The tail of the spatial kernel final

advertisement
1
1
The shape of the spatial kernel and its implications for biological invasions in patchy
2
environments.
3
Tom Lindström, Nina Håkansson and Uno Wennergren
4
Abstract
5
Ecological and epidemiological invasions occur in a spatial context. In the study presented we
6
tested how these processes relate to the distance dependence of spread or dispersal between
7
spatial entities such as habitat patches or infective units. The distance dependence was
8
described by a spatial kernel which can be characterized by its shape, quantified by kurtosis, and
9
width, quantified by the kernel variance. We also introduced a method to analyze or generate
10
non randomly distributed infective units or patches as point pattern landscapes. The method is
11
based on Fourier transform and consists of two measures in the spectral representation;
12
Continuity that relates to autocorrelation and Contrast that refers to difference in density of
13
patches, or infective units, in different areas of the landscape. The method was also used to
14
analyse some relevant empirical data where our results are expected to have implications for
15
ecological or epidemiological studies. We analyzed the distributions of large old trees (Quercus
16
and Ulmus) as well as the distributions of farms (both cattle and pig) in Sweden. We tested the
17
invasion speed in generated landscapes with different amount of Continuity and Contrast. The
18
results showed that kurtosis, i.e. the kernel shape, was not important for predicting the invasion
19
speed in randomly distributed patches or infective units. However, depending on the
20
assumptions of dispersal, it may be highly important when the distribution of patches or
21
infective units deviates from randomness, in particular when the Contrast is high. Hence speed
2
22
of invasions and spread of diseases depends on its spatial context through the spatial kernel
23
intertwined to the spatial structure. This implies high demands on the empirical data; it
24
requires knowledge of both shape and width of the spatial kernel as well as spatial structure of
25
patches or infective units.
26
27
1. Introduction
28
Both ecological and epidemiological studies are concerned with invasion of organisms. The
29
mechanism and dynamics of invasion are essential components in numerous specific topics.
30
These include recolonization of habitats (Lubina & Levin 1988, Seabloom et al. 2003), migration
31
in response to climate variations (Clark 1998, Walters et al. 2006), spread of human and
32
livestock diseases (Fergusson 2001, Boender et al. 2007) and invasion of alien species (Skellam
33
1951, Urban 2008). The rate of the invasion will largely be determined by the dispersal or spatial
34
contact pattern that allows for transmission. Commonly, this is described with a spatial kernel
35
(Clark 1998, Tildesley et al. 2008). In this paper we investigate the role of kernel characteristics
36
and how this is affected by the spatial arrangement of the habitats or infective units.
37
38
The kernel may be characterized by the 2nd and 4th moment (Clark 1998, Mollison 1991,
39
Lindström et al. 2008). The 2nd moment is more commonly known as variance (ν) or squared
40
displacement and is a measure of the width of the kernel. Kurtosis (κ), a dimensionless quantity
41
defined as the 4th moment dived by the square of the 2nd moment, describes the shape. For
3
42
animal and plant dispersal a random walk or correlated random walk might be assumed which
43
will result in a kernel according to a Gaussian distribution (Turchin 1998), where κ=3 or κ=2 for
44
one and two dimensional kernels respectively. In this study we will consider two dimensional
45
kernels since most ecological an epidemiological dynamics occur in at least two dimensional
46
landscapes. These kernels are also often denoted dispersal kernels in ecological studies yet we
47
will use the notation spatial kernel throughout the paper.
48
49
For epidemiological studies, assumptions regarding the kurtosis of the kernel should be made
50
from knowledge of how transmission occurs. If transmission arises through direct contact the
51
kernel should be based on the movement behaviour of the hosts. For many pathogens however,
52
transmission is mediated via a vector. If the movement of the vector resembles a random walk
53
it may be a fair assumption to model transmission with a Gaussian kernel (Gerbier et al. 2008)
54
but outbreak data (Fergusson et al. 2001) and studies of pathways that may mediate
55
transmission (Lindström et al. 2009) often reveal highly leptokurtic distributions. Empirical
56
studies show that dispersal most commonly deviate from Gaussian distributions. Usually a
57
leptokurtic (κ>2) distribution is observed for both plants (Kot et al. 1996, Skarpaas & Shea 2007)
58
and animals (Schweiger et al. 2004, Walters et al. 2006), implying a peak in density at short
59
distances but at the same time a fat tail, indicating fairly frequent long distance dispersers. A
60
number of explanations has been proposed that explain leptokurtic dispersal, including
61
population differences in dispersal abilities (Fraser & Bernatchez 2001), temporal variation in
62
the diffusion constant (Yamamura et al. 2007) and loss of individuals during dispersal (Schneider
63
1999).
4
64
65
If a Gaussian dispersal in a homogenous and continuous space is assumed, the invasion can be
66
modeled as a reaction diffusion process and the speed of the invasion will be proportional to
67
the variance (Skellam 1951). Deviations from Gaussian kernels may still tend to the same speed,
68
i.e. determined by the variance of the kernel, as long as the tail is exponentially bounded
69
(Mollison 1977, Clark 1998). If however the density in the distributions tail is higher than an
70
exponentially decreasing function (for which κ>4 in two dimensions) the invasion speed is
71
expected to accelerate. Recent work has moved further from invasion speed assuming
72
homogenous and continuous space and turned the attention to heterogeneous landscapes
73
(Smith et al. 2002, Urban et al. 2008). In this paper we take another step and focus on invasion
74
of organisms in environments where the habitats or infective units are best represented as
75
discrete entities with a fixed spatial location. Examples of this are studies of livestock epidemics
76
(Keeling 2001, Boender et al. 2007) and ecological invasions where habitats and infective units
77
are considered as isolated patches surrounded by a hostile matrix (as is done in metapopulation
78
studies). Throughout the paper we will refer to such habitats and infective units as patches.
79
80
Our aim in this paper is to explore the role of kurtosis, κ, and variance, ν, of spatial kernels on
81
the speed of biological invasion in patchy environments. Such results may support studies and
82
the predictive power of estimated speed of invasions. We expect that this may also depend on
83
the spatial pattern of focal entities. We therefore introduce a method to incorporate spatial
84
aggregation in point patterns using spectral density. Hence we test whether the role of kurtosis
5
85
and variance is dependent on spatial structures as aggregation in patchy landscapes. To
86
exemplify what spatial patterns may be found, we also analyze relevant point pattern data with
87
a method developed from the analysis given by Mugglestone & Renshaw (2001).
88
89
2. Materials and Methods
90
2.1 Kernel variance and kurtosis
91
In this study we modeled the spatial kernel with a generalized normal distribution (Nadarajah
92
2005). In Lindström et al. (2008) this is extended to two dimensions for symmetrical kernels.
93
Kernel density is given by
94
P D  
e
 a
d
b
S
(2.1)
95
Where d is the distance and S is a normalizing constant which in two dimensions is given by
96
S
97
Parameters a and b determines ν and κ of the kernel. For two dimensions these are given by
98
4
 
b
  a2  
2
 
b
b
.
2a1 b 
(2.2)
(2.3)
6
99
6 2
    
b
b
     2
  4 
   
  b 
(2.4)
100
101
Hence the kernel density P(D) can be completely defined by kurtosis and variance; examples are
102
given in figure 1.
103
104
2.2 Generating and analyzing neutral point pattern landscapes
105
Keith (2000) defined neutral landscapes for lattices as models where the value at any point in
106
the landscape can be considered random and pointed out that this does not exclude models
107
with spatial autocorrelation. Such landscapes are then intrinsically stationary while completely
108
random surfaces with no autocorrelation are second-order stationary (Cressie, 1993). This
109
neutral landscape definition may also be applied to point pattern landscapes where the
110
distribution of points may deviate from random as long as the exact position of a point cannot
111
be predicted. We used a set of such landscapes to test the effect of kernel characteristic under
112
different patterns of spatial aggregation. We refer to these as Neutral Point Pattern Landscapes
113
(NPPL). We will use spectral density functions to handle the autocorrelation and in the first
114
section we show how spectral density functions can be applied to point pattern.
115
7
116
2.2.1 Spectral density and point pattern
117
Spectral density has been used frequently for time series and lattice data (see Mugglestone &
118
Renshaw 1996 for relevant references). The basis of the spectral method follows from the
119
Fourier theorem where it is stated that continuous time series or surfaces (and also higher
120
dimensions); can be represented by a combination of sine waves with different frequencies and
121
amplitudes. Analysis of the time series or surfaces can then instead be performed on the sine
122
waves. Here we present two measurements required to capture the spatial point pattern.
123
124
Continuity (γ) is a measure of spatial autocorrelation over multiple scales. It is a measure of the
125
Power Spectral Density Function (PSDF). Large values of γ means that nearby areas have similar
126
density. It is a measure of the relationship between frequencies and amplitudes assuming a
127
linear relationship after logarithmic transformation of frequencies and amplitudes. Hence γ is
128
given by the slope from a linear regression fitted to the log(frequency) vs. log(amplitude). In the
129
analysis of time series this measure is termed 1/f noise. We are interested in analyzing and
130
generating point pattern data and hence there is a methodological part for the transformation
131
between point pattern and the continuous representation. Mugglestone & Renshaw (1996)
132
have described an efficient way to calculate the PSDF for point pattern data (equation 2.5).
133

 

PSDFPP    cos( K  p xi )     sin( K q yi ) 
 i
  i

134
(μp,μq) are frequencies for Fourier transform of grid data (Mugglestone & Renshaw 1996), (xi,yi)
135
are coordinates in the point pattern, and K is a constant that determines the number of
2
2
(2.5)
8
136
frequencies. Mugglestone and Renshaw (1996) argue that no more frequencies should be used
137
than there are points in the point pattern, N, to keep them independent. Therefore we chose K
138
to be
139
γpp for the point pattern using the same method as for continous data, but with the power
140
spectral density function for point pattern data PSDFPP (see Mugglestone & Renshaw 1996).
N p 2 rounded down. And the number of frequencies used is nF =2K×2K. We measure
141
142
The second measure is Contrast (δ), which is a normalized measure of density dispersion. Large
143
values of δ reflect a large difference between sparse and dense areas. We measure δ in the
144
frequency domain as the coefficient of variation, CV, over point patterns. To formulate an
145
equation for that measure we first have to relate it to variance in the continuous case. That is
146
how variance in a continuous landscape can be measured using spectral representation. Note
147
that when time series and surfaces are represented by sine functions it is solely the amplitudes
148
of the sine functions that determines the variance, as shown in equation 2.6 below. As an
149
example we use a time series α(t) with its Fourier transform A(μ) where μ is the frequency of a
150
sine function.
151
var  (t )  
152
The mean is represented by the amplitude in origin and hence the coefficient of variation,
153
standard deviation divided by the mean, is
154
CV 
1
M2
 | A( ) |
2

1
1
A2 (0) 
mean( PSDF  PSDF (origo ))
2
M
M
M
1
mean( PSDF  PSDF (origo ))
PSDF (origo ) M
(2.6)
(2.7)
9
155
Switching to point pattern by equation 2.5 we end up with an equation of the Contrast
156
measure:
157
δ = CVPP 
NF
PSDFPP (origo )
1
mean( PSDFPP  PSDFpp (origo ))
NF
(2.8)
158
159
2.2.2 Generating neutral point pattern landscapes
160
To get NPPL with given characteristics we generated lattice landscapes of size m×m. The density
161
defines the probability of a point in the landscapes. We first generated 2-dimensional 1 / f
162
noise (denoted LG) using a method similar to that presented by Halley et al. (2004). Hence this is
163
still a representation of a lattice landscape not a point pattern. The values in LG are normally
164
distributed and since this may include negative values it is not suitable for describing
165
probabilities. While this could be solved by truncating we found that it would not allow for
166
generation of sufficiently high values of δ. We therefore transformed LG using spectral mimicry.
167
This method is defined by Cohen et al. (1999) and has been used when applying Fourier series
168
to time series analysis. Cohen et al. presents the method for transformation to a series with
169
normally distributed values with a specific mean and variance. We instead transformed LG to LΓ
170
using a Γ distribution (which contains no values <0) with mean=1/m2 and coefficient of variation
171
δL. Point locations were distributed according to the probabilities given by LΓ. Examples of the
172
method are given in figure 2.
173
L
-
10
174
While γ and δ of the spectral point pattern is determined by γL and δL, they are altered by both
175
the Γ-transformation of the grid values and the distribution of points. Hence we measured these
176
quantities in the generated landscape (see method given above). The relationship between
177
spectral point pattern values of γ and δ used in the study and the γL and δL required to generate
178
them was found iteratively. Furthermore, we found that the linear relationship in the power
179
spectra was maintained better for large grids (values of m) and we used m=2000. The
180
autocorrelation parameter Continuity generates a general aggregation pattern while the
181
variance within the system is reflected by the Contrast parameter. Some examples of the NPPL
182
generated with the method can be found in figure 2. High Contrast parameter will impose more
183
isolated clusters of aggregated points onto the aggregation structure defined by the Continuity
184
parameter; compare the two rightmost examples in figure 2.
185
186
2.2.3 Analysing neutral point pattern landscapes
187
We also tested the NPPL model by analyzing empirical point pattern data. The empirical data
188
consists of tree distributions and the distributions of farms in Sweden. The tree data was
189
provided by the Östergötland County Administrative Board. It is the result of a massive
190
inventory of large and old trees, (Länsstyrelsen Östergötland 2009).The locations and
191
production type of farms was supplied by the Swedish Board of Agriculture, and more details on
192
the data can be found in Nöremark et al. (2009) and Lindström et al. (2009).
193
11
194
2.3 Simulation
195
The effect of κ and ν on invasion was estimated by simulating invasions in NPPL with discrete
196
time and the combination of parameters given by table 1. Some combinations of δ and γ were
197
not possible to generate (see figure 3-5). Starting at a random patch, we simulated invasions
198
with 200 replicates of each parameter combination, for both absolute and relative distance
199
dependence. To reduce edge effects, we arranged the landscape such that the starting point
200
was located in the centre of the NPPL, which is possible due the periodic nature of the Fourier
201
transform.
202
203
2.3.1 Probability of colonization
204
The probability of colonization from one occupied patch to an unoccupied one can be modeled
205
differently, corresponding to different assumptions regarding dispersal and contact. These
206
probabilities are used for simulation studies of invasion and spread of disease. First, one may
207
assume that the probability is only dependent on the distance between the two patches, dij, in
208
which case the probability, P, of patch i becoming occupied by dispersal from patch j within one
209
time step is given by
210
P(Ot+1(i) = 1| Ot(j) = 1, Ot(i)= 0) = RP(dij)
(2.9)
211
where Ot(k) is equal to one if patch k is occupied at time t and equal to zero if it is unoccupied,
212
and R is a measure of growth rate. This modeling approach assumes that the probability of
213
colonization from one occupied patch to an unoccupied one is independent of the existence and
12
214
position of other patches. We will refer to this as absolute distance dependence. Alternatively
215
one may assume that the colonization potential of all occupied patches is the same. In that case,
216
equation 2.9 and also equation 2.1, for colonization from patch j is normalized by summation
217
over all patches k≠j:
218
S  e
N 1
( d kj / a )b
(2.10)
k 1
219
where N is the number of patches. We refer to this as relative distance dependence and all
220
patches will have the same colonizational potential regardless whether it’s an isolated patch or
221
positioned within a dense area.
222
223
2.3.2 Simulation outputs and analysis
224
Our interest was to estimate the importance of κ for biological invasions. Two measurements of
225
invasion speed were analyzed. First we investigate the time, Τl, to reach fixed proportions, pl, of
226
occupied patches. We used pl = 10%, 50% and 90 %, to get estimates at different stages of the
227
invasion. Secondly, we also analyzed the speed, Ψ, of spatial spread, defined as
228
  dl t
(2.11)
l
229
where dl is a fixed distance and tl is the number of time steps required to reach that distance. In
230
this paper we present the results for dl=0.25 (given relative to the unit square). At this distance,
231
the influence of the edge effect is considered very small. For Ψ, we analyzed the results of both
232
absolute and relative distance dependence.
13
233
234
The results were analyzed with an ANOVA (type three) for each combination of landscape
235
parameters, with the output parameters as dependent variable and ν and κ as categorical
236
predictors. Since the outputs showed non normal residuals, a Box-Cox transform (Box & Cox
237
1964) was performed for each analysis. The exact value of γ and δ varies between replicates and
238
therefore were included as continuous co-variables. The relative effect of kurtosis was
239
calculated Eκ=MSκ/(MSκ + MSν) where MSκ and MSν are the mean sum of squares of κ and ν,
240
respectively.
241
242
3. Results
243
3.1 Simulations of invasion in neutral point pattern landscapes
244
Our results show that kurtosis of the dispersal kernel is generally a factor that has significant
245
effect on the speed of invasion (figure 3) compared to its variance, but the effect varied
246
dependent on the spatial structure of patches. Black areas in figures 3-5 indicate low
247
importance of kurtosis, and this is consistently found for random NPPLs (δ=1 and γ=0),
248
indicating that the shape of the kernel is of little importance in when patches are randomly
249
distributed. The trend was found for both densities tested but more prominent for dense NPPLs.
250
Also, the relative importance changes during the course of invasion (figure 5) with the most
251
prominent effect found during the initial phases of invasion. The general pattern is that the
252
Contrast was the characteristic that mainly shifted the importance of kurtosis while Continuity
14
253
had less effect. This can be seen in figure 3 and 5 as a more evident shift left-right than up-
254
down.
255
256
Figure 4 show that kurtosis had less importance when invasion was modeled with relative
257
distance dependence (as described in section 2.3.1). We have analyzed invasions with relative
258
distance dependence for both speed and times to fixed proportions, yet all these results also
259
showed no effect of kurtosis and hence are only represented by figure 4. The results suggest
260
that in studies of organisms corresponding to these assumptions it may be sufficient to estimate
261
the variance and disregard the kernel kurtosis.
262
263
3.2 Examples of real neutral point pattern landscapes (NPPL)
264
To demonstrate the use of NPPL and what characteristics may be found in areas where our
265
results have impact we analyzed relevant data with the method given in section 2.2, figure 6.
266
We analyzed the distribution of two tree species, oak (Quercus) and elm (Ulmus). Especially old
267
trees of these species are important habitats for saproxylic insects. Many of these are
268
endangered and limited dispersal has been proposed to be a major explanation (Ranius 2006,
269
Hedin et al. 2008). Both tree species are also host for many lichens (Jüriado et al. 2009) and
270
Ulmus is in addition relevant for epidemiological studies because of the spread of Dutch elm
271
disease (Ophiostoma ulmi) (Gilligan & van den Bosch 2008). We also examined the spatial
272
distribution of pig and cattle farms in southern Sweden. The spatial distributions of farms are
15
273
known to be essential for possible outbreak of livestock diseases (Boender et al. 2007). The
274
distribution of the analyzed data and their estimated values of Contrast (δ) and Continuity (γ)
275
(figure 6) indicate that the NPPL model is applicable also for analyzing empirical data. All the
276
relations, 1/fγ, in the spectral representation of the point patterns are consistent with the linear
277
assumptions of the assumptions of γ. The Continuities in the point patterns are all fairly close to
278
one, but the Contrast measures are more variable ranging from 1.29 for cattle farms to 4.9 for
279
elm trees.
280
281
Discussion
282
Ecological and epidemiological processes occur in a spatial context. Our understanding of, and
283
possibility to predict and control, those processes are dependent on how well we may describe
284
this spatial context. This includes both the spatial environment and the spatial behaviour of the
285
process itself. In our work we have studied invasion, using a patchy landscape as the spatial
286
environment and a family of spatial kernels for modeling of the spatial behaviour. The novel
287
part of our study is to release commonly used assumption of homogeneous and continuous
288
spatial structures and instead focusing on the process in patchy landscapes. We thereby include
289
the interplay between the spatial kernel and patchy landscapes. Our result indicates that
290
depending on the assumptions of distance dependence, this interplay may be very strong and
291
whether kurtosis has evident effect on the invasion speed depends on the spatial structure of
292
the patches. More specifically the importance of kurtosis of the spatial kernel is measured
293
relative the variance of the spatial kernel. Kurtosis is a measure of the shape of the spatial
16
294
kernel and thereby our results emphasize the importance of correct representation of this
295
kernel feature. Since a vast area of topics such as colonization of habitats, migrations in
296
response to climate variations, and spread of diseases occurs in a spatial context where spatial
297
structure is an obvious component (Kareiva & Wennergren 1995) we expect that our results
298
may have implications on direct applications and on future research and investigations. That the
299
importance of kurtosis differs depending on landscape structure implies that both speed of
300
invasion, and the methodology to estimate it, may differ between landscapes. In some
301
landscapes it may suffices with the variance of the spatial kernel while other landscapes enforce
302
assessments of the kurtosis of the kernel. Furthermore it also stresses the importance of
303
developing empirical methods that correctly captures landscape structure. In this study we
304
evaluated both a direct spatial measure of speed and time to a specified proportion colonized,
305
which corresponds to slightly different questions regarding invasions. The trends are similar
306
(figures 3 and 5) and hence our results have implications for studies focusing on either of these
307
measures.
308
309
In this study we have used an admittedly simplified colonization model to represent both
310
ecological invasions and spread of disease. That analogy between colonization in a
311
metapopulation and spread of disease has been discussed and used in disease modeling, for
312
example by Vernon & Keeling (2009) in their study on spread of disease in a network
313
representation. As pointed out in their study, the assumptions of a simplified colonization
314
model may be to crude to capture the dynamics of any real invasion, but it allows for testing the
315
effect contact of the contact structure. Our aim was to reduce the system such that the main
17
316
characteristics in the study was landscape and dispersal and we excluded recovery/extinction
317
and within patch dynamics such as density dependence. We argue that our results regarding the
318
importance of kurtosis and the interaction with landscape features would hold also for more
319
realistic models.
320
321
Both variance and kurtosis relates to long distance dispersal (LDD). Studies of LDD commonly
322
defines this (see e.g. Nathan 2006) as either dispersal events beyond some fixed distance or
323
some percentile of the tail. As these distances or percentiles are chosen by the researchers, the
324
measures of LDD are to some extent subjective and comparison between studies may be
325
problematic. We argue that dispersal is better described by analysis of the spatial kernel and its
326
characteristics. From analysis of dispersal in continuous space it has been shown in several
327
studies (for example Yamamura 2004, van den Bosch et al. 1990, Kot et al. 1996) that the fat tail
328
of the spatial kernel, reflected by kurtosis in our study, has an impact on invasion speed. To our
329
knowledge this is however the first study that focuses on the importance of the kernel
330
characteristics for invasions in patchy environments, using both random and non random
331
distribution of patches. By describing the kernel by variance and kurtosis, and test the effect of
332
these, it is possible to analyze if and when these characteristics are important to estimate. This
333
is made possible by the use of a kernel function where these characteristics are possible to
334
control. In previous studies (Lindström et al. 2008) we have studied the effect of kernel
335
characteristics on the population distribution and found that kurtosis was not important and
336
hence did not have to be estimated. Here we find that the importance is dependent on the
337
spatial characteristics.
18
338
339
An invasion in a patchy landscape may spread over the landscape fairly different compared to
340
the diffusive processes in a more homogenous landscape. In a patchy landscape, as illustrated in
341
figures 2 and 6, some of the last patches to be colonized are not necessarily the most distant to
342
the initial point. This is why we chosen to evaluate both a direct spatial measure of speed and
343
time to specified proportion colonized. For example, the spatial speed measure applies to when
344
a disease will reach a specific area or country, while the proportion colonized applies to how
345
much will be infected within an area.
346
347
We also used two different assumptions regarding colonization from an occupied patch,
348
referred to as absolute or relative distance dependence. These correspond to different
349
assumptions of the organisms dispersal pathways. Absolute distance dependence would best
350
describe a colonization process of an organism with large amount of propagules and passive
351
dispersal (e.g. by wind). In such case, the probability of one patch colonizing another is
352
independent of the probability of colonization of other patches. The relative distance
353
dependence corresponds to colonization by actively dispersing individuals without mortality or
354
disease spread between farms via human activities if the number of contacts of infected
355
premises is independent of its location (e.g. the number of animal transports may be expected
356
to be the same for geographically isolated farms and those in dense areas – Lindström et al.
357
2009). Many colonization processes would be a mixture of the two mechanisms.
358
19
359
The review of Hawkes (2009) introduces a set of principles regarding the relation between
360
movement behaviour, dispersal and population processes. It is pointed out that a more complex
361
spatial kernel is expected when there is individual variation in movement behaviour within the
362
population and such variation also promotes leptokurtic spatial kernels. Another principle of
363
Hawkes (2009) is that for some species we may even anticipate changes of movement
364
behaviour during dispersal as a result of spatial structure. Such a feedback will interweave the
365
landscape and the spatial kernel even more. This feedback principle somewhat twist the
366
question into what movement behaviour is optimal in a given spatial structure? Our results may
367
then give some guidance since the importance of kurtosis reflects possible selection pressure in
368
relation to invasion.
369
370
The interaction between landscape structure and spatial kernel emphasizes the need of reliable
371
estimates of these features. Kernels with variable kurtosis and variance, such as the two
372
dimensional generalized normal distribution given by Lindström et al. (2008), may form a basis
373
for assessing the spatial kernel while the Neutral Point Pattern Landscape method that we
374
introduced here may capture these structures for point patterns representation of non random
375
landscapes. Spectral representation has become increasingly important to spatial data analysis.
376
It is especially advantageous when studying spatial dependence in point pattern processes since
377
it may capture more complex dependencies than other techniques, even anisotropy
378
(Schabenberger & Gotway 2005). Keitt (2000) introduced spectral methods to landscape
379
ecology and presented neutral landscapes for lattice models. By developing the point pattern
380
representation by Mugglestone & Renshaw (1996) and the spectral mimicry of time series by
20
381
Cohen et al. (1999) we introduce the neutral point pattern landscape model (NPPL). The
382
methodology includes both a continuity measure related to autocorrelation and a contrast
383
measure that relates to the variance in the landscape. The contrast measure can be viewed as a
384
measure of proportion of points within aggregates, see rightmost examples in figure 2 while
385
continuity is a measure of spatial autocorrelation. This effectively means that the structure is
386
locally similar and may explain the decrease in the effect of kurtosis with higher continuity. For
387
random landscapes, the effect of kurtosis is very small and for high continuity the distribution of
388
patches may locally resemble a random distribution patches. High contrasts result in groups of
389
locally connected but regionally isolated patches and colonization between such isolated groups
390
are rare when dispersal is limited (i.e. low variance). The occurrence of rare but long distant
391
events (described by the tail of the leptokurtic kernels) enables such events.
392
393
The distributions of trees and Swedish farms indicate that the NPPL analyses may capture
394
important structures in vast areas of empirical data. The analysis reveals contrast measures in
395
the interval 1.3-4.9 which indicate landscape structures where kurtosis clearly matters in our
396
sumulations. The analyses also show that the distribution of trees seem to have a higher
397
contrast measure than farms. While it is not the aim of this paper to compare these two
398
systems, our results does suggest that kurtosis is more important in studies focusing on
399
dispersal between trees than for epidemiological studies on disease spread between farms.
400
Also, the analyzed examples show that the assumptions of a linear relationship between
401
log(frequency) and log(amplitude) appears to be a god fit. This means that there is a spatial self
402
similarity over scales, which is the definition of a fractal process (Halley et al. 2004). There are
21
403
however many underlying processes for the distributions of these point patterns. Because of
404
this, and reminded of Bakers lemma given in Halley et al. (2004) as “Even an elephant appears
405
linear if plotted on log–log axes”; we refrain from drawing conclusions on the fractal properties
406
of these distributions. Instead we conclude that the analyzed patterns justify the assumptions of
407
the NPPL used in this study.
408
409
The impact of the spatial aspect in ecological and epidemiological theory is especially apparent
410
and obvious in the light of invasions and spread of disease. The spatial aspect has two
411
components, the landscape and the dispersal of organisms. We have showed that these two are
412
entwined when the landscape structure is complex and it’s usually not enough to assess the
413
variance of the dispersal kernel. Instead the specific shape of the spatial kernel becomes
414
important. Yet, its importance is dependent on the landscape structure and thereby there is a
415
need to measure this structure. These theoretical results point out that studying ecological and
416
epidemiological spread in a spatial context puts a lot of pressure on empirical details on
417
dispersal, contact patterns and landscape structures.
418
419
Acknowledgement
420
We thank the Swedish Contingency Agency (MSB) for funding and also both the County Administration
421
Board of Östergötland and the Swedish Board of Agriculture for supplying data.
422
22
423
References
424
Boender, G. J., Meester, R., Gies, E. & De Jong, M. C. M. 2007 The local threshold for
425
geographical spread of infectious diseases between farms. Prev. Vet. Med. 82, 90-101.
426
Box, G. E. P. & Cox, D. R. 1964 An analysis of transformations. J. Roy. Stat. Soc. B. 26, 211-252.
427
Clark, J. S. 1998 Why trees migrate so fast: confronting theory with dispersal biology and the
428
paleorecord. Amer. Nat. 152, 204-224.
429
Cohen J. E., Newman C. M., Cohen A. E., Petchey O. L. & Gonzalez A. 1999 Spectral mimicry: A
430
method of synthesizing matching time series with different Fourier spectra. Circ. Syst. Signal
431
Process. 18, 431-442.
432
Cressie N. 1993 Statistics for spatial data revised edition. Chapter 2.2.1. USA: John Wiley & Sons,
433
Inc.
434
Fraser, D. J. & Bernatchez, L. 2001 Adaptive evolutionary conservation: towards a unified
435
concept for defining conservation units. Mol. Ecol. 10, 2741-2752.
436
Ferguson, N. M., Donnelly, C. A. & Anderson, R. M. 2001 The foot-and-mouth epidemic in Great
437
Britain: Pattern of spread and impact of interventions. Science 292, 1155-1160.
438
Gerbier, G., Baldet, T., Tran, A., Hendrickx, G., Guis, H., Mintiens, K., Elbers, A. & Staubach, C.
439
2008 Modelling local dispersal of bluetongue serotype 8 using Random walk. Prev. Vet. Med. 87,
440
119–130.
23
441
Gilligan, C. A. & van den Bosch, F. 2008 Epidemiological models for invasion and persistence of
442
pathogens. Annu. Rev. Phytopathol. 46, 385-418.
443
Halley, J. M., Hartley, S., Kallimanis, A. S., Kunin, W. E., Lennon, J. J. & Sgardelis, S. P. 2004 Uses
444
and abuses of fractal methodology in ecology. Ecol. Lett. 7, 254-271.
445
Hawkes, C. 2009 Linking movement behaviour, dispersal and population processes: is individual
446
variation a key? J. Anim. Ecol. 78, 894-906.
447
Hedin, J., Ranius, T., Nilsson, S. G. & Smith, H. G. 2008 Restricted dispersal in a flying beetle
448
assessed by telemetry. Biodivers. Conserv. 17, 675–684.
449
Jüriado, I., Liira, J. & Paal, J. 2003 Epiphytic and epixylic lichen species diversity in Estonian
450
natural forests. Biodivers. Conserv. 12, 1587–1607.
451
Kareiva, P. & Wennergren, U. 1995. Connecting landscape patterns to ecosystem and
452
population processes. Nature 373, 299-302.
453
Keeling, M. J., Woodhouse, M. E., Shaw, D. J. & Matthews, L. 2001 Dynamics of the 2001 UK foot
454
and mouth epidemic: Stochastic dispersal in a dynamic landscape. Science 294, 813-817.
455
Keitt, T. H. 2000 Spectral representation of neutral landscapes. Landscape Ecol. 15, 479-493.
456
Kot, M., Lewis, M. A. & van den Driessche, P. 1996 Dispersal data and the spread of invading
457
organisms. Ecology 77, 2027-2042.
458
Lindström, T., Håkansson, N., Westerberg, L. & Wennergren, U. 2008 Splitting the tail of the
459
displacement kernel shows the unimportance of kurtosis. Ecology 89, 1784–1790.
24
460
Lindström, T., Sisson, S. A., Nöremark, M., Jonsson, A. & Wennergren, U. 2009 Estimation of
461
distance related probability of animal movements between holdings and implications for
462
disease spread modeling. Prev. Vet. Med.91, 85-94.
463
Lubina, J. A. & Levin S. A. 1988 The Spread of a Reinvading Species: Range Expansion in the
464
California Sea Otter. Amer. Nat. 131, 526-543.
465
Länsstyrelsen Östergötland 2009. Skyddsvärda träd i Östergötland – 1997-2008. Rapport
466
2008:13
467
Mollion, D. 1977 Spatial contact models for ecological and epidemic spread. J. Roy. Stat. Soc. B.
468
39, 283-326.
469
Mollion, D. 1991 Dependence of epidemic and population velocities on basic parameters. Math.
470
Biosci. 107, 255-287.
471
Mugglestone, M. A. & Renshaw, E. 1996 A practical guide to the spectral analysis of spatial point
472
processes. Comput. Sta.t Data An. 21, 43-65.
473
Mugglestone, M. A. & Renshaw, E. 2001 Spectral tests of randomness for spatial point patterns.
474
Environ. Ecol. Stat. 8, 237-251.
475
Nadarajah, S. 2005 A generalized normal distribution. Appl. Statist. 32, 685-694.
476
Nathan, R. 2006 Long-distance dispersal of plants. Science 313, 786-788.
25
477
Nöremark, M., Håkansson, N., Lindström, T., Wennergren, U. & Sternberg Lewerin, S. 2009
478
Spatial and temporal investigations of reported movements, births and deaths of cattle and pigs
479
in Sweden. Acta Vet. Scand. 51:37.
480
Ranius, T. 2006 Measuring the dispersal of saproxylic insects: a key characteristic for their
481
conservation. Popul. Ecol. 48, 177–188.
482
Schabenberger O. & Gotway C. 2005. Statistical methods for spatial data analysis. Chapter 2.5.7.
483
London: Chapman & Hall.
484
Schneider, J. C. 1999 Dispersal of a highly vagile insect in a heterogeneous environment. Ecology
485
80, 2740-2749.
486
Schweiger, O., Frenzel, M. & Durka, W. 2004 Spatial genetic structure in a metapopulation of
487
the land snail Cepaea nemoralis (Gastropoda: Helicidae). Mol. Ecol. 13, 3645–3655.
488
Seabloom, E. W., Borer, E. T., Boucher, V. L., Burton, R. S., Cottingham, K. L., Goldwasser, L.,
489
Gram, W. K., Kendall, B. E. & Micheli, F. 2003 Competition, seed limitation, disturbance, and
490
reestablishment of California native annual forbs. Ecol. Appl. 13, 575-592.
491
Skarpaas, O. & Shea, K. 2007 Dispersal patterns, dispersal mechanisms, and invasion wave
492
speeds for invasive thistles. Am. Nat. 170, 421-430.
493
Skellam, J. G. 1951 Random dispersal in theoretical populations. Biometrika 38, 196-218.
494
Smith, D. L., Lucey, B., Waller, L. A., Childs, J. E. & Real, L. A. 2002 Predicting the spatial dynamics
495
of rabies epidemics on heterogeneous landscapes. Proc. Natl. Acad. Sci. 99, 3668-3672.
26
496
Tildesley, M. J., Deardon, R., Savill, N. J., Bessell, P. R., Brooks, S. P., Woolhouse, M. E., Grenfell,
497
B. T. & Keeling, M. J. 2008 Accuracy of models for the 2001 foot-and-mouth epidemic. Proc. R.
498
Soc. B. 275, 1459-1468.
499
Turchin, P. 1998 Quantitative Analysis of Movement, Sinauer Associates, Sunderland, MA.
500
Urban, M. C., Phillips, B. L., Skelly, D. K. & Shine, R. 2008 A toad more traveled: The
501
heterogeneous invasion dynamics of cane toads in australia. Am. Nat. 171, E134-E148.
502
Van den Bosch, F., Metz, J. A. J. & Diekmann O. 1990 The velocity of spatial population
503
expansion. J. Math. Biol., 28, 529-565.
504
Vernon, M. C. & Keeling, M. J. 2009 Representing the UK’s cattle herd as static and dynamic
505
networks. Proc. Roy. Soc. B. 276, 469-476.
506
Walters, R. J., Hassall, M., Telfer, M. G., Hewitt, G. M. & Palutikof, J. P. 2006 Modelling dispersal
507
of a temperate insect in a changing climate. Proc. R. Soc. B., 273, 2017-2023.
508
Yamamura K. 2004 Dispersal distance of corn pollen under fluctuating diffusion coefficient.
509
Popul. Ecol. 46, 87-101.
510
Yamamura K., Moriya, S., Tanaka, K., & Shimizu, T. 2007 Estimation of the potential speed of
511
range expansion of an introduced species: characteristics and applicability of the gamma model.
512
Popul. Ecol. 49, 51-62.
27
513
514
Table 1. Input parameters of the simulations and values used
515
516
Figure 1. (a) Probability densities at distance from source for κ=4 and ν=0.0025 (dashed),
517
ν=0.005 (solid) and ν=0.01 (dotted) respectively. (b) Probability densities at distance from source
518
for ν=0.005 and κ=2 (dashed), κ=4 (solid) and κ=6 (dotted) respectively. Embedded axis’ shows
519
same as major axes but at larger distances and with logarithmic y-axis.
520
521
Figure 2. Examples of (first row) spatial distributions of patches used in the simulation study and
522
(second row) their corresponding spectral densities with estimated Continuity (γ) and Contrast
523
(δ).
524
525
Figure 3. The relative importance of κ for the speed of spatial spread with absolute distance
526
dependence under different landscape parameters (Contrast - δ and Continuity - γ). Black
527
indicates that κ is unimportant and instead the variance of the dispersal kernel determines the
528
speed while white areas indicate that κ is highly important. The relative importance was
529
calculated as Eκ=MSκ/(MSκ + MSν) where MSκ and MSν are the mean sum of squares of κ and ν,
530
respectively, from ANOVAs for each combination of δ and γ. Areas where the grid appears (for
531
low δ and high γ) are point pattern landscape not possible to generate with present method.
532
28
533
Figure 4. The relative importance of kurtosis, κ, for the speed of spatial spread with relative
534
distance dependence under different landscape parameters (Contrast - δ and Continuity - γ).
535
Black indicates that κ is unimportant and instead the variance of the dispersal kernel determines
536
the speed while white areas indicate that κ is highly important. The relative importance was
537
calculated as Eκ=MSκ/(MSκ + MSν) where MSκ and MSν are the mean sum of squares of κ and ν,
538
respectively, from ANOVAs for each combination of δ and γ. Areas where the grid appears (for
539
low δ and high γ) are point pattern landscape not possible to generate with present method.
540
541
Figure 5. The relative importance of κ for time of invasion to reach proportions (pl=0.1, 0.5 and
542
0.9) of occupied patches with absolute distance dependence under different landscape parameters
543
(Contrast - δ and Continuity - γ). Black indicates that κ is unimportant and instead the variance of
544
the dispersal kernel determines the speed while white areas indicate that κ is highly important.
545
The relative importance was calculated as Eκ=MSκ/(MSκ + MSν) where MSκ and MSν are the
546
mean sum of squares of κ and ν, respectively, from ANOVAs for each combination of δ and γ.
547
Areas where the grid appears (for low δ and high γ) are point pattern landscape not possible to
548
generate with present method.
549
550
Figure 6. Observed spatial distribution of Np patches of (top row, left to right) Quercus and
551
Ulmus large trees and pig and cattle farms. Large tree defined as more than 0.7 m in breast height
552
diameter or with visible holes. In the second row we show corresponding speqtral densities and
553
the regression line of 1/fγ given the estimated Conitnuity (γ) and Contrast (δ). X and Y axis of top
554
row are given as 106 m.
Download