Top Banner
Please cite this article in press as: Ramirez-Villegas, J., et al. Using species distributions models for designing conservation strategies of Tropical Andean biodiversity under climate change. Journal for Nature Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007 ARTICLE IN PRESS G Model JNC-25349; No. of Pages 14 Journal for Nature Conservation xxx (2014) xxx–xxx Contents lists available at ScienceDirect Journal for Nature Conservation j o ur nal homepage: www.elsevier.de/jnc Using species distributions models for designing conservation strategies of Tropical Andean biodiversity under climate change Julian Ramirez-Villegas a,b,c,, Francisco Cuesta d , Christian Devenish e,f , Manuel Peralvo d , Andy Jarvis a,b , Carlos Alberto Arnillas g,h a CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Cali AA6713, Colombia b Decision and Policy Analysis (DAPA), International Center for Tropical Agriculture (CIAT), Cali AA6713, Colombia c Institute for Climate and Atmospheric Science (ICAS), School of Earth and Environment, University of Leeds, Leeds LS2 9JT, UK d Biodiversity Department Consorcio para el Desarrollo Sostenible de la Ecorregión Andina (CONDESAN), Ecuador e BirdLife International Americas Secretariat, Ecuador f School of Science and the Environment, Manchester Metropolitan University, UK g Centro de Datos para la Conservación, Universidad Agraria La Molina, Peru h University of Toronto-Scarborough, Department of Physical and Environmental Sciences, Toronto, M1C 1A4, ON, Canada a r t i c l e i n f o Article history: Received 14 June 2013 Received in revised form 24 March 2014 Accepted 24 March 2014 Keywords: Andes Biodiversity Climate change Climatic niche Conservation Maxent Threats a b s t r a c t Biodiversity in the Tropical Andes is under continuous threat from anthropogenic activities. Projected changes in climate will likely exacerbate this situation. Using species distribution models, we assess pos- sible future changes in the diversity and climatic niche size of an unprecedented number of species for the region. We modeled a broad range of taxa (11,012 species of birds and vascular plants), including both endemic and widespread species and provide a comprehensive estimation of climate change impacts on the Andes. We find that if no dispersal is assumed, by 2050s, more than 50% of the species studied are projected to undergo reductions of at least 45% in their climatic niche, whilst 10% of species could be extinct. Even assuming unlimited dispersal, most of the Andean endemics (comprising 5% of our dataset) would become severely threatened (>50% climatic niche loss). While some areas appear to be climatically stable (e.g. Pichincha and Imbabura in Ecuador; and Nari ˜ no, Cauca, Valle del Cauca and Putu- mayo in Colombia) and hence depict little diversity loss and/or potential species gains, major negative impacts were also observed. Tropical high Andean grasslands (páramos and punas) and evergreen mon- tane forests, two key ecosystems for the provision of environmental services in the region, are projected to experience negative changes in species richness and high rates of species turnover. Adapting to these impacts would require a landscape-network based approach to conservation, including protected areas, their buffer zones and corridors. A central aspect of such network is the implementation of an integrated landscape management approach based on sustainable management and restoration practices covering wider areas than currently contemplated. © 2014 Elsevier GmbH. All rights reserved. Introduction Despite ambitious goals to significantly reduce the rate of biodiversity loss by 2010 (CBD 2007), biodiversity continues to be severely threatened (Ramirez-Villegas et al. 2012; Sachs et al. 2009). These threats include over exploitation of natural resources (e.g. water, agricultural soils), habitat loss and degrada- tion, and invasive species (Butchart et al. 2010; Kim & Byrne 2006). Biodiversity loss has been increasing since the second half of the Corresponding author at: Centro Internacional de Agricultura Tropical (CIAT), Km 17 recta Cali-Palmira, Colombia. Tel.: +57 (2) 445 0100 x 3656. E-mail addresses: [email protected], [email protected] (J. Ramirez-Villegas). 20th century, and is likely to continue into the future (Kim & Byrne 2006; MEA 2005). With climate change entailing likely increases in temperature and regional and seasonal changes in precipitation (Knutti & Sedlacek 2013), ecosystems and their services are likely to suffer additional stresses (Chen et al. 2009; Feeley & Silman 2010; Fuhrer 2003; IPCC 2007). The Tropical Andes tops the list of worldwide hotspots for species diversity and endemism (Fjeldså et al. 1999; Gentry 1995; Sklenár & Ramsay 2001). For this reason, the region is considered a key priority for biodiversity conservation (Brooks et al. 2006; Myers et al. 2000). At the same time, the Tropical Andes have been iden- tified as one of the most severely threatened natural areas globally (Jetz et al. 2007; Mittermeier et al. 1997). During the last century, concentration of human population and associated demands for http://dx.doi.org/10.1016/j.jnc.2014.03.007 1617-1381/© 2014 Elsevier GmbH. All rights reserved.
14

Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

Mar 01, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

J

Us

JAa

b

c

d

e

f

g

h

a

ARRA

KABCCCMT

I

btertB

K

h1

ARTICLE IN PRESSG ModelNC-25349; No. of Pages 14

Journal for Nature Conservation xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Journal for Nature Conservation

j o ur nal homepage: www.elsev ier .de / jnc

sing species distributions models for designing conservationtrategies of Tropical Andean biodiversity under climate change

ulian Ramirez-Villegasa,b,c,∗, Francisco Cuestad, Christian Devenishe,f, Manuel Peralvod,ndy Jarvisa,b, Carlos Alberto Arnillasg,h

CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Cali AA6713, ColombiaDecision and Policy Analysis (DAPA), International Center for Tropical Agriculture (CIAT), Cali AA6713, ColombiaInstitute for Climate and Atmospheric Science (ICAS), School of Earth and Environment, University of Leeds, Leeds LS2 9JT, UKBiodiversity Department – Consorcio para el Desarrollo Sostenible de la Ecorregión Andina (CONDESAN), EcuadorBirdLife International – Americas Secretariat, EcuadorSchool of Science and the Environment, Manchester Metropolitan University, UKCentro de Datos para la Conservación, Universidad Agraria La Molina, PeruUniversity of Toronto-Scarborough, Department of Physical and Environmental Sciences, Toronto, M1C 1A4, ON, Canada

r t i c l e i n f o

rticle history:eceived 14 June 2013eceived in revised form 24 March 2014ccepted 24 March 2014

eywords:ndesiodiversitylimate changelimatic nicheonservationaxent

hreats

a b s t r a c t

Biodiversity in the Tropical Andes is under continuous threat from anthropogenic activities. Projectedchanges in climate will likely exacerbate this situation. Using species distribution models, we assess pos-sible future changes in the diversity and climatic niche size of an unprecedented number of species for theregion. We modeled a broad range of taxa (11,012 species of birds and vascular plants), including bothendemic and widespread species and provide a comprehensive estimation of climate change impactson the Andes. We find that if no dispersal is assumed, by 2050s, more than 50% of the species studiedare projected to undergo reductions of at least 45% in their climatic niche, whilst 10% of species couldbe extinct. Even assuming unlimited dispersal, most of the Andean endemics (comprising ∼5% of ourdataset) would become severely threatened (>50% climatic niche loss). While some areas appear to beclimatically stable (e.g. Pichincha and Imbabura in Ecuador; and Narino, Cauca, Valle del Cauca and Putu-mayo in Colombia) and hence depict little diversity loss and/or potential species gains, major negativeimpacts were also observed. Tropical high Andean grasslands (páramos and punas) and evergreen mon-tane forests, two key ecosystems for the provision of environmental services in the region, are projected

to experience negative changes in species richness and high rates of species turnover. Adapting to theseimpacts would require a landscape-network based approach to conservation, including protected areas,their buffer zones and corridors. A central aspect of such network is the implementation of an integratedlandscape management approach based on sustainable management and restoration practices coveringwider areas than currently contemplated.

ntroduction

Despite ambitious goals to significantly reduce the rate ofiodiversity loss by 2010 (CBD 2007), biodiversity continueso be severely threatened (Ramirez-Villegas et al. 2012; Sachst al. 2009). These threats include over exploitation of natural

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

esources (e.g. water, agricultural soils), habitat loss and degrada-ion, and invasive species (Butchart et al. 2010; Kim & Byrne 2006).iodiversity loss has been increasing since the second half of the

∗ Corresponding author at: Centro Internacional de Agricultura Tropical (CIAT),m 17 recta Cali-Palmira, Colombia. Tel.: +57 (2) 445 0100 x 3656.

E-mail addresses: [email protected], [email protected] (J. Ramirez-Villegas).

ttp://dx.doi.org/10.1016/j.jnc.2014.03.007617-1381/© 2014 Elsevier GmbH. All rights reserved.

© 2014 Elsevier GmbH. All rights reserved.

20th century, and is likely to continue into the future (Kim & Byrne2006; MEA 2005). With climate change entailing likely increasesin temperature and regional and seasonal changes in precipitation(Knutti & Sedlacek 2013), ecosystems and their services are likely tosuffer additional stresses (Chen et al. 2009; Feeley & Silman 2010;Fuhrer 2003; IPCC 2007).

The Tropical Andes tops the list of worldwide hotspots forspecies diversity and endemism (Fjeldså et al. 1999; Gentry 1995;Sklenár & Ramsay 2001). For this reason, the region is considered akey priority for biodiversity conservation (Brooks et al. 2006; Myers

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

et al. 2000). At the same time, the Tropical Andes have been iden-tified as one of the most severely threatened natural areas globally(Jetz et al. 2007; Mittermeier et al. 1997). During the last century,concentration of human population and associated demands for

Page 2: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

2 r Natu

gonbw2ehPsatce

becKtcewcpaTt

cTwtpjasw

S

cVNtot(bo(tCsdeceO

M

a

ARTICLENC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal fo

oods and services in the inter-Andean valleys and the inner slopesf the Andean ridges, has transformed a significant portion of theatural landscape causing habitat loss and degradation followedy species extinction and disruption of ecosystem functions (e.g.ater-flow regulation), especially in the Northern Andes (Bruinsma

003; Wassenaar et al. 2007; Armenteras et al. 2011; Rodríguezt al. 2013). Resource-base over-exploitation of natural resourcesas led to a severe land degradation process (Poulenard et al. 2001;oulenard et al. 2004), increasing the pressure on the goods andervices provided by these ecosystems (Rundel & Palma 2000). Inddition, the Andes are expected to undergo severe stresses overhe next 100 years as a result of on-going land use change pro-esses, but also of climate change (Beaumont et al. 2011; Malcolmt al. 2006).

Addressing potential impacts from climate change is importantecause the environmental impacts of human activities (Biesmeijert al. 2006; MEA 2005) could be exacerbated by the likely rapidhanges in the climate system during the 21st century (IPCC 2007;nutti & Sedlacek 2013). Warren et al. (2013) estimated that, in

he absence of any climate change mitigation strategy, large rangeontractions for ca. 60% of plants and 35% of animals could bexpected globally. Understanding and quantifying the extent athich climate change could threaten Andean species is therefore

ritical since many of the species in the region occur in low denseopulations with narrow distribution patterns (i.e. endemics) with

high level of replacement within the environmental gradients.hese characteristics make the Andean biota particularly sensitiveo climate change disruptions.

Our primary objective was to assess the likely impacts of climatehange on the distributions of vascular plant and bird species of theropical Andes. Using species distributions modelling techniques,e assessed the potential climatic niche of 11,012 species, and

hen projected them under the SRES-A2 emission scenario for twoeriods: 2020s (2010-2039) and 2050s (2040-2069). Future pro-

ected changes in species assemblages, including richness, turnovernd range size were assessed. Lastly, the projected impacts inelected groups of species of Andean origin were analyzed. Finally,e discuss future strategies to reduce expected biodiversity loss.

tudy area

The study area (Tropical Andes hereafter) comprises all inter-onnected areas above altitudes of 500 m within the countries ofenezuela, Colombia, Ecuador, Peru and Bolivia, plus the Sierraevada de Santa Marta in Colombia, delimited using data from

he SRTM digital elevation model (Farr et al., 2007). Extendingver 1.5 million km2 from 11◦ N to 23◦ S, the Tropical Andes arehe longest and widest mountain region in the tropics (Fig. 1)Clapperton 1993; Fjeldså & Krabbe 1990). The morphological andioclimatic heterogeneity of the Andes have led to the formationf an enormous diversity of microhabitats favouring speciationMittermeier et al. 1997; Young et al. 2002). Moreover, their loca-ion between the lowlands of the Amazon, La Chiquitanía and Elhaco to the east and the Chocó, Tumbes-Guayaquil and the aridystems of the Sechura Desert to the west, has created complexynamics of species exchange and isolation (Bass et al. 2010; Youngt al. 2002). The Tropical Andes harbours more than 45,000 vas-ular plant (20,000 endemics) and 3400 vertebrate species (1567ndemics) in just 1% of the Earth’s land mass (Lamoreux et al. 2006;lson et al. 2001).

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

ethods

We modeled the climatic niches of 11,012 species (1555 birdsnd 9457 plants) using species distributions models. We modeled

PRESSre Conservation xxx (2014) xxx–xxx

the climate-constrained present-day distributions of all species,and projected them onto two different future periods (2020s and2050s) and two contrasting dispersal scenarios. The approachimplemented here aims to evaluate the likely impacts of climatechange on the widest array possible of Andean plant and birdspecies by mid-2020s and mid-2050s and comprises the followingfour steps:

1. Assembling of species occurrence data2. Generation of climate surfaces3. Maximum entropy species distribution modeling4. Analysis of projected climate change impacts on species assem-

blages

Based on these analyses, we delineate conservation recommen-dations for the 2020s and 2050s.

Species datasets

Presence data for 11,012 species (1555 birds and 9457 plants)were sourced from three databases. CONDESAN, the Centro deDatos para la Conservación de la Universidad Nacional Agraria LaMolina (CDC-UNALM), and a previous global study (Warren et al.2013) (hereafter W2013). From the three sources, we extracted alloccurrences in the five Tropical Andean countries (i.e. Venezuela,Colombia, Ecuador, Peru and Bolivia) of all vascular plant clades(Magnoliophyta, Pteridophyta, Pinophyta, Psilophyta, Cycadophyta,Gnetophyta, Lycopodiophyta) and bird (class Aves, phylum Chordata)species with at least one record within the study area (Fig. 1B). Byincluding these three sources of data we ensured the inclusion ofcommon and widespread species (see Warren et al. 2013) as wellas narrow-range Andean endemics and imperil species (also seesection Results, sub-section “Species datasets” for details).

CONDESAN’s database consisted of data from multiple sources.Vascular plant specimen data were obtained from the Mis-souri Botanical Garden’s Vascular Tropicos (VAST) nomenclaturaldatabase (MBG 2004), the Herbarium of the National Science Insti-tute in Colombia (ISN) and the Catholic University Herbarium (QCA)in Ecuador. Bird species data were obtained from databases belong-ing to the Chicago Field Museum of Natural History, Academy ofNatural Sciences of Philadelphia, California Academy of Sciencesand the Berkeley Museum of Natural History and cross-checkedwith BirdLife International database (version 2012). Additional datawere obtained from private databases (Juan Fernando Freile forAntpittas, Paul Hamec for Dendroica cerulea; Cal Dodson-LorenaEndara for orchid’s records and James Luteyn’s database stored atthe New York Botanical Garden site for Ericaceae) and publishedliterature (Casares et al. 2003; Renjifo et al. 2002; Schuchmannet al. 2001). The CDC-UNALM database was produced from thereview of papers and reports during the last 25 years. It also com-prises field reports obtained by its own research as well as dataprovided by other national (i.e. Peruvian) researchers. The W2013database was originally sourced from the Global Biodiversity Infor-mation Facility (GBIF, available at http://data.gbif.org). Warren et al.(2013) thoroughly checked the GBIF plant and animal databasefor location errors following the methodology of Ramirez-Villegaset al. (2012), whereby the consistency of the location data is ver-ified at both geographic (using coastal and country borders) andenvironmental (using outlier-removal tests) levels. We carefullychecked bird species names using BirdLife’s taxonomy databaseas a reference. Plant taxonomy was verified using The Plant List(http://www.theplantlist.org, see Warren et al. 2013).

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

Climate data

Current climate data were derived from WorldClim (Hijmanset al. 2005). WorldClim is a global gridded dataset of monthly

Page 3: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx 3

Fig. 1. Study area. (A) Elevation (in metres) across the Tropical Andes countries overlaid wiin 0.5◦ cells and key sites with high projected impacts (mentioned throughout the text).

Table 1List of bioclimatic variables used in the modelling.

ID Variable name Units

P1 Annual mean temperature ◦CP4 Temperature seasonality (standard

deviation)

◦C

P5 Maximum temperature of warmestmonth

◦C

P6 Minimum temperature of coldestmonth

◦C

P12 Annual precipitation mmP15 Precipitation seasonality (coefficient of

variation)%

P16 Precipitation of Wettest quarter mmP17 Precipitation of Driest quarter mm

◦ −1

cai(teUiiqiodeidae

Io Ombrothermic index mm CIod2 Ombrothermic index of the driest

2-months of the driest quartermm ◦C−1

limatological means of maximum, minimum and mean temper-ture and total precipitation developed through Thin Plate Splinenterpolation of long-term (i.e. 1950–2000) weather station recordsFig. 1A). There is a generally dense distribution of weather sta-ions across the core of our geographic analysis domain (Hijmanst al. 2005), though weather station density decreases southwards.sing the monthly WorldClim data we derived 10 ‘bioclimatic’

ndices (Busby 1991; Rivas-Martinez 2004a, 2004b) (Table 1). Thesendices describe annual and seasonal trends and allow for an ade-uate characterization of the species bioclimatic niches. These

ndices are important limiting factors for growth and developmentf species, and have been used extensively for predicting speciesistributions using presence-only data (Elith et al. 2006; Grahamt al. 2008; Warren et al. 2013). For the Andes, the 10 bioclimatic

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

ndices chosen cover aspects of both average and extreme con-itions of a year. In addition, the use of the ombrothermic indexllows for differentiating climate conditions between and acrosscosystems (Rivas-Martinez 2004a, 2004b).

th locations of weather stations in WorldClim. (B) Number of modelling occurrences

We obtained future climate projections from the CMIP3 (Cou-pled Model Inter-comparison Project phase 3) web data portal(https://esg.llnl.gov:8443/index.jsp) (Meehl et al. 2007). We down-loaded monthly time series of temperature and precipitation datafor the baseline period (20th century) and projections of future cli-mate for the 21st century for the SRES-A2 emission scenario for 24different Intergovernmental Panel on Climate Change (IPCC) cou-pled GCMs (Table 2). We chose SRES-A2 because we considered thefull-mitigation SRES-B1 unlikely, and because differences betweenSRES-A2 and SRES-A1B and SRES-A1FI by 2050s are negligible(Hawkins & Sutton 2009). Based on the availability of maximumand minimum temperature data, we further selected a subset ofnine GCMs (Table 2).

Using the complete GCM time series, for each of the GCMs,months and variables, we calculated the 30 year running aver-age over the baseline period (1961–1990) and two future periods:2020s (2010–2039) and 2050s (2040–2069), representing the earlyand mid-21st century. We then calculated the anomalies (deltas)of each GCM future scenario with respect to the baseline period(average 1961–1990 climate) for each month, variable and period.

Given the significant heterogeneity in Andean climates, coarsescale GCM grids fail to represent the diversity of niches wherespecies are distributed, hence we increased the resolution of theGCM data by means of empirical downscaling with the deltamethod (Ramirez-Villegas & Jarvis 2010). For each month, vari-able, and period, the respective set of GCM deltas was averaged(i.e. ensemble mean). Temperature anomalies were directly added,whilst precipitation anomalies were added as a relative factorto the value in WorldClim in order to avoid precipitation valuesbelow zero due to the differences between the GCM simulatedand WorldClim observed baseline. For each of the future periods,

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

we calculated the same bioclimatic indices as for current climatedata (Table 1). This yielded climate scenarios for each of the futureperiods as an average trend of the set of available GCMs on theSRES-A2 emission scenario.

Page 4: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

4 J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx

Table 2List of all and available GCMs and principal characteristics (resolutions).

Model Country Atmosphereb Oceanb A2a

BCCR-BCM2.0 Norway T63, L31 1.5 × 0.5, L35 ACCCMA-CGCM3.1 (T47) Canada T47 (3.75 × 3.75), L31 1.85 × 1.85, L29CCCMA-CGCM3.1 (T63) Canada T63 (2.8 × 2.8), L31 1.4 × 0.94, L29CNRM-CM3 France T63 (2.8 × 2.8), L45 1.875 × (0.5–2), L31CSIRO-Mk3.0 Australia T63, L18 1.875 × 0.84, L31 ACSIRO-Mk3.5 Australia T63, L18 1.875 × 0.84, L31 AGFDL-CM2.0 USA 2.5 × 2.0, L24 1.0 × (1/3–1), L50 AGFDL-CM2.1 USA 2.5 × 2.0, L24 1.0 × (1/3–1), L50 AGISS-AOM USA 4 × 3, L12 4 × 3, L16GISS-MODEL-EH USA 5 × 4, L20 5 × 4, L13GISS-MODEL-ER USA 5 × 4, L20 5 × 4, L13IAP-FGOALS1.0-G China 2.8 × 2.8, L26 1 × 1, L16INGV-ECHAM4 Italy T42, L19 2 × (0.5–2), L31INM-CM3.0 Russia 5 × 4, L21 2.5 × 2, L33 AIPSL-CM4 France 2.5 × 3.75, L19 2 × (1–2), L30MIROC3.2-HIRES Japan T106, L56 0.28 × 0.19, L47MIROC3.2-MEDRES Japan T42, L20 1.4 × (0.5–1.4), L43 AMIUB-ECHO-G Germany/Korea T30, L19 T42, L20MPI-ECHAM5 Germany T63, L32 1 × 1, L41MRI-CGCM2.3.2A Japan T42, L30 2.5 × (0.5–2.0)NCAR-CCSM3.0 USA T85L26, 1.4 × 1.4 1 × (0.27–1), L40 ANCAR-PCM1 USA T42 (2.8 × 2.8), L18 1 × (0.27–1), L40 AUKMO-HADCM3 UK 3.75 × 2.5, L19 1.25 × 1.25, L20UKMO-HADGEM1 UK 1.875 × 1.25, L38 1.25 × 1.25, L20

a A: Monthly maximum and minimum temperature available.ded. V

d solutiov

oat

S

2ewWscdPbnP

bbe2tdmo

o2euews<(

b Horizontal (T) resolution indicates number of cells in which the globe was diviivided. When a model is developed with different latitudinal and longitudinal realue.

We used the ensemble mean (rather than individual GCMs)wing to processing and storage needs, and given the consider-ble number of species being modeled and the resolution at whichhe models were projected (2.5 arc-min).

pecies distribution models (SDMs)

Species distributions were modeled using Maxent (Phillips et al.006; Phillips & Dudík 2008), a robust bioclimatic envelope mod-lling techniques (Smith et al., 2013). We modeled only speciesith at least 10 distinct locations (Ramírez-Villegas et al. 2010;isz et al. 2008), as a compromise between model quality and

ufficient coverage of limited-range species. Maxent models thelimate-constrained distribution of a species using presence-onlyata and a set of environmental descriptors (Elith et al. 2010;hillips et al. 2006). Maxent has been tested extensively and haseen found to suitably perform as a state-of-the-art modelling tech-ique both under current and future conditions (Costa et al. 2010;hillips 2008; Smith et al. 2013).

Here, we followed a similar methodology to that employedy Warren et al. (2013), whereby default features optimized toroad species groups were used to construct Maxent models forach species (Phillips 2008; Phillips et al. 2006; Phillips & Dudík008). For each species we drew 10,000 pseudo-absences fromhe countries where the species was reported (according to ouratabase). This was done to avoid over-fitting of the models whilstaintaining a good discrimination between presence and absence

f the species (Isaac et al. 2009; VanDerWal et al. 2009).Most niche modelling techniques are sensitive to the number

f predictors used and Maxent is no exception (Braunisch et al.013; Dormann 2007; Phillips 2008). Excess predictors in a Max-nt model can cause over-fitting and hence bias the responsesnder future scenarios by over-weighting certain drivers over oth-rs (Warren & Seifert 2010). Hence, following Warren et al. (2013),

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

e reduced the number of predictors in the Maxent model forpecies with low numbers of occurrences. For those species with40 unique data points, a set of six climate predictors was usedi.e. annual mean temperature [P1], temperature seasonality [P4],

ertical (L) resolution indicates the number of layers in which the atmosphere wasns, the respective cellsizes (LonxLat) in degrees are provided instead of a unique

annual total precipitation [P12], precipitation seasonality [P15],ombrothermic index [Io], and ombrothermic index of the driest 2months of the driest quarter [Iod2]), whilst for taxa with >40 uniquedata points, the complete set of 10 predictors (i.e. annual meantemperature [P1], temperature seasonality [P4], maximum tem-perature of warmest month [P5], minimum temperature of coldestmonth [P6], annual total precipitation [P12], precipitation season-ality [P15], precipitation of wettest quarter [P16], precipitation ofdriest quarter [P17], ombrothermic index [Io], and ombrothermicindex of the driest 2 months of the driest quarter [Iod2]) was used(see Table 1). This choice was a compromise between having overlycomplex Maxent models for species with low numbers of occur-rences and having overly simplistic models for species with verylarge numbers of occurrences.

Maxent models were fitted using cross-validation (10 itera-tions), each one randomly dropping 10–20% input points. We thenassessed the model skill using the Area under the ROC (ReceiverOperating Characteristic) Curve of the test data (AUCTest), cal-culated as the average AUCTest of the 10 runs. Despite knownlimitations (Lobo et al. 2008; Warren et al. 2013), AUCTest is auseful metric for selecting Maxent models of appropriate complex-ity (Warren & Seifert 2010) and is a widely used model accuracyand selection criterion (Braunisch et al. 2013; Graham et al. 2008;VanDerWal et al. 2009). The procedure applied here allowed usto discard species with models showing low predictive skill: onlymodels with 10-fold average test AUCTest ≥ 0.7 were projected ontothe future climatic periods.

We then projected the fitted models onto both the continuousWorldClim current climate surfaces and the downscaled surfacesof future climate conditions (2020s and 2050s). We then binned theprobability distributions using the ‘prevalence threshold’ (Liu et al.2005, 2013). This threshold is defined as the average probabilityover all input data points used to fit the model (i.e. training presencepoints). To reduce commission (i.e. straying too far from the actual

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

niche of a taxon) or omission (i.e. missing major species populationsdue to lack of observations), the current climate distributions ofeach species were further clipped within a 300 km buffer aroundthe respective input occurrence points (also see Warren et al. 2013).

Page 5: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

Natu

fitm2datttnWliba

A

taedsesWd

s

Taht(a

A

frps(abn(crts

R

S

ost

ARTICLENC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for

For future climatic scenarios, species distribution maps wererst binned using the prevalence threshold, and then fur-her limited using two assumptions about species’ dispersion

echanisms (Jarvis et al. 2008; Thomas et al. 2004; Thuiller et al.005): (1) no dispersal and (2) unlimited dispersal. For the noispersal scenario, the projected future distributions were notllowed to stray away from the current-climate distribution. Forhe unlimited dispersal scenario, all future suitable areas outsidehe current-climate distribution were considered of the future dis-ribution. This implies that a species can migrate and occupy anyew site that becomes suitable under future climatic conditions.e acknowledge that unlimited dispersal is unrealistic (particu-

arly for plants), but we use this scenario to illustrate the likelympacts of climate change on diversity even when the best possi-le conditions are assumed (e.g. through use of assisted migration,lso see section “Management and conservation implications”).

ssessment of climate change impacts in species assemblages

Species richness was calculated using the binned species dis-ributions as the total number of species in a given site (i.e. pixel)nd then used to calculate changes in species richness as the differ-nce between future species richness and current species richnessivided by current species richness. Additionally, we calculated thepecies turnover for the unlimited dispersal scenario (Broennimannt al. 2006). This index arises from a modification of the ‘classical’pecies turnover (beta-diversity) indicators (Lennon et al. 2001;

hittaker 1960) which are computed in geographic space using aefined spatial neighbourhood (Broennimann et al. 2006) (Eq. (1)).

pecies turnover = 100 × species gain + species lossinitial species richness + species gain

(1)

his turnover index has a lower limit of zero when the ‘species gain’nd the ‘species loss’ are zero (both of which are very unlikely toappen with a large set of species), and an upper limit of 100, whenhe whole set of species changes from one time period to the otheri.e. either the species gain or loss equals the initial species richnessnd there is no loss or gain, respectively).

ssessment of individual species responses to climate change

To estimate the sensitivity to climate change at the species levelor both migration scenarios and periods, we intersected the cur-ent and future climatic niches and calculated the climatic nicheersistence. This is defined as the percentage of area that remainsuitable in relation to the total area in the current climatic nicheLoehle & LeBlanc 1996; Peterson et al. 2001). Climatic niche lossnd gain were first calculated as the percentage area predicted toecome unsuitable or suitable respectively in the future climaticiche in relation to the total area in the current climatic nicheBroennimann et al. 2006). The species range change was then cal-ulated as the difference between climatic niche gain and loss. Thisepresents the percentage of range expansion or contraction in rela-ion to the current climatic niche for each species under the futurecenarios.

esults

pecies datasets

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

Our final modelling dataset comprised 478,301 vascular plantccurrences for 9457 species and 88,636 bird occurrences for 1555pecies (Fig. 1B). The W2013 dataset provided the greatest propor-ion of occurrences, with 93% of all locality points used, and holding

PRESSre Conservation xxx (2014) xxx–xxx 5

data for 9371 vascular plants species and 1429 birds. The databasefrom CDC-UNALM provided 4.14% of the occurrence points usedfor 186 vascular plant and 1316 bird species. CONDESAN’s datasetcontributed 2.9% of the occurrences representing 501 birds and237 vascular plants. Despite the majority of records were from theW2013, the CDC-UNALM and CONDESAN datasets provided criticaloccurrence data for rare, endemic and narrow-range species thatwere poorly (if at all) represented in the W2013 database (see e.g.Supplementary Figure S1 in Warren et al. 2013).

Performance of species distribution models

Almost half of the plant (48%) and bird species (44%) had anaverage test AUC > 0.9, suggesting a good aptitude of the mod-els to discriminate the species’ fundamental climatic niche. Theaverage test AUC of all plant species was 0.874 (median = 0.894,SD = 0.088), while that of bird species was 0.872 (median = 0.889,SD = 0.076) (Fig. 2). Cross-validated runs indicated that variabilityof AUC ranged from 0 to 13.7% for training-sets and from 0 to 38.8%for evaluation sets. Relatively unstable test statistics were foundfor species with very low number of data points (high variability inAUC across repetitions), both in training and test sets.

Maxent models performance as measured by the average AUCwas relatively similar for birds (BD) and vascular plants (VP), onaverage (Fig. 2). Average training VP AUC ranged from 0.433 to0.999, whilst test AUC varied from 0.28 to 0.999. In a few cases (<500for plants and <50 for birds) the AUC statistic fell below the 0.7threshold for model quality, probably owing to a combination of alimited number of species records and an asymmetric spatial distri-bution (i.e. high spatial autocorrelation). Less than 1% of the wholeset of plant and bird species had an AUC value equal to or worsethan random discrimination of presences and absences (AUC ≤ 0.5).All species with average test AUC below 0.7 were removed fromany further analyses (see section “species distributions models(SDMs)”). Based on a sufficiently high AUC (i.e. >0.7), a total of 9062vascular plant and 1456 bird species (95.7 and 96.6%, respectively)were used in all following analyses.

Shifts in species richness and community turnover

Current species richness ranged from 0 to 452 species for birdsand from 0 to 1535 species for vascular plants per pixel of 25 km2

(Fig. 3). The highest concentration of plants is located on the outerslopes of the Western and Eastern Andean chain, between 1500and 3000 m in altitude, primarily in the Andes of Colombia, Ecuadorand Venezuela, on the inner slopes of the Central Chain of Colombia(upper Magdalena river basin) and in the montane forests along theEastern ridge between Peru and Bolivia (Fig. 3A). Diversity of birdsis particularly high throughout the Peruvian Andes, in the mon-tane forests along the Eastern ridge (Range = 141–452), and in themontane forests of the north-Western chain of Ecuador (Fig. 3B).

Patterns of changes in species richness show important dif-ferences depending on the dispersal thresholds and the periodanalyzed (2020 or 2050, Fig. 4). The unlimited dispersal scenarioprojects an upslope migration of both plant and bird species sug-gesting important changes in the configuration of the diversitypatterns of Andean biota. On the other hand, the no-dispersal sce-narios show a significant reduction in species richness for bothplant and bird species with major changes by 2050. The maxi-mum richness values in the no dispersal scenario by 2050 periodare 1244 for plant species (mean = 163 ± 178) and 295 for birds(mean = 29 ± 36) per 25 km2 pixel (Fig. 4). Areas showing the largest

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

decreases in species richness are located along the montane forestsof the Eastern Andes of Bolivia and Peru between 500 and 1200 m,on the outer slopes of the Eastern Andean foothills in Colombia andEcuador, and on the Pacific slope of Northern Ecuador and southern

Page 6: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

6 J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx

F AUC) f( of 10

Cht

w

Fc

ig. 2. Evaluation of Maxent models. Distribution of the area under the ROC curve (grey bars) and test (black bars) sets. AUC values of individual species are averages

olombia (Fig. 4). Conversely, the areas with minor changes are the

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

ighlands of Peru and Bolivia (Altiplano) and the pacific slope ofhe Peruvian Andes.

Negative changes in species richness are also observed evenhen unlimited dispersal is considered. Loss of diversity is

ig. 3. Modeled current species richness for (A) vascular plants and (B) birds in the Tropiounts of species occurring in a 25 km2 pixel.

or (A). All vascular plants. (B) All birds. Training AUC values are plotted for trainingcross-validated runs with 10–20% of the input points drawn randomly.

observed from north to south of the Andes, although some par-

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

ticular areas are worthy of more attention; areas below altitudesof 1500 m in the east Peruvian Andean mountains (i.e. central andeastern Huanuco, Pasco and Junin) seem to be severely impacted(>60% loss in species richness), and the same pattern is observed

cal Andes as derived by the sum of binned species distributions models. Values are

Page 7: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx 7

Fig. 4. Spatial patterns of changes in species richness for birds and vascular plants under both migration scenarios and time periods. Values are percentage change in speciesrichness from the present-day value shown in Fig. 3. (For interpretation of the references to color in text, the reader is referred to the web version of the article.)

iCts

taSSt

I

spfodasoapbtb(cu

n the border between Ecuador and Peru, and in Narino, Valle delauca, and Putumayo in Colombia. These changes may be attributedo the eastern margins of the mountain chain being less climaticallyuitable in warmer climates.

The projected changes in community turnover are concentratedo a large extent in the High Andes of Bolivia and Peru, as wells in the foothills of the Sierra de la Macarena, Sierra Nevada deanta Marta and around the Magdalena river basin in Colombia.ignificant shifts are also evident in the Venezuelan Andes alonghe Merida chain (Fig. 5).

ndividual species responses

Increases are projected in average climatic niche size for allpecies under the unlimited dispersal assumptions for the 2020seriod (Fig. 6A). As expected, more severe impacts are projectedor the 2050s, and this is reflected in a less pronounced increasef range size in the unlimited dispersal scenario and a strongerecrease in the non-dispersal scenario (Fig. 6A and B). Consideringn unlimited dispersal scenario, the rates of climatic niche expan-ion seem to be high, with most of the species being highly favouredr barely affected by climate change if migration in fact occursnd other non-abiotic factors remain stable (e.g. land-use patterns,ests and diseases), particularly for birds. Some 45% (n = 655) ofird and 41% (n = 3715) of vascular plant species modeled are likelyo experience an increase in their climatic niches of 100% or more

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

y 2050s (Fig. 6A). By contrast, only a limited proportion of species<10%) is expected to experience no increase or a net loss in theirlimatic niche size. Our estimates indicate that even assumingnlimited dispersal some species are expected to undergo range

contraction (even to the extent of extinction), thus highlightingspecific sensitivities to climate change.

In a no dispersal scenario, the differences between periodsbecome more evident (Fig. 6B). Whilst by 2020s the maximumchanges in range size are reductions of 50% and 80% for birds andvascular plants, respectively, by the 2050s, species within bothgroups are projected to experience 100% range reduction, indicat-ing likely extinctions for a vast number of species.

To illustrate species-specific responses under future climate,we further selected and analyzed two contrasting genera for eachspecies group (plants and birds). These genera were selectedbecause they are of relatively recent origin (during the Pleistocene,ca. 1–3 million years ago), include species that are endemic to theAndes, and are classified vulnerable or critically endangered byIUCN (Tables 3 and 4). Many of the species of the genera Grallariaand Eriocnemis (class: Aves) are projected to expand their nicheby more than 100% if dispersal was assumed. In particular, thespecies E. cupreoventris and E. nigrivestis were found to increasetheir niche considerably by 2020 and 2050. In the case of no-dispersal, however, these species depict range contractions of 69and 65% (respectively) by 2050. Similar responses were found formost species of the genus Grallaria, notably G. alleni, G. haplonota,G. gigantea, and G. hypoleuca, for which range contractions of 59,83, 54, and 63% are projected by 2050s (no dispersal), respectively(Table 3).

Similar responses are reported for the plant genera Polylepis and

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

Gynoxis. Species such as P. lanuginosa and P. tomentella showedsignificant increases in range size in both future scenarios (unlim-ited migration), but rather large decreases in range size underno-migration assumptions. By contrast, some species of these

Page 8: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

8 J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx

Fig. 5. Species turnover for birds and vascular plants, for both periods. Community turnover can only be calculated for scenarios that somehow assume migration as thiscalculation requires that species can move to more suitable environments whenever possible. Values are percentages of change in community turnover as calculated by Eq.( detail

gr(usu

1) (see section “Assessment of climate change impactsin species assemblages” for

enera (e.g. P. incana, P. reticulata, G. buxifolia, and G. caracensis)eport range contractions for both dispersal scenarios and periods

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

Table 4). These species that respond negatively even under whennlimited dispersal is allowed can be considered of very high sen-itivity, and perhaps also be prioritized for further research tonderstand such sensitivities.

s).

Discussion

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

Changes in species distribution patterns

Our results suggest that impacts of climate change over theAndean biota could be extremely severe. This finding is in

Page 9: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx 9

F white

f ve been

aSa(Awtmodsa

TC

S

A

ig. 6. Climate change impacts on individual species. Change in range size for birds (or the SRES-A2 emission scenario and both periods (2020s and 2050s) (outliers ha

= 1456 and n = 9062 for birds and vascular plants, respectively.

greement with previous studies for the Andean region (Feeley &ilman 2010; Feeley et al. 2011; Tovar et al. 2013), other tropicalreas (Hole et al. 2009; Miles et al. 2004; Still et al. 1999), or globallyWarren et al. 2013). The effects of climate change on the Tropicalndes can be synthesized at two different levels: the extent of thehole Tropical Andes (regional level), and at the species level. At

he regional level, the inner and outer Andean foothills (800–1500etres) are likely to be the most affected due to a high amount

f species loss. In addition, the spatial patterns of species turnover

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

emonstrate a bimodal response. First, an upslope shift of severalpecies from mid elevations to the high Andes is expected. Second,

large west and southward displacement of species from the upper

able 3hange in distributional range for the Andean bird genera Eriocnemis and Grallaria.

Species IUCN 2010 categorya Endemic to Andesb

Eriocnemis alinae LC – 23Eriocnemis cupreoventris NT – 19Eriocnemis derbyi NT – 25Eriocnemis luciani LC – 28Eriocnemis mosquera LC – 12Eriocnemis nigrivestis CR EC 17Eriocnemis vestita LC – 28Grallaria alleni VU B1a + b(i,ii,iii) – 18Grallaria erythroleuca LC PE 21Grallaria flavotincta LC 13Grallaria gigantea VU B1a + b(i,ii,iii) – 12Grallaria guatimalensis LC – 2Grallaria haplonota LC – 7Grallaria hypoleuca LC – 14Grallaria nuchalis LC – 19Grallaria quitensis LC – 22Grallaria ruficapilla LC – 12Grallaria rufocinerea VU B1a + b(i,ii,iii) – 22Grallaria rufula LC – 23Grallaria squamigera LC – 20Grallaria watkinsi LC – 6

pecies in bold depict range contractions (either by 2020 or 2050) regardless of migratioa Status of the species according to the IUCN red list of threatened species: LC, least concedditional criteria as in http://www.iucnredlist.org/static/categories criteria 3 1b Country where endemic, if endemic to the Andes. EC, Ecuador; PE, Peru; BO, Bolivia.c Range change under different periods and for two dispersal scenarios. Full, unlimited

bars) and vascular plants (grey bars) for (A) unlimited dispersal and (B) no dispersal,n removed from the plot for easier visualization). Box plots were constructed with

areas of the northern portion of the study area (i.e. Merida, Perijáand Santa Marta) towards lower latitudes and a significant climaticniche reduction of mountain-top endemics is also projected.

The areas that would be most affected by high absolute speciesturnover rates and the subsequent change in the composition ofcommunities are the montane dry forest, the Santa Marta massif,the Mérida ridge, the inner slopes of the Central and Eastern ridgesof the Colombian Andes and the Altiplano of Peru and Bolivia(>3800 m).

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

At the species level, the biophysical impacts of exposure to cli-mate change are projected to be highly variable. In this study, thetwo contrasting dispersal scenarios show extremes of a spectrum

Elevation range (m) Range change (%)c

2020 2050

Full Null Full Null

00–2800 −16.8 −23.3 −32.6 −37.050–3000 149.4 −44.8 101.2 −68.600–3600 −31.3 −45.3 18.0 −48.300–3800 41.8 −13.6 −9.4 −30.300–3600 −17.8 −20.8 −34.0 −37.900–3500 261.4 −30.0 92.1 −65.000–3500 8.4 −29.7 −1.7 −52.000–2500 46.7 −31.5 3.1 −59.150–3000 38.8 −21.6 −12.5 −46.900–2350 50.9 −16.7 −8.4 −47.600–2600 >500 −26.1 >500 −54.000–3000 10.0 −31.3 2.4 −50.800–2000 11.0 −55.1 −18.9 −82.700–2300 170.3 −12.7 71.1 −63.000–3150 73.1 −10.3 25.9 −36.400–4500 −8.4 −38.2 −48.5 −66.600–3600 28.6 −15.5 18.1 −35.000–3150 11.1 −31.1 60.7 −42.200–3650 30.8 −25.9 10.4 −52.900–3800 5.6 −26.0 −21.8 −50.700–1700 43.6 −20.8 33.7 −49.9

n assumptions.rn; NT; near-threatened; VU, vulnerable; EN, endangered; CR, critically endangered.

dispersal; Null, no dispersal.

Page 10: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ARTICLE IN PRESSG ModelJNC-25349; No. of Pages 14

10 J. Ramirez-Villegas et al. / Journal for Nature Conservation xxx (2014) xxx–xxx

Table 4Change in distributional range for the Andean plant genera Gynoxis and Polylepis.

Species IUCN 2010 categorya Endemic to Andesb Elevation range (m) Range change (%)c

2020 2050

Full Null Full Null

Gynoxis acostae LC EC 2700–4300 >500 −36.8 >500 −84.0Gynoxis asterotricha n/a – 3100–4100 >500 −21.0 >500 −65.5Gynoxis baccharoides VU D(ii) – 3300–4200 233.3 −41.4 109.6 −69.2Gynoxis buxifolia n/a – 2500–4100 −12.8 −21.9 −52.1 −56.9Gynoxis caracensis LC PE 2800–4335 −13.3 −69.0 −39.6 −81.3Gynoxis cuicochensis NT EC 2500–4050 90.9 −21.7 53.8 −39.3Gynoxis fuliginosa n/a – 2700–4150 −7.3 −26.7 −35.1 −52.6Gynoxis hallii LC EC 2500–4100 266.4 −17.7 198.6 −39.6Gynoxis miniphylla NT EC 3100–4000 223.8 −36.6 44.6 −64.4Gynoxis oleifolia LC PE 3380–4900 −58.8 −81.6 −90.1 −94.5Gynoxis parvifolia n/a – 2900–4100 >500 −22.1 >500 −42.5Gynoxis psilophylla n/a BO 2800–3900 >500 −7.6 >500 −14.6Gynoxis reinaldii n/a – 2400–3300 165.2 −44.9 226.1 −64.5Gynoxis sodiroi VU B1ab(iii) EC 2900–4286 55.5 −15.8 21.4 −37.6Polylepis incana No – 2450–3800 −39.1 −64.8 −55.8 −83.3Polylepis lanuginosa VU B1abIII EC 2600–3630 >500 −26.1 >500 −49.1Polylepis pauta No – 2700–4200 8.3 −59.7 −61.1 −87.5Polylepis reticulata VU A4c EC 3200–4450 −28.9 −52.3 −31.3 −81.3Polylepis sericea No – 2500–3900 −39.1 −63.6 −52.6 −83.8Polylepis besseri No – 2500–4100 12.8 −24.5 8.4 −32.4Polylepis racemosa No – 2900–4500 23.8 −16.4 30.2 −31.5Polylepis tomentella No – 2800–4700 71.9 −7.2 59.0 −16.2Polylepis weberbaueri No – 2700–4800 −38.0 −60.3 −46.7 −73.0

Species in bold depict range contractions (either by 2020 or 2050) regardless of migration assumptions.a Status of the species according to the IUCN red list of threatened species: LC, least concern; NT, near-threatened; VU, vulnerable; EN, endangered; CR, critically endangered.

Alivia.mited

ol(iabbpp(esedommd

AJatMdftac

saecl

dditional criteria as in http://www.iucnredlist.org/static/categories criteria 3 1.b Country where endemic, if endemic to the Andes. EC, Ecuador; PE, Peru; BO, Boc Range change under different periods and for two dispersal scenarios. Full, unli

f projected responses by species to climate change. For plants, it isikely that the true response lies nearer the no-dispersion scenariosee also Feeley et al. 2011), whereas for birds the response mayn some cases resemble that of the full-dispersion scenario. Over-ll, we report that plant species may be more negatively affected inoth magnitude and direction of range change impacts than birds inoth periods. The same pattern holds for both migration scenarios,robably due to a greater proportion of endemic and narrow-rangelant species and/or the presence of isolated (meta) populationsFig. 6A) (also see Ramirez-Villegas et al. 2012), and perhaps to somextent also due to incompleteness of samples for some species. Yetpecies interactions might have a prominent role in this point. Forxample, species interactions can slow climate tracking and pro-uce more extinctions than predicted by climatic niche modelsnly (Urban et al. 2013); or on the contrary, broad-ranging animalsight transport seeds enabling long-distance dispersal, as docu-ented before during the last de-glaciation period, in which trees

ispersed at rates of 100–1000 m year−1 (Clark 1998).The projected alteration of the spatial distribution patterns of

ndean assemblages (Feeley & Silman 2010; Feeley et al. 2011;etz et al. 2007) suggest the appearance of novel communitiesdapted to non-analogous climatic conditions, which could affecthe functioning of Andean ecosystems (Williams & Jackson 2007).

any shrubby and epiphyte species (e.g. Solanaceae, Bromeliaceae)epend on their specialized symbiotic interactions with animalsor seed dispersion and pollination. Climate change effects onhese organisms could cause spatial, temporal, or physiologicalsynchronies between mutualistic species, producing changes inommunity composition and structure (Zavaleta et al. 2003).

Our estimates are thus useful in gauging general trends and pos-ible impacts, although it is very likely that individual responses

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

t the species or community level will be determined by species’cological traits (i.e. dispersal capacity), species interactions (i.e.ompetition) and/or by their physiological response to stresses,eading (in some cases) to different outcomes. If species are

dispersal; Null, no dispersal.

sufficiently mobile they may be able to track the geographic dis-placement of their climatic niches, or if species are capable of rapidevolutionary change or have a wide range of abiotic tolerances,they may adjust to changing ecological conditions and landscapes(Broennimann et al. 2006). According to Travis (2003) and Opdamand Wascher (2004), the exact nature of a species’ response to dif-ferent rates of climate change depends upon colonization abilityand how much of a generalist the species is. For species with lowercolonization ability and for specialist species, the threshold occursat a lower climate change signal. In a human dominated world,however, natural or semi-natural ecosystems are embedded intracts of unsuitable landscape, and populations of species restrictedto those habitat types are spatially dissected. By consequence, whatis ascribed as a shifting species range is in fact the complex resultof extinction of (meta) populations at the warm range limit (thatsurpasses thresholds of species adaptability), and colonization andgrowth of (meta) populations into regions that newly came withinthe cold range limit (that enters the range of species adaptabil-ity). Hence, for understanding the potential risks of climate changeto a species, we must consider the dynamics of the populationsconstituting the geographical range in connection to the spatialfeatures of the landscapes across the range (also see section “Man-agement and conservation implications”). Human land-use may beespecially important in the Andes where anthropogenic activitiesabove tree line and in the piedmont may create a hard barrier toupward migrations, imperilling Andean biodiversity (Feeley 2010;Feeley et al. 2011); therefore, the incorporation of a coupled modelthat integrates climate change scenarios together with land coverchange dynamics is a priority task to analyze specific responses ofthe Andean biota to these drivers of change.

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

Species extinction risks

Climatic fluctuations during the Pliocene-Pleistocene periodstrongly influenced the origin and spatial arrangement of the

Page 11: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

Natu

mectttfaejTmc

trnplwtopeaewan(aotak

M

ac(Ap1a

srohartvtidummspcs

ARTICLENC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for

ajority of Andean species used in this study (Luteyn 2002; Youngt al. 2002; García-Moreno et al. 1999). During periods of intenselimatic change in the Pleistocene, epiphyte-laden evergreen vege-ation remained only where conditions remained stable, suggestinghat ecologically stable areas may have existed during the glacia-ions as small pockets within surrounding drier pieces of montaneorest (Arctander & Fjeldså 1997; Fjeldså 1995; Roy et al. 1997). As

consequence, many of these surviving species present in thesecosystems are endemic, with narrow habitat tolerances in con-unction with a restricted distribution range (Kattan et al. 2004).hese patterns and conditions constitute a perfect scenario to pro-ote higher rates of species loss and turnover under projected

limate anomalies such as those projected in the present study.In this context, reductions in the size of the climatic niche such as

hose herein projected imply that a number of species may becomeestricted to a few sites. Species with small range sizes are vul-erable to smaller stochastic events as these could affect a largerroportion of the species’ total population, especially in fragmented

andscapes (With & King 1999). As a result of this, extinction risksill likely intensify for a large portion of the taxa analyzed here, par-

icularly at long lead times (2050s in this study). Our study, as manythers, assumes that species will die out within regions that areredicted to become climatically unsuitable for them (Ohlemüllert al. 2006), and takes no account of species- or population-leveldaptive responses that may reduce negative effects (see e.g. Hartet al. 2004). Despite that, our results may be conservative given thate (1) did not include habitat loss data for the Tropical Andes in the

nalysis (Leisher et al. 2013; Ramirez-Villegas et al. 2012), (2) didot consider potential impacts of changing interannual variabilitye.g. frequency or intensity of drought or heat waves) in our models,nd (3) did not model any secondary effects such as pests, diseasesr important species-level interactions required for survival. Fur-hermore, the rather low generation times of many vascular plantsnd some bird species will probably preclude adaptation rates fromeeping pace with human induced climate change.

anagement and conservation implications

In conservation planning, irreplaceability (commonly measureds singularity) and vulnerability (measured through threat pro-esses) are among the most important dimensions to analyzeBrooks et al. 2006). Several authors have depicted the Tropicalndes as being within the most vulnerable regions with high irre-laceability (Brooks et al. 2006; Kattan et al. 2004; Mittermeier et al.997), placing the region extremely important for conservationction.

The question of whether the current protected area system isufficient given the challenges of climate change is a critical one. Aegional analysis by Ramirez-Villegas et al. (2012) showed that 8ut of 16 conservation areas in South America are in the Andeanighlands. According to the present study, negatively impactedreas (orange to red areas in Fig. 4) could lose up to 60% of speciesichness and suffer up to 100% changes in community makeup,hus, affecting ecosystem functioning as well as ecosystem ser-ices to human society (Gamfeldt et al. 2008). There is no questionhat these projected impacts will affect conservation planning dur-ng the 21st century, and hence further research should focus oneveloping a better understanding of conservation effectivenessnder future climates for the Andes (Araújo et al. 2004). Tropicalountain systems such as the Andes are highly variable in cli-ate, and therefore, offer a wide range of adaptation pathways for

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

pecies, further increasing their value for conservation. The hereinrojected changes in range sizes, species richness and communityomposition are useful metrics in evaluating tools for conservation,uch as for adjusting extinction risk assessments, delimitation of

PRESSre Conservation xxx (2014) xxx–xxx 11

priority conservation areas and conservation targets within pro-tected areas.

Using these results to identify priority areas at a medium tolarge scale could be particularly useful, given that diversity can-not always be easily captured in a single site-specific targeting ofconservation in the Andes, requiring instead, conservation actionsspread throughout entire biomes (Fjeldså et al. 2005; Ramirez-Villegas et al. 2012). In this context, based on Opdam and Wascher(2004) we propose three major components for a conservationstrategy in a warmer Tropical Andes. Firstly, a focus on landscapeconditions for biodiversity, where populations potentially canrespond to large-scale changes and disturbances. These conditionsshould allow populations to respond to large-scale disturbances. Ifspecies distributions patterns change more dynamically in spaceand time, local conservation management for single species willbe less effective. Secondly, we propose to shift in strategy fromprotected areas towards landscape networks including protectedareas, connecting zones and intermediate landscapes. Thirdly, wepropose a shift from a defensive conservation strategy towards alandscape development strategy. A static approach of establishingisolated reserves surrounded by a highly unnatural landscape is notan effective strategy under a climate change scenario. Given theintense land use changes in the Andes, the sensitivity of Andeanspecies to climatic changes, and the fact we are globally alreadycommitted to at least +2 ◦C warming, we must accept that conser-vation of biodiversity is only effective if we dynamically integrate itin the development of the entire landscape, based on coalitions withother functions such as the identification of key areas for provisionof ecosystem services, heterogeneity, and landscape permeability(Brooks et al. 2006).

Regional policy and planning should aim at improving landscapeconnectivity. Amongst the most evident conservation planningstrategies is the establishment of reserves. Particularly under cli-mate change, the inclusion of new areas seems to be a relevant,albeit challenging, task (Hannah et al. 2007). Land tenure issues,poverty, development gaps between rural and urban areas, thedemand for natural resources, and an economic model orientedtowards extraction (e.g. mining) make the establishment of newconservation areas difficult in the Andes. In the absence of suchpossibilities, the appropriate articulation of national reserves withother conservation sub-systems such as protective forests, indige-nous territories, civil society reserves, and sub-national protectedareas could be an appropriate mechanism of action. In addition, sig-nificant attention should be paid to the design (or adjustment) ofthe Andean protected area system. We recommend the followingcriteria be taken into account:

• Maintain the connectivity across the elevation, moisture andedaphic gradient (Killeen & Solórzano 2008). These gradientsare critical for maintaining beta diversity and response capacity(Thuiller et al. 2008).

• Incorporate ecotone diversity in the design of conservationareas. The landscapes within these areas are characterized byhabitat mosaics that reflect differences in soil humidity, pro-ductivity, among others. These mosaics are occupied by speciesassembled in communities that reflect the presence of micro-environmental constraints in an area where climate stress isthe overriding macro-environmental characteristic. These popu-lations may have genetic traits distinct from core populationspre-adapting them to the physiological stress of climate change(Killeen & Solórzano 2008). In the Tropical Andes the preserva-

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

tion of the ecotone between the montane forest and grasslandsecosystems is a fundamental adaptation measure to buffer themassive upward displacement of species ranges in response toincreased warming (Feeley et al. 2011b).

Page 12: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

1 r Natu

Ibtaaetgotph

F

pm2Atbpe2mndmaliwe

A

IiaiMmpATaaPA

R

A

A

ARTICLENC-25349; No. of Pages 14

2 J. Ramirez-Villegas et al. / Journal fo

The identification of climatically stable areas as potential biolog-ical refugia through bioclimatic envelope model (see e.g. darkgreen areas in Fig. 4 combined with dark areas in Fig. 3) whichcould act as connectors and/or corridors between current andfuture areas of high biodiversity (Vos et al. 2008).

mprovement of landscape connectivity through the creation ofiological corridors is probably the most frequent recommenda-ion in the scientific literature (Heller & Zavaleta 2009). We suggestn optimization of spatial configuration of such corridors and anssessment of the risks of these turning into channels for dis-ase transmission and/or movement of invasive species. In additiono these, a better land use planning through better and targetedovernment-level policies is warranted in order to reduce the risksf deforestation, loss of pollination services and genetic erosion inhe agricultural frontier, while at the same time bolstering the dis-ersion and population breeding between (and within) remainingabitat patches (Opdam & Wascher 2004).

inal remarks

Several sources of uncertainty may influence the results werovide here. These include the primary biodiversity data, the cli-ate data and the climate envelope modelling (Braunisch et al.

013; Pearson et al. 2006; Ramirez-Villegas & Challinor 2012).lthough these uncertainties are carried into the analysis, we argue

hat our results provide important insight on a globally importantiodiversity hotspot. Importantly, our results agree and partly com-lement with previous regional and global studies (see Warrent al. 2013; Still et al. 1999; Thomas et al. 2004; Feeley & Silman010). Improvement to our modelling approach for future studiesay be warranted through achieving better spatial representative-

ess of both species and climate observations, the use of abundanceata (in addition to presence-only data), better constraining speciesigration patterns, the inclusion of changes interannual variability

nd their effects on species distributions, the use of higher reso-ution climate models that resolve local climatic change patternsn a more detailed manner, information on species interactions as

ell as a detailed assessment of relevant local processes drivingxtinctions.

cknowledgments

The authors thank Héctor Tobón and Daniel Amariles, from thenternational Center for Tropical Agriculture (CIAT) for their helpn programming the automated data cleansing algorithms. Authorslso thank Johannes Signer, from the International Center for Trop-cal Agriculture (CIAT) for his help in some of the processing, and

aría Teresa Becerra and Wouter Buytaert for their useful com-ents and improvement on earlier versions of this manuscript. This

roject was funded by the Andean Regional Program of the Spanishgency for International Cooperation and Development (AECID),he Mountain Partnership and the Swiss Agency for Developmentnd Cooperation (SDC) through their regional program, ECOBONAnd CIMA. JRV and AJ were partly supported by the CGIAR Researchrogram on Climate Change, Agriculture and Food Security (CCAFS).uthors thank two anonymous reviewers for their comments.

eferences

raújo, M. B., Cabeza, M., Thuiller, W., Hannah, L., & Williams, P. H. (2004). Would

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

climate change drive species out of reserves? An assessment of existing reserve-selection methods. Global Change Biology, 10, 1618–1626.

rctander, P., & Fjeldså, J. (1997). Andean Tapaculos of the Genus Scytalopus (Rhic-ocryptidae): A study of speciation using DNA sequence data. Basel: BirkhauserVerlag.

PRESSre Conservation xxx (2014) xxx–xxx

Armenteras, D., Rodríguez, N., Retana, J., & Morales, M. (2011). Understanding defor-estation in montane and lowland forests of the Colombian Andes. RegionalEnvironmental Change, 11, 693–705.

Bass, M. S., Finer, M., Jenkins, C. N., Kreft, H., Cisneros-Heredia, D. F., McCracken, S.F., et al. (2010). Global conservation significance of Ecuador’s Yasuní NationalPark. PLoS ONE, 5, e8767.

Beaumont, L. J., Pitman, A., Perkins, S., Zimmermann, N. E., Yoccoz, N. G., & Thuiller, W.(2011). Impacts of climate change on the world’s most exceptional ecoregions.Proceedings of the National Academy of Sciences, 108(6), 2306–2311.

Biesmeijer, J. C., Roberts, S. P. M., Reemer, M., Ohlemüller, R., Edwards, M., Peeters,T., et al. (2006). Parallel declines in pollinators and insect-pollinated plants inBritain and the Netherlands. Science, 313, 351–354.

Braunisch, V., Coppes, J., Arlettaz, R., Suchant, R., Schmid, H., & Bollmann, K. (2013).Selecting from correlated climate variables: A major source of uncertainty forpredicting species distributions under climate change. Ecography,

Broennimann, O., Thuiller, W., Hughes, G., Midgley, G. F., Alkemade, J. M. R., & Guisan,A. (2006). Do geographic distribution, niche property and life form explainplants’ vulnerability to global change? Global Change Biology, 12, 1079–1093.

Brooks, T. M., Mittermeier, R. A., da Fonseca, G. A. B., Gerlach, J., Hoffmann, M., Lam-oreux, J. F., et al. (2006). Global biodiversity conservation priorities. Science, 313,58–61.

Bruinsma, J. (2003). World agriculture: Towards 2015/2030; an FAO perspective. Lon-don: Earthscan Publications Ltd.

Busby, J. R. (1991). BIOCLIM—A bioclimate analysis and prediction system. PlantProtection Quarterly, 6, 8–9.

Butchart, S. H. M., Walpole, M., Collen, B., van Strien, A., Scharlemann, J. P. W.,Almond, R. E. A., et al. (2010). Global biodiversity: Indicators of recent declines.Science, 328, 1164–1168.

Casares, C., Boyla, K., & Davidson, I. (2003). Threatened birds of the America –The ICBP/UICN Red Data Book. Washington and London: Smithsonian Institu-tion Press. In cooperation with the International Council for Bird Preservation,Cambridge, UK (1150 pp.).

CBD. (2007). Cross-roads of life on earth – Exploring means to meet the 2010 bio-diversity target. In Solution-oriented scenarios for Global Biodiversity Outlook 2Montreal, Canada, (90 pp.).

Chen, I.-C., Shiu, H.-J., Benedick, S., Holloway, J. D., Chey, V. K., Barlow, H. S., et al.(2009). Elevation increases in moth assemblages over 42 years on a tropicalmountain. Proceedings of the National Academy of Sciences, 106, 1479–1483.

Clapperton, C. M. (1993). Quaternary geology and geomorphology of South America.Amsterdam and New York: Elsevier.

Clark, J. S. (1998). Why trees migrate so fast: Confronting theory with dispersalbiology and the paleorecord. American Naturalist, 152, 204–224.

Costa, G., Nogueira, C., Machado, R., & Colli, G. (2010). Sampling bias and the useof ecological niche modeling in conservation planning: A field evaluation in abiodiversity hotspot. Biodiversity and Conservation, 19, 883–899.

Dormann, C. F. (2007). Promising the future? Global change projections of speciesdistributions. Basic and Applied Ecology, 8, 387–397.

Elith, J., Graham, C., Anderson, R., Dudík, M., Ferrier, S., Guisan, A., et al. (2006).Novel methods improve prediction of species distributions from occurrencedata. Ecography, 29, 129–151.

Elith, J., Phillips, S. J., Hastie, T., Dudík, M., Chee, Y. E., & Yates, C. J. (2010). A statisticalexplanation of MaxEnt for ecologists. Diversity and Distributions, 17, 43–57.

Farr, T. G., Rosen, P. A., Caro, E., Crippen, R., Duren, R., Hensley, S., et al.(2007). The shuttle radar topography mission. Reviews of geophysics, 45(2)http://dx.doi.org/10.1029/2005RG000183

Feeley, K. J. (2010). The conservation value of secondary forests for tropical nocturnalbird species. Animal Conservation, 13, 16–18.

Feeley, K. J., & Silman, M. R. (2010). Land-use and climate change effects on pop-ulation size and extinction risk of Andean plants. Global Change Biology, 16,3215–3222.

Feeley, K. J., Silman, M. R., Bush, M. B., Farfan, W., Garcia Cabrera, K., Malhi, Y.,et al. (2011). Upslope migration of andean trees. Journal of Biogeography, 38,783–791.

Fjeldså, J. (1995). Geographical patterns of neoendemic and older relict species of Andeanforest birds: The significance of ecologically stable areas. New York, USA.

Fjeldså, J., Lambin, E., & Mertens, B. (1999). Correlation between endemism and localecoclimatic stability documented by comparing Andean bird distributions andremotely sensed land surface data. Ecography, 22, 63–78.

Fjeldså, J., Álvarez, M. D., Lazcano, J. M., & León, B. (2005). Illicit crops and armedconflict as constraints on biodiversity conservation in the Andes region. AMBIO:A Journal of the Human Environment, 34, 205–211.

Fjeldså, J., & Krabbe, N. (1990). Birds of the high Andes: A manual to the birds of thetemperate zone of the Andes and Patagonia, South America. Copenhagen, Denmark:Apollo Books (880 pp.).

Fuhrer, J. (2003). Agroecosystem responses to combinations of elevated CO2, ozone,and global climate change. Agriculture, Ecosystems & Environment, 97, 1–20.

Gamfeldt, L., Hillebrand, H., & Jonsson, P. R. (2008). Multiple functions increasethe importance of biodiversity for overall ecosystem functioning. Ecology, 89,1223–1231.

García-Moreno, J., Arctander, P., & Fjeldså, J. (1999). Strong diversification at thetreeline among Metallura Hummingbirds. Auk, 116, 702–711.

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

Missouri Botanical Garden (2004) VAScular Tropicos (VAST) nomenclaturaldatabase, available at http://www.tropicos.org. Missouri Botanical Garden, SaintLouis, Missouri, USA.

Gentry, A. H. (1995). Patterns of diversity and floristic composition in Neotropical mon-tane forests. Bronx: New York Botanical Garden.

Page 13: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

Natu

G

H

H

H

H

H

H

I

I

J

J

K

K

K

K

L

L

L

L

L

L

L

L

M

M

M

M

M

M

O

O

O

P

ARTICLENC-25349; No. of Pages 14

J. Ramirez-Villegas et al. / Journal for

raham, C. H., Elith, J., Hijmans, R. J., Guisan, A., Townsend Peterson, A., Loiselle, B.A., et al. (2008). The influence of spatial errors in species occurrence data usedin distribution models. Journal of Applied Ecology, 45, 239–247.

annah, L., Midgley, G., Andelman, S., Araújo, M., Hughes, G., Martinez-Meyer, E.,et al. (2007). Protected area needs in a changing climate. Frontiers in Ecology andthe Environment, 5, 131–138.

arte, J., Ostling, A., Green, J. L., & Kinzig, A. (2004). Biodiversity conservation: Cli-mate change and extinction risk. Nature, 430.

awkins, E., & Sutton, R. (2009). The potential to narrow uncertainty inregional climate predictions. Bulletin of the American Meteorological Society, 90,1095–1107.

eller, N. E., & Zavaleta, E. S. (2009). Biodiversity management in the face of climatechange: A review of 22 years of recommendations. Biological Conservation, 142,14–32.

ijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G., & Jarvis, A. (2005). Very high res-olution interpolated climate surfaces for global land areas. International Journalof Climatology, 25, 1965–1978.

ole, D. G., Willis, S. G., Pain, D. J., Fishpool, L. D., Butchart, S. H. M., Collingham, Y. C.,et al. (2009). Projected impacts of climate change on a continent-wide protectedarea network. Ecology Letters, 12, 420–431.

PCC. (2007). IPCC fourth assessment report: Climate change 2007 (AR4). Geneva,Switzerland: IPCC.

saac, J. L., Vanderwal, J., Johnson, C. N., & Williams, S. E. (2009). Resistance andresilience: Quantifying relative extinction risk in a diverse assemblage of Aus-tralian tropical rainforest vertebrates. Diversity and Distributions, 15, 280–288.

arvis, A., Lane, A., & Hijmans, R. J. (2008). The effect of climate change on crop wildrelatives. Agriculture, Ecosystems & Environment, 126, 13–23.

etz, W., Wilcove, D. S., & Dobson, A. P. (2007). Projected impacts of climate andland-use change on the global diversity of birds. PLoS Biology, 5, e157.

attan, G. H., Franco, P., Rojas, V., & Morales, G. (2004). Biological diversification ina complex region: A spatial analysis of faunistic diversity and biogeography ofthe Andes of Colombia. Journal of Biogeography, 31, 1829–1839.

illeen, T. J., & Solórzano, L. A. (2008). Conservation strategies to mitigate impactsfrom climate change in Amazonia. Philosophical Transactions of the Royal SocietyB: Biological Sciences, 363, 1881–1888.

im, K., & Byrne, L. (2006). Biodiversity loss and the taxonomic bottleneck: Emergingbiodiversity science. Ecological Research, 21, 794–810.

nutti, R., & Sedlacek, J. (2013). Robustness and uncertainties in the new CMIP5climate model projections. Nature Climate Change, 3, 369–373.

amoreux, J. F., Morrison, J. C., Ricketts, T. H., Olson, D. M., Dinerstein, E., McKnight,M. W., et al. (2006). Global tests of biodiversity concordance and the importanceof endemism. Nature, 440, 212–214.

eisher, C., Touval, J., Hess, S., Boucher, T., & Reymondin, L. (2013). Land and forestdegradation inside protected areas in Latin America. Diversity, 5, 779–795.

ennon, J. J., Koleff, P., Greenwood, J. J. D., & Gaston, K. J. (2001). The geographi-cal structure of British bird distributions: Diversity, spatial turnover and scale.Journal of Animal Ecology, 70, 966–979.

iu, C., Berry, P. M., Dawson, T. P., & Pearson, R. G. (2005). Selecting thresholds ofoccurrence in the prediction of species distributions. Ecography, 28, 385–393.

iu, C., White, M., & Newell, G. (2013). Selecting thresholds for the predictionof species occurrence with presence-only data. Journal of Biogeography, 40,778–789.

obo, J. M., Jiménez-Valverde, A., & Real, R. (2008). AUC: A misleading measure of theperformance of predictive distribution models. Global Ecology and Biogeography,17, 145–151.

oehle, C., & LeBlanc, D. (1996). Model-based assessments of climate change effectson forests: A critical review. Ecological Modelling, 90, 1–31.

uteyn, J. (2002). Diversity, adaptation, and endemism in neotropical Ericaceae:Biogeographical patterns in the Vaccinieae. Botanical Review, 68, 55–87.

alcolm, J. R., Canran, L., Neilson, R. P., Hansen, L., & Hannah, L. (2006). Global warm-ing and extinctions of endemic species from biodiversity hotspots. ConservationBiology, 20(2), 538–548.

EA. (2005). Millennium ecosystem assessment: Ecosystems and human well-being:Biodiversity synthesis. Washington, DC, USA: World Resources Institute.

eehl, G. A., Covey, C., Taylor, K. E., Delworth, T., Stouffer, R. J., Latif, M., et al. (2007).The WCRP CMIP3 Multimodel Dataset: A new era in climate change research.Bulletin of the American Meteorological Society, 88, 1383–1394.

iles, L., Grainger, A., & Phillips, O. (2004). The impact of global climate change ontropical forest biodiversity in Amazonia. Global Ecology and Biogeography, 13(6),553–565.

ittermeier, R. A., Gil, P. R., & Mittermeier, C. G. (1997). Megadiversity: Earth’s bio-logically wealthiest nations. Cemex: Conservation International.

yers, N., Mittermeier, R. A., Mittermeier, C. G., da Fonseca, G. A. B., & Kent, J. (2000).Biodiversity hotspots for conservation priorities. Nature, 403, 853–858.

hlemüller, R., Gritti, E. S., Sykes, M. T., & Thomas, C. D. (2006). Towards European cli-mate risk surfaces: The extent and distribution of analogous and non-analogousclimates 1931–2100. Global Ecology and Biogeography, 15, 395–405.

lson, D. M., Dinerstein, E., Wikramanayake, E. D., Burgess, N. D., Powell, G. V. N.,Underwood, E. C., et al. (2001). Terrestrial ecoregions of the world: A new mapof life on earth. BioScience, 51, 933–938.

pdam, P., & Wascher, D. (2004). Climate change meets habitat fragmentation: Link-

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

ing landscape and biogeographical scale levels in research and conservation.Biological Conservation, 117, 285–297.

earson, R. G., Thuiller, W., Araújo, M. B., Martinez-Meyer, E., Brotons, L., McClean,C., et al. (2006). Model-based uncertainty in species range prediction. Journal ofBiogeography, 33, 1704–1711.

PRESSre Conservation xxx (2014) xxx–xxx 13

Peterson, A. T., Sánchez-Cordero, V., Soberón, J., Bartley, J., Buddemeier, R. W., &Navarro-Sigüenza, A. G. (2001). Effects of global climate change on geographicdistributions of Mexican Cracidae. Ecological Modelling, 144, 21–30.

Phillips, S. J. (2008). Transferability, sample selection bias and background data inpresence-only modelling: A response to Peterson et al. (2007). Ecography, 31,272–278.

Phillips, S. J., Anderson, R. P., & Schapire, R. E. (2006). Maximum entropy modelingof species geographic distributions. Ecological Modelling, 190, 231–259.

Phillips, S. J., & Dudík, M. (2008). Modeling of species distributions with Maxent:New extensions and a comprehensive evaluation. Ecography, 31, 161–175.

Poulenard, J., Podwojewski, P., Janeau, J., & Collinet, J. (2001). Runoff and soil ero-sion under rainfall simulation of Andisols from the Ecuadorian Páramo: Effectof tillage and burning. Catena, 45, 185–207.

Poulenard, J., Michel, J. C., Bartoli, F., Portal, J. M., & Podwojewski, P. (2004). Waterrepellency of volcanic ash soils from Ecuadorian páramo: Effect of water con-tent and characteristics of hydrophobic organic matter. European Journal of SoilScience, 55, 487–496.

Ramirez-Villegas, J., & Challinor, A. (2012). Assessing relevant climate data for agri-cultural applications. Agricultural and Forest Meteorology, 161, 26–45.

Ramirez-Villegas, J., & Jarvis, A. (2010). Disaggregation of global circulation modeloutputs. Decision and policy analysis working paper no. 2, decision and policy anal-ysis working paper. Cali, Colombia: International Center for Tropical Agriculture(CIAT).

Ramirez-Villegas, J., Jarvis, A., & Touval, J. (2012). Analysis of threats to South Amer-ican flora and its implications for conservation. Journal for Nature Conservation,20, 337–348.

Ramírez-Villegas, J., Khoury, C., Jarvis, A., Debouck, D. G., & Guarino, L. (2010). A gapanalysis methodology for collecting crop gene pools: A case study with Phaseolusbeans. PLoS ONE, 5, e13497.

Renjifo, L. M., Franco-Maya, A. M., Amaya-Espinel, J. D., Kattan, G., & López-Lanús, B.(2002). Libro rojo de aves de Colombia. Bogota Colombia: Instituto de Investiga-ciones de Recursos Biológicos Alexander von Humboldt y Ministerio del MedioAmbiente.

Rivas-Martinez, S. (2004a). Global bioclimatics (Clasificación Bioclimática de laTierra). Version 2004. Phytosociological Research Center. Available from:http://www.globalbioclimatics.org

Rivas-Martinez, S. (2004b). Global bioclimatics (Clasificación Bioclimática de la Tierra).Version 27-08-2004.

Rodríguez, E., Armenteras, P., & Alumbreros. (2013). Land use and land cover changein the Colombian Andes: Dynamics and future scenarios. Journal of Land UseScience, 8, 154–174.

Roy, M. S., da Silva, J. M. C., Arctander, P., Garcia-Moreno, J., & Fjeldså, J. (1997).The speciation of South American and African birds in montane regions. In D. P.Mindell (Ed.), Avian molecular evolution and systematics. San Diego, US: AcademicPress.

Rundel, P. W., & Palma, B. (2000). Preserving the unique puna ecosystems of theAndean Altiplano. Mountain Research and Development, 20, 262–271.

Sachs, J. D., Baillie, J. E. M., Sutherland, W. J., Armsworth, P. R., Ash, N., Beddington, J.,et al. (2009). Biodiversity conservation and the millennium development goals.Science, 325, 1502–1503.

Schuchmann, K.-L., Weller, A.-A., & Heynen, I. (2001). Systematics and biogeographyof the Andean genus Eriocnemis (Aves: Trochilidae). Journal of Ornithology, 142,433–481.

Sklenár, P., & Ramsay, P. M. (2001). Diversity of zonal páramo plant communities inEcuador. Diversity and Distributions, 7, 113–124.

Smith, A. B., Santos, M. J., Koo, M. S., Rowe, K. M. C., Rowe, K. C., Patton, J. L., et al.(2013). Evaluation of species distribution models by resampling of sites sur-veyed a century ago by Joseph Grinnell. Ecography.

Still, C. J., Foster, P. N., & Schneider, S. H. (1999). Simulating the effect of climatechange on tropical montane cloud forests. Nature, 398, 608–610.

Thomas, C. D., Cameron, A., Green, R. E., Bakkenes, M., Beaumont, L. J., Collingham,Y. C., et al. (2004). Extinction risk from climate change. Nature, 427, 145–148.

Thuiller, W., Albert, C., Araújo, M. B., Berry, P. M., Cabeza, M., Guisan, A., et al. (2008).Predicting global change impacts on plant species’ distributions: Future chal-lenges. Perspectives in Plant Ecology, Evolution and Systematics, 9, 137–152.

Thuiller, W., Lavorel, S., Araújo, M. B., Sykes, M. T., & Prentice, I. C. (2005). Climatechange threats to plant diversity in Europe. Proceedings of the National Academyof Sciences of the United States of America, 102, 8245–8250.

Tovar, C., Arnillas, C. A., Cuesta, F., & Buytaert, W. (2013). Diverging responses ofTropical Andean biomes under future climate conditions. PLoS ONE, 8, e63634.

Travis, J. M. J. (2003). Climate change and habitat destruction: A deadly anthro-pogenic cocktail. Proceedings of the Royal Society of London Series B: BiologicalSciences, 270, 467–473.

Urban, M. C., Zarnetske, P. L., & Skelly, D. K. (2013). Moving forward: dispersal andspecies interactions determine biotic responses to climate change. Annals of theNew York Academy of Sciences, 1297(1), 44–60.

VanDerWal, J., Shoo, L. P., Graham, C., & Williams, S. E. (2009). Selecting pseudo-absence data for presence-only distribution modeling: How far should you strayfrom what you know? Ecological Modelling, 220, 589–594.

Vos, C. C., Berry, P., Opdam, P., Baveco, H., Nijhof, B., O’Hanley, J., et al. (2008).Adapting landscapes to climate change: Examples of climate-proof ecosys-

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

tem networks and priority adaptation zones. Journal of Applied Ecology, 45,1722–1731.

Warren, D. L., & Seifert, S. N. (2010). Ecological niche modeling in Maxent: The impor-tance of model complexity and the performance of model selection criteria.Ecological Applications, 21, 335–342.

Page 14: Ramirez-Villegas et al 2014 Using-species-distributions-models-climate-change-impacts-Andean-biodiversity

ING ModelJ

1 r Natu

W

W

W

W

W

Andean South America: An introduction. Botanical Review, 68, 4–21.Zavaleta, E. S., Shaw, M. R., Chiariello, N. R., Mooney, H. A., & Field, C. B. (2003).

Additive effects of simulated climate changes, elevated CO2, and nitrogen depo-

ARTICLENC-25349; No. of Pages 14

4 J. Ramirez-Villegas et al. / Journal fo

arren, R., VanDerWal, J., Price, J., Welbergen, J. A., Atkinson, I., Ramirez-Villegas,J., et al. (2013). Quantifying the benefit of early climate change mitigation inavoiding biodiversity loss. Nature Climate (change advance online publication).

assenaar, T. D., Ferreira, S. M., & Van Aarde, R. J. (2007). Flagging aberrant sites andassemblages in restoration projects. Restoration Ecology, 15, 68–76.

hittaker, R. H. (1960). Vegetation of the Siskiyou mountains, Oregon and California.Ecological Monographs, 30, 279–338.

Please cite this article in press as: Ramirez-Villegas, J., et al. Using specTropical Andean biodiversity under climate change. Journal for Nature

illiams, J. W., & Jackson, S. T. (2007). Novel climates, no-analog communities, andecological surprises. Frontiers in Ecology and the Environment, 5, 475–482.

isz, M. S., Hijmans, R. J., Li, J., Peterson, A. T., Graham, C. H., Guisan, A., et al.(2008). Effects of sample size on the performance of species distribution models.Diversity and Distributions, 14, 763–773.

PRESSre Conservation xxx (2014) xxx–xxx

With, K. A., & King, A. W. (1999). Extinction thresholds for species in fractallandscapes [Umbrales de Extinción para Especies en Paisajes Fraccionados].Conservation Biology, 13, 314–326.

Young, K., Ulloa, C., Luteyn, J., & Knapp, S. (2002). Plant evolution and endemism in

ies distributions models for designing conservation strategies of Conservation (2014), http://dx.doi.org/10.1016/j.jnc.2014.03.007

sition on grassland diversity. Proceedings of the National Academy of Sciences, 100,7650–7654.