:111-123.<br />

Fjeldså, J., and M. Kessler. 1996. Conserving the biological diversity of Polylepis woodlands of<br />

the highland of Peru and Bolivia: a contribution to sustainable natural resource<br />

management in the Andes. Copenhagen: NORDECO.<br />

Fjeldså, J., E. Lambin, and B. Mertens. 1999. Correlation between endemism and local<br />

ecoclimatic stability documented by comparing <strong>Andean</strong> bird distributions and remotely<br />

sensed land surface data. Ecography 22:63-78.<br />

Foden, W. B., G. M. Mace, J.-C. Vie, A. Angulo, S. H. M. Butchart, L. DeVantier, H. T. Dublin,<br />

A. Gutsche, S. Stuart, and E. Turak. 2009. Species susceptibility to climate change<br />

impacts. Pp. 77-88 in Wildlife in a changing world – an analysis of the 2008 IUCN Red<br />

List of threatened species, edited by J.-C. Vie, C. Hilton-Taylor, and S. N. Stuart. Gland:<br />

International Union for Conservation of Nature and Natural Resources. 180 pp.<br />

Francou, B., P. Ribstein, P. Wagnon, E. Ramirez, and B. Pouyaud. 2005. Glaciers of the tropical<br />

Andes, indicators of the global climate variability. Pp. 197-204 in Global change and<br />

mountain regions: a state of knowledge overview, edited by U. Huber, K. M. Harald, and<br />

M. A. Reasoner. Dordrecht: Springer.<br />

Gade, D. 1999. Nature and culture in the Andes. Madison: University of Wisconsin Press.<br />

Gascon, C., T. E. Lovejoy, R. O. Bierregaard, J. R. Malcolm, P. C. Stouffer, H. L. Vasconcelos,<br />

W. F. Laurance, B. Zimmerman, M. Tocher, and S. Borges. 1999. Matrix habitat and<br />

species richness in tropical forest remnants. Biological Conservation 91:223-229.<br />

Grau, H. R., and T. M. Aide. 2007. Are rural-urban migration and sustainable development<br />

compatible in mountain systems? <strong>Mountain</strong> Research and Development 27:119-123.<br />

Grau, H. R., T. M. Aide, J. K. Zimmerman, J. R. Thomlinson, E. Helmer, and X. Zou. 2003. The<br />

ecological consequences of socioeconomic and land use changes in post agricultural<br />

Puerto Rico. BioScience 53:1159-1168.<br />

Green, R. E., S. J. Cornell, J. P. W. Scharlemann, and A. Balmford. 2005. Farming and the fate<br />

of wild nature. Science 307:550-555.<br />

Guarnizo, C. E., A. Amezquita, and E. Bermingham. 2009. The relative roles of vicariance<br />

versus elevational gradients in the genetic differentiation of the high <strong>Andean</strong> tree frog,<br />

Dendropsophus labialis. Molecular Phylogenetics and Evolution 50:84-92.<br />

Halloy, S. R. P., A. Seimon, K. Yager, and A. Tupayachi Herrera. 2005. Multidimensional<br />

(climate, biodiversity, socio-economics, agriculture) context of changes in land use in the<br />

Vilcanota watershed, Peru. Pp. 323-337 in Land use changes and mountain biodiversity,<br />

edited by E. M.Spehn, M. Liberman Cruz, and C. Körner. Boca Raton FL: CRC Press<br />

LLC.<br />

Hannah, L., R. Dave, P. P. Lowry, S. Andelman, M. Andrianarisata, L. Andriamaro, A.<br />

Cameron, R. Hijmans, C. Kremen, J. MacKinnon, H. H. Randrianasolo, S.<br />

Andriambololonera, A. Razafimpahanana, H. Randriamahazo, J. Randrianarisoa, P.<br />

Razafinjatovo, C. Raxworthy, G. E. Schatz, M. Tadross, and L. Wilmee. 2008. Climate<br />

change adaptation for conservation in Madagascar. Biology Letters 4:590-594.<br />

Hannah, L., G. Midgley, S. Andelman, M. Araujo, G. Hughes, E. Martinez-Meyer, R. Pearson,<br />

and P. Williams. 2007. Protected area needs in a changing climate. Frontiers in Ecology<br />

and the Environment 5:131-138.<br />

39


Heller, N. E., and E. S. Zavaleta. 2009. Biodiversity management in the face of climate change: a<br />

review of 22 years of recommendations. Biological Conservation 142:14-32.<br />

Hoegh-Guldberg, O., L. Hughes, S. McIntyre, D. B. Lindenmayer, C. Parmesan, H. P.<br />

Possingham, and C. D. Thomas. 2008. Assisted colonization and rapid climate change.<br />

Science 321:345-346.<br />

Hoffmann, D. 2009. Municipal protected areas as policy instruments to enhance ecosystem<br />

resilience. Lessons from Latin American experiences. IOP Conference Series: Earth and<br />

Environmental Sciences 6:312019.<br />

Hole, D. G., B. Huntley, J. Arinaitwe, S. H. M. Butchart, Y. C. Collingham, L. D. Fishpool, D. J.<br />

Pain, and S. G. Willis. In press. Towards a management framework for key biodiversity<br />

networks in the face of climate change. Conservation Biology.<br />

Hole, D. G., S. G. Willis, D. J. Pain, L. D. Fishpool, S. H. M. Butchart, Y. C. Collingham, C.<br />

Rahbek, and B. Huntley. 2009. Projected impacts of climate change on a continent-wide<br />

protected area network. Ecology Letters 12:420-431.<br />

Holmgren, J. 2004. Prediction of tree height, basal area and stem volume in forest stands using<br />

airborne laser scanning. Scandinavian Journal of Forest Research 19:543-553.<br />

Holmgren, J., and A. Persson. 2004. Identifying species of individual trees using airborne laser<br />

scanner. Remote Sensing of Environment 90:415-423.<br />

IPCC. 2007. Climate change 2007: the physical science basis. Contribution of Working Group I<br />

to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change,<br />

edited by S. Soloman, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M.<br />

Tignor, and H. L. Miller. Cambridge: Cambridge University Press.<br />

Jiguet, F., R. Julliard, C. D. Thomas, O. Dehorter, S. E. Newson, and D. Couvet. 2006. Thermal<br />

range predicts bird population resilience to extreme high temperatures. Ecology Letters<br />

9:1321-1330.<br />

Keith, D. A., H. R. Akcakaya, W. Thuiller, G. F. Midgley, R. G. Pearson, S. J. Phillips, H. M.<br />

Regan, M. B. Araujo, and T. G. Rebelo. 2008. Predicting extinction risks under climate<br />

change: coupling stochastic population models with dynamic bioclimatic habitat models.<br />

Biology Letters 4:560-563.<br />

Kessler, M. 2002. The “Polylepis problem”: where do we stand? Ecotropica 8:97-110.<br />

Killeen, T. J., and L. A. Solorzano. 2008. Conservation strategies to mitigate impacts from<br />

climate change in Amazonia. Philosophical Transactions of the Royal Society of London<br />

B, Biological Sciences 363:1881-1888.<br />

Lobell, D. B., M. B., Burke, C. Tebaldi, M. D. Mastrandrea, W. P. Falcon, and R. L. Naylor.<br />

2008. Prioritizing climate change adaptation needs for food security in 2030. Science<br />

319:607-610.<br />

Margules, C. R., and R. L. Pressey. 2000. Systematic conservation planning. Nature 405:243-<br />

253.<br />

Martens, W. J. M. 1998. Health impacts of climate change and ozone depletion: an ecoepidemiological<br />

modelling approach. Environmental Health Perspectives 106(Suppl.<br />

1):241-251.<br />

MEA, 2005. Ecosystems and human well-being: synthesis. Washington, DC: Island Press.<br />

Mittermeier, R. A., P. Robles Gil, M. Hoffmann, J. Pilgrim, T. Brooks, C. G. Mittermeier, J.<br />

Lamoreux, and G. A. B. da Fonseca. 2004. Hotspots revisited. Mexico City: CEMEX.<br />

Naesset, E., and R. Nelson. 2007. Using airborne laser scanning to monitor tree migration in the<br />

boreal-alpine transition zone. Remote Sensing of Environment 110:357-369.<br />

40


Ohlemuller, R., E. S. Gritti, M. T. Sykes, and C. D. Thomas. 2006. Towards European climate<br />

risk surfaces: the extent and distribution of analogous and non-analogous climates 1931-<br />

2100. Global Ecology and Biogeography 15:395-405.<br />

Parmesan, C. 2006. Ecological and evolutionary responses to recent climate change. Annual<br />

Review of Ecology, Evolution, and Systematics 37:637-669.<br />

Patz, J. A., M. Hulme, C. Rosenzweig, T. D. Mitchell, R. A. Goldberg, A. K. Githeko, S. Lele,<br />

A. J. McMichael, and D. Le Sueur. 2002. Climate change - regional warming and malaria<br />

resurgence. Nature 420:627-628.<br />

Pearson, R. G., and T. P. Dawson. 2003. Predicting the impacts of climate change on the<br />

distribution of species: are bioclimate envelope models useful? Global Ecology and<br />

Biogeography 12:361-371.<br />

Postigo, J., K. R. Young, and K. A. Crews. 2008. Change and continuity in a pastoralist<br />

community in the high Peruvian Andes. Human Ecology 36:535-551.<br />

Pounds, J. A., M. R. Bustamante, L. A. Coloma, J. A. Consuegra, M. P. L. Fogden, P. N. Foster,<br />

E. La Marca, K. L. Masters, A. Merino-Viteri, R. Puschendorf, S. R. Ron, G. A. Sanchez-<br />

Azofeifa, C. J. Still, and B. E. Young. 2006. Widespread amphibian extinctions from<br />

epidemic disease driven by global warming. Nature 439:161-167.<br />

Poveda, G., W. Rojas, M. L. Quinones, I. D. Velez, R. I. Mantilla, D. Ruiz, J. S. Zuluaga, and G.<br />

L. Rua. 2001. Coupling between annual and ENSO timescales in the malaria-climate<br />

association in Colombia. Environmental Health Perspectives 109:489-493.<br />

Rees, W. G. 2007. Characterization of arctic treelines by LiDAR and multispectral imagery.<br />

Polar Record 227:345-352.<br />

Rodrigues, A. S. L., S. J. Andelman, M. I. Bakarr, L. Boitani, T. M. Brooks, R. M. Cowling, L.<br />

D. C. Fishpool, G. A. B. da Fonseca, K. J. Gaston, M. Hoffmann, J. S. Long, P. A.<br />

Marquet, J. D. Pilgrim, R. L. Pressey, J. Schipper, W. Sechrest, S. N. Stuart, L. G.<br />

Underhill, R. W. Waller, M. E. J. Watts, and X. Yan. 2004. Effectiveness of the global<br />

protected area network in representing species diversity. Nature 428:640-643.<br />

Ruiz, D., G. Poveda, I. D. Velez, M. L. Quinones, G. L. Rua, L. E. Velasquez, and J. S. Zuluaga.<br />

2006. Modelling entomological-climatic interactions of Plasmodium falciparum malaria<br />

transmission in two Colombian endemic-regions: contributions to a national malaria early<br />

warning system. Malaria Journal 5:66.<br />

Schroth, G., P. Laderach, J. Dempewolf, S. Philpott, J. Haggar, H. Eakin, T. Castillejos, J. G.<br />

Moreno, L. S. Pinto, R. Hernandez, A. Eitzinger, and J. Ramirez-Villegas. 2009. Towards<br />

a climate change adaptation strategy for coffee communities and ecosystems in the Sierra<br />

Madre de Chiapas, Mexico. Mitigation and Adaptation Strategies for Global Change<br />

1:605-625.<br />

Schulte, M. 1996. Tecnología agrícola altoandina. La Paz: Plural Editores – CID.<br />

Scott, D., J. R. Malcolm, and C. Lemieux. 2002. Climate change and modelled biome<br />

representation in Canada's national park system: implications for system planning and<br />

park mandates. Global Ecology and Biogeography 11:475-484.<br />

Seimon, A. 2001. Dilemmas of transport and energy development in Peru's Cordillera Carabaya.<br />

Pp. 26-27 in <strong>Mountain</strong>s of the world: mountains, energy, and transport, edited by M.<br />

Price, T. Kohler, T. Wachs, and A. Zimmermann. Berne: <strong>Mountain</strong> Agenda and the<br />

Swiss Agency for Development. 53 pp.<br />

Seimon, T. A., A. Seimon, P. Daszak, S. R. P. Halloy, L. M. Schloegel, C. A. Aguilar, P. Sowell,<br />

A. D. Hyatt, B. Konecky, and J. E. Simmons. 2007. Upward range extension of <strong>Andean</strong><br />

41


anurans and chytridiomycosis to extreme elevations in response to tropical deglaciation.<br />

Global Change Biology 13:288-299.<br />

Soule, M. E., and J. W. Terborgh (eds.). 1999. Continental conservation: scientific foundations<br />

of regional reserve networks. Washington, DC: Island Press.<br />

Sutherland, W.J. 2006. Predicting the ecological consequences of environmental change: a<br />

review of the methods. Journal of Applied Ecology 43:599-616.<br />

Terborgh, J. W. 1999. Requiem for nature. Washington, DC: Island Press.<br />

The Economics of Ecosystems and Biodiversity. 2008. The economics of ecosystems and<br />

biodiversity: an interim report. European Communities. 68 pp.<br />

Trenel, P., M. M. Hansen, S. Normand, and F. Borchsenius. 2008. Landscape genetics, historical<br />

isolation and cross-<strong>Andean</strong> gene flow in the wax palm, Ceroxylon echinulatum<br />

(Arecaceae). Molecular Ecology 17:3528-3540.<br />

Vargas, F. H., R. C. Lacy, P. J. Johnson, A. Steinfurth, R. J. M. Crawford, P. D. Boersma, and D.<br />

W. Macdonald. 2007. Modelling the effect of El Niño on the persistence of small<br />

populations: the Galapagos penguin as a case study. Biological Conservation 137:138-<br />

148.<br />

Vergara, W., A. Deeb, A. Valencia, R. Bradley, B. Francou, A. Zarzar, A. Grünwaldt, and S.<br />

Haeussling. 2007. Economic impacts of rapid glacier retreat in the Andes. EOS,<br />

Transactions, American Geophysical Union 88:261-263.<br />

Vuille, M., B. Francou, P. Wagnon, I. Juen, G. Kaser, B. G. Mark, and R. S. Bradley. 2008.<br />

Climate change and tropical <strong>Andean</strong> glaciers: past, present and future. Earth-Science<br />

Reviews 89:79-96.<br />

Warner, K., C. Ehrhart, A. de Sherbinin, S. Adamo, and T. Chai-Onn. 2009. In search of shelter<br />

– mapping the effects of climate change on human migration and displacement. Available<br />

at: http://www.careclimatechange.org/.<br />

Williams, J. W., and S. T. Jackson. 2007. Novel climates, no-analog communities, and<br />

ecological surprises. Frontiers in Ecology and the Environment 5:475-482.<br />

Williams, J. W., S. T. Jackson, and J. E. Kutzbach. 2007. Projected distributions of novel and<br />

disappearing climates by 2100 AD. Proceedings of the National Academy of Sciences of<br />

the United <strong>State</strong>s of America 104:5738-5742.<br />

Williams, P., L. Hannah, S. Andelman, G. Midgley, M. Araujo, G. Hughes, L. Manne, E.<br />

Martinez-Meyer, and R. Pearson. 2005. Planning for climate change: identifying<br />

minimum-dispersal corridors for the Cape Proteaceae. Conservation Biology 19:1063-<br />

1074.<br />

Wilme, L., S. M. Goodman, and J. U. Ganzhorn. 2006. Biogeographic evolution of Madagascar's<br />

microendemic biota. Science 312:1063-1065.<br />

Winterhalder, B. P., and R. B. Thomas. 1978. Geoecology of southern highland Peru: a human<br />

adaptation perspective. Occasional Paper 27. Boulder: Institute of Arctic and Alpine<br />

Research, University of Colorado.<br />

Young, K. R. 2007. Causality of current environmental change in tropical landscapes.<br />

<strong>Geography</strong> Compass 1:1299-1314.<br />

Young, K. R. 2008. Stasis and flux in long-inhabited locales: change in rural <strong>Andean</strong> landscapes.<br />

Pp. 11-32 in Land-change science in the tropics: changing agricultural landscapes,<br />

edited by A. Millington and W. Jepson. New York: Springer.<br />

42


Young, K. R., and B. Leon. 2007. Tree-line changes along the Andes: implications of spatial<br />

patterns and dynamics. Philosophical Transactions of the Royal Society of London B,<br />

Biological Sciences 362:263-272.<br />

Young, K. R., and J. K. Lipton. 2006. Adaptive governance and climate change in the tropical<br />

highlands of western South America. Climatic Change 78:63-102.<br />

Zhang, Y., M. Xu, J. Adams, and X. Wang. 2009. Can Landsat imagery detect tree line<br />

dynamics? International Journal of Remote Sensing 30:1327-1340.<br />

43


Appendix 2.1. Summary information on the current PA systems in three of the four tropical<br />

<strong>Andean</strong> countries<br />

1. What is the focus of the current PA system – i.e., is it selected on biodiversity targets or<br />

just the cheapest/unusable land?<br />

Bolivia<br />

In general, on biodiversity targets, but not necessarily in the best portions of the priority<br />

ecosystems in terms of conservation status and threat levels. Most recent PAs were created on<br />

the basis of socio-political opportunities (political will). Gap analysis at national level requires<br />

downscaling in order to better prioritize where future PAs need to be created.<br />

Peru<br />

At present what is being used are “marxan” type approximations, which are looking for sectors<br />

with maximum protection at minimum cost.<br />

Ecuador<br />

Apparently the design of PA systems is changing, not only looking for the opportunity to create<br />

spaces that might contain representative samples of biodiversity, but also to include concrete<br />

targets, such as a certain key or endangered species, certain natural monuments or specific<br />

ecosystem services. “Using the cheapest lands” has not been a common criteria for selection it<br />

seems.<br />

2. How well does the current PA system likely represent current biodiversity pattern and<br />

process?<br />

Bolivia<br />

At species level, about 70-80% covered by national PAs (without considering municipal and<br />

departmental level PAs or indigenous and community conservation areas or indigenous<br />

communal lands (TCOs). There is a bias on representation of lowlands (and <strong>Andean</strong> foothills).<br />

Peru<br />

It seems that there is a lack of representation of ecological processes in the western part.<br />

Ecuador<br />

There is no definition (base line) about what to understand as “biodiversity pattern and process”.<br />

Both GAP analysis in Ecuador (marine and terrestrial) consider maintaining representative<br />

samples of all ecosystems – vegetation formations of the country as main criteria (rather than<br />

processes).<br />

3. What are the likely key gaps under climate change?<br />

Bolivia<br />

Altitudinal ranges and intersections between ecosystems (e.g. bosque seco chiquitano –<br />

pantanal).<br />

44


Peru<br />

Difficult to say, as there are not many climate change scenarios available for analysis of<br />

representativness in mountainous conditions. It seems the problem is more serious on the western<br />

side, because of the lower rate of PA coverage and the high number of population, especially in<br />

the valleys. This population generates a high demand of hydrological resources, that are being<br />

extracted from the high parts of the (Eastern slope?) watersheds.<br />

Ecuador<br />

Lack of information (e.g., a base line of how biodiversity would be affected by climate change).<br />

Existing climate models are not detailed enough, especially for the <strong>Andean</strong> region, and are not an<br />

adequate basis to define strategies.<br />

What would be needed is a balance between practical mitigation actions and local adaptation<br />

measures, combining the following: biodiversity management, management of hydrological<br />

resources, risk management, agro-ecology, food security, poverty reduction strategies, conflict<br />

management, capacity development, territorial approach, among others.<br />

45


Appendix 2.2. Definitions of key terms<br />

Biological diversity: Variability among living organisms from all sources including, inter alia,<br />

terrestrial, marine and other aquatic ecosystems and the ecological complexes of which they are<br />

part; this includes diversity within species, between species and of ecosystems. (CBD,Art.2)<br />

Planning: Refers to a set of activities which leads to identifying i) issues, and goals; and to ii)<br />

formulating strategies and plans to be followed for the achievement of concrete goals.<br />

Conservation: Protection and management of natural resources. The CBD identifies 2<br />

biodiversity conservation options: i)"In-situ conservation" (in their natural habitats), ii)"Ex-situ<br />

conservation" (outside their natural habitats.) (CBD, Art. 2)<br />

Natural resources management: Set of activities and actions or inactions carried out to either:<br />

exploit, use, conserve or preserve natural resources. These set of actions or inactions can lead to<br />

either positive or negative impacts on biodiversity.<br />

Question: * whether conservation is a component of management/ or whether management is a<br />

component of conservation.<br />

Climate change: A change of climate which is attributed directly or indirectly to human activity<br />

that alters the composition of the global atmosphere and which is in addition to natural climate<br />

variability observed over comparable time periods. (UNFCCC, Art.1)<br />

46


Global Change Biology (2007) 13, 288–299, doi: 10.1111/j.1365-2486.2006.01278.x<br />

Upward range extension of <strong>Andean</strong> anurans and<br />

chytridiomycosis to extreme elevations in response<br />

to tropical deglaciation<br />

TRACIE A. SEIMON* 1 , ANTON SEIMONw 1 , PETER DASZAKz, STEPHAN R.P. HALLOY§,<br />

LISA M. SCHLOEGELz, CÉSAR A. AGUILAR} , PRESTON SOWELLk, ALEX D. HYATT**,<br />

BRONWEN KONECKYww and J O H N E . S I M M O N S zz<br />

*Columbia University, 630 W 168th Street, New York, NY 10032, USA, wThe Earth Institute, Columbia University, 61 Rt. 9W,<br />

Palisades, NY 10964, USA, zConsortium for Conservation Medicine, 460 West 34th Street, New York, NY 10001, USA, §Instituto<br />

de Ecología, Universidad Mayor de San Andrés, La Paz, Bolivia, }Departamento de Herpetología, Museo de Historia Natural<br />

Universidad Nacional Mayor de San Marcos, Av. Arenales 1256, Jesús María, Apdo 14-0434, Lima 14, Perú, kStratus Consulting,<br />

1881 9th Street Suite 201, Boulder, CO, USA, **CSIRO, Livestock Industries, Australian Animal Health Laboratory, 5<br />

Portarlington Road, Private Bag 24, Geelong, Vic. 3220, Australia, wwBarnard College, Columbia University, 3001 Broadway, New<br />

York, NY 10027, USA, zzNatural History Museum and Biodiversity Research Center, 1345 Jayhawk Blvd, University of Kansas,<br />

Lawrence, KS 66045, USA<br />

Introduction<br />

Abstract<br />

High-alpine life forms and ecosystems exist at the limits of habitable environments, and<br />

thus, are especially sensitive to environmental change. Here we report a recent increase<br />

in the elevational limit of anurans following glacial retreat in the tropical Peruvian<br />

Andes. Three species have colonized ponds in recently deglaciated terrain at new record<br />

elevations for amphibians worldwide (5244–5400 m). Two of these species were also<br />

found to be infected with Batrachochytrium dendrobatidis (Bd), an emerging fungal<br />

pathogen causally associated with global amphibian declines, including the disappearance<br />

of several Latin American species. The presence of this pathogen was associated<br />

with elevated mortality rates of at least one species. These results represent the first<br />

evidence of upward expansion of anurans to newly available habitat brought about by<br />

recent deglaciation. Furthermore, the large increase in the upper limit of known Bd<br />

infections, previously reported as 4112 m in Ecuador, to 5348 m in this study, also expands<br />

the spatial domain of potential Bd pathogenicity to encompass virtually all high<br />

elevation anuran habitats in the tropical Andes.<br />

Keywords: alpine biodiversity, amphibians, amphibian decline, chytridiomycosis, climate change,<br />

deglaciation, ecological succession, Pleurodema, Telmatobius, tropical andes<br />

Received 18 April 2006; revised version received 27 July 2006 and accepted 14 June 2006<br />

Global environmental changes such as climatic warming<br />

and emerging infectious diseases are impacting<br />

animal biodiversity in the Neotropics, and are having<br />

particularly adverse consequences for amphibian<br />

populations. Climate warming has been shown to influence<br />

species distribution and abundance, and cause<br />

upward migration of species in tropical mountains<br />

Correspondence: Tracie A. Seimon, tel. 1 1845 596 6883,<br />

e-mail: tad2105@columbia.edu<br />

1 Contributed equally to this work.<br />

(Parmesan, 1996; Pounds et al., 1999). Investigators in<br />

Costa Rica have demonstrated upward migration of a<br />

range of bird populations in response to an elevational<br />

increase in the adiabatically produced cloud base (and<br />

resulting cloud forest fauna), a phenomenon that is also<br />

associated with declines in anuran and lizard populations<br />

(Pounds et al., 1999). Infectious diseases have<br />

increasingly been reported to be associated with population<br />

declines of amphibian species (Daszak et al.,<br />

2000), and climate-induced ecological changes can alter<br />

host–parasite dynamics so that infectious diseases exert<br />

a greater impact on certain species (Pounds et al., 1999).<br />

Chytridiomycosis, an emerging fungal disease of<br />

r 2006 The Authors<br />

288 Journal compilation r 2006 Blackwell Publishing Ltd


amphibians caused by Batrachochytrium dendrobatidis<br />

(Bd), is causally linked to a loss of amphibian biodiversity<br />

in tropical montane amphibian species (Lips et al.,<br />

2006). Recently, Pounds et al. (2006) introduced the<br />

chytrid-thermal-optimum hypothesis and proposed<br />

that large-scale warming is likely a key factor in driving<br />

epidemic disease and the disappearance of anuran<br />

species in Central and South America. Their data show<br />

that extinctions in the Neotropics occur where climatic<br />

conditions may be more favorable for disease development,<br />

and that elevations below 1000 m, and above<br />

3500 m, would likely fall outside of the optimum thermal<br />

growth conditions to support disease outbreaks.<br />

As most amphibian surveys have been conducted at<br />

elevations below 4000 m, relatively little is known about<br />

survival of amphibians above this elevation, the impact<br />

that climatic warming is having upon these populations,<br />

or whether chytridiomycosis may also be affecting<br />

these communities. Climatic warming is also<br />

implicated in ongoing tropical deglaciation (i.e. the<br />

reduction in high-altitude glacial ice volume over recent<br />

decades in low latitude mountains), the impact of<br />

which has yet to be ascertained for anurans inhabiting<br />

high alpine zones.<br />

The principal goal of this study was to investigate if<br />

anurans in high <strong>Andean</strong> environments are being impacted<br />

by recent environmental changes. Our field<br />

study site is in the Cordillera Vilcanota, a heavily<br />

glacierized range in southern Peru within a region of<br />

the tropical Andes where significant changes in temperature<br />

and environmental responses have been recorded<br />

in recent decades. Documented changes include<br />

significant tropospheric warming, averaging 0.33 1C per<br />

decade (Vuille et al., 2003), a corresponding rise in<br />

freezing level (Diaz & Graham, 1996), and a rapid<br />

melting of glaciers since the Little Ice Age (LIA:<br />

ca. 1520–1880; Thompson, 1992; Francou et al., 2003).<br />

We conducted eight multidisciplinary field investigations<br />

in the Cordillera Vilcanota between 2000 and 2005,<br />

including surveys of anuran populations at six sites<br />

from 2003–2005, to study rapid ecological successions<br />

by high-<strong>Andean</strong> flora and fauna as ice retreats. Our<br />

study area offers an unusual opportunity to evaluate<br />

ecological observations in contexts of 75 years of landscape<br />

change. The upper reaches of the watershed are<br />

heavily glacierized, but over recent decades have experienced<br />

marked retreat of its ice margins concurrent<br />

with region wide climatic warming. An air photo series<br />

from 1931 and subsequent aerial and satellite imagery<br />

from 1962, 1980, and 2005 provide a temporally resolved<br />

portrayal of ice margins during this period of<br />

glacial retreat, and when combined with field surveys,<br />

allows for the monitoring of species migration into<br />

newly available habitat exposed by deglaciation.<br />

Materials and methods<br />

Geographic referencing<br />

Topographic map quads 28t-I-SE and 28t-II-NE<br />

[1 : 25 000 scale, 25 m elevation contour interval; Instituto<br />

Geográfico Militar (IGM), Lima, 1972] served as the<br />

reference base for this study. All GPS field measurements<br />

were made with a Garmin XL12 set to the<br />

Provisional South America 1956 datum for consistency<br />

with the topographic map series. Elevation observations<br />

specified in the text are from GPS and are consistent,<br />

though slightly more elevated, than interpolated<br />

values indicated on the topographic maps (mean difference<br />

124.7 m, standard deviation 7.95 m, N 5 18<br />

points); only points from within areas that were unglaciated<br />

in 1962, when data for topographic mapping<br />

was originally obtained, were compared to eliminate<br />

elevation differences related to changes in ice thickness.<br />

Ice recession analysis<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299<br />

FROG MIGRATION AND DEGLACIATION 289<br />

The temporal evolution of glacial recession was examined<br />

by pairing ground-based observations with aerial<br />

photography (available since 1931) and satellite<br />

imagery (since 1966). Only images exhibiting clearly<br />

defined glacial margins (unobscured by recent snowfall)<br />

were considered for analysis. Due to the variety of<br />

image types, view angles and resolutions, the isoline<br />

placements may have positional errors of up to 50 m.<br />

The exception is the 1962 isoline, which was precisionmapped<br />

by the IGM, and therefore, considered an<br />

accurate representation of glacial extent.<br />

LIA. The position of the LIA ice extent maximum was<br />

inferred by the visual discontinuity in substrate tone<br />

related to the abrupt change in abundance of lichens<br />

across this former glacial trimline. This ecotone is<br />

clearly evident in all imagery in the vicinity of site D,<br />

where its position was locally confirmed with multiple<br />

GPS ground measurements.<br />

1931. The 1931 isoline was determined by analysis of<br />

oblique aerial photographs taken during an overflight<br />

of the Cordillera Vilcanota by the Shippee–Johnson<br />

expedition on approximately 1 June 1931 (date inferred<br />

from Shippee, 1933). Examination of the series identified<br />

11 large-format (17 23 cm 2 ) glass plate negatives (nos.<br />

O86–O93 and O101–O103) displaying the pass region<br />

containing Sites D–F. These negatives were scanned by<br />

personnel at the American Museum of Natural History<br />

(AMNH) and then photogrammetrically analyzed to<br />

determine aircraft position and altitude for each<br />

image. Ice margin positions were then established


290 T. A. SEIMON et al.<br />

through photogrammetric referencing of image content<br />

to the topographic base map.<br />

1962. The 1962 glacial margin isoline is reproduced<br />

from the 1 : 25 000 topographic maps. We confirmed<br />

the absence of seasonal snow by examining vertical<br />

image stereopairs used by the IGM to generate the<br />

maps.<br />

1980. A declassified CORONA satellite black and white<br />

positive transparency (R0744 232 009) taken on August 3,<br />

1980 was used to derive the 1980 isoline. The Sibinacocha<br />

watershed is located in the far upper left corner of the<br />

scene, and thus contains some geometric distortion of<br />

landscape features owing to the obliquity of the view<br />

angle. The image was cropped, scanned into Photoshop s<br />

and rescaled to the base map projection. The brightness<br />

gradient of the ice margin was then used as a first-guess<br />

for the isoline location. Manual corrections were then<br />

applied where geometric distortion required local<br />

realignment of image content with geo-referenced<br />

landscape features.<br />

2005. Numerous aerial images, including repeated<br />

scenes from the 1931 Shippee–Johnson series, were<br />

taken during an overflight of the Cordillera Vilcanota<br />

on March 15, 2005. This image series was utilized to<br />

locate the 2005 ice margin position; ground-truth GPS<br />

measurements made during fieldwork on March 20,<br />

2005 were used to refine ice margin positions near<br />

sampling sites in the pass region.<br />

Field surveys<br />

Multiday field surveys of amphibian populations in the<br />

Cordillera Vilcanota were conducted during two dry<br />

seasons (August 6–17, 2003; July 13–25, 2004) and two<br />

wet seasons (March 3–7, 2003; March 14–21, 2005).<br />

Visual encounter surveys by two to four researchers<br />

identified habitats supporting anuran populations.<br />

Researchers surveyed each site for 1–5 h, overturning<br />

rocks and sifting ponds with nets to locate hidden<br />

specimens, and collecting environmental data [water<br />

temperature, pH, vegetation types (aquatic and terrestrial),<br />

GPS coordinates and elevation, and photographs].<br />

Soil temperatures were measured by Hobo<br />

data loggers (Onset Computer Corporation, Bourne,<br />

MA, USA). Water temperature was measured using a<br />

digital thermometer (Oregon Scientific, Tualatin, OR,<br />

USA). As a proxy for population density we recorded<br />

the number of specimens observed and express these<br />

as the number of individuals/the number of surveyors<br />

the number of hours of survey (survey hours). Captured<br />

specimens were photographed and measured for<br />

snout-vent length (SVL), then released. Voucher specimens<br />

were collected in July 2002, August 2003, and<br />

March 2005 for species identification and Bd analysis.<br />

All voucher specimens are housed at the Museo de<br />

Historia Natural de la Universidad Nacional San Antonio<br />

Abad de Cusco.<br />

Botanical diversity was assessed to characterize<br />

habitat occupied by amphibians and other fauna. Plant<br />

species and richness are based on observations during<br />

two expeditions, July 2002 and August 2003, in which<br />

plants, reptiles, amphibians, and some invertebrates<br />

were collected. The main sampling area is 300 m from<br />

site D at a monitoring site for the Global Research<br />

Initiative in Alpine Areas (GLORIA), but also includes<br />

general area sampling outside the LIA boundary.<br />

Although these months are mid-winter (dry season)<br />

there was considerable snow cover most of the time in<br />

2002 and over parts of the area (particularly south<br />

slopes) in 2003. Many plants are dormant at this time<br />

of the year but some were flowering. Five days in 2002<br />

and 8 days in 2003 were spent in the Sibinacocha<br />

watershed above 4880 m. In 2004, two days were spent<br />

sampling the GLORIA site at 5250 m, while 3 days were<br />

spent there in 2003 (GLORIA methodology is described<br />

in Pauli et al., 2004). Further sampling in 2005 has not<br />

been analyzed for this account. The detailed description<br />

of this GLORIA site will be made available on the<br />

GLORIA website (http://www.gloria.ac.at/res/gloria_<br />

home/) once it is completed.<br />

Plants were identified to the highest level possible<br />

in the field. Vouchers collected were deposited at the<br />

Herbario Vargas of the Universidad Nacional San<br />

Antonio Abad del Cusco.<br />

Tissue preservation, histology, and polymerase chain<br />

reaction (PCR) analysis<br />

Live frogs were collected from the above referenced<br />

sites A–D, quickly euthanized and preserved in a 70%<br />

ethanol solution. Larger dead frogs were subsequently<br />

injected with 5 cm 3 of 70% ethanol. Ventral hind limb<br />

skin was excised, stored in 70% ethanol, and transferred<br />

to neutral-buffered 10% formalin before embedding in<br />

paraffin. Tissues were then sectioned at 6 mm thickness<br />

for histological staining with hematoxylin and eosin (H<br />

and E), and imaged using an Olympus IX-70 inverted<br />

light microscope (New Jersey Scientific, Middlebush,<br />

NJ, USA) equipped with an Olympus LCPlanF1<br />

20 objective, Imaging software (Roper Scientific, Tucson,<br />

AZ, USA), and a Cool Snap CCD camera (RS<br />

Photometrics, Tucson, AZ, USA). Ten fields of view<br />

( 20) were examined on two slides from different<br />

planes of section for each individual. Presence of a<br />

single Bd sporangium denoted a positive individual.<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


(a)<br />

(b)<br />

(c)<br />

Site D<br />

LIA margin<br />

Site E<br />

deglaciated corridor<br />

Site F<br />

1931<br />

2005<br />

2005<br />

Infectivity was scored by averaging the highest observed<br />

number of cell layers of Bd per field, for 10<br />

fields. All specimens in 2005 were analyzed for chytridiomycosis<br />

using dry skin swabs (MW100–100, Medical<br />

Wire and Equipment Co. (UK), Wiltshire, UK) of collected<br />

frogs in the field. Collected frogs were swabbed<br />

on the dorsal and ventral surface 20 times, and five<br />

times between the webbing of the limbs, and then<br />

released. The swab was then placed in a sterile plastic<br />

vial and stored in a backpack for the duration of the<br />

trip. Swabs were then stored at 20 1C until analysis.<br />

The swabs were analyzed in triplicate by Taqman realtime<br />

PCR assay (Boyle et al., 2004), CSIRO Livestock<br />

Industries, Australian Animal Health Laboratory,<br />

Australia. All specimens were collected under<br />

the authority of Universidad Nacional San Antonio<br />

Abad de Cusco.<br />

N<br />

(d)<br />

Results<br />

0273 0274 0275 0276<br />

Fig. 1 Glacial recession since the Little Ice Age. (a) Aerial perspective from the north at 6600 m from June 1931 showing a 5400 m pass at<br />

upper-center of image completely icebound (red arrow). (b) Repeat image of the same scene on March 15, 2005, revealing significant<br />

glacial recession and development of an ice-free corridor (red arrow) across the pass. (c) Aerial perspective on March 15, 2005 looking SW<br />

from 6600 m across the recently deglaciated corridor atop the pass. Annotations identify LIA margin and anuran sampling sites within<br />

deglaciated terrain; arrow on lower right points N. (d) Topographic map section (100 m contour interval) of the pass region showing<br />

glacial margin positions at the ca. 1850–1900 LIA maximum (blue), 1931 (cyan), 1962 (green), 1980 (gold) and 2005 (red). Glacial coverage<br />

in 2005 is shaded light blue. 1-km UTM grid is overlaid for reference [Photo credit for (a): Negative No. O-103 Photo: Shippee–Johnson<br />

Collection. Department of Library Services, American Museum Natural History (AMNH)].<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299<br />

FROG MIGRATION AND DEGLACIATION 291<br />

8480<br />

8479<br />

8478<br />

8477<br />

8476<br />

8475<br />

Landscape and biotic changes are both strongly evident<br />

in the watershed of Laguna Sibinacocha (13.851S,<br />

71.051W; 4900 m m.s.l., 33 km 2 ), the largest high-alpine<br />

lake in the Andes, and a principal source to the<br />

Urubamba tributary of the Amazon River. A 5400 m<br />

pass 5 km NNW of Lake Sibinacocha was thickly icebound<br />

when first photographed in 1931, but by 2005,<br />

deglaciation had exposed much of the rock substrate<br />

leading to the pass summit (Fig. 1a and b). The glacier<br />

change analysis reveals an ice-free corridor across the<br />

pass became established around 1980, and by 2005 had<br />

expanded to 238 m wide (measured in the field by<br />

Global Positioning System, GPS), an average rate of<br />

9.52 m yr 1 (Fig. 1c and d). Thus, an ecological corridor<br />

exists today between watershed habitats previously


292 T. A. SEIMON et al.<br />

N<br />

Ausangate, 6372 m<br />

Chillca<br />

0260<br />

A1<br />

Rio Chillcamayo<br />

A2<br />

Rio Jampamayo<br />

km<br />

0 4<br />

0264<br />

Uyuni, 4422 m<br />

5960 m<br />

0268<br />

Quebrada Killita<br />

6093 m<br />

segregated by an ice barrier for centuries, perhaps<br />

millennia.<br />

We studied anuran populations at six locations (sites<br />

A–F) above 4400 m within and close to the Sibinacocha<br />

watershed, including three (sites D–F) within the recently<br />

deglaciated terrain above 5244 m north of Lake<br />

Sibinacocha (Fig. 1c and d, and Fig. 2). Three species of<br />

three genera were identified: Pleurodema marmorata, Bufo<br />

spinulosus, and Telmatobius marmoratus. Site D, a cluster<br />

of tarns at 5244 m within the LIA moraine harbored all<br />

three species, but in adjacent water bodies there was a<br />

considerable variation in relative species abundance.<br />

Each species had 100% representation in one of the five<br />

ponds surveyed, while the other two ponds showed<br />

cohabitation by P. marmorata and T. marmoratus (Fig. 3a<br />

and b). The LIA moraine abruptly delimits sparsely<br />

vegetated terrain on its uphill side from more biodiverse<br />

alpine habitat just meters away. The area containing the<br />

ponds exhibits scattered lichen cover on rocks (o5%)<br />

B<br />

Condor Pass, 5027m<br />

Rio Chuamayu<br />

F<br />

GLORIA<br />

5812 m<br />

5348 m<br />

5360–5400 m<br />

E<br />

Quebrada Puca Orjo<br />

D<br />

5244 m<br />

C<br />

4900 m<br />

Laguna Sibinacocha<br />

Fig. 2 Location of anuran sampling sites A–F in the Cordillera Vilcanota of southern Peru. Glaciers are shown in light grey with<br />

mountain summits marked by triangles. Sites A–F are circled. The Global Research Initiative in Alpine Areas (GLORIA) site and villages<br />

of Uyuni and Chillca are shown. The dashed box indicates the domain of Fig. 1d.<br />

0272<br />

0276<br />

0280<br />

8480<br />

8476<br />

8472<br />

8468<br />

8464<br />

and patchy vascular plant growth ( 12 species). Outside<br />

the LIA boundary, over 39 species of lichens and bryophytes<br />

cover 75–90% of available rock surfaces, and over<br />

54 species of vascular plants have been recorded (see<br />

‘Materials and methods’) (Halloy et al., 2005). The ponds<br />

at site D are evident below the retreating ice margin in<br />

1931 aerial photographs, so presumably formed relatively<br />

soon after the onset of ice recession, estimated<br />

between 1850 and 1900 (Fig. 3c). The ponds are perennial,<br />

being sustained by snowfall and glacial melt<br />

throughout the year, and appear to exist today in a<br />

near-pristine natural state. Ponds and their margins<br />

have been colonized by Distichia muscoides (cushion<br />

bogs) and aquatic Myriophyllum and Potamogeton.<br />

Higher still than site D are much more recently<br />

developed ponds at sites E and F (Fig. 4a). The ice<br />

recession analysis shown in Fig. 1d indicate that deglaciation<br />

of sites E1–5 (5369–5376 m) occurred between<br />

1931 and 1962, and sites E6 and E7 (5381–5400 m) within<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


N<br />

(a) (b)<br />

(c)<br />

Pond 6<br />

n = 45<br />

Pond 5<br />

n = 16<br />

Pond 3<br />

n = 45<br />

Pond 2<br />

n = 20<br />

Pond 1<br />

n = 50<br />

the last 10 years. Site F (5348 m) developed within the<br />

past 35–40 years (Fig. 4b and c). In March 2005, Pleurodema<br />

tadpoles and adults were found in both seasonal<br />

and permanent ponds surveyed within the pass corridor<br />

at sites E and F (Fig. 4d–f). Diurnal warming had<br />

melted nocturnal ice cover by the time of observation<br />

(1200–1300 local time). Tadpoles were observed in<br />

dense clusters of up to 950 50 individuals, mostly in<br />

1<br />

3<br />

2<br />

5<br />

6<br />

2005 Survey at Site D<br />

0% 20% 40% 60% 80% 100%<br />

Bufo Telmatobius Pleurodema<br />

Fig. 3 Three species of anurans inhabiting permanent ponds in deglaciated terrain (5244 m). (a) Aerial view on 15 March 2005 of high<br />

ponds at site D identified by number. Dotted line represents 42.2 m. (b) Relative abundances of tadpole species within individual ponds<br />

at site D during the wet season of 2005. Percentages are derived from the number of tadpoles observed/survey hour (n 5 individuals<br />

counted). (c) Cropped section of 1931 air photo showing ponds at site D (5244 m, solid arrow) below receding ice margin. Margin of the<br />

LIA maximum is indicated (dotted arrow); view is looking N. [Negative No. O-86 Photo: Shippee–Johnson Collection. Department of<br />

Library Services, AMNH].<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299<br />

FROG MIGRATION AND DEGLACIATION 293<br />

1931<br />

water of 1–2 cm depth, where temperatures ranged<br />

from 18.2 1C in E4, to 22.2 1C in E2 (Fig. 4e and f). These<br />

findings of recently established anuran populations are<br />

consistent with our surveys on the vigorous colonization<br />

by plants such as Senecio graveolens, Calamagrostis<br />

spp., Ephedra rupestris, and Valeriana cf. nivalis, recorded<br />

at elevations of 5400–5513 m within areas of recent<br />

glacial recession (data not shown).


294 T. A. SEIMON et al.<br />

(a) (d)<br />

(b)<br />

(c)<br />

E 1-5, 5369 m<br />

1970<br />

2003<br />

E 6, 5381 m<br />

E 7, 5400 m<br />

F , 5348 m<br />

Repeated population surveys at Sites A2 and D1<br />

revealed a significant reduction in T. marmoratus tadpoles<br />

from August 2003 to July 2004 at both sites (Fig.<br />

5a). Adult Telmatobius were present at A2 in July 2002,<br />

but were not found on subsequent visits (not shown).<br />

At site D1, a 90% decrease in the number of live postmetamorphic<br />

Telmatobius was observed from August<br />

2003 to July 2004 (Fig. 5b). No tadpoles were recorded<br />

(e)<br />

(f)<br />

Fig. 4 Sites within recently deglaciated terrain above 5300 m with the highest anuran communities. (a) Near-vertical aerial perspective<br />

of high pass in March 2005. The locations of the seasonal ponds (sites E1–7) and perennial pond (site F) are circled. Minimum width of<br />

the ice-free corridor is 238 m. (b) Pond at site F (circled) in 1970 emerging from disintegrating ice field [photo by J. Ricker]. (c) Repeat of<br />

1970 image from August 14, 2003 showing site F (circled), marked glacial recession, and upward advance of vegetation. (d) Seasonal<br />

pond (site E7) on the pass summit at 5400 m, the highest site where Pleurodema tadpoles were found (inset) Photos: March 20, 2005.<br />

Receding glacier in the background is 45 m from the pond; Senecio graveolens is also pictured (lower right). (e) Pleurodema tadpole cluster<br />

(black area, 950 50 individuals) observed in a 1m 2 area of site E1 (5369 m). (f) Enlargement of area indicated by small rectangle in (e)<br />

showing tadpoles clustering in 1–2 cm of water.<br />

in 2004 when 24 dead adult frogs were found within<br />

4 survey hours (6 h 1 ), indicating a recent die-off, and<br />

consistent with the observed absence of breeding.<br />

Reports from local highland inhabitants suggested<br />

that anuran populations had been disappearing rapidly.<br />

As Bd has been linked to the rapid decline of amphibian<br />

populations worldwide and catastrophic loss of amphibian<br />

biodiversity (Daszak et al., 1999, 2003; La Marca<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


(a)<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Tadpoles / survey hour<br />

(c)<br />

Telmatobius<br />

Pleurodema<br />

A1/A2<br />

2002: 3/4<br />

2003: 2/7<br />

B<br />

_<br />

2003: 1/4<br />

Site A2<br />

Site D1<br />

August -2003 July-2004 August-2003 July-2004<br />

C<br />

2003: 1/2<br />

Year:# individuals positive / # individuals tested<br />

(d) (e)<br />

(f)<br />

s<br />

s<br />

Sp<br />

E<br />

z<br />

FROG MIGRATION AND DEGLACIATION 295<br />

_<br />

(b)<br />

Individuals / survey hour<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

D<br />

2003: 1/11<br />

(g)<br />

Live postmetamorphosis<br />

Dead postmetamorphosis<br />

Live tadpoles<br />

_<br />

E<br />

_<br />

_<br />

F<br />

_<br />

2005: 1/2<br />

Fig. 5 Pathogenic Batrachochytrium dendrobatidis (Bd) afflicting high-<strong>Andean</strong> Telmatobius and Pleurodema communities. (a) Comparison of<br />

Telmatobius tadpole populations recorded during the dry seasons of 2003 and 2004 at sites A2 and D1. (b) Numbers of postmetamorphic live<br />

Telmatobius, post-metamorphic dead Telmatobius, andTelmatobius tadpoles are compared between dry seasons 2003 vs. 2004 at site D1. (c)<br />

Summary of the Bd results for sites A–F of Pleurodema and Telmatobius. Specimens that tested positive from sites A–F are scored for infectivity<br />

in Table 1. (d) Histological hematoxylin and eosin stained section from the ventral hindlimb skin of Telmatobius marmoratus (2003 site D1). The<br />

majority of the sporangia (s) within the stratum corneum are empty. Significant hyperkeratosis (three to five cell layers) is evident in the<br />

epithelial layer, E. Scale bar represents 25 mm.(e)PhotographofhealthyadultfrogofthespeciesT. marmoratus collected at site D1 in 2003. (f)<br />

Histological hematoxylin and eosin stained section from the ventral hindlimb skin of Pleurodema marmorata (2003 site B). Lesion displays light<br />

hyperkeratosis (1–2 cell layers) and is focal. Zoospores (z) are evident in the majority of sporangia identifiable as small dark basophilic<br />

masses, some divided by a septum (sp). Scale bar represents 25 mm. (g) Adult P. marmorata that tested positive for Bd shown in (f).<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


296 T. A. SEIMON et al.<br />

Table 1 Results for Batrachochytrium dendrobatidis positive specimens collected at sites A–F<br />

Year and location Species Infectivity Signs of illness<br />

2003 Site A1 Pleurodema marmorata 2.5 Healthy and active<br />

2003 Site A1 P. marmorata 0.6 Healthy and active<br />

2002 Site A2 Telmatobius marmoratus 2.4 Lethargic<br />

2002 Site A2 T. marmoratus Not determined Dead<br />

2002 Site A2 T. marmoratus Not determined Active<br />

2003 Site B P. marmorata 0.2 Healthy and active<br />

2003 Site C T. marmoratus 0.8 Irregular spotting<br />

2003 Site D T. marmoratus 1.3 Signs of Red-leg/blood in hindlimbs under skin<br />

2005 Site F P. marmorata Not determined Healthy and active<br />

Increased severity of infection between individuals is indicated by the higher infectivity score (see ‘Materials and methods’).<br />

Samples labelled ‘not determined’ could not be scored due to poor tissue preservation, insufficient material for microscopic<br />

examination, or samples were analyzed by polymerase chain reaction analysis.<br />

et al., 2005; Lips et al., 2005, 2006), we tested for the<br />

presence of this fungus at our six study sites. We<br />

analyzed skin samples from specimens collected in<br />

2002 (Seimon et al., 2005) and 2003 at sites A–D, and<br />

skin swabs from uncollected specimens at sites E and F<br />

in July 2005; results are summarized in Fig. 5c. Dead<br />

specimens found in July of 2004 were not tested for Bd<br />

as these specimens had been decomposing for at least<br />

several days, and perhaps weeks, before collection. Two<br />

species tested positive for Bd: P. marmorata and<br />

T. marmoratus (Fig. 5d–g). We did not find any evidence<br />

of Bd in B. spinulosus (N 5 2). The fungus was found to<br />

be present at five sites, including sites D and F within<br />

recently deglaciated terrain. No anurans tested positive<br />

at the highest site, site E, in 2005. All Pleurodema specimens<br />

collected were energetic and did not exhibit signs<br />

of sloughing skin, a clinical sign of chytridiomycosis. In<br />

contrast, Telmatobius specimens collected exhibited<br />

characteristics consistent with clinical and pathological<br />

signs of fatal chytridiomycosis in captive and wild<br />

caught amphibians (Pessier et al., 1999; Nichols et al.,<br />

2001), including: lethargy upon handling; had more<br />

marked hyperplasia of the keratinaceous layer (Fig. 5d<br />

and f) and more intense infection (i.e. greater numbers<br />

of fungal sporangia scored for infectivity). A single<br />

specimen was found dead. These results are summarized<br />

in Table 1.<br />

Discussion<br />

Rapid deglaciation over recent decades associated with<br />

strong regional warming has exposed new terrain that<br />

is now suitable for anurans, whereas before the 20th<br />

century the ice advances of the LIA had depressed the<br />

elevational limit of potential habitat. Our findings on<br />

anuran distributions are consistent with a broader<br />

pattern of ecological succession in the <strong>Andean</strong> region<br />

and worldwide, where species’ ranges have expanded<br />

vertically in response to warmer temperatures in highalpine<br />

zones (Parmesan & Yohe, 2003; Halloy et al.,<br />

2005). At lower elevations in Ecuador, recent observations<br />

of anuran communities have also identified new<br />

records in elevational range of Eleutherodactylus and<br />

Hyla species (Bustamante et al., 2005). Our observations<br />

provide the first direct evidence that anurans are now<br />

colonizing terrain exposed by recent deglaciation at the<br />

upper limits of the biome. Furthermore, glacial recession<br />

in the Cordillera Vilcanota has opened a direct<br />

ecological corridor within the past 25 years between<br />

watershed habitats previously segregated by ice. The<br />

opening of this corridor may have long-term ecological<br />

consequences, such as distributional overlap of species,<br />

which may have genetically diverged, or facilitation of<br />

the spread of chytridiomycosis through host-species<br />

colonization, and is the subject of our future studies.<br />

Our observations represent a new elevation record for<br />

amphibian species, exceeding the unsubstantiated account<br />

of 5200 m purported for Scutiger alticola in Tibet<br />

(Swan, 1990). In Latin America, records of up to 4800 m<br />

have been found recently in Atelopus species (La Marca<br />

et al., 2005); up to 5000 m for Pleurodema, up to 4500 m<br />

for Bufo, and up to 5000 m, for Telmatobius (Pefaur &<br />

Duellman, 1980; Diaz-Páez & Ortiz, 2003).<br />

Our observations indicate that Pleurodema, Bufo, and<br />

Telmatobius have colonized ponds at 5244 m within the<br />

deglaciated area, while Pleurodema has colonized newly<br />

formed ponds at even higher elevations, to 5400 m<br />

within the last 10 years. To survive at such elevations<br />

these species must possess adaptations to extreme environmental<br />

conditions, which include the hypoxic lowpressure<br />

of the mid-troposphere ( 52% of mean sea<br />

level pressure at 5400 m), intense solar radiation<br />

(1000 W m 2 ; Hastenrath, 1978), low absolute humidity,<br />

diurnal freeze–thaw cycles, and occasional deep snow<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


cover that can persist for several weeks. Although direct<br />

measurements for atmospheric temperatures are not<br />

available from our field sites, during 2002 soil surface<br />

temperatures recorded at 5200 m at a site 160 m south of<br />

site D showed minima as low as 13.5 1C, and diurnal<br />

variability of up to 30 1C (data not shown). These results<br />

suggest that these anurans have capabilities for surviving<br />

frost. In line with these observations, B. spinulosus<br />

has been shown to survive body temperatures as low as<br />

5 1C with no noticeable impact on vitality (Halloy &<br />

Gonzalez, 1993). This adaptation has also been documented<br />

in lizards such as Liolaemus huacahuasicus in<br />

Argentina (Halloy & Gonzalez, 1993), of a genus we<br />

have observed to 5450 m in the Sibinacocha area. The<br />

low-pressure environment at high-altitudes appears to<br />

enhance the freezing tolerance of these animals (Halloy<br />

& Gonzalez, 1993).<br />

The upward range extension of amphibians parallels<br />

the upward shifts of other organisms. We have documented<br />

upward expansion of plants, which serve as<br />

anuran habitats and food sources, and insects, which act<br />

as pollinators for plants and food for amphibians.<br />

Glacial recession and snow melt also provide water<br />

sources for the development of year-round and seasonal<br />

ponds, terrestrial and aquatic vegetative growth, aquatic<br />

microorganisms, amphipods, diptera, butterflies,<br />

bivalves, and insect larvae, and a drinking source for<br />

mammals such as vicuña (Vicugna vicugna), viscacha<br />

(Lagidium spp.), fox (Dusicyon spp.), puma (Felis<br />

concolor), and other cats (Oncifelis spp. or Oreailurus<br />

spp.).<br />

We show chytridiomycosis in anuran communities at<br />

much higher elevations in the Neotropics than previously<br />

recognized. The existence of severe Bd infections<br />

in amphibians at site D1 in August 2003 correlates<br />

with the subsequent die-off observed there in July 2004,<br />

and the apparent disappearance of both post-metamorphic<br />

and pre-metamorphic populations of T. marmoratus<br />

by March 2005. As our quantitative surveys<br />

span less than 2 years we cannot rule out the possibility<br />

that the disappearance of Telmatobius at sites D1 and A2<br />

are part of an interannual fluctuation within these<br />

populations. However, given the high degree of infectivity<br />

observed in histological sections, the abundance<br />

of dead and sick specimens collected within the last 2<br />

years, of which several were Bd positive, and that<br />

Telmatobius species are disappearing in Ecuador (Merino-Viteri<br />

et al., 2005) and Bolivia (Ergueta & Morales,<br />

1996), we believe it likely that T. marmoratus is in decline<br />

due to chytridiomycosis. In contrast to Telmatobius,<br />

Pleurodema populations appear to be relatively robust<br />

at the highest elevations, and the Bd-positive specimens<br />

examined did not appear sick. Furthermore, Pleurodema<br />

also tested positive on the more isolated north side of<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299<br />

FROG MIGRATION AND DEGLACIATION 297<br />

the pass at site F, which deglaciated 35–40 years ago,<br />

raising important questions on how Bd arrived at this<br />

remote location that should be addressed in future<br />

research. The Sibinacocha watershed is home to several<br />

species of waterfowl, mammals (described above), and<br />

domesticated animals such as alpaca and llama that<br />

travel through the area, and thus could potentially<br />

facilitate the spread of Bd. Further studies will also be<br />

needed to ascertain whether Pleurodema exhibits an<br />

innate capacity to survive chytridiomycosis, and is<br />

therefore a potential vector for this disease.<br />

The discovery of chytridiomycosis in a high alpine<br />

setting in the Andes is at odds with elevational (much<br />

higher) and climatic (much colder) characteristics of<br />

environments where chytridiomycosis have been reported<br />

elsewhere in the Neotropics. Niche modeling<br />

analysis based on observations from 44 localities predicts<br />

the most likely occurrence of Bd in the Andes is<br />

above 1000 m, with a median elevation for pathogenic<br />

Bd found at 1714 m and maximum at 4112 m (Ron,<br />

2005). Our observations demonstrate that Bd is occupying<br />

an area of the biome that was previously unstudied.<br />

By confirming the presence of chytridiomycosis 1236 m<br />

above the previously reported maximum, our findings<br />

also significantly increase the spatial domain of potential<br />

pathogenicity because much of the central <strong>Andean</strong><br />

region, which includes the vast Altiplano plateau, lies<br />

within this elevational range. Using linear regression<br />

analysis, Pounds et al. (2006) predicted that above<br />

3500 m in the Neotropics, daily minimum and maximum<br />

temperatures would fall well below the optimal<br />

temperature range of 17–25 1C for Bd growth, and<br />

become increasingly inhospitable at higher elevations.<br />

However, our results clearly indicate that neither temperatures<br />

far below freezing nor large diurnal variations<br />

are limiting factors for Bd pathogenicity. Although<br />

atmospheric temperature means in tropical alpine<br />

zones fall well below the window of optimal Bd growth,<br />

our observations suggest that intense solar heating<br />

during midday hours temporarily raises water temperatures<br />

in shallow ponds to within this range, even<br />

at extreme elevations around glacial margins above<br />

5300 m. Thus, solar heating may provide a thermal<br />

opportunity to allow Bd to survive and grow in these<br />

environments: this conjecture could be investigated in a<br />

controlled laboratory setting. The infectivity of the Bd<br />

within these extreme environments suggests that frogs<br />

responding to climatic warming are unlikely to escape<br />

this pathogen by moving to higher, colder, and more<br />

pristine areas. Thus, it is likely that Bd may drive<br />

changes in species assemblages in the high Andes by<br />

promoting susceptible species loss – as already seen<br />

with the disappearance of three species of Telmatobius in<br />

Ecuador (Merino-Viteri et al., 2005) – or by promoting


298 T. A. SEIMON et al.<br />

the expansion of resistant species. There are 55 known<br />

species of Telmatobius (De la Riva et al., 2005), yet given<br />

the efficiency of pathogenic Bd to propagate, our results<br />

suggest an ominous future for the sustainability of this<br />

genus. Furthermore, this study offers an example of<br />

how the expansion of biota into previously unsuitable<br />

habitat can bring with it the introduction of new diseases.<br />

In this case it is chytridiomycosis, however, with<br />

other fauna it may encompass other pathogens of<br />

unknown host ranges and impacts.<br />

Acknowledgements<br />

We gratefully acknowledge the contributions of our teams of<br />

local arrieros, Aleyda Curo, Julia Rosen, and Karina Yager for<br />

field support, Alfredo Tupayachi for help with plant identifications,<br />

Edgar Lehr and Laszlo Nagy for comments on the manuscript,<br />

the Department of Library Services at the AMNH, New<br />

York, for assistance with the Shippee–Johnson image collection,<br />

Carsten Peter and National Geographic Magazine, and John<br />

Ricker for 1970 images. This study was supported by a Columbia<br />

University-Lamont Climate Center grant, a NASA-ESS Fellowship<br />

to A.S., a National Science Foundation IRCEB grant (DEB-<br />

02133851) to P.D., and by core funding to the Consortium<br />

for Conservation Medicine from the V. Kann Rasmussen<br />

Foundation.<br />

References<br />

Boyle D, Boyle DB, Olsen V, Morgan JHA (2004) Rapid quantitative<br />

detection of chytridiomycosis (Batrachochytrium dendrobatidis)<br />

in amphibian samples using real-time Taqman PCR<br />

assay. Diseases of Aquatic Organisms, 60, 141–148.<br />

Bustamante M, Ron S, Coloma L (2005) Cambios en la diversidad<br />

en siete comunidades de anuros en los Andes de Ecuador.<br />

Biotropica, 37, 180–189.<br />

Daszak P, Berger L, Cunningham A et al. (1999) Emerging<br />

infectious diseases and amphibian population declines. Emerging<br />

Infectious Diseases, 5, 735–748.<br />

Daszak P, Cunningham AA, Hyatt AD (2000) Emerging infectious<br />

diseases of wildlife – threats to biodiversity and human<br />

health. Science, 287, 443–449.<br />

Daszak P, Cunningham AA, Hyatt AD (2003) Infectious disease<br />

and amphibian population declines. Diversity and Distribution,<br />

9, 141–150.<br />

De la Riva I, Aparicio J, Rios JN (2005) New species of Telmatobius<br />

(Anura: leptodactylidae) from humid paramo of Peru<br />

and Bolivia. Journal of Herpetology, 39, 409–416.<br />

Diaz HF, Graham NE (1996) Recent changes in tropical freezing<br />

heights and the role of sea surface temperature. Nature, 383,<br />

152–155.<br />

Diaz-Páez H, Ortiz JC (2003) Evaluación del estado de conservación<br />

de los anfibios en Chile. Revista Chilena de Historia Natural,<br />

76, 509–525.<br />

Ergueta P, Morales C (1996) Libro rojo de los vertebrados de Bolivia.<br />

Centro de los datos para la conservacion, La Paz, Bolivia.<br />

Francou B, Vuille M, Wagnon P et al. (2003) Tropical climate<br />

change recorded by a glacier in the central Andes during the<br />

last decades of the twentieth century: Chacaltaya, Bolivia, 16<br />

degrees South. Journal of Geophysical Research, 108, 4154.<br />

Halloy SRP, Gonzalez JA (1993) An inverse relationship between<br />

frost survival and atmospheric pressure. Artic and Alpine<br />

Research, 25, 117–123.<br />

Halloy SRP, Seimon A, Yager K et al. (2005) Multidimentional<br />

(climate, biodiversity, socioeconiomics, agriculture) context of<br />

changes in land use in the Vilcanota Watershed, Peru. In: Land<br />

Use Changes and <strong>Mountain</strong> Biodiversity (eds Spehn EM, Liberman<br />

Cruz M, Koerner C), pp. 323–337. CRC Press LLC,<br />

Ft. Lauderdale.<br />

Hastenrath S (1978) Heat-budget measurements on the Quelccaya<br />

ice cap, Peruvian Andes. Journal of Glaciology, 20, 85–97.<br />

La Marca E, Lips K, Lotters S et al. (2005) Catastrophic population<br />

declines and extinctions in neotropical Harlequin frogs<br />

(Bufonidae: Atelopus). Biotropica, 37, 190–201.<br />

Lips K, Brem F, Brenes R et al. (2006) Emerging infectious disease<br />

and the loss of biodiversity in a Neotropical amphibian community.<br />

Proceedings of the National Academy of Sciences USA,<br />

103, 3165–3170.<br />

Lips K, Burrowes P, Mendelson J et al. (2005) Amphibian declines<br />

in Latin America: widespread population declines, extinctions,<br />

and impacts. Biotropica, 37, 163–165.<br />

Merino-Viteri A, Coloma LA, Almendariz A (2005) Los Telmatobius<br />

(Leptodactylidae) de los Andes de Ecuador y su disminucion<br />

poblacional. In: Studies on the <strong>Andean</strong> frogs of the<br />

genera Telmatobius and Batrachophrynus. Monographias de Herpetologia,<br />

Vol. 7 (eds Lavilla EO, De la Riva I), pp. 9–38. Asociación<br />

Herpetológica Española, Valencia, Spain.<br />

Nichols DK, Lamirande EW, Pessier AP et al. (2001) Experimental<br />

transmission of cutaneous chytridiomycosis in dendrobatid<br />

frogs. Journal of Wildlife Diseases, 37, 1–11.<br />

Parmesan C (1996) Climate and species’ range. Nature, 382, 765–766.<br />

Parmesan C, Yohe G (2003) A globally coherent fingerprint of climate<br />

change impacts across natural systems. Nature, 421, 37–42.<br />

Pauli H, Gottfried M, Hohenwallner D et al. (2004) The Gloria<br />

Field Manual, Multi-Summit Approach. Directorate-General for<br />

Research, European Commission. Office for Official Publications<br />

of the European Communities: Luxembourg.<br />

Pefaur J, Duellman W (1980) Community structure in high<br />

<strong>Andean</strong> herpetofaunas. Transactions of the Kansas Academy of<br />

Sciences, 83, 45–65.<br />

Pessier AP, Nichols DK, Longcore JE et al. (1999) Cutaneous<br />

chytridiomycosis in poison dart frogs (Dendrobates spp.) And<br />

White’s tree frogs (Litoria caerulea). Journal of Veterinary<br />

Diagnostic Investigation, 194–199.<br />

Pounds AJ, Bustamante M, Coloma L et al. (2006) Widespread<br />

amphibian extinctions from epidemic disease driven by global<br />

warming. Nature, 39, 161–167.<br />

Pounds AJ, Fogdon MPL, Cambell JH (1999) Biological response<br />

to climate change on a tropical mountain. Nature, 398, 611–615.<br />

Ron S (2005) Predicting the distribution of the amphibian<br />

pathogen Batrachochytrium dendrobatidis in the new world.<br />

Biotropica, 37, 209–221.<br />

Seimon TA, Hoernig G, Sowell P et al. (2005) Identification of<br />

chytridiomycosis in Telmatobius marmoratus at 4450 m in the<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299


Cordillera Vilcanota of southern Peru. In: Studies on the <strong>Andean</strong><br />

frogs of the genera Telmatobius and Batrachophrynus. Monografías<br />

de Herpetología, Vol. 7 (eds Lavilla EO, De la Riva I), pp. 275–<br />

283. Asociación Herpetológica Española. Monografias de Herpetologia,<br />

Valencia, Spain.<br />

Shippee R (1933) Air adventures in Peru. National Geographic<br />

Magazine, Vol. 64, No. 1, pp. 81–120, January 1933.<br />

r 2006 The Authors<br />

Journal compilation r 2006 Blackwell Publishing Ltd, Global Change Biology, 13, 288–299<br />

FROG MIGRATION AND DEGLACIATION 299<br />

Swan L (1990) The highest life. Himalayan Journal, 46, 125–133.<br />

Thompson LG (1992) Ice core evidence from Peru and China.<br />

In: Climate Since A.D. 1500 (eds Bradley RS, Jones PD),<br />

pp. 517–548. Routledge, London.<br />

Vuille M, Bradley R, Werner M et al. (2003) 20th century climate<br />

change in the tropical Andes: observations and model results.<br />

Climate Change, 59, 75–99.


NEAR-SURFACE FACETED CRYSTALS AND AVALANCHE DYNAMICS IN HIGH-ELEVATION,<br />

TROPICAL MOUNTAINS OF BOLIVIA<br />

Douglas Hardy 1 , Mark W. Williams 2 , and Carlos Escobar 3<br />

1 Climate Systems Research Center and Department of Geosciences, University of Massachusetts, Amherst MA<br />

2 Department of <strong>Geography</strong> and Institute of Arctic and Alpine Research, University of Colorado, Boulder CO<br />

3 Bolivian Association of <strong>Mountain</strong> Guides, La Paz, Bolivia<br />

ABSTRACT: The importance of near-surface faceted crystals in forming weak layers associated with snow<br />

avalanches has recently received greater attention. However, there is still much to be learned concerning the<br />

formation and growth of these crystal types, their geographical extent, and related avalanche activity. Here we report<br />

on a spatially extensive avalanche cycle that occurred during late September 1999 at high-elevations in the Bolivian<br />

Andes, claiming two lives. Climbers released one slide at about 5,300 m in the Cordillera Apolobamba (on El<br />

Presidente), and four days later snow scientists servicing a high-elevation meteorological site triggered another at<br />

6,300 m near the summit of Illimani (Cordillera Real). Both slab avalanches followed lateral fracture propagation<br />

through 25-50 cm of relatively new snow; deeper pockets existed due to wind redistribution. Analysis of a snowpit<br />

on Illimani, from a nearby and safe location, showed that the avalanche ran on a thick layer of near-surface faceted<br />

crystals overlying the austral winter dry-season snow surface. Average size of the crystals was 5-7 mm, and<br />

individual crystals exceeded 10 mm in diameter. We evaluate local and regional meteorological information in an<br />

effort to understand what caused the growth of these large crystals and the resultant snowpack instability. Insights<br />

are offered regarding the avalanche hazard due to near-surface faceted crystal growth in high-elevation areas of the<br />

Tropics, where avalanches are not generally recognized as a significant hazard during the climbing season.<br />

KEYWORDS: avalanches, avalanche formation, snow, snow stratigraphy, mountains<br />

1. INTRODUCTION<br />

In late September 1999, climbers at high-elevations<br />

in the Bolivian Andes released at least two slab<br />

avalanches, as snowfall and wind loading onto the dryseason<br />

snow surface led to instability over a large<br />

region. Avalanches are not among the widely<br />

recognized hazards of climbing in Bolivia, because the<br />

climbing season is heavily concentrated in the dry<br />

southern (austral) winter months of May through<br />

September. In the most recent publication on climbing<br />

in Bolivia, Brain (1999) states that “serious accidents<br />

are rare in part because of the consistently stable and<br />

good weather during the climbing season". Avalanches<br />

receive no mention in this climbing guide, consistent<br />

with the opinion of a guide-in-training recently that<br />

“[avalanches] don’t happen in Bolivia” (Arrington,<br />

1999).<br />

On 25 September 1999, two climbers triggered a<br />

slab avalanche at an elevation of 5,200 m on Cerro<br />

Presidente, Cordillera Apolobamba, Bolivia (Table 1).<br />

Two members of the party witnessed the slide but were<br />

not involved. There was one partial burial and one<br />

complete burial, killing both climbers including the<br />

aforementioned Yossi Brain. Four days after the<br />

avalanche on Cerro Presidente, we triggered a slab<br />

release at 6,300 m near the summit of Illimani<br />

(Cordillera Real), while servicing a high-elevation<br />

-----------------------------------------<br />

* Corresponding author address: Douglas R. Hardy,<br />

Dept. of Geosciences, University of Massachusetts,<br />

Amherst MA 01003-5820; tel: 802-649-1829; fax:<br />

413-545-1200; email: dhardy@geo.umass.edu<br />

meteorological station. Both slab avalanches followed<br />

lateral fracture propagation through 25-75 cm of<br />

relatively new snow, with deeper pockets due to wind<br />

redistribution (Table 1). Meteorological, weather, and<br />

snow conditions were similar between Illimani and the<br />

Apolobamba Range where the fatalities occurred.<br />

Table 1. Characteristics of two avalanches in the<br />

Bolivian Andes, September 1999<br />

Cerro<br />

Presidente a<br />

Nevado<br />

Illimani<br />

Date 9-25-99 9-30-99<br />

Time 0830 1700<br />

Elevation (m) 5,200 6,300<br />

Incline (°) 30-40 30-40<br />

Type Slab Slab<br />

Width (m) 100 100<br />

Avalanche Trigger climbers snow scientists<br />

Activity Climbing Climbing<br />

Party size 2 3<br />

Partial Burials 1 0<br />

Complete Burials 1 0<br />

Fatalities 2 0<br />

a For further details see: http://www.csac.org/Incidents/<br />

1999-00/19990925-Bolivia.html


Analysis of a snowpit on Illimani, from a nearby<br />

and safe location, showed that the avalanche ran on<br />

well-developed faceted crystals located just below the<br />

new snow. We evaluate snowpit observations, along<br />

with local and regional meteorological information, in<br />

an effort to understand what caused the growth of these<br />

large crystals and the resultant snowpack instability.<br />

Insights are offered regarding the potential avalanche<br />

hazard in high-elevation areas of the Tropics, where<br />

avalanches are not generally recognized as a significant<br />

hazard during the climbing season.<br />

The growth of near-surface faceted crystals, and<br />

their role in the formation of weak layers within the<br />

snowpack, has recently received attention by Birkeland<br />

et al. (1996, 1998). Furthermore, Birkeland (1998)<br />

proposed a terminology and identified the predominant<br />

processes associated with the formation of weak layers<br />

in mountain snowpacks caused by near-surface faceted<br />

crystals. This work has built upon considerable work<br />

investigation the growth of faceted snow crystals in<br />

response to vapor pressure gradients resulting from<br />

temperature gradients, primarily in basal snowpack<br />

layers (e.g., Akitaya, 1974; Marbouty, 1980; Colbeck,<br />

1982; Sturm and Benson, 1997). Research focusing<br />

upon the growth of faceted crystals close to the surface<br />

has also demonstrated the importance of these<br />

gradients. Armstrong (1985) for example, determined<br />

that faceted crystal growth is initiated when the vapor<br />

pressure gradient exceeds 5 hPa m -1 . Birkeland et al.<br />

(1998) document bi-directional gradients five times<br />

larger, resulting in very rapid growth of near-surface<br />

faceted crystals. In the latter case study, a significant<br />

weak layer formed in less than 48 hours. Of great<br />

importance to the understanding of near-surface faceted<br />

crystal growth at high elevations is the mathematical<br />

treatment of the issue by Colbeck (1989), who<br />

determined that formation could occur due to either<br />

temperature cycling or solar radiation input – but that<br />

“…solar input definitely increases the sub-surface<br />

growth rate”.<br />

2. SITE DESCRIPTION<br />

The University of Massachusetts maintains two<br />

high-elevation meteorological stations near the summits<br />

of Illimani and Sajama in Bolivia (Figure 1), as part of<br />

a project to better understand the climatic signal<br />

recorded by tropical ice cores (Hardy et al., 1998). The<br />

meteorological station on Illimani was not fully<br />

functional in September; we were there to service the<br />

station and collect snow samples. Meteorological<br />

records from Sajama are used in our analysis of the<br />

September avalanche cycle to supplement the Illimani<br />

observations.<br />

Nevado Illimani (6458 m) is the highest peak in the<br />

Cordillera Real (Figure 2). The meteorological station<br />

on Illimani is located approximately 200 m below the<br />

80° W<br />

Illimani<br />

Sajama<br />

Chacaltaya<br />

a<br />

60° W 40° W<br />

10° N<br />

0°<br />

10° S<br />

20° S<br />

Figure 1: Map showing the location of Bolivia,<br />

Nevados Illimani and Sajama, and Chacaltaya. The<br />

University of Massachusetts maintains high-elevation<br />

meteorological stations near the summit of both<br />

mountains.<br />

summit (16°39’ S; 67°47’ W at 6,265 m (20,555 ft)), in<br />

a large bowl oriented to the southwest.<br />

Nevado Sajama is within the Cordillera Occidental,<br />

on the eastern side of the Altiplano (Figure 1). The<br />

meteorological station on Sajama is located about 27 m<br />

below the summit (18°06’ S; 68°53’ W at 6,515 m<br />

(21,376 ft)).<br />

Figure 2: Nevado Illimani as seen from a commercial<br />

airline flight between La Paz, the capital of Bolivia, and<br />

Santa Cruz. An oval on the image, east of the summit,<br />

indicates the 14-day old avalanche path. Illimani (6,458<br />

m) is the highest peak in the Cordillera Real, a massive<br />

mountain with three summits over 20,000 feet, visible<br />

from hundreds of miles out on the Altiplano to the west<br />

and from far out into the Amazon Basin on the east.


The Bolivian Andes experience a marked<br />

seasonality in precipitation, with an extended summer<br />

wet season and a dry winter. The 30-year record<br />

fromChacaltaya, just west of Illimani (Figure 1),<br />

illustrates the pronounced dry period June-August,<br />

when less than 5% of the annual precipitation is<br />

delivered (Figure 3). Precipitation over the entire<br />

Bolivian Altiplano (including Sajama) originates almost<br />

exclusively from the east, and annual totals decrease to<br />

the west of the Cordillera Real.<br />

Monthly Precipitation (mm)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

oct nov dec jan feb mar apr may jun jul aug sep<br />

Figure 3: Monthly precipitation at Chacaltaya, Bolivia<br />

(5,220 m), 1967 – 1997. Whisker plot boxes illustrate<br />

the median (line) and enclose the 25 th and 75 th<br />

percentiles. Error bars enclose the 10th and 90th<br />

percentiles, with extremes indicated by circles. Annual<br />

precipitation averaged 560 mm during this period.<br />

Mean annual temperature (MAT) near the Illimani<br />

weather station is –7.6°C, as indicated by the borehole<br />

temperature at 10 m (Zweifel, personal<br />

communication). MAT on Sajama is 3.0°C lower by<br />

our station measurements and by the 10m borehole<br />

temperature (Zagoradnov, personal communication).<br />

3. METHODS<br />

On 30 September 1999 a snowpit was dug and<br />

analyzed in a location adjacent to the avalanche we<br />

initiated on 29 September 1999. We measured snow<br />

properties using standard protocols (e.g., Williams et<br />

al., 1999). We used a 100 cc stainless steel Taylor-<br />

LaChapelle cutter and an electronic scale (± 1g) to<br />

measure snow density. Snowpack temperatures were<br />

measured with 20-cm long dial stem thermometers.<br />

Grain type, size, and snowpack stratigraphy were also<br />

recorded, following the protocols in Colbeck et al.<br />

(1990). The chemical and dust content of individual<br />

strata in the snowpack were also analyzed, providing<br />

additional information on the snowpack history (Hardy<br />

et al., submitted).<br />

Here we analyze the climatological and snow<br />

conditions that led to the avalanche cycle using<br />

information from the high-elevation meteorological<br />

stations on Nevado Illimani and Sajama. For this report,<br />

we emphasize measurements of air temperature and<br />

snow accumulation.<br />

4. RESULTS and DISCUSSION<br />

4.1 Avalanche on Nevado Illimani<br />

We released a slab avalanche near the summit of<br />

Nevado Illimani on 29 September 2000, similar to the<br />

slab release on Cerro Presidente 4 days earlier (Table<br />

1). We were at about 6,300 m on the afternoon of 29<br />

September en route to a weather station on Nevado<br />

Illimani in the Cordillera Real. Snow was well-sintered<br />

and supported steps and front-pointing on the 45-50°<br />

west-facing approach to a saddle below the summit of<br />

Illimani Sur. We began descending from the saddle<br />

below the summit of Illimani Sur to the weather station,<br />

on a very gentle slope of perhaps 10 degrees. As the<br />

slope began to steepen the walking became more<br />

difficult in deepening snow. We started to post-hole<br />

with every step. Twice the snow whoomphed on the<br />

very gradual slope, a phenomena one of us (CE) had<br />

experienced only five times in 13 years of climbing in<br />

Bolivia. We were in a dense white-out at the time,<br />

proceeding toward the meteorological station by<br />

reckoning. Suddenly a larger whoomph (or “firn<br />

quake”) occurred, obviously representing a widespread<br />

collapse of the snowpack. Simultaneously, we actually<br />

heard a fracture propagate to our left, followed by a dull<br />

roar. Visibility was minimal, but a fresh fracture line<br />

was visible in the snowpack underneath our feet and<br />

extending to our left, where we had heard what we<br />

thought was an avalanche release. We retreated back to<br />

the saddle and set up camp, as in the whiteout it was<br />

impossible to assess the danger of additional slides.<br />

Shortly thereafter the clouds thinned and the<br />

avalanche we released became visible (Figure 2) about<br />

200 meters to our left. The slab originated from a<br />

distinct fracture, 1-2 m thick. We estimate the slide path<br />

at 100-200 m in both width and length, although we did<br />

not have time to inspect it more closely. The runout<br />

zone of the avalanche was within 100 m of our<br />

meteorological station and snowpit.<br />

4.2 Snow Accumulation and Snowpit Stratigraphy<br />

Negative sea-surface temperature anomalies in the<br />

central equatorial Pacific (i.e., La Niña) are known to<br />

enhance wet-season precipitation on the Altiplano<br />

(Vuille et al., 2000), and a large La Niña event peaked<br />

during the 1998-99 wet season. At the Illimani weather<br />

station, 240 cm of snow accumulated by early April of<br />

1999, which buried the snow depth sensor. Visits to the<br />

station on 7 April and 17 June revealed no change in<br />

surface height, although without the sensor we have no<br />

record of variability between those dates. The snow<br />

surface on 29 September was 25-30 cm above that of 17<br />

June, representing 270 cm of accumulation since the<br />

previous November.<br />

A 180-cm deep snowpit was dug on 30 September<br />

1999, the day after the avalanche, on a slope angle of


about 10 degrees. Stratigraphic analysis and sample<br />

collection were conducted from about 1300 to 1600 hrs<br />

on that day. Weather was clear, bright, and brisk, with<br />

the sun almost directly overhead. Snow temperatures<br />

were about 0°C in the first 9 cm, then about -3.0°C for<br />

the remainder of the snowpack (Figure 4). Density<br />

generally increased with depth, from about 210 kg m -3<br />

at the snow surface to 410 kg m -3 at a depth of 180 cm,<br />

with the exception of several ice lenses.<br />

9wc<br />

6cl<br />

9sc<br />

6cl<br />

6cl<br />

8il<br />

6cl<br />

4sf<br />

4sf<br />

Density (kg m-3 180<br />

0 300 600 900 -3 0 1 2 3 4 5<br />

) ° C Grain size (mm)<br />

0<br />

20<br />

40<br />

60<br />

80<br />

100<br />

120<br />

140<br />

160<br />

Figure 4: Snowpit profile sampled at 6,300 m on 30<br />

September 1999, near the summit of Nevado Illimani,<br />

Bolivia. Note the presence of near-surface faceted<br />

crystals from 27-64 cm.<br />

The stratigraphy of the snowpit contained some<br />

interesting surprises. Snow grains in the first 19 cm<br />

were rounded, suggesting the presence of some liquid<br />

water and wet snow metamorphism (e.g. Colbeck,<br />

1979). The small grain size of 0.3 mm along with the<br />

absence of polycrystalline grains suggests that<br />

conditions consisted of snow with low water content<br />

and little if any percolation. A thin ice layer at 19-20<br />

cm was consistent with either subsurface melting and<br />

refreezing (e.g., Liston, 1999), or refreezing of draining<br />

meltwater. Snow grains from 20-27 cm were similar to<br />

grains at 0-19 cm, though slightly larger at 0.5 mm in<br />

diameter.<br />

Depth (cm)<br />

Below a depth of 27 cm, the stratigraphy suggested<br />

a different snow environment. Snow grains at a depth of<br />

27-37 cm were heavily faceted. Many of the grains<br />

showed well-developed cups and scroll patterns similar<br />

to kinetic growth or depth hoar grains that develop at or<br />

near the bottom of the snowpack (Colbeck, 1983).<br />

Average grain size was 3-5 mm, with several grains<br />

larger than 10 mm. Dust was visible throughout this<br />

layer, with much higher concentrations in the upper 3<br />

cm of this layer. Sintering was particularly poor in this<br />

layer. The layer from 37-64 cm was similar to the<br />

above layer, with the exception that faceting was<br />

evident but not as well-developed. Grain sizes of 2-4<br />

mm were smaller than in the layer above. There was no<br />

visible dust in this layer. A well-bonded ice lens was at<br />

65-73 cm, composed of melt-freeze grains frozen<br />

together.<br />

The avalanche that we triggered appeared to run on<br />

the faceted grains located at a depth of 27-37 cm in our<br />

snowpit. We believe that the faceted grains we report<br />

for layers from 27-64 cm are near-surface faceted<br />

crystals, formed by the process Birkeland (1998) terms<br />

diurnal re-crystalization. To our knowledge, these nearsurface<br />

faceted grains are the largest ever reported,<br />

either in the avalanche literature or informally among<br />

avalanche professionals. Their presence suggests<br />

unique meteorological conditions at high elevations in<br />

the tropical Andes.<br />

4.3 Meteorological Conditions<br />

4.3.1 Precipitation<br />

Direct measurements of precipitation on Illimani<br />

are not available through the 1999 dry season. From<br />

observations, we know that 25-30 cm of snow<br />

accumulated sometime between 17 June and 29<br />

September, and snowpit evidence indicates that this<br />

increment fell within days to weeks prior to 29<br />

September. Such winter precipitation events have been<br />

recorded in prior years on Illimani, including two<br />

snowfalls in August 1997 (19 and 12 cm), 15 cm in<br />

September of that year, and one 10 cm event during<br />

June of 1998.<br />

Bolivian station data show that a widespread<br />

regional event occurred on 17 September 1999 (7.4 to<br />

7.9 mm at La Paz, Cochabamba, and Oruro). On<br />

Sajama, after a prolonged period without precipitation,<br />

19 cm of accumulation was recorded on 18-19<br />

September and 23 cm on 28-30 September. These are<br />

relatively large events for the more arid Sajama, and the<br />

accumulation was not blown away as typically occurs<br />

with winter precipitation events on Sajama.<br />

One additional piece of indirect evidence for the<br />

timing of precipitation comes from Illimani. During the<br />

dry (low humidity) evening of 22-23 September, snow<br />

was evidently drifting in response to wind speeds of 4-7


Table 2. Mean monthly temperatures and incoming solar radiation (K-down) on Sajama, June - September 1999.<br />

Minimum Maximum Daily Maximum<br />

Aspirated Daily Aspirated Daily Temperature Daily<br />

T<br />

T<br />

Range<br />

K-down<br />

Month (deg. C) (deg. C) (deg. C) (W m -2 )<br />

June 1999 -16.6 -9.6 7.0 817<br />

July 1999 -16.6 -10.4 6.3 763<br />

August 1999 -15.9 -8.1 7.8 861<br />

September 1999 -15.2 -6.9 8.3 950<br />

m s -1 , because the reflected solar radiometer was buried<br />

beneath the snow surface on 23 September. Indeed, for<br />

the two week period prior to the Illimani avalanche,<br />

winds were consistently from the northwest (blowing<br />

downslope) at speeds varying between 2 and 8 m s -1 . In<br />

summary, considerable snowfall apparently occurred<br />

within the period 8-12 days prior to avalanches in the<br />

Cordilleras Apolobamba and Real, accompanied by<br />

sufficient wind to cause redistribution of the new snow.<br />

4.3.2 Solar radiation and air temperature<br />

Solar radiation receipt remains high during the<br />

austral winter at high elevations in the Andes, with<br />

daily irradiance maxima typically greater than 800 W<br />

m -2 (Hardy et al., 1998). However, the mean<br />

temperature on Sajama during the winter is only –<br />

12.8°C (1996-99), due to the low air density at high<br />

elevation. In contrast to mid-latitude mountain<br />

environments, the average diurnal temperature range of<br />

7.8°C is larger than the annual temperature range.<br />

Monthly measurements on Sajama from June through<br />

September 1999 show increasing solar radiation, and<br />

increases in both mean air temperature and the daily<br />

temperature range (Table 2). Illimani temperatures are<br />

generally slightly higher, due to the 250 m elevation<br />

difference and higher humidity, but unavailable for this<br />

time period.<br />

5. SUMMARY<br />

The near-surface faceted crystals observed on<br />

Illimani in September 1999 appear to be both better<br />

developed (i.e., large) and form a thicker layer than any<br />

reported previously. While the best-developed grains<br />

were in a layer 10 cm thick, those in the 27 cm thick<br />

layer below were also faceted. This thicker layer may<br />

have been briefly exposed to the surface prior to the<br />

additional 10 cm of accumulation, however, the lack of<br />

dust in this layer suggests that the faceting occurred<br />

after the layer was buried (i.e., 10 – 37 cm deep).<br />

Conditions unique to high-elevation tropical mountains<br />

favoring the growth of near-surface faceted crystals<br />

include: (1) higher input of incoming solar radiation,<br />

due to less atmosphere; (2) greater absorption of solar<br />

radiation at the snow surface during the austral winter,<br />

due to the relatively high dust concentration (enhancing<br />

the temperature gradient); and (3) greater longwave<br />

radiation loss due to rapid cooling of the dry, thin<br />

atmosphere in the evening.<br />

Development of near-surface faceted crystals<br />

during the dry season, at high elevations in the Bolivian<br />

Andes, may present a greater avalanche hazard than is<br />

generally recognized. Two documented incidents in<br />

September 1999 occurred when ‘early season’ snow<br />

was redistributed onto leeward slopes, burying a thick,<br />

well-developed layer of near-surface faceted crystals.<br />

This layer probably formed through the process<br />

Birkeland (1998) termed diurnal recrystallization. In<br />

this region of the Andes, our results suggest that ideal<br />

meteorological conditions for faceted crystal growth<br />

persist through the entire climbing season of June-<br />

August, and probably occur most years in highelevation<br />

areas. Indeed, we have observed buried nearsurface<br />

faceted crystals within a previous years<br />

snowpack on Sajama. Regional precipitation decidedly<br />

increases during September, and relatively large<br />

snowfall events are not uncommon in our short record<br />

from Illimani. Avalanches may occur when synoptic<br />

weather conditions result in dry-season snow events<br />

and high wind speeds that produce a slab over this weak<br />

layer. La Nina conditions, such as during 1998-99, may<br />

favor this situation. Although the resulting instability<br />

should be relatively easy to recognize due to audible<br />

collapse upon loading (whoomphing), climbers must be<br />

alert to the potential threat. Currently, climbers and<br />

other users of the Bolivian high-mountain environment<br />

are not expecting to encounter avalanche hazards<br />

during the climbing season, and many have little<br />

experience in evaluating avalanche dangers.<br />

6. ACKNOWLEDGEMENTS<br />

Funding was provided by a NOAA Global<br />

Programs Office grant to the University of<br />

Massachusetts, a Fulbright Fellowship and Faculty<br />

Fellowship from the University of Colorado-Boulder<br />

(MW), as well as NSF Hydrology and International<br />

Programs. Thanks to Karl Birkeland for initially


stimulating our interest in near-surface faceted crystals<br />

at high elevations. We are grateful for excellent<br />

logistical support from Nuevo Horizontes in La Paz.<br />

This work has also benefited greatly from the on-going<br />

involvement of Carsten Braun and Mathias Vuille at the<br />

University of Massachusetts.<br />

7. REFERENCES<br />

Akitaya, E., 1974: Studies on depth hoar. Contr. Inst.<br />

Low Temp. Sci., A-26, 1-67.<br />

Armstrong, R.L., 1985: Metamorphism in a<br />

subfreezing, seasonal snow cover: The role of<br />

thermal and vapor pressure conditions. Ph.D<br />

dissertation, Dept. Geogr., University of Colorado,<br />

Boulder, CO.<br />

Arrington, V., 1999: Tragedy strikes in the<br />

Apolobamba. Bolivian Times, 7 October 1999, see<br />

also: http://www.latinwide.com/boltimes/<br />

edit9940/tapa.htm<br />

Birkeland, K.W., R.F. Johnson, and D.S. Schmidt,<br />

1996: Near-surface faceted crystals: Conditions<br />

necessary for growth and contribution to avalanche<br />

formation, southwest Montana, U.S.A. Proc. 1996<br />

Int. Snow Sci. Wksp., Banff, Alberta, Canada, 75-<br />

79.<br />

Birkeland, K.W., R.F. Johnson, and D.S. Schmidt,<br />

1998: Near-surface faceted crystals formed by<br />

diurnal recrystallization: A case study of weak<br />

layer formation in the mountain snowpack and its<br />

contribution to snow avalanches. Arctic Alpine<br />

Res., 30, 200-204.<br />

Birkeland, K.W., 1998: Terminology and predominant<br />

processes associated with the formation of weak<br />

layers of near-surface faceted crystals in the<br />

mountain snowpack. Arctic Alpine Res., 30, 193-<br />

199.<br />

Brain, Y., 1999: Bolivia: A Climbing Guide. The<br />

<strong>Mountain</strong>eers, Seattle, 224p.<br />

Colbeck, S.C., 1979: Grain clusters in wet snow. J.<br />

Colloid Interface Sci., 72, 371-384.<br />

Colbeck, S.C., 1982: Growth of faceted crystals in a<br />

snow cover. CRREL Rep. 82-29, U.S. Army Cold<br />

Regions Res. Eng. Lab., Hanover, NH.<br />

Colbeck, S.C., 1983: Theory of metamorphism of dry<br />

snow. J. Geophys. Res., 88, 5475-5482.<br />

Colbeck, S.C., 1989: Snow-crystal growth with<br />

varying surface temperatures and radiation<br />

penetration. J. Glaciol., v. 35, 23-29.<br />

Colbeck, S. C., E. Akitaya, R. Armstrong, H. Gubler, J.<br />

Lafeuille, K. Lied, D. McClung, and E. Morris,<br />

1990: The international classification for seasonal<br />

snow on the ground, pp. 1-23, National Snow and<br />

Ice Data Center, Boulder,CO.<br />

Hardy, D.R., M. Vuille, C. Braun, F. Keimig, and R.S.<br />

Bradley, 1998: Annual and daily meteorological<br />

cycles at high altitude on a tropical mountain. Bull.<br />

Amer. Meteorol. Soc., 79, 1899-1913.<br />

Hardy, D.R., M.W. Williams, and C. Escobar: Nearsurface<br />

faceted crystals, avalanche dynamics and<br />

climate in high-elevation tropical mountains of<br />

Bolivia. submitted to Cold Regions Sci. and Tech.,<br />

Oct. 2000.<br />

Liston, G.E., J. -G. Winther, O. Bruland, H. Elvehøy,<br />

and K. Sand, 1999: Below-surface ice melt on the<br />

coastal Antarctic ice sheet. J. Glaciol, 45, 273-<br />

285.<br />

Marbouty, D., 1980: An experimental study of<br />

temperature-gradient metamorphism. J. Glaciol.,<br />

26, 303-312.<br />

Sturm, M. and C.S. Benson, 1997: Vapor transport,<br />

grain growth, and depth-hoar development in the<br />

subarctic snow. J. Glaciol., 43, 42-59.<br />

Vuille, M., R.S. Bradley, and F. Keimig, F., 2000:<br />

Interannual climate variability in the Central Andes<br />

and its relation to tropical Pacific and Atlantic<br />

forcing. J. Geophys. Res., 105, 12,447-12,460.<br />

Williams, M.W., D. Cline, M. Hartman, and T.<br />

Bardsley, 1999: Data for snowmelt model<br />

development, calibration, and verification at an<br />

alpine site, Colorado Front Range. Water Resour.<br />

Res., 35, 3205-3209.


Global Change Biology (2010) 16, 3223–3232, doi: 10.1111/j.1365-2486.2010.02203.x<br />

Nonlinear climate change and <strong>Andean</strong> feedbacks: an<br />

imminent turning point?<br />

M. B. BUSH, J. A. HANSELMAN 1 and W. D . G O S L I N G 2<br />

Department of Biological Sciences, Florida Institute of Technology, 150 W. University Blvd., Melbourne, FL 32901, USA<br />

Abstract<br />

A 370 000-year paleoecological record from Lake Titicaca provides a detailed record of past climate change in which<br />

interglacial periods are seen to have some elements of commonality, but also some key differences. We advance a<br />

conceptual feedback model to account for the observed changes that includes previously ignored lake effects. Today<br />

Lake Titicaca serves to warm the local environment by about 4–5 1C and also to increase rainfall. We observe that as<br />

water levels in the lake are drawn down due to warm, dry, interglacial conditions, there is a possible regional cooling<br />

as the lake effect on local microclimates diminishes. Positive feedback mechanisms promote drying until much of the<br />

lake basin is reduced to salt marsh. Consequently, the usual concept of upslope migration of species with warming<br />

would not be applicable in the Altiplano. If, as projected, the next century brings warmer and drier conditions than<br />

those of today, a tipping point appears to exist within ca. 1–2 1C of current temperatures, where the relatively benign<br />

agricultural conditions of the northern Altiplano would be replaced by inhospitable arid climates. Such a change<br />

would have profound implications for the citizens of the Bolivian capital, La Paz.<br />

Keywords: aridity, charcoal, conservation, fossil pollen, grayscale, Lake Titicaca, positive feedback, warming<br />

Received 21 October 2009; revised version received 16 December 2009 and accepted 17 December 2009<br />

Introduction<br />

For the last 5000 years, inhabitants of the Bolivian<br />

Altiplano have taken advantage of the relatively mild,<br />

moist climate adjacent to Lake Titicaca (Erickson, 1988).<br />

Lying at 3810 m elevation, Lake Titicaca is the world’s<br />

highest great lake and produces a local microclimate or<br />

lake effect that allows cultivation of crops, such as<br />

maize, at unusually high elevations. However, paleoecological<br />

data show that the size of this and other vast<br />

paleolakes of the Altiplano have been volatile, exhibiting<br />

rapid changes in lake level with presumed concomitant<br />

changes in microclimate. The 370 thousand year<br />

(ka)-long dataset from Lake Titicaca (Peru/Bolivia)<br />

(Hanselman et al., 2005, in press), reveals that environmental<br />

changes associated with the warming of marine<br />

isotope stage (MIS) 9 and 5e are more extreme than of<br />

the Holocene. The trajectory of <strong>Andean</strong> warming (ca.<br />

1 1C in the last 30 years) Andes suggests a minimum<br />

warming of 1–2 1C by mid century (Christensen et al.,<br />

2007) producing temperatures probably equivalent to<br />

1<br />

Present address: J. A. Hanselman, Department of Biology, Westfield<br />

<strong>State</strong> College, Westfield, MA, USA.<br />

2 Present address: W. D. Gosling, Department of Earth and Environmental<br />

Sciences, CEPSAR, The Open University, Milton Keynes,<br />

UK.<br />

Correspondence: M. B. Bush, tel. 1 1 321 674 7166, e-mail:<br />

mbush@fit.edu<br />

those of MIS 9 and 5e (Van der Hammen & Hooghiemstra,<br />

2003) by ca. 2050.<br />

A common expectation has developed among biologists<br />

that, as the world warms, species will migrate<br />

upslope and poleward (Parmesan & Yohe, 2003). Such<br />

changes are already documented on tropical mountains<br />

particularly where ecotones are strong and range shifts<br />

readily detectable, such as the lower limit of cloud<br />

forest (Pounds et al., 1999; Pounds, 2001). Assumptions<br />

that species will continue to migrate upslope have<br />

raised concerns that as the bioclimatic envelopes defining<br />

the potential range of a species lift off the top of<br />

mountains, these upper-elevation species will go extinct<br />

(Peters & Darling, 1985; Thomas et al., 2004; Malcolm<br />

et al., 2006). Although such extinction may be the case in<br />

some settings, an alternative scenario is one in which,<br />

rather than marching steadily upslope in response to<br />

linear climate change, the system flips to a new state.<br />

Nonlinear changes in climate have been suggested to<br />

underlie millennial-scale climate pulses such as Dansgaard-Oeschger<br />

cycles and Heinrich events (Broecker &<br />

Denton, 1989; Stocker, 2000; Bond et al., 2001), and are<br />

clearly capable of causing rapid and profound switches<br />

of climatic direction. Such changes would require an<br />

entirely different societal response to conserve livelihoods,<br />

agricultural production, and biota, compared<br />

with that of incremental linear changes.<br />

Thus, estimates of future climate change need to take<br />

into consideration phenomena operating from global to<br />

r 2010 Blackwell Publishing Ltd 3223


3224 M. B. BUSH et al.<br />

local scales, as their interaction, through feedback mechanisms,<br />

could result in abrupt tipping points that lead<br />

to nonlinear responses in temperature or precipitation.<br />

The paleoecological record becomes our best resource<br />

for studying such complex responses, and attention<br />

needs to focus on times when climate changed rapidly,<br />

and temperatures were warmer-than-modern.<br />

Here we describe, climatic events and feedbacks<br />

associated with past interglacials that were probably<br />

about 1–3 1C warmer-than-present in the Andes. Twice,<br />

a tipping point was reached that led to a profound<br />

alteration of regional ecosystems.<br />

The present and past climates of Lake Titicaca<br />

Lake Titicaca (Bolivia/Peru, 15.5–171S, 68.5–701W) lies at<br />

3810 m above sea level in the Altiplano (Fig. 1) and has a<br />

modern climate that is cool and dry. Cloudiness is much<br />

greater during the wet season than the dry season, with<br />

just 6 h of bright sunlight in January compared with<br />

9.6 h in July (http://waterwiki.net/index.php/Lake_Titicaca-Poopo).<br />

A lake effect is evident in local temperatures<br />

with averages at Titicaca about 5 1C higher than<br />

areas remote from the lake. Roche et al. (1992) mapped<br />

temperatures around Titicaca and concluded that the<br />

lake warming effect extended for tens of kilometers<br />

around the lake. Similarly, the lake influences precipitation<br />

patterns. Average annual rainfall within the Altiplano<br />

ranges from ca. 200 mm yr 1 in the south to<br />

400 mm yr 1 in the north, with ca. 890 mm yr 1 over<br />

the main basin of Titicaca. About 80% of the precipitation<br />

falls during the period of the South American<br />

summer monsoon (SASM) (December–March) when<br />

the winds are dominated by an easterly flow bringing<br />

Amazonian moisture to the Altiplano (Roche et al., 1992).<br />

Over Lake Titicaca itself, precipitation is supplemented<br />

by night-time convective rains in the wet season.<br />

Climatic oscillations between glacial and interglacial<br />

landscapes were evident in the 370 ka paleoecological<br />

record from Lake Titicaca (Hanselman et al., 2005, in<br />

press). During glacials, cold temperatures resulted in<br />

very little pollen production as the shoreline of Lake<br />

Titicaca became a glacial foreland. The ice masses lay<br />

within 100 m vertically of the lake, but never reached<br />

the shore (Seltzer, 1990; Seltzer et al., 2002). During the<br />

glacial periods, combinations of meltwater and changes<br />

in the precipitation : evaporation (P : E) ratio of open<br />

water bodies resulted in both higher and lower lake<br />

levels than those of today. During highstand events, the<br />

increased overflow of Lake Titicaca into the Rio Desaguadero<br />

contributed to the flooding of the southern<br />

Altiplano (Ybert, 1992; Clapperton, 1993; Placzek et al.,<br />

2006). Considerable debate surrounds the timing of<br />

highstands that resulted in a series (temporally) of<br />

paleolakes with shorelines as much as 120 m above<br />

the modern level of the Salar de Uyuni (e.g. Hastenrath<br />

& Kutzbach, 1985; Baker et al., 2001a; Placzek et al., 2006;<br />

Gosling et al., 2008). At the dry extreme, during each of<br />

the four interglacials heightened concentrations of carbonate<br />

were deposited in the sediments of Lake Titicaca<br />

relative to modern values. Carbonate was deposited as<br />

the level of Titicaca fell below its outlet and the P : E<br />

ratio declined causing salts in the lake to become more<br />

concentrated (Baker et al., 2005). In the Holocene, the<br />

peak calcium carbonate (CaCO3) concentration in sediments<br />

was about 54%, but during prior interglacials it<br />

was as high as 100%, perhaps suggesting longer or more<br />

intense periods of evaporative dominance. These long<br />

oscillations appeared to have followed the pattern of<br />

orbital precession, a finding that was consistent between<br />

<strong>Andean</strong> and Amazonian lakes and regional<br />

speleothem records (Hooghiemstra et al., 1993; Baker<br />

et al., 2001a; Bush et al., 2002; Fritz et al., 2004; Wang<br />

et al., 2004; Cruz et al., 2005).<br />

Although long-term precessional variability underlay<br />

the lowstand of the mid-Holocene (Baker et al., 2001a),<br />

shorter-term events, ranging from annual to millennial,<br />

were superimposed on this basic pattern (Ekdahl et al.,<br />

2008; Hillyer et al., 2009). Understanding the mid-Holocene<br />

dry event provides insights into changes associated<br />

with MIS 9 and 5e. The mid-Holocene dry<br />

event has been a consistent feature of paleoecological<br />

records from western Amazonia and the Andes. As<br />

knowledge of the event deepened it became apparent<br />

that it should be described as a period of increased<br />

drought probability, rather than a single drought<br />

(Hillyer et al., 2009). Between 9 and 4.4 ka, the net effect<br />

was a drier climate that lowered lake level, but this was<br />

not a time of uniform aridity. Rather, the climate had<br />

shifted toward a drier tendency, marked in the paleoecological<br />

record by sediment chemistry and fossil diatoms<br />

associated with shallow, ephemeral, or saline<br />

lakes. However, the same proxies indicate brief but<br />

frequent wetter periods in which lake levels rose<br />

(Hillyer et al., 2009).<br />

The mid-Holocene drying lowered lake level at Titicaca<br />

by ca. 85 m (Seltzer et al., 1998; Baker et al., 2001b).<br />

Once the lake level had dropped below the outflow to<br />

the southern basin of Huinaimarca (a 25 m lowering),<br />

further loss of water from the system was due to<br />

evaporation. A hydrological model of this event suggested<br />

that P : E ratios could account for the observed<br />

lowering of lake level if windspeed was held constant,<br />

temperature increased by 1 1C and precipitation fell by<br />

40%. Tweaking of the parameters in the hydrological<br />

simulations demonstrated that lake level was not<br />

strongly responsive to variations in temperature (Cross<br />

et al., 2000).<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232


Fig. 1 Sketch map of the Altiplano in Peru/Bolivia showing the modern lakes (gray shading) and salars (hashing). Also shown on the<br />

main diagram is the highstand associated with paleolake Tauca (short dashed line, after Placzek et al., 2006), and, on the inset, the<br />

lowstands of the Holocene (dashed and dotted line) and MIS 5e (long dashed line, after Baker et al., 2005). Location of core site LT01-2B<br />

marked with star on inset.<br />

Whatever had caused the compound dry events<br />

comprising the mid-Holocene drought ended ca.<br />

5.2 ka, when lake levels started to rebound (Baker<br />

et al., 2001b; Paduano et al., 2003). Most lakes in the<br />

Andes and Amazon were filled close to modern levels<br />

by 4.4 ka (Seltzer et al., 2000; Abbott et al., 2003). This<br />

wet event was apparently due to increased precipitation<br />

that corresponded both to the increased intensity of El<br />

Niño/southern oscillation (ENSO) events and reduced<br />

summer insolation to the northern tropical Atlantic<br />

Ocean (Moy et al., 2002; Riedinger et al., 2002; Rein,<br />

2007).<br />

Thus, the long-term pattern provided by precessional<br />

change was overlain by regional and local factors that<br />

produced sharply defined changes in lake level. Similar<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232<br />

interactions of factors operating on a range of spatial<br />

and temporal scales would be expected to influence<br />

past and future climate.<br />

Methods<br />

ANDEAN TIPPING POINTS 3225<br />

Analysis of the fossil pollen data from core LT01-2B followed<br />

standard protocols and has been described in detail elsewhere<br />

(Hanselman et al., 2005). The chronology of the core was<br />

established via a combination of isotopic dating using 14 C<br />

A.M.S and U/Th, and wiggle matching to the Vostok ice core<br />

record (Petit et al., 1999) and is described in Fritz et al. (2007,<br />

2008).<br />

Grayscale analysis, in which the relative darkness (blackest<br />

5 0, whitest 5 255) of sediment bands was recorded. This


3226 M. B. BUSH et al.<br />

technique was applied to the MIS 5e section of the core based<br />

on high-resolution images collected on a Geotek core-logger at<br />

the LacCore facility, and analyzed using IMAGEJ (Rasband,<br />

2005).<br />

Results<br />

The most striking features of the Lake Titicaca paleoecological<br />

record are the differences between glacial<br />

and interglacial conditions (Fig. 2). During glacials very<br />

little pollen was deposited and there was an almost<br />

complete absence of fire from the system (Hanselman<br />

et al., in press). However, during each of the interglacials<br />

a progression from Polylepis woodland to Puna<br />

vegetation is apparent and fire becomes a regular<br />

feature of the landscape. Changing lake levels are<br />

indicated by the presence of aquatic taxa such as<br />

Myriophyllum and Isöetes. In general, Isöetes could not<br />

withstand the cold peaks of glacials nor the warm<br />

peaks of the interglacials, and therefore is at peak<br />

abundance at the transition between stages. Similarly,<br />

Polylepis occurs primarily in the transitional period<br />

(Gosling et al., 2009) and declines as charcoal concentrations,<br />

i.e. fire events, increase in abundance and<br />

frequency.<br />

Another marked pattern is the difference in the pollen<br />

signatures between interglacials. In MIS 9 and 5e,<br />

following peaks of aquatic taxa and Polylepis, Amaranthaceae<br />

(cf. Chenopodium) dominate the interglacial<br />

flora. In contrast to this pattern, MIS 7 and 1 the glacial<br />

termination and inception stages are similar to those of<br />

MIS 9 and 5e, but there is no Amaranthaceae-dominated<br />

stage in the middle of the interglacial.<br />

The grayscale analysis of sediments ascribed to MIS<br />

5e revealed increasing contrast as the matrix of the<br />

sediment changed from gray to black with sharply<br />

defined white laminae composed of CaCO3 (Fig. 3).<br />

Discussion<br />

During the early phases of both MIS 9 and 5e the lack of<br />

charcoal in sediments, the upslope migration of <strong>Andean</strong><br />

taxa, and occupation of the Altiplano by fire-sensitive<br />

trees such as Polylepis, indicate warm, moist conditions<br />

(Gosling et al., 2009). A simple model of warming<br />

would predict that the vertical migration of trees would<br />

continue until they fully occupied the lake basin (Fig. 4).<br />

However, long before the peak of the interglacial, a<br />

sudden transition was evident in which trees were<br />

replaced by Amaranthaceae and the 280 m modern<br />

water depth of Lake Titicaca fell so much that large<br />

beds of Myriophyllum formed in shallow water (or on<br />

exposed mudflats). As the lake became more saline this<br />

genus was replaced by Amaranthaceae (Hanselman<br />

et al., in press).<br />

The lake sediment deposited at the inferred peak of<br />

MIS 5e was an almost pure precipitate of CaCO3. The<br />

mid-Holocene warming, which resulted in a 85 m lowering<br />

of lake level was sufficient to deposit some carbonate,<br />

but the change in conditions was insufficient to<br />

generate a comparable change in vegetation (Paduano<br />

et al., 2003). Cross et al. (2000) estimated that a 40%<br />

reduction in precipitation for about 4000 years was sufficient<br />

to achieve the Holocene drawdown. Here we add to<br />

those insights with a conceptual feedback model that<br />

highlights the importance of local effects in this system.<br />

A few other pieces of information are important for<br />

our model. First, following the transition to an Amaranthaceae-dominated<br />

system, climate was not constant<br />

and wetter events punctuated the dry state. Evidence<br />

for the wet episodes comes from peaks of Asteraceae<br />

pollen percentages and concentrations rising and falling<br />

in counterpoint to Amaranthaceae. We interpret the<br />

Amaranthaceae to indicate dry times and low lake level,<br />

with Asteraceae representing wetter cycles. Indeed, the<br />

sediment throughout this section of the core is finely<br />

Fig. 2 The fossil record of selected pollen and spores from Lake Titicaca core LT01-2B. Note the peak of Myriophyllum in MIS 9 was<br />

truncated, actual peak was 6000% of dry land pollen (after Hanselman et al. in press).<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232


Fig. 3 Grayscale analysis of sediments attributed to MIS 5e in the Lake Titicaca LT01-2B core. An image of the sediment is shown<br />

directly beneath the grayscale values for that section.<br />

laminated (Fig. 3). Dark organic material alternating<br />

with carbonate-rich layers, indicate a lake oscillating<br />

between states and rather rapid and strong climate<br />

change.<br />

These data suggest that more than a simple precessional<br />

pattern is represented. ENSO is thought to have<br />

been operating in a broadly similar temporal pattern<br />

to today during much of MIS 5e (Tudhope et al., 2001),<br />

and so it is possible that when the lake is at very<br />

low levels the influence of long ENSO-like cycles on<br />

<strong>Andean</strong> precipitation become evident. Alternatively,<br />

and perhaps more likely, the oscillations could reflect<br />

thermal changes in the Atlantic Ocean. Baker et al.<br />

(2005) suggested that Bond Cycles were evident in the<br />

Holocene portion of the Lake Titicaca record with cold<br />

conditions in the North Atlantic corresponding to wet<br />

conditions on the Altiplano, and warm conditions corresponding<br />

to dry times. This suggestion is complimentary<br />

to the hypothesis that weakened trade winds<br />

resulting from warming of the subtropical North Atlantic<br />

would weaken the transport mechanism that pumps<br />

Atlantic moisture into Amazonia, the South American<br />

low level jet (SALLJ), and induce aridity in Amazonia<br />

and the Andes (Marengo et al., 2008; Zeng et al., 2008).<br />

This mechanism probably caused the Amazon drought<br />

of AD 2005, and here we suggest that it may also have<br />

been important at longer timescales in response to<br />

persistent thermal dipoles in the Atlantic Ocean (Bond<br />

et al., 2001; Baker et al., 2005).<br />

Conceptual models of nonlinear climate change in the<br />

Altiplano<br />

Our conceptual model starts with precessional forcing<br />

that promotes warm conditions in western Amazonia<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232<br />

and the Andes. The large-scale presence of ice changes<br />

the feedbacks, and so glacial and interglacial versions<br />

are presented.<br />

Interglacial<br />

ANDEAN TIPPING POINTS 3227<br />

Warming during the deglacial period increased convection<br />

and exported moisture from Amazonia via SASM<br />

to the Altiplano. A combination of increased precipitation<br />

and runoff from meltwater facilitated formation of<br />

a paleolake. Hence the largest paleolakes would be<br />

predicted to occur within the deglacial period. Periodically,<br />

late Pleistocene paleolakes occupied much of the<br />

Altiplano, with paleolake Tauca (Fig. 1) forming during<br />

the last deglaciation. Lake Tauca occupied 450 000 km 2<br />

(i.e. ca. 6 the area of modern Titicaca) and would have<br />

exerted a considerable lake effect on the climate of the<br />

southern Altiplano. However, this warm wet interval<br />

was cut short by an ‘oceanic forcing’ (Fig. 5a), which<br />

could have been due to warming of either the eastern<br />

equatorial Pacific (EEP) inducing El Niño, a warming in<br />

the subtropical Atlantic weakening the SALLJ, or some<br />

combination of these phenomena. If the SASM weakened<br />

(December–March) and skies were less cloudy,<br />

i.e. more direct sunlight than the present average<br />

6 h day 1 , evaporation would increase. As drought<br />

gripped the Altiplano a positive feedback mechanism<br />

could have exacerbated this effect.<br />

The paleolake would have begun to contract and as it<br />

did so the warm, moist, lake effect would have been<br />

reduced. A decreased lake area would lead to reduced<br />

humidity and thereby increased evaporation, which<br />

would have reinforced reduction of lake area. Although<br />

humidity provokes convective rains over Titicaca, in<br />

general nights over the adjacent land are cloudless. If


3228 M. B. BUSH et al.<br />

(a)<br />

(b)<br />

(c)<br />

Fig. 4 Schematic diagram showing (a) the modern system of<br />

Titicaca, (b) a hypothetical migration response due to warming<br />

assuming that species would migrate upslope, and (c) the outcome<br />

of warming in prior interglacials (MIS 5e and 9) in which<br />

forest did not migrate to fill the basin, but the ecosystem dried<br />

out to become a salt marsh.<br />

nightly low temperatures over the land warmed, the<br />

temperature differential between the lake and the land<br />

would have lessened (currently 4 1C with the lake being<br />

warmer than the land at night). One consequence of<br />

reducing that differential would have been to weaken<br />

the lake breeze (2–4 m s 1 as an average modern flow)<br />

that promotes night-time convective rain over the lake,<br />

further reducing inputs of water and thereby accentuating<br />

the drawdown of the lake.<br />

The lowstand phase would be reversed by external<br />

forcing such as a change in the dominant phase of<br />

ENSO or in the temperature regime of the subtropical<br />

Atlantic.<br />

Glacial<br />

The cool conditions of the glacial probably reduced the<br />

importance of convective activity (Fig. 5b). As the ice<br />

mass grew so the influence of katabatic winds coming<br />

from the ice would have increased their microclimatic<br />

influence. The cold descending air would have encountered<br />

a relatively warm lake and created fog in the base<br />

of the valleys. Thus, the system would have been wet,<br />

cold, and light-limited for much of the year. Fog over<br />

the lake surface and generally cold air would combine<br />

to reduce evaporation. Warmer oscillations would have<br />

added meltwater to accentuate highstands. The feedback<br />

loop that linked precession to height of lake can<br />

serve to draw down the lake in the absence of ice (i.e. in<br />

an interglacial), but it requires the meltwater component<br />

to boost lake level from this forcing to create a<br />

highstand. Thus once the ice mass was established<br />

precessional patterns became evident, but as the ice<br />

mass melted, there was a much weaker precessional<br />

signature of wetting.<br />

Lake level effects and climate change<br />

It is an important realization that the modern conditions<br />

of the Altiplano are part of a cycle from high to low lake<br />

level change and that modern conditions are not a<br />

‘norm.’ The cycle does not follow a regular rhythm as<br />

it is subject to multiple forcings, some of which are, at<br />

best, quasiperiodic, i.e. Bond Cycles. Nor are the cycles<br />

of lake level variability of equal amplitude, i.e. they<br />

produce high or lowstands of different magnitudes.<br />

Thus during MIS 5e, the cycle was strong taking the<br />

Altiplano to a drier extreme. In looking for an approximate<br />

analog for MIS 5e conditions in Lake Titicaca,<br />

some similarities might exist with Lake Poopó, which is<br />

a shallow saline lake that receives about 390 mm of<br />

precipitation annually. As the lake receives some input<br />

from Titicaca via the Rio Desaguadero, 390 mm becomes<br />

a minimum amount of precipitation that can sustain an<br />

open body of water on the Altiplano. For Lake Titicaca<br />

to flip to a saline lake, such as Lake Poopó, a reduction<br />

of 450% precipitation might be needed.<br />

We hypothesize that as evaporation increased and<br />

lake level dropped, it triggered the positive feedback of<br />

aridification, in which temperatures dropped due to<br />

loss of lake effect, and reduced lake surface area resulted<br />

in less convective rain falling basinwide. Just a<br />

25 m lowering of lake level ceases drainage into the<br />

Huinaimarca sub-basin. That sub-basin would dry out<br />

(this is known from paleoecological data; Gosling et al.,<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232


Fig. 5 Conceptual model of feedbacks that result in nonlinear responses in climate change in the Altiplano of Peru/Bolivia. (a)<br />

Interglacial feedbacks, note that deglacial feedbacks are grayed and (b) glacial feedbacks.<br />

2008), the overflow into the Desaguadero River would<br />

have stopped and the downstream system of Lake<br />

Poopó would have dried up. Consequently, a warmer<br />

interglacial instead of stimulating local temperature<br />

increase, and increased convective precipitation, might<br />

at its peak have induced falling temperatures and<br />

aridity in this basin.<br />

Some quantification of when this tipping point<br />

occurred can be made based on moist air adiabatic<br />

lapse rates and observations of tree line. In the absence<br />

of anthropogenic clearance, modern tree-line in the<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232<br />

ANDEAN TIPPING POINTS 3229<br />

Andes would probably lie between 3300 and 3700 m<br />

elevation (Wille et al., 2002; Di Pasquale et al., 2008), i.e.<br />

300–500 m below Lake Titicaca. Moist air adiabatic lapse<br />

rates in this section of the Andes are ca. 5.2 1C per<br />

1000 m of vertical ascent. As the trees never fully<br />

invaded the shoreline, the tip to drier conditions must<br />

have occurred within 1–2 1C of modern temperatures.<br />

Thus, anthropogenic climate change could drive<br />

this system past that the tipping point, which judging<br />

from the abrupt rise in Amaranthaceae pollen in MIS 9<br />

and 5e, comes with very little warning.


3230 M. B. BUSH et al.<br />

This pattern of aridity was observed in both MIS 9<br />

and 5e. MIS 7 does not appear to have been quite so<br />

extreme and its cycle toward aridity was curtailed<br />

before the most extreme lake draw down was reached.<br />

This observation is entirely consistent with marine and<br />

ice core records that show MIS 7 to have been a<br />

protracted, multipeak interglacial, but with less extreme<br />

conditions than MIS 9 and 5e. If, as is projected, climates<br />

warm by 3–6 1C this century, the positive feedback<br />

toward aridity might be renewed. Although the last<br />

1000 years have probably been the wettest of the<br />

Holocene in the Andes, warming of the subtropical<br />

north Atlantic may have already begun, and has been<br />

suggested to have induced the AD 2005 Amazonian<br />

drought (Marengo et al., 2008; Zeng et al., 2008). Indeed,<br />

aerosols, e.g. SO2 and fine particulate carbon, may have<br />

mitigated subtropical Atlantic warming, and as the<br />

loading of these pollutants is reduced, warming may<br />

accelerate (Cox et al., 2008). If that trend becomes<br />

persistent it would probably increase fire frequency,<br />

decrease forest cover, and result in reduced moisture<br />

transport into the Andes (da Silva et al., 2008; Cochrane<br />

& Barber, 2009). Similarly, if the sea surface temperature<br />

(SST) changes predicted by the HADCM3LC happen,<br />

then a warmer EEP would probably result in <strong>Andean</strong><br />

drought. Of these two processes ENSO variation appears<br />

to be dominant, as effects of subtropical Atlantic<br />

temperature are clearest when ENSO is weak (Zeng<br />

et al., 2008). However, the systems are not mutually<br />

exclusive and perhaps the fluctuating pulses of wet and<br />

dry conditions within the overall dry interglacial conditions<br />

reflect a synergy between these systems.<br />

One factor that will differ substantially between past<br />

lowstands on the Altiplano and the modern state is that<br />

it will occur in a landscape containing a capital city (La<br />

Paz, Bolivia) and approximately 2 200 000 human inhabitants.<br />

Ecosystem and societal adaptation<br />

The predictable ecosystem response to a lowering of<br />

Lake Titicaca and desiccation of the Altiplano would be<br />

a stalling of the anticipated upslope migration of forest<br />

and the expansion of xerophytic species. Erickson<br />

(1988) has argued that humans overcome climate<br />

change in the Andes through cultural adaptation and<br />

landscape manipulation. While this appears to be true<br />

of the relatively subtle changes in climate of the last<br />

3000 years, the response of humans to the millennial<br />

scale mid-Holocene droughts in the Altiplano was to<br />

abandon the region (Núñez et al., 2002). Speculating<br />

about how humans could maintain crop productivity<br />

and water supply is beyond the scope of this paper,<br />

though the twin obstacles of increasingly concentrated<br />

salts in surface waters and decreased water availability<br />

(both from reduced precipitation and loss of ice caps)<br />

would be substantial hurdles to overcome.<br />

In the seminatural systems of the Andes, populations<br />

of species that were stressed by the drier conditions<br />

would be predicted to go locally extinct, survive in<br />

moist microrefugia, or undergo selection to genotypes<br />

that could withstand drought. Emphasizing the conservation<br />

of potential microrefugia and ecosystem connectivity<br />

may become an important component of<br />

strategies to safeguard some at-risk species.<br />

Biomass and fuel load are inextricably linked in the<br />

high Andes, and fire is a landscape transformer. Plantations<br />

of trees that trap moisture could mitigate some<br />

drought effects, as shading helps to retain soil moisture<br />

and organic material, plus the structure of the plants<br />

serves to intercept moisture from ground-level cloud.<br />

Thus such stands, while increasing biomass, could<br />

reduce fire risk. However, Eucalyptus, which is the most<br />

widely planted tree in the high Andes, wicks prodigious<br />

quantities of moisture out of the soil, effectively<br />

drying the landscape, and is highly combustible (Luzar,<br />

2007). Eucalyptus plantations increase the probability<br />

that the wettest locations would become fire-prone<br />

and unsuitable for many of the indigenous elements<br />

that might otherwise occupy these settings. A more<br />

desirable candidate for afforestation is Polylepis, but<br />

control of fire would be essential for their ability to<br />

mature (Cierjacks et al., 2008).<br />

Conclusion<br />

Cycles of lake level have characterized at least the last<br />

370 ka on the Peru/Bolivian Altiplano. Orbital forcing<br />

has been a basic pacemaker of events, but superimposed<br />

on this pattern are changes in moisture availability<br />

driven by oceanic processes, which are themselves partially<br />

controlled by orbital patterns. Nonlinear feedbacks<br />

appear to have induced rapid landscape change.<br />

Changes in lake area have significant climatic effects on<br />

the Altiplano and positive feedbacks are envisaged that<br />

lead both to high and low lake stands.<br />

In MIS 9 and 5e, a trend toward warm wet conditions<br />

was initiated by changes in precession, but suddenly<br />

altered by a change in the moisture balance. We hypothesize<br />

that oceanic forcing induced droughts that<br />

were then accentuated through positive feedback mechanisms,<br />

leading to a profound change in local ecosystems.<br />

Eventually, outside forcing broke the feedback<br />

cycle and once again reversed these trends.<br />

On the Altiplano, contracting lake area during interglacial<br />

warm periods may have initiated feedback effects<br />

so strong that they reversed the thermal trend of<br />

the progression toward warming. A similar turning<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232


point where the dominant signature shifting from<br />

warming to aridity may occur again as regional temperatures<br />

rise 1–2 1C. Despite this turning point potentially<br />

reducing local temperatures, it offers no solace to<br />

conservation of natural resources as the accompanying<br />

droughts would cause major dislocations of natural and<br />

human communities.<br />

A key lesson to learn from this study is that although<br />

species have already started to expand their ranges<br />

upslope on tropical mountains this pattern could be<br />

abruptly halted, even reversed, as a result of nonlinear,<br />

local, responses to ongoing global warming. A critical<br />

difference between MIS 9 and 5e in South America<br />

relative to the Holocene is the presence of a human<br />

population. Our prognoses of climate change are of<br />

direct relevance to the livelihoods of more than 2<br />

million people living in the Peruvian and Bolivian<br />

Andes, as they will influence their choice of land use,<br />

occupancy of marginal lands, and which kind of lands<br />

become marginal.<br />

The modern relevance of this study is emphasized by<br />

the recent push to use the Andes for carbon sequestration.<br />

The need to understand the future capability of the<br />

land to support forest or maintain soil moisture vital for<br />

carbon sequestration is critical to the long-term success<br />

of such ventures. In addition to its global benefits,<br />

agroforestry may play a role in delaying or alleviating<br />

the fullest local effects of these droughts. However,<br />

choice of species to be planted and control of fire are<br />

as important as the area that becomes forested.<br />

Acknowledgements<br />

This work was supported by Grant NSF-ATM 0317539. Andy<br />

Lloyd of the Open University is thanked for preparing Fig. 1.<br />

This work is a product of the Andes Biodiversity and Ecosystem<br />

Research Group (ABERG) and is publication #6 of the Florida<br />

Institute for Technology, Institute for Research on Global Climate<br />

Change.<br />

References<br />

Abbott MB, Wolfe BB, Wolfe AP et al. (2003) Holocene paleohydrology and glacial<br />

history of the central Andes using multiproxy lake sediment studies. Palaeogeography,<br />

Palaeoclimatology, Palaeoecology, 194, 123–138.<br />

Baker PA, Fritz SC, Garland J, Ekdahl E (2005) Holocene hydrologic variation at Lake<br />

Titicaca, Bolivia/Peru, and its relationship to North Atlantic climate variation.<br />

Journal of Quaternary Science, 20, 655–662.<br />

Baker PA, Rigsby CA, Seltzer GO, Fritz SC, Lowenstein TK, Bacher NP, Veliz C (2001a)<br />

Tropical climate changes at millennial and orbital timescales on the Bolivian<br />

Altiplano. Nature, 409, 698–701.<br />

Baker PA, Seltzer GO, Fritz SC et al. (2001b) The history of South American tropical<br />

precipitation for the past 25,000 years. Science, 291, 640–643.<br />

Bond G, Kromer B, Beer J et al. (2001) Persistent solar influence on North Atlantic<br />

climate during the Holocene. Science, 294, 2130–2136.<br />

Broecker WS, Denton GH (1989) The role of ocean–atmosphere reorganizations in<br />

glacial cycles. Geochimica et Cosmochimica Acta, 53, 2465–2501.<br />

Bush MB, Miller MC, de Oliveira PE, Colinvaux PA (2002) Orbital forcing signal in<br />

sediments of two Amazonian lakes. Journal of Paleolimnology, 27, 341–352.<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232<br />

ANDEAN TIPPING POINTS 3231<br />

Christensen JH, Hewitson B, Busuioc A et al. (2007) Regional climate projections. In:<br />

Climate Change 2007: the Physical Science Basis. Contribution of Working Group I to the<br />

Fourth Assessment Report of the Intergovernmental Panel on Climate. Change. (eds<br />

Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, Tignor M, Miller<br />

HL), pp. 847–940. Cambridge University Press, New York.<br />

Cierjacks A, Salgado S, Wesche K, Hensen I (2008) Post-fire population dynamics of<br />

two tree species in high-altitude Polylepis forests of Central Ecuador. Biotropica, 40,<br />

176–182.<br />

Clapperton CW (1993) Quaternary Geology and Geomorphology of South America. Elsevier,<br />

Amsterdam.<br />

Cochrane MA, Barber CP (2009) Climate change, human land use and future fires in<br />

the Amazon. Global Change Biology, 15, 601–612.<br />

Cox PM, Harris PP, Huntingford C et al. (2008) Increasing risk of Amazonian drought<br />

due to decreasing aerosol pollution. Nature, 453, 212–215.<br />

Cross SL, Baker PA, Seltzer GO, Fritz SC, Dunbar RB (2000) A new estimate of the<br />

Holocene lowstand level of Lake Titicaca, central Andes, and implications for<br />

tropical palaeohydrology. The Holocene, 10, 21–32.<br />

Cruz FW Jr, Burns SJ, Karmann I et al. (2005) Insolation-driven changes in atmospheric<br />

circulation over the past 116,000 years in subtropical Brazil. Nature, 434, 63–66.<br />

da Silva RR, Werth D, Avissar R (2008) Regional impacts of future land-cover changes<br />

on the Amazon basin wet-season climate. Journal of Climate, 21, 1153–1170.<br />

Di Pasquale G, Marziano M, Impagliazzo S, Lubritto C, De Natale A, Bader MY (2008)<br />

The Holocene treeline in the northern Andes (Ecuador): first evidence from soil<br />

charcoal. Palaeogeography, Palaeoclimatology, Palaeoecology, 259, 17–34.<br />

Ekdahl EJ, Fritz SC, Baker PA, Rigsby CA, Coley K (2008) Holocene multidecadal- to<br />

millennial-scale hydrologic variability on the South American Altiplano. The<br />

Holocene, 18, 867–876.<br />

Erickson CL (1988) Raised field agriculture in the Lake Titicaca Basin: putting ancient<br />

agriculture back to work. Expedition, 30, 8–16.<br />

Fritz SC, Baker PA, Lowenstein TK et al. (2004) Hydrologic variation during the last<br />

170,000 years in the southern hemisphere tropics of South America. Quaternary<br />

Research, 61, 95–104.<br />

Fritz SC, Baker PA, Seltzer GO, Ballantyne A, Tapia P, Cheng H, Edwards RL (2008)<br />

Corrigendum to ‘‘Quaternary glaciation and hydrologic variation in the South<br />

American tropics as reconstructed from the Lake Titicaca drilling project’’ [Quaternary<br />

Research 68 (2007) 410–420]. Quaternary Research, 69, 342.<br />

Fritz SC, Baker PA, Seltzer GO, Ballantyne A, Tapia PM, Cheng H, Edwards RL (2007)<br />

Quaternary glaciation and hydrologic variation in the South American tropics<br />

as reconstructed from the Lake Titicaca drilling project. Quaternary Research, 68,<br />

410–420.<br />

Gosling WD, Bush MB, Hanselman JA, Chepstow-Lusty AJ (2008) Glacial–interglacial<br />

changes in moisture balance and the impact on vegetation in the southern hemisphere<br />

tropical Andes (Bolivia/Peru). Palaeogeography, Palaeoclimatology, Palaeoecology, 259,<br />

35–50.<br />

Gosling WD, Hanselman JA, Knox C, Valencia BG, Bush MB (2009) Long term drivers<br />

of change in Polylepis woodland distribution in the central Andes. Journal of<br />

Vegetation Science, 20, 1041–1052.<br />

Hanselman JA, Bush MB, Gosling WD, Collins A, Knox C (in press) A 370,000-year<br />

record of vegetation and fire history around Lake Titicaca (Bolivia/Peru). Palaeogeography,<br />

Palaeoclimatology, Palaeoecology.<br />

Hanselman JA, Gosling WD, Paduano GM, Bush MB (2005) Contrasting pollen<br />

histories of MIS 5e and the Holocene from Lake Titicaca (Bolivia/Peru). Journal of<br />

Quaternary Science, 20, 663–670.<br />

Hastenrath S, Kutzbach J (1985) Late Pleistocene climate and water budget of the<br />

South American Altiplano. Quaternary Research, 24, 249–256.<br />

Hillyer R, Valencia BG, Bush MB, Silman MR, Steinitz-Kannan M (2009) A 24,<br />

900-year paleolimnological history from the Peruvian Andes. Quaternary Research,<br />

71, 71–82.<br />

Hooghiemstra H, Melice JL, Berger A, Shackleton NJ (1993) Frequency spectra and<br />

paleoclimatic variability of the high-resolution 30–1450 ka Funza I pollen record<br />

(Eastern Cordillera, Colombia). Quaternary Science Reviews, 12, 141–156.<br />

Luzar J (2007) The political ecology of a ‘‘forest transition’’: Eucalyptus forestry in the<br />

Southern Peruvian Andes. Ethnobotany Research and Applications, 5, 85–93.<br />

Malcolm JR, Liu C, Neilson RP, Hansen L, Hannah L (2006) Global warming and<br />

extinctions of endemic species from biodiversity hotspots. Conservation Biology, 20,<br />

538–548.<br />

Marengo JA, Nobre CA, Tomasella J et al. (2008) The Drought of Amazonia in 2005.<br />

Journal of Climate, 21, 495–516.<br />

Moy CM, Seltzer GO, Rodbell DT, Anderson DM (2002) Variability of El Nino/<br />

Southern Oscillation activity at millennial timescales during the Holocene epoch.<br />

Nature, 420, 162–164.


3232 M. B. BUSH et al.<br />

Núñez L, Grosjean M, Cartajena I (2002) Human occupations and climate change in<br />

the Puna de Atacama, Chile. Science, 298, 821–824.<br />

Paduano GM, Bush MB, Baker PA, Fritz SC, Seltzer GO (2003) A vegetation and fire<br />

history of Lake Titicaca since the Last Glacial Maximum. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology, 194, 259–279.<br />

Parmesan C, Yohe G (2003) A globally coherent fingerprint of climate change impacts<br />

across natural systems. Nature, 421, 37–42.<br />

Peters RL, Darling JDS (1985) The greenhouse effect and nature reserves. BioScience, 35,<br />

707–711.<br />

Petit JR, Jouzel J, Raynaud D et al. (1999) Climate and atmospheric history of the past<br />

420,000 years from the Vostok ice core, Antarctica. Nature, 399, 429–436.<br />

Placzek C, Quade J, Patchett PJ (2006) Geochronology and stratigraphy of late<br />

Pleistocene lake cycles on the southern Bolivian Altiplano: implications for causes<br />

of tropical climate change. Geological Society of America Bulletin, 118, 515–532.<br />

Pounds JA (2001) Climate and amphibian declines. Nature, 410, 639–640.<br />

Pounds JA, Fogden MPL, Campbell JH (1999) Biological response to climate change on<br />

a tropical mountain. Nature, 398, 611–615.<br />

Rasband WS (2005) ImageJ 1.32j. National Institute of Health, USA.<br />

Rein B (2007) How do the 1982/83 and 1997/98 El Niños rank in a geological record<br />

from Peru? Quaternary International, 161, 56–66.<br />

Riedinger MA, Steinitz-Kannan M, Last WM, Brenner M (2002) A 6100 14C yr record<br />

of El Nino activity from the Galapagos Islands. Journal of Paleolimnology, 27, 1–7.<br />

Roche MA, Bourges JC, Mattos R (1992) Climatology and hydrology of the Lake<br />

Titicaca basin. In: Lake Titicaca: A Synthesis of Limnological Knowledge (eds Dejoux C,<br />

Iltis A), pp. 63–83. Kluwer Academic Press, Boston.<br />

Seltzer G, Rodbell D, Burns S (2000) Isotopic evidence for late Quaternary climatic<br />

change in tropical South America. Geology, 28, 35–38.<br />

Seltzer GO (1990) Recent glacial history and paleoclimate of the Peruvian–Bolivian<br />

Andes. Quaternary Science Reviews, 9, 137–152.<br />

Seltzer GO, Cross S, Baker P, Dunbar R, Fritz S (1998) High-resolution seismic<br />

reflection profiles from Lake Titicaca, Peru/Bolivia. Evidence for Holocene aridity<br />

in the tropical Andes. Geology, 26, 167–170.<br />

Seltzer GO, Rodbell DT, Baker PA, Fritz SC, Tapia PM, Rowe HD, Dunbar RB (2002)<br />

Early deglaciation in the tropical Andes. Science, 298, 1685–1686.<br />

Stocker T (2000) Past and future reorganizations in the climate system. Quaternary<br />

Science Reviews, 19, 301–319.<br />

Thomas CD, Cameron A, Green RE et al. (2004) Extinction risk from climate change.<br />

Nature, 427, 145–148.<br />

Tudhope AW, Chilcott CP, McCulloch MT et al. (2001) Variability in the El Nino-<br />

Southern Oscillation through a glacial–interglacial cycle. Science, 291, 1511–1517.<br />

Van der Hammen T, Hooghiemstra H (2003) Interglacial–glacial Fuquene-3 pollen<br />

record from Colombia: an Eemian to Holocene climate record. Global and Planetary<br />

Change, 36, 181–199.<br />

Wang X, Auler AS, Edwards RL et al. (2004) Wet periods in northeastern Brazil over the<br />

past 210 kyr linked to distant climate anomalies. Nature, 432, 740–743.<br />

Wille M, Hooghiemstra H, Hofstede R, Fehse J, Sevink J (2002) Upper forest line<br />

reconstruction in a deforested area in northern Ecuador based on pollen and<br />

vegetation analysis. Journal of Tropical Ecology, 18, 409–440.<br />

Ybert JP (1992) Ancient lake environments as deduced from pollen analysis. In: Lake<br />

Titicaca. A Synthesis of Limnological Knowledge (eds DeJoux C, Iltis A), pp. 49–62.<br />

Kluwer Academic Publishers, Boston.<br />

Zeng N, Yoon JH, Marengo JA, Subramaniam A, Nobre CA, Mariotti A, Neelin JD<br />

(2008) Causes and impacts of the 2005 Amazon drought. Environmental Research<br />

Letters, 3, 014002, doi: 10.1088/1748-9326/1083/1081/014002.<br />

r 2010 Blackwell Publishing Ltd, Global Change Biology, 16, 3223–3232

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!