Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186...

11
Ecological Applications, 19(5), 2009, pp. 1176–1186 Ó 2009 by the Ecological Society of America Predicting Argentine ant spread over the heterogeneous landscape using a spatially explicit stochastic model JOEL P. W. PITT, 1,3 SUE P. WORNER, 1 AND ANDREW V. SUAREZ 2 1 Bio-Protection Research Centre, Lincoln University, P.O. Box 84, Lincoln 7647, New Zealand 2 Department of Integrative Biology, University of Illinois, 681 Morrill Hall, 505 S. Goodwin Avenue, Urbana, Illinois 61801 USA Abstract. The characteristics of spread for an invasive species should influence how environmental authorities or government agencies respond to an initial incursion. High- resolution predictions of how, where, and the speed at which a newly established invasive population will spread across the surrounding heterogeneous landscape can greatly assist appropriate and timely risk assessments and control decisions. The Argentine ant (Linepithema humile) is a worldwide invasive species that was inadvertently introduced to New Zealand in 1990. In this study, a spatially explicit stochastic simulation model of species dispersal, integrated with a geographic information system, was used to recreate the historical spread of L. humile in New Zealand. High-resolution probabilistic maps simulating local and human-assisted spread across large geographic regions were used to predict dispersal rates and pinpoint at-risk areas. The spatially explicit simulation model was compared with a uniform radial spread model with respect to predicting the observed spread of the Argentine ant. The uniform spread model was more effective predicting the observed populations early in the invasion process, but the simulation model was more successful later in the simulation. Comparison between the models highlighted that different search strategies may be needed at different stages in an invasion to optimize detection and indicates the influence that landscape suitability can have on the long-term spread of an invasive species. The modeling and predictive mapping methodology used can improve efforts to predict and evaluate species spread, not only in invasion biology, but also in conservation biology, diversity studies, and climate change studies. Key words: Argentine ant; heterogeneous landscape; invasive species; Linepithema humile; long- distance dispersal; New Zealand; pest risk assessment; spatially explicit model; spread rate. INTRODUCTION Increase in world tourism and trade has been linked to a rise in the number of species unintentionally introduced to new environments (Levine and D’Antonio 2003). The establishment of a new species in an area where it is not normally found is often associated with severe ecological and economic consequences (Mack et al. 2000, Pimentel et al. 2000). If that species is determined to have the potential to spread over a large area and have a negative impact, then eradication or control actions must be quickly prioritized to minimize damage. Once an exotic species establishes a reproducing population in a new region, the next stage of the invasion process is its spread across the landscape (Hastings 1996). A model that can estimate the rate of spread and the direction of an invasion would greatly assist relevant authorities in the design of sampling programs for detection and in the monitoring of spread and eradication attempts. However, modeling spread is difficult for two reasons. First, data for model param- eterization are not usually available to estimate the rate of spread of a newly detected species. Second, many species can spread by multiple methods, for example, by natural means, often a diffusion-like process, and also by large jumps, often mediated by humans, a process often referred to as stratified diffusion (Hengeveld 1989). If such components are characteristic of a new invader, strategies for slowing its spread must account for the effect of each component on both temporal and spatial aspects of dispersal. Previous methods for modeling spread have provided theoretical insights into the potential speed at which a population front might travel and have highlighted important factors that impact the rate of spread (for an extensive review see Hastings et al. [2005]). Those factors include dispersal kernel shapes that describe the distance that propagules travel and Allee effects that can limit spread rates and constrain population fronts that otherwise are predicted to accelerate indefinitely. To date, most models focus on an abstract environment that is typically homogenous and constrained to one dimen- Manuscript received 24 September 2008; revised 24 October 2008; accepted 5 November 2008. Corresponding Editor: M. P. Ayers. 3 Present address: Fruition Technology, 103 Maupuia Road, Wellington 6022 New Zealand. E-mail: [email protected] 1176

Transcript of Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186...

Page 1: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

Ecological Applications, 19(5), 2009, pp. 1176–1186� 2009 by the Ecological Society of America

Predicting Argentine ant spread over the heterogeneous landscapeusing a spatially explicit stochastic model

JOEL P. W. PITT,1,3 SUE P. WORNER,1 AND ANDREW V. SUAREZ2

1Bio-Protection Research Centre, Lincoln University, P.O. Box 84, Lincoln 7647, New Zealand2Department of Integrative Biology, University of Illinois, 681 Morrill Hall, 505 S. Goodwin Avenue, Urbana, Illinois 61801 USA

Abstract. The characteristics of spread for an invasive species should influence howenvironmental authorities or government agencies respond to an initial incursion. High-resolution predictions of how, where, and the speed at which a newly established invasivepopulation will spread across the surrounding heterogeneous landscape can greatly assistappropriate and timely risk assessments and control decisions.

The Argentine ant (Linepithema humile) is a worldwide invasive species that wasinadvertently introduced to New Zealand in 1990. In this study, a spatially explicit stochasticsimulation model of species dispersal, integrated with a geographic information system, wasused to recreate the historical spread of L. humile in New Zealand. High-resolutionprobabilistic maps simulating local and human-assisted spread across large geographic regionswere used to predict dispersal rates and pinpoint at-risk areas. The spatially explicit simulationmodel was compared with a uniform radial spread model with respect to predicting theobserved spread of the Argentine ant. The uniform spread model was more effective predictingthe observed populations early in the invasion process, but the simulation model was moresuccessful later in the simulation. Comparison between the models highlighted that differentsearch strategies may be needed at different stages in an invasion to optimize detection andindicates the influence that landscape suitability can have on the long-term spread of aninvasive species.

The modeling and predictive mapping methodology used can improve efforts to predictand evaluate species spread, not only in invasion biology, but also in conservation biology,diversity studies, and climate change studies.

Key words: Argentine ant; heterogeneous landscape; invasive species; Linepithema humile; long-distance dispersal; New Zealand; pest risk assessment; spatially explicit model; spread rate.

INTRODUCTION

Increase in world tourism and trade has been linked to

a rise in the number of species unintentionally introduced

to new environments (Levine and D’Antonio 2003). The

establishment of a new species in an area where it is not

normally found is often associated with severe ecological

and economic consequences (Mack et al. 2000, Pimentel

et al. 2000). If that species is determined to have the

potential to spread over a large area and have a negative

impact, then eradication or control actions must be

quickly prioritized to minimize damage.

Once an exotic species establishes a reproducing

population in a new region, the next stage of the

invasion process is its spread across the landscape

(Hastings 1996). A model that can estimate the rate of

spread and the direction of an invasion would greatly

assist relevant authorities in the design of sampling

programs for detection and in the monitoring of spread

and eradication attempts. However, modeling spread is

difficult for two reasons. First, data for model param-

eterization are not usually available to estimate the rate

of spread of a newly detected species. Second, many

species can spread by multiple methods, for example, by

natural means, often a diffusion-like process, and also

by large jumps, often mediated by humans, a process

often referred to as stratified diffusion (Hengeveld 1989).

If such components are characteristic of a new invader,

strategies for slowing its spread must account for the

effect of each component on both temporal and spatial

aspects of dispersal.

Previous methods for modeling spread have provided

theoretical insights into the potential speed at which a

population front might travel and have highlighted

important factors that impact the rate of spread (for an

extensive review see Hastings et al. [2005]). Those factors

include dispersal kernel shapes that describe the distance

that propagules travel and Allee effects that can limit

spread rates and constrain population fronts that

otherwise are predicted to accelerate indefinitely. To

date, most models focus on an abstract environment that

is typically homogenous and constrained to one dimen-

Manuscript received 24 September 2008; revised 24 October2008; accepted 5 November 2008. Corresponding Editor: M. P.Ayers.

3 Present address: Fruition Technology, 103 MaupuiaRoad, Wellington 6022 New Zealand.E-mail: [email protected]

1176

Page 2: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

sion (e.g., Kot et al. 1996) or if not, they are constrained

with respect to scale. More realistic models are needed to

progress practical understanding of the dispersal and

spread of organisms and to provide decision support for

those involved with containing or eradicating invasive

species. Moreover, a model using an incorrect spatial

dimension can produce misleading results. A good

example is a study by Petrovskii (2005) in which it was

shown that the invasion of a predator or infectious

disease that has a patchy spatial distribution can persist

in two-dimensional space, but will go extinct in a

corresponding one-dimensional system. Of critical im-

portance also is that an abstract, one-dimensional model

excludes spatial patterning that arises from the interac-

tion of a population with the landscape. Turner et al.

(1993) were among the first to suggest that an essential

component to the progress of predictive spread models is

to utilize the spatial heterogeneity of the natural

landscape, and Worner (1994) called for models of

species establishment and distribution to be integrated

with geographic information system (GIS) technology.

In this study we incorporate realistic landscapes into a

spread model by modeling dispersal processes within a

GIS. The model uses concepts taken from traditional

theoretical population and spread models and applies

them to the simulation of the dispersal of the Argentine

ant (Linepithema humile, Mayr) in New Zealand.

The Argentine ant, Linepithema humile, is a world-

wide pest that is cited as one of the six worst invasive ant

species (Holway et al. 2002). When this species was first

recorded in New Zealand in 1990 (Green 1990), there

was no attempt to control it as it was considered already

well established. Argentine ants are considered a

successful ‘‘tramp’’ ant species (Passera 1994) in part

due to a strong tendency to move and associate with

humans (Suarez et al. 2001), its unicoloniality (individ-

uals mix freely among physically separate colonies)

(Holway 1998), strong interspecific aggression (Holway

1999), polygyny (Keller and Passera 1990), and dispersal

by budding (a queen supported by as few as 10 workers

can establish a new colony) (Hee et al. 2000).

Linepithema humile is a threat to New Zealand’s

biodiversity because in addition to potential negative

impacts on wildlife, it readily displaces other ant species

(reviewed in Holway et al. 2002). The displacement of

existing ant species can cause complex mutualisms to be

disrupted (Bond and Slingsby 1984, Lach 2003) as well

as disrupt other ecosystem processes (Harris 2002).

Because L. humile was considered well established and

there were limited means to control it, the species was

largely left to spread unhindered and provides a good

example to study the spread of an invasive species that is

not confounded by an eradication attempt. Internation-

al data on L. humile distribution and spread were used to

parameterize a model that was then used to simulate L.

humile spread from its initial site of invasion in New

Zealand. While the distances of jump dispersal events

and their frequency in New Zealand were used for

comparison with International data, the exact spatial

locations from New Zealand occurrence data were kept

for model validation.

METHODOLOGY

Model design

A modular dispersal framework was used to model

temporally discrete dispersal processes within GIS.

Raster maps were used to represent population distri-

butions and the modeling framework utilized open-

source software and was implemented using the

programming languages Python (available online)4 and

C, within the open-source GIS GRASS (Neteler and

Mitasova 2004). The L. humile model uses a raster map

for each year to represent either the presence or absence

of the species in a raster cell. The modular framework

was designed as a generic simulation model that could

be used for any species.

In this study, the dispersal of L. humile is character-

ized by stratified diffusion (Shigesada et al. 1995) with

the invasion proceeding from several locations or foci.

Such patchy distributions are often the result of the

interaction of more than one mode of dispersal and

Barber (1916), Holway (1995), and Suarez et al. (2001)

indicate the two most influential dispersal modes for L.

humile are local spread by budding and jump dispersal

facilitated by human transport. Providentially, splitting

the dispersal process into two components addresses the

frequent problem of finding an appropriate dispersal

kernel to represent dispersal distances. In our model, the

two dispersal modes are represented in the model by

local and kernel modules, respectively (Fig. 1a, b).

For any time step, the local module processes a raster

map comprised of cells. For each occupied raster cell (in

which the species is present), the model updates

neighboring cells such that each becomes occupied,

leading to contiguous spread of the population. Line-

pithema humile has been shown to have an average

budding spread rate of 150 m/yr for regions where

habitat and climate are not limiting (Suarez et al. 2001).

To represent this spread rate at the correct resolution,

the model used a raster resolution of 150 m and the local

module used a von Neumann neighborhood to represent

local dispersal (Fig. 1a).

The kernel module represents jump and long-distance

dispersal and is based on the Cauchy probability

distribution to represent the probability a dispersal

event travels a given distance (Hastings et al. 2005). For

each occupied raster cell, the kernel module first samples

a Poisson distribution, with mean k, to determine the

number of long-distance dispersal events that arise from

that cell. Then, for each of these events, the module

samples the Cauchy probability distribution for the

dispersal distance of each event. After the distance of the

dispersal event is determined, the model samples a

4 hhttp://www.python.org/i

July 2009 1177MODELING L. HUMILE SPREAD IN NEW ZEALAND

Page 3: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

uniform distribution in the range [0, 2p) to establish the

angle or direction of that event.

Habitat suitability and survival

To link the dispersal model to the real landscape

where habitat suitability is variable, another module,

survival, controls the probability an occupied cell

becomes extinct. The probability of extinction is based

on a suitability map that indicates the relative suitability

of cells within the region being studied. The suitability

map could potentially be constructed using any number

of methods for modeling potential distributions of

species based on environmental conditions (e.g., Worner

1988, Stockwell 1999, Sutherst et al. 1999, Thuiller 2003,

Guisan and Thuiller 2005, Gevrey and Worner 2006,

Pitt et al. 2007). For L. humile, however, we combined

expert knowledge about the suitability of various land

cover types for persistence of populations of this species

(Harris 2002) with a degree-day analysis that measures

the cumulative amount of heat required for continued

development at any location (Hartley and Lester 2003).

For land cover, the New Zealand Land Cover

Database version 2 was used (available online).5 Land

cover types were divided into three categories, from

unsuitable to highly suitable (H0, H1, H2), with a

separate category for urban areas (HU). A category for

urban areas was used because L. humile has been found

to survive at lower ambient temperatures in urban

environments, because of a close association with

human activity and the warm microclimates created by

that activity (Suarez et al. 2001). The land cover

suitability map is shown in Fig. 2a. Thirty years of

historical daily minimum and maximum temperatures

interpolated as a grid with a latitudinal and longitudinal

resolution of 0.058 were used to create a map of the

annual average degree-days for L. humile development

above a threshold of 15.98C (Hartley and Lester 2003).

This map was transformed to the New Zealand Map

Grid (NZMG) projection and interpolated, using

inverse distance squared weighting, to a resolution of

150 m. Based on the degree-days required for develop-

ment, the degree-day map was divided into four

categories of total degree-days from low to high

suitability (T0 , 200, T1 ¼ 201–300, T2 ¼ 301–400, T4

. 401 degree-days) (Fig. 2b).

The land cover and degree-day suitability maps (Fig.

2a, b) were combined (Fig. 2c) using the scheme

presented by Harris (2002) as shown in Table 1. Harris

(2002) used mean annual temperature classified into

suitability levels, whereas in this study, degree-days have

been substituted, because they are biologically more

relevant and usually more accurate. The result was a

map classified into regions of low, medium, and high

suitability, with additional categories for regions with

unsuitable land cover or that were too cold. Each

category was assigned a survival probability indicating

the probability of occupants surviving to the subsequent

year, with progressively lower probabilities for less

suitable categories (Table 1).

The overall suitability map (Fig. 2c) was constrained

predominantly by the degree-day map, which resulted in

most of the central and southern parts of the North

Island classified as too cool for L. humile development.

Much of Northland, along with Auckland and other

northern cities and towns, showed medium or high

suitability. Northeastern and some eastern coasts

showed low suitability.

Model calibration

To calibrate the kernel module, data from the

historical spread of L. humile in the United States was

used (Suarez et al. 2001). The data included average

rates of spread as a consequence of the budding process

across multiple invasion fronts as well as the dates of

first detection within U.S. counties from 1891 to 1999.

Nearest-neighbor distances among established popula-

tions were used to fit the (Cauchy) probability distribu-

tion used in the kernel module to represent the frequency

distribution of long-distance dispersal events. The

nearest-neighbor distances were calculated each year as

the distance between the centroids of new counties

FIG. 1. The three modules represented in the Linepithemahumile (Argentine ant) dispersal simulation: (a) local dispersalto neighboring sites, using a von Neumann neighborhood torepresent the dispersal ability of the ant; (b) long-distancedispersal; and (c) stochastic survival associated with occupiedcells based on an underlying suitability map.

5 hhttp://www.mfe.govt.nz/issues/land/land-cover-dbase/i

JOEL P. W. PITT ET AL.1178 Ecological ApplicationsVol. 19, No. 5

Page 4: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

invaded by L. humile and the centroids of counties that

had previous records of presence. For counties that were

noncontiguous, the average of the region centroids,

weighted by area, was used. The number of new

occurrences each year, divided by the number of

preexisting occurrences, was used to estimate k, the

frequency of long-distance dispersal events as described

by the Poisson distribution. Because of the scale over

which the data were measured (over individual coun-

ties), the estimate of k that represents the number of new

foci per year was likely to be a conservative underesti-

mate. However, any underestimation of k would be

offset in this model by local spread represented in the

model by the local module that generates new sites

neighboring those that already exist.

The frequency and distance of long-distance dispersal

events are often difficult to quantify (Higgins and

Richardson 1999), but good estimates are essential for

accurate model construction (Shigesada and Kawasaki

1997). Despite the fact that the New Zealand L. humile

occurrence data is point-based, compared with the U.S.

data, which was recorded over geographic regions, the

frequency of long-distance dispersal events and the

frequency of dispersal distances could still be calculated.

Sites that were within 300 m/yr of any of their neighbors

were assumed to have established by budding and were

ignored as per Ward et al. (2005), who previously had

investigated the statistics of L. humile human-assisted

dispersal distances in New Zealand.

Frequencies of L. humile’s long-distance dispersal

events within the United States and New Zealand were

examined to determine whether they occurred at similar

scales. A qualitative assessment of dispersal events

showed that, for both New Zealand and the United

States, the majority of dispersal events occurred over

relatively short distances, and the distribution in each

country included several large-distance movements (Fig.

3). The U.S. data had a higher proportion of dispersal

distances distributed between 100 and 700 km. These

large dispersal distances were probably the result of two

factors. First, North America is part of a large

continent, so large human-assisted dispersal distances

are possible, compared with New Zealand, which is a

small island nation. Second, for the U.S. data, distances

might be overestimated because of the size and shape of

the counties on which they were based. For example, a

dispersal event may cross the short distance from one

side of a border between two counties, but the distance

FIG. 2. Suitability maps for Linepithema humile (Argentine ant) in the North Island of New Zealand, based on (a) land coverclassification, (b) degree-days above a threshold of 15.98C, and (c) a combination of both.

TABLE 1. Scheme used for combining land cover and degree-day suitability maps.

Combination Result Survival probability

T3 þ H2, T3 þ HU, T2 þ HU high suitability 100T3 þ H1 moderate suitability 80T1 þ HU, T1 þ H1, T2 þ H1, T1 þ H2, T2 þ H2 low suitability 50H0 þ T* unsuitable habitat 10H* þ T0 too cold 10

Notes: Land cover types were divided into three categories, from unsuitable to highly suitable(H0, H1, H2), with a separate category for urban areas (HU). Based on the degree-days requiredfor development, the degree-day map was divided into four categories of total degree-days fromlow to high suitability (T0 , 200, T1 ¼ 201–300, T2¼ 301–400, T4 . 401).

July 2009 1179MODELING L. HUMILE SPREAD IN NEW ZEALAND

Page 5: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

recorded will be the larger distance between the

centroids of the two county regions.

We used the U.S. frequencies for parameter estima-

tion of the Cauchy distribution. The Cauchy distribu-

tion was chosen to represent the distances of dispersal

events as it is ‘‘fat-tailed,’’ allowing for rare events at

extreme distances that have been shown to be an

important characteristic of the dispersal of many species

(Hengeveld 1989, Higgins and Richardson 1999, Clark

et. al 2001, Suarez et al. 2001) and even that of humans

(Brockmann et al. 2006). Maximum likelihood estima-

tion (MLE) using the simplex search method was used

(Lagarias et al. 1998) to estimate the Poisson k and the

Cauchy c. Maximum likelihood estimation estimated c,the shape parameter for the Cauchy distribution, as 8.37

3 104 (Table 2).

A comparison of the frequency of dispersal events per

site for each country showed a greater frequency of a

low number of dispersal events per site in the United

States, although both countries showed averages per site

less than 1 (Fig. 4). The greater frequency of low

numbers of dispersal events per site in the United States

is best explained by the longer time frame over which

dispersal was recorded. Also, as the species became

established in more counties it became harder for

dispersing ants to find unoccupied counties in which to

establish. Maximum likelihood estimation estimated kfor the Poisson distribution that represents the frequen-

cy of dispersal events as 0.298 (Table 2).

Simulation

For this study, the simulation was constrained to the

North Island of New Zealand, an area of 113 729 km2,

at a raster resolution of 150 m. Within the GIS, the

simulation was carried out on the 1949 New Zealand

Geodetic Datum and the NZMG projection. The time

step was one year, and within each time step the L.

humile distribution map was processed by the model

modules in the order: local, kernel, and survival. At the

end of each time step the distribution map is saved for

later analysis. The simulation was run from 1990 to

2005, starting with the three sites discovered in 1990.

To measure the uncertainty of prediction to variation

in model parameter values, the parameters for the

Poisson mean k and the scale parameter, c, of the

Cauchy distribution were simulated over their estimated

mean value and their 95% confidence interval (CI)

limits, for a total of nine combinations of parameter

values. Both the kernel and survival modules involve

random sampling from appropriate probability distri-

butions and are therefore stochastic processes. Thus,

each parameter combination was simulated 100 times

from 1990 to 2005 to give 900 realizations of spread.

To obtain an average representation of L. humile

spread, all maps for a given year were averaged to create

an occupancy map that represents the probability a cell

is occupied at a given time. This probability is simply the

proportion of times a cell was occupied over the 100

simulations. This occupancy map was then masked by

excluding areas below a given threshold or a very low

probability of occupancy.

To evaluate the performance of the L. humile

simulation model against a simple spread model,

uniform radial spread from the mean center of the three

initial invasion sites was also modeled. A linear increase

FIG. 3. Comparison of the frequency of Linepithema humile(Argentine ant) dispersal event distances per year in NewZealand and the United States.

TABLE 2. Maximum likelihood estimation (MLE)-derivedparameters for the probability distributions used forsimulating long-distance human-mediated dispersal eventswith the kernel module.

Parameter 95% lower bound Estimate 95% upper bound

Cauchy c 72 700 83 700 94 600Poisson k 0.199 0.298 0.428

JOEL P. W. PITT ET AL.1180 Ecological ApplicationsVol. 19, No. 5

Page 6: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

in the square root of area is equivalent to a constant

spread rate for a circular area. The rate of radial increase

for the uniform spread model was calculated by

assuming the square root of the increase in area of the

simulation model over time was approximately linear

and by calculating an approximate slope. The rate of

increase of the uniform spread model radius was given by

ffiffiffiffiffiffiffiffiffiffiffi

At1=pp

�ffiffiffiffiffiffiffiffiffiffiffi

At0=pp

t1 � t0

where At is the area encompassed by the simulation

model at time t, and t1 is some time after t0. This formula

meant that both models encompassed approximately the

same total area at any time step. Simulation results

indicated that certain phases in the increase in occupancy

area had the square root of their area increase at an

approximately linear rate.

RESULTS

The percentage of observed L. humile occurrence sites

in New Zealand for each respective year that fell within

the occupancy envelope was calculated (Fig. 5a). The

model at all occupancy thresholds predicted a high

percentage of sites early in the simulation, before

dropping to ;40% of observed sites at 1993–1994. The

percentage of observed sites predicted by the model

(occupancy threshold . 0) increased in 1996. The

percentage of predicted sites at .5% and 10% occupan-

cy follow, both peaking in ca. 1999. The percentage of

FIG. 5. Performance of the simulation model (k [thefrequency of new dispersal events] ¼ 0.298, c [the scaleparameter of the Cauchy distribution representing dispersaldistances] ¼ 83 700) with respect to the extent that theprobability envelope includes observed Linepithema humile(Argentine ant) occurrence sites in a given year, 1990–2005.(a) The percentage of sites included. (b) The difference inperformance (percentage of sites predicted) between thesimulation model (number of sites within the probabilityenvelope) and the uniform spread model (number of siteswithin the uniform radial spread). When the change inpercentage is positive the simulation model performs betterthan the uniform spread model and vice versa when the changeis negative.

FIG. 4. Comparison of long-distance dispersal event fre-quency for Linepithema humile (Argentine ant) per number ofpreexisting sites in New Zealand and the United States.

July 2009 1181MODELING L. HUMILE SPREAD IN NEW ZEALAND

Page 7: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

predicted sites for all occupancy thresholds drop around

2000 but thereafter increase.

If the performance of the simulation model is

compared to the simple uniform spread model, the

benefits of using a simple or more complex model at

different phases of the invasion become apparent (Fig.

5b). If both models predict the same number of sites,

then the graph would follow the x-axis at zero. While the

.0% and 5% occupancy envelope predicts the observed

sites much earlier than the uniform radial spread model

in all but one year, the higher occupancy thresholds

appeared to perform poorly compared with the uniform

radial spread model initially, despite covering a similar

total area. This was caused by the slow ‘‘spread’’ of high-

occupancy sites (Fig. 6) and lack of agreement between

replicates at the beginning of the simulation. Clearly the

uniform spread model gave better prediction early in the

invasion, while the simulation model improved perfor-

mance as the invasion developed over time.

Model performance was assessed over all parameter

combinations by the percentage of sites predicted by the

model (in other words, included in the occupancy

envelope). In general, the same pattern was observed.

Prediction was good at the start of the invasion but

degraded around 1992–1993, improving around 1998,

and decreasing around 2001 before increasing again.

Increasing c generally decreased model performance.

Increasing k to its 95% CI upper bound flattened the

peaks and lows and generally increased predictive

performance. A comparison of the simulation model

with the uniform spread model over various parameter

values and combinations showed little difference to what

was observed when their mean values were used except

for the higher occupancy levels (10% and 50%) that

performed relatively better after 2000 at the upper 95%

CI of k.

Hotspots, or those areas with high values in the

occupancy envelope, are of particular interest to

agencies charged with eradicating or monitoring inva-

sive species. In this study, hotspots included regions near

the invasion epicenter, within and near Auckland city

and nearby cities such as Whangarei and Hamilton,

both of which have L. humile infestations (Fig. 7).

Towns near Auckland, such as Pukekohe and Waiuku,

were also indicated as hotspots, but to date have no

recorded occurrences. Great Barrier Island, Little

Barrier Island, Ponui Island, and Tiri Matangi Island

were also indicated as hotspots. Great Barrier Island

and Tiri Matangi Island have had occurrences recorded,

most likely from human-assisted dispersal. Both islands

have undergone poison baiting treatments to eradicate

L. humile (J. Boow, personal communication; C. Green,

personal communication). The model indicated one large

hotspot covering most of the Hauraki Plains, which has

no recorded occurrences of L. humile. The land cover

map indicates a large area of scrubland that is a highly

suitable habitat for L. humile (Suarez et al. 1998). The

whole of the Hauraki Plains also has a sufficient number

of degree-days for complete L. humile development.

No simulation for any combination of variables

reached the southern end of the North Island, despite

that L. humile has been established there since 2000. This

model result is probably due to the large area of

unsuitable habitat between the invasion epicenter and

the south of the North Island. The occurrences much

further south likely arose from individuals hitch-hiking

on road or rail networks. Over the entire simulation area

(Fig. 8), high-suitability regions were more quickly

occupied than low-suitability regions, as expected. Some

low-suitability regions eventually did show significant

occupancy probabilities despite the high probability of

extinction through their proximity to high-suitability

regions that provided high propagule pressure.

While there were differences in model output in

response to variation in its parameters, the result after

some time had elapsed was a qualitative pattern similar

to the underlying landscape suitability map. This result

reinforces the observation that landscape heterogeneity

can often have a large stabilizing effect on ecological

models (Kuno 1981, Ruxton and Rohani 1999, Gardner

and Gustafson 2004). The simulation was most affected

by differences in k, as that parameter controlled the

number of events, with a cumulative effect so that more

events earlier in the simulation led to more occupied

sites from which successive events could occur.

Fig. 6 shows how the square root of area ðffiffiffi

ApÞ of the

occupancy envelope increased through time. The rate

thatffiffiffi

Ap

increased for all occupied sites with no

occupancy threshold accelerated until 2000–2002 before

the rate slowed slightly. The graphs for various

occupancy thresholds seem to show approximately three

phases with different rates of increase. The different

rates seem to be due to the shape of the initial invasion

area being an isthmus between the main land mass of the

FIG. 6. The increase in area for the entire probabilityenvelope of simulation (k [the frequency of new dispersalevents] ¼ 0.298, c [the scale parameter of the Cauchydistribution representing dispersal distances] ¼ 83 700). Areaincrease is shown for three occupancy thresholds (0.5, 0.1, and0.05).

JOEL P. W. PITT ET AL.1182 Ecological ApplicationsVol. 19, No. 5

Page 8: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

North Island and Northland, with less area to spread

than inland areas. However, once the invasion has

spread far enough, it can reach these inland regions

containing more area available for occupation.

DISCUSSION

Lonsdale (1999) suggests that more robust relation-

ships in the pattern of invasive species spread are likely

to emerge at broad spatial scales. Clearly, simulating the

spread of a species over large spatial scales, such as in

this study, is important to help progress theory

concerning the dispersal and spread of organisms over

the heterogeneous environment. We showed that it is

possible to estimate how and where an invasive species

might spread over a large spatial scale, by calibrating a

stochastic and spatially explicit presence/absence model

to observed data and combining that with expert

knowledge.

When simulation results were compared using a

uniform circular spread model with a constant radial

FIG. 7. Snapshots of the dispersal simulation (k [the frequency of new dispersal events]¼ 0.298, c [the scale parameter of theCauchy distribution representing dispersal distances] ¼ 83 700) for years 1999, 2001, 2003, and 2005. The map area is coloredaccording to when the probability envelope exceeds the thresholds (red . 0.5, blue . 0.1, green . 0.05, yellow , 0.05). The coloredcircle outlines indicate the boundary of a uniform spread model, the area of which is equivalent to the area encompassed by thethresholded probability envelope. Red dots with black outlines indicate observed Linepithema humile (Argentine ant) occurrencesites.

July 2009 1183MODELING L. HUMILE SPREAD IN NEW ZEALAND

Page 9: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

spread rate from the invasion epicenter, the uniform

spread model was more accurate early in the invasion.

The simulation model, however, predicted the observed

long-distance events much more quickly as the invasion

proceeded. The most demanding and costly aspect of

monitoring, controlling, or eradicating an invasive

species is the extent of the area to be searched. Clearly

the ability to quickly identify hotspots for searching far

from the perceived epicenter of an invasion would be

very useful. In the case of L. humile, the simulation

model matched the actual occurrence data for five

hotspots (Auckland, Hamilton, Whangerei, Great Bar-

rier Island, and Tiri Matangi Island) and has indicated

that several other towns and the Hauraki Plains could

potentially harbor populations of the ant, despite the

fact that none have yet been reported in these regions.

The simulations in this study underestimated the rate

of spread of the Argentine ant in New Zealand; no

simulation reached the southern end of the North Island

where the species has been established since 2000.

Models often underestimate the rate of spread of

invading organisms, and this failure is attributed to the

inability to accurately measure infrequent long-distance

dispersal (Andow et al. 1990, Liebhold et al. 1992,

Buchan and Padilla 1999, Neubert and Caswell 2000).

Despite this result, this study clearly highlights the

importance of including long-distance dispersal when it

has been shown to occur, since without it, species spread

would progress much more slowly. Indeed, within the

time frame of the simulation, 1990 to 2005, L. humile

spread due to population diffusion (via budding) would

not have spread beyond 2 km from each of the three

initial occurrence sites. Even theoretical models such as

integro-difference equation models of spread show that

the long-distance component of dispersal ultimately

decides invasion speed, even when long-distance dis-

persal is rare (Kot et al. 1996, Lewis 1997, Neubert and

Caswell 2000).

In this study we noted that maps from early

simulations using our spatially explicit stochastic simu-

lation model had little agreement. That is an expected

result however, as it is difficult to predict what will

happen very early in an invasion because of the inherent

randomness of the process, especially with respect to

long-distance dispersal (Higgins and Richardson 1999).

On the other hand, as the invasion progresses, the

simulation occupancy envelope tends to conform to the

underlying suitability map, suggesting that perhaps only

a suitability map is required. However, the ability to

estimate the rate at which the invasion occupies suitable

regions gives authorities, charged with the task of

monitoring or eradicating an exotic species, critical

dynamic information that is superior to just an estimate

of potential distribution such as that provided by

current distribution models. The latter, however, are

useful for pre-border pest risk assessment, as they can be

used to indicate the initial site at which an exotic species

might establish, given a pathway of arrival into the

country. However, this simulation model could be used

to explore invasions from a number of potential

disembarkation points over a wide region. Mooij and

DeAngelis (2003) suggest that because spatially explicit

dispersal models utilize landscape details, they suffer less

from uncertainty than simpler models. Despite the

FIG. 8. Snapshots of the entire simulation region (k¼ 0.298, c ¼ 83 700). For further explanation, see Fig. 7 legend.

JOEL P. W. PITT ET AL.1184 Ecological ApplicationsVol. 19, No. 5

Page 10: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

variability of output of the simulation model in response

to parameter change, output maps were qualitatively

similar and in fact decision makers are likely to be

interested in the relative probability of occupancy

between regions that indicate those most at risk rather

than the absolute occupancy value.

Situations in which the model presented here would be

a poor choice include where established populations are

confined to distinct and explicit patches of suitable

habitat, such that all other areas outside of these patches

is unsuitable. In such a situation, a metapopulation

model (Hanski et al. 2000) or a stochastic patch

occupancy model (Moilanen 2004) would be more

appropriate. Another case is when a species spreads

through a monocultural environment, such as in certain

agricultural systems. Here the assumption of a homo-

geneous environment is appropriate and mathematical

models such as partial differential equations (PDEs) and

integro-difference equations (IDEs) would have benefits

not available to a simulation over a complex landscape.

This model is capable of generating realistic realiza-

tions of spread that can be used to design appropriate,

balanced sampling programs to detect or monitor an

invasive species especially at low densities. Designs with

different sample sizes and spatial and temporal patterns

can be tested over a realistic landscape. Similarly,

experiments using different eradication treatments,

particularly investigating the pattern in which they are

applied, can be carried out. Such experiments are rarely

possible in reality. Moody and Mack (1988) point out

that for species spreading by stratified dispersal,

decisions whether to control small, distant foci or the

main source population can greatly affect the impact

and cost of control. A modular model means modifica-

tions, such as linking human-mediated dispersal to

transport networks, wind dispersal of airborne life

stages, local population growth, or cellular-automata-

based rules, are easily implemented. The modular nature

also forces consideration of the various behaviors

underlying the patterns observed in species spread (Pitt

2008). Elucidating the underlying mechanisms driving

such patterns makes models less arbitrary and links

them to explicit spatial scales (Grimm et al. 1996).

In conclusion, stochastic, spatially explicit dispersal

models integrated with GIS are required to incorporate

ecological theory about dispersal with heterogeneous

landscapes. Not only will they help progress our

understanding concerning species spread, they allow

predictions to be made about the direction and rate of

the spread. The predictions not only inform invasion

biology, they provide essential information that can

assist the prevention of further spread by an invasive

population or for efficient eradication attempts. Such

models will also assist climate change studies by

simulating the movement of species into new areas that

become climatically suitable. Lastly, while maps are very

useful communication tools, allowing model results to

be easily conveyed to policy and decision makers,

representing uncertainty is a problem. Stochastic simu-

lations, as have been used in this study, provideprobabilistic estimates that encapsulate some of the

uncertainty involved in the prediction of invasive spread.

ACKNOWLEDGMENTS

We thank the Bio-Protection Research Centre, LincolnUniversity, for funding the research, Stephen Hartley forclarification and explanation of his degree-day calculation,and the National Institute of Water and Atmospheric Researchfor providing the historical daily minimum and maximumtemperature grid.

LITERATURE CITED

Andow, D. A., P. M. Kareiva, S. A. Levin, and A. Okubo.1990. Spread of invading organisms. Landscape Ecology 4:177–188.

Barber, E. R. 1916. The Argentine ant: distribution and controlin the United States. U.S. Department of Agriculture,Washington, D.C., USA.

Bond, W., and P. Slingsby. 1984. Collapse of an ant–plantmutualism: the Argentine ant (Iridomyrmex humilis) andmyrmecochorous Proteaceae. Ecology 65:1031–1037.

Brockmann, D., L. Hufnagel, and T. Geisel. 2006. The scalinglaws of human travel. Nature Letters 439:462–465.

Buchan, L. A. J., and D. K. Padilla. 1999. Estimating theprobability of long-distance overland dispersal of invadingaquatic species. Ecological Applications 9:254–265.

Clark, J. S., M. Lewis, and L. Horvath. 2001. Invasion byextremes: population spread with variation in dispersal andreproduction. American Naturalist 157:537–554.

Gardner, R. H., and E. J. Gustafson. 2004. Simulating dispersalof reintroduced species within heterogeneous landscapes.Ecological Modelling 171:339–358.

Gevrey, M., and S. P. Worner. 2006. Prediction of globaldistribution of insect pest species in relation to climate byusing an ecological informatics method. Journal of EconomicEntomology 99:979–986.

Green, O. R. 1990. Entomologist sets new record at Mt Smartfor Iridomyrmex humilis established in New Zealand. Weta13:14–16.

Grimm, V., K. Frank, F. Jeltsch, R. Brandl, Uchmann, J., andC. Wissel. 1996. Pattern-orientated modelling in populationecology. Science of the Total Environment 183:151–166.

Guisan, A., and W. Thuiller. 2005. Predicting species distribu-tion: offering more than simple habitat models. EcologyLetters 8:993–1009.

Hanski, I., J. Alho, and A. Moilanen. 2000. Estimating theparameters of survival and migration of individuals inmetapopulations. Ecology 81:239–251.

Harris, R. J. 2002. Potential impact of the Argentine ant(Linepithema humile) in New Zealand and options for itscontrol. Science for Conservation 196. Department ofConservation, Wellington, New Zealand.

Hartley, S., and P. J. Lester. 2003. Temperature-dependentdevelopment of the Argentine ant, Linepithema humile(Mayr) (Hymenoptera: Formicidae): a degree-day modelwith implications for range limits in New Zealand. NewZealand Entomologist 26:91–100.

Hastings, A. 1996. Models of spatial spread: a synthesis.Biological Conservation 78:143–148.

Hastings, A., et al. 2005. The spatial spread of invasions: newdevelopments in theory and evidence. Ecology Letters 8:91–101.

Hee, J. J., D. A. Holway, A. V. Suarez, and T. J. Case. 2000.Role of propagule size in the success of incipient colonies ofthe invasive Argentine ant. Conservation Biology 14:559–563.

Hengeveld, R. 1989. Dynamics of biological invasions. Chap-man and Hall, London, UK.

July 2009 1185MODELING L. HUMILE SPREAD IN NEW ZEALAND

Page 11: Predicting Argentine ant spread over the …...Ecological Applications, 19(5), 2009, pp. 1176–1186 2009 by the Ecological Society of America Predicting Argentine ant spread over

Higgins, S. I., and D. M. Richardson. 1999. Predicting plantmigration rates in a changing world: the role of long-distancedispersal. American Naturalist 153:464–475.

Holway, D. 1995. Distribution of the Argentine ant (Line-pithema humile) in northern California. Conservation Biology9:1634–1637.

Holway, D. A. 1998. Factors governing rate of invasion: anatural experiment using Argentine ants. Oecologia 115:206–212.

Holway, D. A. 1999. Competitive mechanisms underlying thedisplacement of native ants by the invasive Argentine ant.Ecology 80:238–251.

Holway, D. A., L. Lach, A. V. Suarez, and N. D. Tsutsui. 2002.The causes and consequences of ant invasions. AnnualReview of Ecology and Systematics 33:181–233.

Kot, M., M. A. Lewis, and P. van den Driessche. 1996.Dispersal data and the spread of invading organisms.Ecology 77:2027–2042.

Kuno, E. 1981. Dispersal and the persistence of populations inunstable habitats: a theoretical note. Oecologia 49:123–126.

Lach, L. 2003. Invasive ants: Unwanted partners in ant–plantinteractions? Annals of the Missouri Botanical Garden 90:91–108.

Lagarias, J. C., J. A. Reeds, M. H. Wright, and P. E. Wright.1998. Convergence properties of the Nelder-Mead simplexmethod in low dimensions. SIAM Journal of Optimization 9:112–147.

Levine, J. M., and C. M. D’Antonio. 2003. Forecastingbiological invasions with increasing international trade.Conservation Biology 17:322–332.

Lewis, M. A. 1997. Variability, patchiness, and jump dispersalin the spread of an invading population. Pages 46–69 in D.Tilman and P. Kareiva, editors. Spatial ecology: the role ofspace in population dynamics and interspecific interactions.Princeton University Press, Princeton, New Jersey, USA.

Liebhold, A., J. Halverson, and G. Elmes. 1992. Gypsy mothinvasion in North America: a quantitative analysis. Journalof Biogeography 19:513–520.

Lonsdale, W. M. 1999. Global patterns of plant invasions andthe concept of invasibility. Ecology 80:1522–1536.

Mack, R. N., D. Simberloff, W. M. Lonsdale, H. Evans, M.Clout, and F. A. Bazzaz. 2000. Biotic invasions: causes,epidemiology, global consequences, and control. EcologicalApplications 10:689–710.

Moilanen, A. 2004. SPOMSIM: software for stochastic patchoccupancy models of metapopulation dynamics. EcologicalModelling 179:533–550.

Moody, M. E., and R. N. Mack. 1988. Controlling the spreadof plant invasions: the importance of nascent foci. Journal ofApplied Ecology 25:1009–1021.

Mooij, W. N., and D. L. DeAngelis. 2003. Uncertainty inspatially explicit animal dispersal models. Ecological Appli-cations 13:794–805.

Neteler, M., and H. Mitasova. 2004. Open source GIS: AGRASS GIS approach. Kluwer Academic, Boston, Massa-chuestts, USA.

Neubert, M. G., and H. Caswell. 2000. Demography anddispersal: calculation and sensitivity analysis of invasionspeed for structured populations. Ecology 81:1613–1628.

Passera, L. 1994. Characteristics of tramp species. Pages 22–43in D. F. Williams, editor. Exotic ants: biology, impact andcontrol of introduced species. Westview, Boulder, Colorado,USA.

Petrovskii, S. V., H. Malchow, F. M. Hilker, and E. Venturino.2005. Patterns of patchy spread in deterministic andstochastic models of biological invasion and biologicalcontrol. Biological Invasions 7:771–793.

Pimentel, D., L. Lach, R. Zuniga, and D. Morrison. 2000.Environmental and economic costs of nonindigenous speciesin the United States. BioScience 50:53–64.

Pitt, J. 2008. Modelling the spread of invasive species acrossheterogeneous landscapes. Thesis. Lincoln University, Lin-coln, New Zealand.

Pitt, J., J. Regniere, and S. Worner. 2007. Risk assessment ofthe gypsy moth, Lymantria dispar (L.), in New Zealand basedon phenology modelling. International Journal of Biomete-orology 51:295–305.

Ruxton, G. D., and P. Rohani. 1999. Fitness-dependentdispersal in metapopulations and its consequences forpersistence and synchrony. Journal of Animal Ecology 67:530–539.

Shigesada, N., and K. Kawasaki. 1997. Biological invasions:theory and practice. Oxford University Press, Oxford, UK.

Shigesada, N., K. Kawasaki, and Y. Takeda. 1995. Modelingstratified diffusion in biological invasions. American Natu-ralist 146:229–251.

Stockwell, D. 1999. The GARP modelling system: problemsand solutions to automated spatial prediction. InternationalJournal of Geographical Information Science 13:143–158.

Suarez, A. V., D. T. Bolger, and T. J. Case. 1998. Effects offragmentation and invasion on native ant communities incoastal southern California. Ecology 79:2041–2056.

Suarez, A. V., D. A. Holway, and T. J. Case. 2001. Patterns ofspread in biological invasions dominated by long-distancejump dispersal: insights from Argentine ants. Ecology 98:1095–1100.

Sutherst, R., G. Maywald, T. Yonow, and P. Stevens. 1999.CLIMEX: predicting the effects of climate on plants andanimals. CSIRO, Collingwood, Australia.

Thuiller, W. 2003. BIOMOD—optimizing predictions ofspecies distributions and projecting potential future shiftsunder global change. Global Change Biology 9:1353–1362.

Turner, M. G., Y. G. Wu, W. H. Romme, and L. L. Wallace.1993. A landscape simulation model of winter foraging bylarge ungulates. Ecological Modelling 69:163–184.

Ward, D. F., R. J. Harris, and M. C. Stanley. 2005. Human-mediated range expansion of Argentine ants Linepithemahumile (Hymenoptera: Formicidae) in New Zealand. Socio-biology 45:1–7.

Worner, S. P. 1988. Ecoclimatic assessment of potentialestablishment of exotic pests. Journal of Economic Ento-mology 81:973–983.

Worner, S. P. 1994. Predicting the establishment of exotic pestsin relation to climate. Pages 11–32 in J. L. Sharp and G. J.Hallman, editors. Quarantine treatments for pests of foodplants. Westview Press, Boulder, Colorado, USA.

JOEL P. W. PITT ET AL.1186 Ecological ApplicationsVol. 19, No. 5