You are on page 1of 46

1 A new general theory of Island Biogeography

3 Gregory Beaugrand1, Loick Kléparski1,2, Christophe Luczak1, Eric Goberville3, Richard R


4 Kirby4,5
5

6 1CNRS, Univ. Lille, Univ. Littoral Côte d’Opale, UMR 8187, LOG, Laboratoire d’Océanologie et

7 de Géosciences, F 62930 Wimereux, France

8 2Marine Biological Association, The Continuous Plankton Recorder (CPR) survey, The

9 Laboratory, Citadel Hill, Plymouth PL1 2PB, UK.

10 3Unité Biologie des Organismes et Ecosystèmes Aquatiques (BOREA), Muséum National


11 d’Histoire Naturelle, CNRS, IRD, Sorbonne Université, Université de Caen Normandie,
12 Université des Antilles, Paris, France

13 4The Secchi Disk Foundation, Kiln Cottage, Gnaton, Yealmpton, Devon PL8 2HU, UK.

14 5 Ronin Institute, Montclair, NJ 07043, USA


15

16 Corresponding author: Gregory.beaugrand@univ-lille.fr

17 Orcid first author: https://orcid.org/0000-0002-0712-5223

18

19 Materials and correspondence


20
21 Correspondence and requests for materials should be addressed to G.B. (gregory.beaugrand@univ-
22 lille.fr)
23

24
25 ABSTRACT

26

27 The Equilibrium Theory of Island Biogeography (ETIB) is a widely applied dynamic theory proposed to
28 explain why islands have coherent differences in species richness. The development of the ETIB was
29 temporarily challenged by the alternative static Theory of Ecological Impoverishment (TEI), which
30 suggests that the number of species on an island is determined by its number of niches or habitats.
31 With no clear evidence relating species richness to the number of niches, the TEI was almost
32 abandoned in favour of the ETIB. Here, we show that the number of climatic niches on islands is an
33 important predictor of the species richness of plants, herpetofauna and land birds, and we propose a
34 new model called the General Theory of Island Biogeography (GTIB) that merges the ETIB and TEI
35 theories. When we test our GTIB for resident land birds in Krakatau Islands it reveals a good
36 correspondence with observed species richness, immigration and extinction rates.
37
38
39 KEYWORDS

40 Equilibrium Theory of Island Biogeography, Area-Species relationship, Theory of Ecological


41 Impoverishment, MacroEcological Theory on the Arrangement of Life, Biodiversity

42
43
44 INTRODUCTION
45
46 The Equilibrium Theory of Island Biogeography (ETIB) proposed by MacArthur and Wilson suggests
47 that immigration and extinction dynamics on an island lead to an equilibrium in species richness S
48 that is influenced by (i) island area A (Arrhenius 1921; Gleason 1922; Darlington 1957; He & Legendre
49 1996; Tjørve 2012) and (ii) its degree of isolation; the former influences ecological factors such as
50 available resources and habitat heterogeneity (MacArthur & Wilson 1967) and the latter affects
51 immigration rates (MacArthur & Wilson 1963, 1967; He & Legendre 1996). Species turnover is a key
52 aspect of the theory, the equilibrium on an island resulting from the continuing variation in the gain
53 and loss of species and in this way, island area also affects positively species richness because a
54 larger island intercepts more immigrating species, a phenomenon known as the target effect
55 (MacArthur & Wilson 1967; Stracey & Pimm 2009). Further, there is a negative relationship between
56 species richness and the distance of an island to the mainland because fewer immigrants arrive as
57 distance increases. MacArthur and Wilson used the equation S=αAβ to model this relationship, with α
58 and β depending upon the taxon and the biogeographic regions (MacArthur & Wilson 1967). In the
59 ETIB (Supplementary table S1 for all acronyms/variables meaning) the number of new species gained
60 by immigration decreases monotonically with species richness, whereas species lost by extinction
61 increases monotonically (Schoener 2009).
62
63 An alternative theory, the Theory of Ecological Impoverishment (TEI), was proposed at approximately
64 the same time as the ETIB (Lack 1970, 1976). The TEI however, differs in the way it interprets the
65 positive and negative effects of area and distance on species richness (Lack 1970; Stracey & Pimm
66 2009). When Lack studied the island biogeography of birds (Lack 1970, 1976), he found there are
67 more visiting birds on an island than breeders which led him to conclude that a newly arriving species
68 makes the choice whether to remain or not. Lack hypothesised that this choice was driven by
69 ecological conditions and he related the species richness of an island to the number of niches or
70 habitats available (Lack 1970, 1976). A recent study in British islands supports Lack’s hypothesis
71 (Stracey & Pimm 2009).
72
73 The ETIB has been extensively tested since its introduction, frequently criticised or even revisited
74 (Diamond 1969; Thornton et al. 1990; Lomolino et al. 2006; Whittaker & Fernandez-Palacios 2007;
75 Whittaker et al. 2007; Whittaker et al. 2008; Lomolino et al. 2009; Schoener 2009; Simberloff 2009;
76 Stracey & Pimm 2009; Warren et al. 2015; Whittaker et al. 2017). Among new models, the island
77 immaturity – speciation pulse model, subsequently termed the general dynamic theory of oceanic
78 island biogeography, has been proposed for volcanic islands to better consider the influence of island
79 geodynamics on biodiversity dynamics, including phases of immigration, extinction and speciation
80 that follow an island’s life cycle (Whittaker et al. 2007; Whittaker et al. 2008; Whittaker et al. 2017).
81
82 Although it seems logical that the species richness of an island should be related to the number of
83 niches available, this prediction has been rarely tested due to the difficulty in estimating the number
84 of niches that an island contains. We have been able to overcome this difficulty using a model we
85 recently developed to reconstruct global biodiversity patterns in the terrestrial and marine realms
86 (Beaugrand et al. 2020) and applying this to estimate the number of climatic niches on islands using
87 temperature and precipitation as a proxy of water availability (Methods). To examine the
88 relationships between species richness, area and distance to mainland, we chose three taxonomic
89 groups: plants (62 islands), herpetofauna (35 islands) and birds (68 islands) (Supplementary tables S2
90 and S3)(Blackburn et al. 2016). We then included the number of climatic niches into an island model
91 and propose a new way to consider the extinction rate to account for higher turnover rates generally
92 observed at the beginning of colonisation (Diamond 1969; Schoener 1983). This model is tested using
93 resident land birds on Krakatau Islands.
94 METHODS

95

96 Biological data

97
98 Data on the biodiversity and characteristics (e.g. species richness, area and distance to continent) of
99 islands for plants and birds originated from Blackburn and colleagues(Blackburn et al. 2016)
100 (Supplementary table S2). Data on the biodiversity and characteristics of islands for herpetofauna were
101 assembled in this work from a variety of sources (Supplementary table S3).
102
103 Resident land bird data of Krakatau islands (Sertung, Panjang Anak and Rakata) with respect
104 immigration, extinction and species richness, originated from Table 1 in Thornton and colleagues
105 (Thornton et al. 1988).
106
107 Environmental data

108 Mean monthly temperature (°C) and precipitation (mm) climatologies (period 1970-2000) were
109 retrieved from the 1-km spatial resolution WorldClim version 2 dataset
110 (http://worldclim.org/version2)(Fick & Hijmans 2017). Climatologies were obtained by performing the
111 thin-plate smoothing spline algorithm implemented in the ANUSPLIN package; more information on
112 the numerical procedures was provided by Hijmans and colleagues (Hijmans et al. 2005) and Fick and
113 Hijmans (Fick & Hijmans 2017).

114 Calculation of the number of climatic niches

115
116 We applied a model developed to reconstruct and investigate large-scale biodiversity patterns in the
117 terrestrial and marine realms (Beaugrand et al. 2020). The model developed as part of the
118 MacroEcological Theory on the Arrangement of Life (METAL) has been fully described and tested in
119 Beaugrand and colleagues (Beaugrand et al. 2013; Beaugrand 2015; Beaugrand et al. 2015). This model
120 generates ecological niches sensu Hutchinson (Hutchinson 1957), which then interacts with the local
121 environmental regime, giving an estimate of the number of niches and therefore species that can occur
122 in a region (Beaugrand et al. 2018; Beaugrand et al. 2020). Although the concept of the niche is
123 multidimensional, we focussed here on climatic niches, which were assessed using temperature and
124 precipitation on each island (Beaugrand et al. 2020). Climatic niches are critical for the long-term
125 establishment of a species (Whittaker 1975) and individuals outside their climatic niches cannot occur
126 for long in a region. Other ecological dimensions (e.g. soil pH and type) also explain species occurrence
127 but we assume here that they play a secondary role at a global scale. The estimation of the mean number
128 of niches on each island was performed following two main steps (Supplementary fig. S2):
129
130 Step 1: building of climatic niches
131
132 We calculated rectangular niches in this model with a 0 corresponding to an absence and a 1 to a presence
133 (Supplementary fig. S2)(Beaugrand et al. 2013). All potential thermal niches ranged from tmin=-1.8°C
134 to tmax=44°C and all precipitation niches ranged from pmin=0 mm to pmax=3000 mm; these thresholds
135 were investigated in previous works and best fit the data in the oceanic and terrestrial realms(Beaugrand
136 et al. 2020). A total of 1,067,175 potential niches was considered in this study (255 precipitation x 4185
137 thermal niches). The mathematics of the model is presented in Supplementary text S1.
138
139 Step 2: niche-climate interaction
140
141 We then tested the pool of potential niches (1,067,175) on each island; niches were selected when
142 climatic conditions were suitable for at least n month(s) (1≤n≤12) and in at least one ~1 km x ~1 km
143 geographical cell. We therefore performed 12 simulations for each island: in the first simulation one
144 niche was considered to be potentially represented on an island if only one month was suitable in any
145 geographical cell. In the second, one niche was considered to be potentially represented on an island if
146 only two months were suitable in any geographical cell. And so on until 12 months. In the twelve
147 simulation, one niche was considered to be potentially represented on an island if all months were
148 suitable in any geographical cell. We then calculated the average of the total number of niches that may
149 be represented on each island from the 12 simulations. The total number of niches was calculated for 56
150 islands for plants (Supplementary table S2), 35 islands for herpetofauna (Supplementary table S2) and
151 62 islands for birds (Supplementary table S2). For some small islands (e.g. St Helena Island), estimation
152 of niches was not possible because of the absence of temperature or precipitation values (Supplementary
153 table S2). Ascension, Macquarie and Saint-Paul Islands were removed because there was no native bird
154 species.
155
156 Statistical analyses
157
158 Relationships between observed richness, area, the number of climatic niches and distance
159 to mainland
160
161 We investigated the relationships between (i) observed species richness and area, (ii) observed species
162 richness and the number of climatic niches, and (iii) observed species richness and distance to mainland
163 for three taxa: plants, herpetofauna and birds. We used scatterplots with latitude as a supplementary
164 variable to examine these relationships (Fig. 1); the colour and the size of the points were proportional
165 to the absolute value of latitude. To examine the magnitude and significance of the relationships, the
166 ordinary linear coefficient of correlation was calculated for all pairs of variables.
167
168 We also calculated the correlations between island area and the number of climatic niches for all islands
169 and taxonomic groups. To examine how latitude may influence these relationships, we also assessed
170 these correlations exclusively for islands between the Tropics of Cancer and the Tropics of Capricorn
171 (i.e. tropical biome).
172
173 Species richness as a function of the number of niches available, area and distance to
174 mainland
175
176 We estimated species richness (Z) using a multiple linear regression based on the number of niches
177 available (M), island area (A) and distance to mainland (d)(Fig. 2a). All variables were log10-
178 transformed. The model, performed individually for each taxonomic group (plant, herpetofauna, bird),
179 was as follows:
180
181 Log10(Z)= ϒ1Log10(M) + ϒ2Log10(A) + ϒ3Log10(d) + β (1)
182
183 With ϒ1, ϒ2 and ϒ3 the multi-regression coefficients of M, A and d, respectively and β the y-intercept
184 (Supplementary table S4). We expected species richness to be proportional to M and A but inversely
185 proportional to d.
186
187 Three multiple linear regression analyses were performed on (i) plants, (ii) herpetofauna and (iii) birds.
188 Then, we applied Equation (1) to estimate species richness on all islands and examined the respective
189 influence of each component in Equation (1) (Fig. 2). We then calculated the overall coefficient of
190 correlation between predicted and observed species richness.
191
192 For the three multiple linear regression analyses, we assessed the contribution of each variable (i.e. ΦM,
193 ΦA, Φd) as follows (Fig. 2b):
194
195 ΦM=ϒ1Log10(M) / Log10(Z)- β (2)
196 ΦA=ϒ2Log10(A) / Log10(Z)- β (3)
197 Φd=ϒ3Log10(d) / Log10(Z)- β (4)
198
199 Model of island biogeography
200
201 Description of the model
202 We designed a dynamic model, based on the same rationale than the Equilibrium Theory of Island
203 Biogeography (ETIB), but jointly considering area (A), distance to mainland (d) and the mean number
204 of climatic niches (M). The number of climatic niches, fixed by METAL, was 105,082 for the whole
205 archipelago, including Rakata, Panjang, Sertung and Anak islands. Our model calculates species
206 richness of an island using long-term immigration rates, short-term and long-term extinction rates. The
207 model was tested using bird data for Krakatau Islands. Species richness on an island at age t was
208 assessed as followed:
209
210 𝐵𝑡+1 = 𝐵𝑡 + 𝑡(𝐼𝑡 − 𝐸𝑡 − 𝐹𝑡 ) (5)
211
212 Where Bt and Bt+1 were species richness at time t and t+1, respectively, with t expressed in year. It and
213 Et were the immigration and the long-term extinction rates at year t, respectively. We included a new
214 term Ft in Equation 5, which was the initial short-term extinction rate. We added this term because, at
215 the beginning of colonisation time, habitats can be rapidly altered and populations remain small, which
216 initially increase the extinction rate (MacArthur & Wilson 1967; Bush & Whittaker 1993).
217
218 Immigration rate It (species.yr-1) on an island was calculated by using a negative exponential function
219 standardised between I0 and Is, with I0 and Is the immigration rates at t=0 and t=s (saturation),
220 respectively:
221
𝐵 𝑏1 𝐼0
−( 𝑡 )
(𝑒 𝐵𝑠 −𝑒 −1 )

222 𝐼𝑡 = 𝐼0 1−𝑒 −1
I0 ≤ It ≤ Is and Bt ≤ Bs (6)

223
224 Where Bt was the species richness at time t and Bs the species richness at saturation (B0=0 at t=0) and
225 Bs assessed as follows:
226
227 Bs=ϕM (7)
228
229 where ϕ was a constant that depends upon the taxonomic group and M the mean number of climatic
230 niches on the island. We assumed here that all niches could be colonised by a unique species after the
231 principle of competitive exclusion (Gause 1934). b1 is a constant (dimensionless) that depends upon
232 the taxonomic group. The constant influences the speed with which the immigration rate diminishes
233 between I0 and Is. At saturation, we assumed that Is=0 species.yr-1 (Fig. 3).
234
235 Long-term extinction rates Gt were estimated as follows:
𝐵𝑡 𝑏2 𝐺𝑆
( )
(𝑒 𝐵𝑠 −1 )

236 𝐺𝑡 = 𝐺𝑠 𝑒 1 −1
G0 ≤ Gt ≤ Gs and Bt ≤ Bs (8)

237
238 With Gs=1 at Bs; b2 is a constant that depends upon the taxonomic group. The constant influences the
239 speed with which the long-term extinction rate increases between G0=0 and Gs.
240
241 Short-term extinction rates Ft was a function of the immigration rate:
242
243 𝐹𝑡 = 𝜎𝐼𝑡 𝑒 −𝑏3 𝐵𝑡 with F0 ≤ Ft ≤ Fs (9)
244
245 With 𝜎 a constant (species-1) that determines the initial rate of short-term extinction rate, b3 a constant
246 (dimensionless) that affects the speed with which short-term extinction rate diminishes and cancels
247 off. Ft is a function of both species richness and the immigration rates at time t. In this study Fs=0.
248
249 Total extinction rate Et is the sum of short (Ft) and long-term (Gt) extinction rates:
250 Et = Ft + Gt with E0 ≤ Et ≤ Es (10)
251
252 Because Gs=1 at Bs, Es=1 as Fs=0.
253
254 Test of the model for Krakatau islands
255
256 The test was conducted by using resident land bird data from the Krakatau Islands (Thornton et al.
257 1988), which was sterilised by a volcanic eruption in 1883. Island colonisation was subsequently
258 investigated in 1908, 1919-1924, 1928-1934, 1951-1952 and 1984-1986.

259 We had to estimate six parameters to model the recolonization of the island: b1 (Equation 6), b2
260 (Equation 8), b3 (Equation 9), I0 (Equation 6), σ (Equation 9) and ϕ (Equation 7). Estimation of the
261 parameters of the model was done by calculating 1,935,360 combinations and by minimising the Root
262 Mean Square Error (RMSE) between modelled species richness, immigration and total extinction rates
263 of resident land bird for different time periods. RMSE is calculated as follows (Chai & Draxler 2014):

∑𝑛
𝑖=1(𝐻𝑖 −𝑂𝑖 )
2
264 𝑅𝑀𝑆𝐸 = √ 𝑛
(11)
265 Where Hi and Oi are modelled and observed species richness, immigration or total extinction rate at
266 observation i, respectively, and n the number of observations. The three RMSEs were subsequently
267 standardised between 0 and 1 and the sum of the three RMSEs, called hereafter total RMSE (range
268 between 0 and 3), were calculated. (We took the three RMSEs to ensure that immigration and
269 extinction rates, as well as species richness, were all best reproduced by the model.) We kept the 1000
270 combinations that had the smallest total RMSE and finally retained the combination with the lowest
271 total RMSE (Fig. 4).

272 Values for b1 were 0.01 0.03, 0.05, 0.08, 1, 1.2, 1.4, 1.6, 1.8, and 2, a total of 10 values; values for b2
273 were 0.1, 1, 2, 3, 4, 5, 8, 10, 25, 40, 50, 60, 75, 100, 110, and 150, a total of 16 values; values for b3
274 were 0.00005, 0.0001, 0.0002, 0.0003, 0.0004, 0.0005, 0.00075, 0.001, 0.005, 0.01, 0.02, 0.04, a total
275 of 12 values; values for I0 were 0.71, 0.8, 0.9, 0.95, 1, 1.05, 1.1, 1.2, 1.6, 2, 2.5, 3, 3.5, 4, 5 and 6, a total
276 of 16 values; values for σ were 0.2, 0.30, 0.35, 0.40, 0.45, 0.5, 0.55, 0.6 and 0.8, a total of 9 values;
277 values for ϕ were 0.0002 0.0003 0.0004 0.0005 0.0006 0.0008 and 0.001, a total of 7 values; IS=0 and
278 G0=0; Gs=Es=1; Fs=0; F0 and therefore E0 were fixed as a function of I0.

279 Ordinary linear coefficient of correlations were calculated between modelled and observed species
280 richness, immigration and total extinction rates (Sokal & Rohlf 1995).

281 Species richness at equilibrium Beq (species) is reached when Et crosses It (Fig. 3). The number of years
282 after the eruption when Beq is reached Teq (year) can be easily assessed graphically or by examination
283 of the matrices (Fig. 3). The year is then deduced by adding Teq to 1883. In practice here, Teq was
284 determined when less than one species remained to reach Beq.
285
286
287
288
289 RESULTS
290
291 Relationships between species richness, area, the number of climatic niches, distance to mainland
292 and latitude
293
294 We first examined the relationships between species richness, area, the number of climatic niches,
295 distance to mainland, and latitude (Fig. 1). The defined niche is the climatic conditions created by
296 temperature and precipitation that enables a species’ individuals to grow and reproduce (Beaugrand
297 et al. 2020); this niche definition is close to the one proposed by Hutchinson(Hutchinson 1957, 1978).
298 The number of climatic niches of an island was assessed using a biodiversity model developed as part
299 of the MacroEcological Theory on the Arrangement of Life (METAL) using temperature and
300 precipitations as an index of water availability (Methods). As we expected, plants, herpetofauna and
301 birds all showed a positive correlation between species richness and area (0.57<r<0.75; Fig. 1a-c). A
302 clear effect of latitude was also evident on the scatterplots where high-latitude islands always
303 exhibited less than expected species richness (see red bullets, Fig. 1a-c, i.e. islands located in the
304 lower triangular part of the diagram). Using the number of climatic niches that an island may contain
305 instead of island area improved the correlations for all taxa other than birds (0.62<r<0.74; Fig. 1d-f).
306 It is noteworthy that high-latitude islands were better distributed on the scatterplots, even for birds,
307 (i.e. distributed along the major axis of the cloud of points), indicating a clear positive influence of
308 the number of climatic niches on species richness for all taxonomic groups. MacArthur and Wilson
309 highlighted that the species-area relationship will hold only for islands belonging to the same
310 biogeographic region (MacArthur & Wilson 1967) and our results concur and therefore, also support
311 Lack’s argument that species richness of an island is influenced by the number of niches
312 available(Lack 1970, 1976), which is influenced by the biogeographic region. As expected, the
313 relationships between species richness and distance to mainland were less consistent and we only
314 found significant negative correlations for plants and birds (-0.60<r<-0.41), not for herpetofauna (r=-
315 0.01; Fig. 1g-i); no effect of island latitude was detected on these scatterplots. We warn that some
316 continental islands in Fig. 1g-I had been colonised before being separated from the continent (e.g.
317 islands from the United Kingdom).
318
319
320
321 Figure 1. Relationships between species richness of an island, its area, the number of climatic
322 niches, distance to mainland and latitude. Scatterplots of observed species richness in each island as
323 a function of area (a-c), mean predicted number of climatic niches (d-f) and distance to mainland (g-i)
324 for plants (a, d, g), herpetofauna (b, e, h) and birds (c, f, i). The size and the colour of the circles are
325 proportional to the absolute value of latitudes (between 0° and 80°). The ordinary linear coefficient
326 of correlation is indicated on each panel. All correlation coefficients, but panel h, were significant at
327 p<0.01. n is the number of couple of points used to calculate the correlation. All variables were log10-
328 transformed.
329
330 We further examined the correlation between area and the number of climatic niches of islands and
331 found them to be positively correlated significantly for all groups, although the correlation was not
332 always strong when all islands (i.e. all latitudes) were considered; plants: linear correlation coefficient
333 r=0.28, probability p=0.04, number of points n=56; herpetofauna: r=0.70, p<0.001, number of points
334 n=35; birds: r=0.52, p<0.001, n=65. As we expected, the correlation improved for all taxonomic
335 groups when the influence of latitude was accounted for. For islands within the Tropic of Cancer and
336 the Tropic of Capricorn; plants: r=0.53, p=0.007, n=24; herpetofauna: r=0.91, p<0.001, n=16; birds:
337 r=0.62, p<0.001, n=32.
338
339 To examine how area, the number of climatic niches and distance to mainland jointly affect species
340 richness for each taxonomic group, we performed three multiple linear regressions (see Methods);
341 these analyses also allowed us to examine the respective linear contribution of each variable (Fig. 2
342 and Supplementary table S4). When combined together, predictions from the three regressions were
343 highly correlated positively with observed species richness (Fig. 2a). Contribution of the mean
344 number of climatic niches was highest (positive contribution), followed by distance to mainland
345 (negative contribution) and to a lesser extent area (positive contribution) (Fig. 2b and Supplementary
346 table S4); the contribution of distance to mainland was greater than area effect (in contrast to Fig. 1)
347 because the consideration of the number of climatic niches diminishes the influence of area alone
348 due to the positive correlation between area and niche number. Interestingly, we found that distance
349 to mainland had a more negative contribution for plants and birds than herpetofauna that has poor
350 dispersal capability, even though there was no clear distinction among taxonomic groups for area
351 and the number of climatic niches (Fig. 2b). In a global-scale analysis, insularity also had a weak
352 influence on herpetofauna spread rates (Liu et al. 2014). To conclude, our analyses suggest that
353 considering the mean number of climatic niches of an island increased the predicted species richness
354 substantially (Fig. 1-2).
355
356 Figure 2. Relationships between observed species richness and richness predicted from a linear
357 multiple regression model using area, the number of climatic niches and distance to mainland. (a)
358 Scatterplot of observed versus predicted species richness from the number of climatic niches M, area
359 A and distance to continent d. The ordinary linear coefficient of correlation is indicated. The
360 correlation was highly significant at p<0. 01. n is the number of couple of points used to calculate the
361 linear correlation. (b) Contribution of the different variables to the estimates of species richness. No
362 distinction was possible among taxa, but herpetofauna for area: distance to mainland had a low
363 negative effect for herpetofauna in contrast to plants and birds.
364
365 A new dynamic model including the number of climatic niches
366
367 Because our results suggested that a consideration of the mean number of climatic niches was an
368 important island property to assess species richness on an island, we built a dynamic model based on
369 the same rationale as the ETIB, but also considering the mean number of climatic niches (M;
370 Methods). We have called our model the General Theory of island Biogeography (GTIB). The rationale
371 of our GTIB, as for the ETIB, can be explained by plotting both immigration and extinction rates as a
372 function of species richness and examines when the two curves cross. In ETIB, the immigration and
373 extinction curves are monotonic (Fig. 3a). In contrast to the ETIB however, our GTIB calculates
374 species richness of an island using short-term Ft and long-term Gt extinction rates (See
375 Supplementary table S1 for variable meaning)(Fig. 3b). The use of short-term extinction rates Ft is
376 justified by the fact that at the beginning of colonisation, populations are small and the environment
377 is changing rapidly due to the effects of colonisation, which increase the likelihood of extinction
378 (Klein 1968; Diamond 1969; Wright 1981). Consequently, the short-term extinction rate Ft diminishes
379 exponentially to a minimum as the long-term extinction rate Gt rises exponentially to Gs (Fig. 3b). The
380 addition of Ft and Gt yields the total extinction rate Et, which varies between E0 and Es; the latter
381 parameter was fixed to 1 in this study. The joint consideration of immigration and short-term
382 extinction rates enables the consideration of the high turnover rate of birds that is observed on some
383 islands (e.g. Channel Islands of California)(Diamond 1969). In our GTIB model, there is therefore no
384 necessary monotonic increase in extinction rate and the shape can sometimes be similar to the
385 shape suggested by Bush and Whittaker, their Figure 2 (Bush & Whittaker 1993); it all depends upon
386 the parameters of the model (Methods and Discussion). Although not explicitly included in our GTIB,
387 an allogenic influence (e.g. wind, oceanic currents, precipitation, heat wave) probably dominates at
388 the beginning of colonisation and an autogenic influence (e.g. density-dependence phenomena and
389 species interaction) is more likely towards species richness at equilibrium Beq (Supplementary table
390 S1). Of course, allogenic or autogenic perturbations may prevent the system reaching Beq. In addition,
391 Beq may be modified by environmental changes originating from geodynamics or climate change (see
392 Methods and Discussion) and our model can therefore account for a time-varying species carrying
393 capacity (Marshall & Quental 2016). The GTIB can therefore be considered as a generalisation of
394 ETIB, the latter being a particular case when short-term extinction rate Ft is nil and for a stable
395 environmental regime, because the environmental regime controls the number of potential niches
396 M, and therefore species richness at saturation Bs (Methods).
397
398
399 Figure 3. Model of island biogeography of (a) MacArthur and Wilson (1963,1967) and (b) its
400 modification proposed here. (a) In ETIB, species richness at equilibrium Beq (Supplementary Table S1)
401 is reached when the monotonic reduction in immigration rate I (blue curve, with I0 ≤ It ≤ Is) crosses
402 the monotonic increase in extinction rate (red curve, with E0 ≤ Et ≤ Es). Note that immigration can be
403 supplemented by speciation at first approximation, especially when distance to mainland is high
404 (Lomolino et al. 2006); see Discussion. (b) In our model, as in ETIB, changes in I is modelled by a
405 negative exponential function standardised between I0 and IS, with I0=1 at t=0 and IS = 0 at saturation
406 as an example. In contrast to ETIB, changes in total extinction rate (red line) are the results of two
407 functions: a negative (i.e. short-term extinction rate Ft at year t with F0 ≤ Ft ≤ Fs) and a positive (i.e.
408 long-term extinction rate Gt with G0 ≤ Gt ≤ Gs) exponential function (red and black dashed lines for
409 both functions). The first negative exponential function that is extended by a black dashed line
410 (toward the right from the red curve) dominates for low values of species richness, i.e. at the
411 beginning of island colonisation; initially the short-term extinction rate was fixed to 0.8 in this
412 example (E0=F0=0.8). The second positive exponential function that is also extended by a black
413 dashed line (toward the left from the red curve) is dominant for higher values of species richness, i.e.
414 from the middle part of island colonisation. In this example, the long-term extinction rate Gt was
415 standardised between G0=0 and GS=Es=1. Black dashed curves are never observed. Our model equals
416 ETIB when the first negative exponential function is nil (i.e. E0=0). Our model is a non-equilibrium
417 model because Beq can be altered when species richness at saturation is modified by an
418 environmental modification that affects both immigration and extinction rates (see Methods).
419
420 We tested our GTIB model using resident bird data from the Krakatau Islands (Anak, Rakata, Sertung,
421 Panjang). Values of the six parameters used in Equations 6-9 were determined using a total of
422 1,935,360 combinations (Methods). We retained the 1000 curves (grey curves in Fig. 4) with the
423 lowest Root Mean Square Error (RMSE) to provide a confidence interval (grey curves in Fig. 4) and
424 chose ultimately the curve with the smallest RMSE (red curve in Fig. 4). Our GTIB reproduced well the
425 dynamics of bird species richness on the island (Fig. 4). Although the number of observations was
426 limited to have unambiguous correlations, especially for immigration (five year/time periods) and
427 total extinction (four years/time periods) rates, ordinary linear correlation coefficients between
428 modelled and observed species richness (six years/time periods), immigration and total extinction
429 rates were r1=0.91 (p<0.001), r2=0.65 (p=0.003) and r3=0.37 (p=0.13), respectively.
430
431 We assessed species richness at equilibrium Beq = 47.94 species (range of values based on 1000
432 curves with smallest RMSE, 47.76-51.36 species); such an equilibrium depends upon the number of
433 climatic niches, which can be readjusted if climate changes. Our estimate of ~48 (47-51) species at
434 equilibrium is higher than values of 30 in MacArthur and Wilson (MacArthur & Wilson 1967), 36-38 of
435 Thornton and colleagues (Thornton et al. 1990) and 40-45 of Mayr (Mayr 1965). The number of years
436 to reach equilibrium Teq was also estimated while there was no alteration in the number of climatic
437 niches (i.e. for a stable environmental regime). We found Teq = 166 years (131-187 years) after 1883,
438 i.e. yeareq=2049 (2025-2133). It is therefore clear that the species richness of resident land birds did
439 not reach an equilibrium in 1933, as proposed early by MacArthur and Wilson (MacArthur & Wilson
440 1967) and already highlighted by Thornton and colleagues (Thornton et al. 1993). As the species
441 richness of the archipelago trends toward an equilibrium that is based upon the current island
442 configuration and climatic regime, it would be interesting to conduct a new inventory of the resident
443 land birds on the Archipelago in order to confirm or refute our estimates.
444
445

446
447
448 Figure 4. Long-term changes in species richness (a), immigration (b) and (c) total extinction rates of
449 land birds in Krakatau Island. On panel (a), levels (minimum, maximum and optimal values) and
450 timing at which species richness flattened off are indicated. On each panel, the red curve denotes the
451 optimal model (i.e. with lowest RMSE) and the grey curves are the 1000 curves with the lowest RMSE
452 out of a total of 1,935,360 possible estimates. Blue circles are observed number of resident land
453 birds carried out on the island. 1883 is the year when the eruption of Krakatoa sterilised the island.
454 Species richness at equilibrium Beq (see Fig. 3) and year at which equilibrium is reached yeareq are
455 indicated. Here Beq = 47.94 species (range of values based on 1000 curves with smallest RMSE, 47.76-
456 51.36 species) and Teq = 166 years (131-187 years) after 1883, or yeareq=2049 (2025-2133).
457 Parameters of the best model were I0=1 (range for the 1000 curves, 0.9-1.1), b1=1.4 (1.4-2), b2=25
458 (25-150) and b3=0.0001 (0.0001-0.005), σ=0.35 (0.3-0.4) and ϕ=0.0005 (0.0005-0.0005). Because Bs=
459 ϕ x M, species at saturation Bs = 52.5 species. Other fixed parameters were M=105,082 niches, E0=0,
460 ES=1 and IS=0.
461
462 DISCUSSION
463
464 We have shown that the number of ecological niches available on an island explains why island area
465 often correlates with species richness (MacArthur & Wilson 1963, 1967; Lomolino et al. 2006) and
466 because island area also correlates positively with the number of climatic niches within a biome (we
467 tested it for the tropical biome). This correlation diminishes when all latitudes are considered
468 together. Our understanding of island biogeography improves substantially by including the number
469 of climatic niches on an island as it enables the theory to be both generalised to all latitudes and
470 more ecologically meaningful. Although the niche dimensions we considered here are important
471 ecologically (Whittaker 1975), we acknowledge that the niche sensu Hutchinson (Hutchinson 1957) is
472 multi-dimensional and so other ecological dimensions should be considered to account for full niche
473 complexity (e.g. pH, soil humidity, soil composition)(Carlquist 1965; Hirzel et al. 2002). Since METAL
474 can generate multidimensional niches, the consideration in the future of more ecological dimensions
475 (ecological dimensions where there is long-term global high spatial resolution data) may improve our
476 current estimates.
477
478 Area also influences positively species interception, a phenomenon called the target effect
479 (MacArthur & Wilson 1963, 1967; Stracey & Pimm 2009). Total area of islands that compose the
480 archipelago was relatively small (24.45 km²) and probably affected immigration rates less than
481 distance to mainland (distance between Krakatau Islands and Sumatra or Java is ~40 km). If our GTIB
482 is used to compare species richness on different islands, I0 could be calculated as a function of area A
483 and distance to mainland d. For example, the following equation could be used. 𝐼0 = 𝑚1 ln(𝐴 + 1) −
484 𝑚2 ln⁡(𝑑 + 1), with m1 and m2 two constants.
485
486 Area might also affect extinction rates; the higher the area, the greater population size and the lower
487 the extinction rate (Simberloff 1976; Schoener 1983). Although Es did not vary here (Es was fixed to 1
488 species.yr-1), Es could also be adjusted to account for the influence of area on extinction rates if our
489 model is used to compare different islands. In our study, we assumed that this effect of area on
490 extinction rates was implicitly considered through Bs, the higher the number of species at saturation,
491 the lower the extinction rate at Bt<<Bs (and Ft≈0; Equation 8, Fig. 3).
492
493 We suggest that considering higher initial extinction rates is important to better reproduce early
494 island colonisation, because this allows a better reproduction of greater species turnover generally
495 observed at the beginning of colonisation (Diamond 1969).
496
497 Our GTIB was well suited to birds because they are mobile and widespread(Thornton et al. 1990). It
498 would be interesting to test our new theory on other taxonomic groups that might exhibit different
499 turnover rates (Schoener 1983). Note that our GTIB can adapt to taxonomic groups for which short-
500 term extinction rates are smaller or even negligible; for example, in Equation 9 (Methods), σ can be
501 chosen low enough to have F0 close to 0 (Fig. 3), which implicates Ft negligible so that Et (Equation 10)
502 is largely driven by Gt (Equation 8). Indeed when σ=0, Et=Gt and we refined the classical dynamic
503 model proposed by MacArthur and Wilson (MacArthur & Wilson 1963), although area A is here
504 replaced by the number of climatic niches M and therefore species richness at saturation Bs.
505
506 The Krakatoa volcano eruption sterilised the island on August 27, 1883 (Guo 2008). The current
507 configuration of the archipelago is therefore young and our GTIB, which does not explicitly consider
508 speciation, reconstructed well the species richness dynamics and associated turnover. On older
509 islands, our GTIB may therefore be less accurate if speciation is not explicitly considered;
510 observations and theoretical models have shown that this process is also important to explain
511 biodiversity dynamics (Whittaker et al. 2007; Whittaker et al. 2008; Whittaker et al. 2017; Veron et
512 al. 2019). Speciation could be integrated in our model in Equation 5 (Methods). Immigration and
513 speciation equal 0 at species richness at saturation Bs, however. Bs was determined by Equation 7
514 through the estimate of ϕ=0.0005. Because M=105,082 niches (determined by METAL), Bs = 52.5
515 species. (For the Krakatau Islands, it is therefore unlikely that the absence of a direct implementation
516 of speciation in the model had an effect on our estimate of Beq=48 (47-51) species because our
517 estimate is only slightly below Bs = 52.5 species.) For more mature or distant islands, we think that
518 speciation should be integrated into the model to explicitly account for high level of endemism
519 observed in some remote islands (MacArthur & Wilson 1967; Gillespie & Roderick 2002; Gillespie
520 2004; Veron et al. 2019).
521
522 An important prediction from our model, due to the fact that Bs strongly influences Beq, is that islands
523 far from mainland should have a greater proportion of endemic species, a prediction that might hold
524 providing island age and dispersal capacity of a taxonomic group are accounted for (Veron et al.
525 2019); this arises because potential niches of islands close to the mainland are rapidly filled with
526 existing species originating nearby in contrast to remote islands where speciation is the only niche-
527 filling alternative to the low immigration rate. This prediction is consistent with some studies that
528 have suggested that patterns of species accumulation through evolution in remote islands is
529 analogous to islands close to continents where species gain takes place through immigration
530 (Gillespie 2004). The author proposed that this might suggest that universal principles may underly
531 processes of community assembly. Studies have found a positive correlation between species
532 richness and the level of endemism in islands (Emerson & Kolm 2005b). Although this may also be
533 explained by some methodological considerations or misinterpretations (Cadena et al. 2005;
534 Emerson & Kolm 2005a), the correlation may be due to higher endemism when the number of
535 available niches, and therefore species richness, is greater. As mentioned by Witt and Maliakal-Witt
536 (Witt & Maliakal-Witt 2007), speciation may be both accelerated and impeded by niche availability.
537 We therefore suggest that the universal mechanism mentioned by Gillespie (Gillespie 2004) may be
538 related to niche availability that fixes the number of species that can establish in an island either by
539 immigration or speciation, a mechanism recently suggested to explain large-scale patterns in
540 biodiversity or niche saturation in the marine and terrestrial realms (Beaugrand et al. 2018;
541 Beaugrand et al. 2020). Distance to mainland is also important because it affects immigration rates
542 and especially I0 in our GTIB and therefore, initial values of It. Among values ranging from 0.71 to 6,
543 the best estimate was I0=1 species.yr-1 (0.9-1.1). Such a value is relatively high, which can be
544 explained by the closeness of the Krakatau Islands to the mainland, i.e. Java and Sumatra (Western
545 Indonesia).
546
547 The general dynamic theory of oceanic island biogeography has significantly increased our
548 knowledge of how species richness and associated biological rates may evolve on volcanic islands
549 (Whittaker & Fernandez-Palacios 2007; Whittaker et al. 2008; Whittaker et al. 2017). Island
550 geodynamics affects local climate and environment that in turn alter biodiversity dynamics. Since we
551 determined a unique number of niches for each island, our model is a simplification of real life and
552 the number of niches will inevitably change as islands evolve in term of elevation, size and
553 configuration, or as climate changes. High-resolution monthly climatologies were the only data
554 available at the time of our analysis but as climatic data becomes more accessible (e.g. on a year-to-
555 year basis) M - and therefore Bs - can be reassessed making Beq a more dynamic equilibrium. Beq is
556 therefore not constant through time in the GTIB in contrast to the ETIB. Not only species richness is
557 likely to fluctuate around Beq through immigration-extinction dynamics (and on more mature or/and
558 remote islands through speciation) but Beq also changes as a function of island geodynamics and
559 climate or environmental change, whether natural or anthropogenic. It follows therefore that
560 equilibrium cannot be reached and that species richness fluctuates around an attractor that is
561 permanently shifting as environmental conditions change (Storch et al. 2021). The GTIB is therefore a
562 nonequilibrium model. Our nonequilibrium model remains a simplification of the reality and some
563 further processes (e.g. speciation) could be implemented in future versions to make it more useful to
564 understand eco-evolutionary dynamics or to consider island geodynamics (Whittaker et al. 2007;
565 Whittaker et al. 2008; Kueffer et al. 2014; Warren et al. 2015; Santos et al. 2016; Whittaker et al.
566 2017).
567
568 CONCLUDING REMARKS
569
570 We propose that our new GITB explains how insular species richness is highly influenced by the
571 number of climatic niches available on an island. This island property enables us to generalise the
572 ETIB to all latitudes, reconciling the theory with TEI; this is even when niche definition in our study
573 was not exactly as envisioned by Lack (Lack 1970, 1976) who defined it in terms of resource(Elton
574 1927). When Mac Arthur and Wilson said “There exists within a given region of relatively uniform
575 climate an orderly relation between the size of a sample area and the number of species found in that
576 area”(MacArthur & Wilson 1967) they realised that their theory was only valid for islands belonging
577 to a similar biome. In our GTIB, the number of climatic niches not only enables the area to be
578 considered but we can also weight the area by the latitudinal influence and this is where our GTIB
579 has better ecological relevance.
580
581 Our GTIB nevertheless remains based on the MacArthur and Wilson’s pioneering theory and while
582 our implementation of a short-term extinction rate into our GTIB is an improvement to account for
583 higher turnover rates observed at the beginning of colonisation, we do not think this alters their
584 original theory drastically; the only difference is that the total extinction rate does not become
585 monotonic. The most important development in our GTIB is, in our opinion, the consideration of the
586 number of niches (here the climatic niches) that can be recalculated as island environment changes,
587 making our GTIB a nonequilibrium model at the time scale of an island’s life cycle. There has been a
588 debate whether or not biodiversification follows an equilibrium model (Sepkoski Jr 1978; Benton &
589 Emerson 2007). A recent molecular phylogenetic survey of the Avian communities at four
590 Macaronesian archipelagos (e.g. Azores, Madeira, the Canary Islands and Cape Verde) has provided
591 evidence that a diversity plateau can be rapidly reached and remain stable for millions of years,
592 supporting an equilibrium (Valente et al. 2017). It remains to be understood if such results can be
593 generalised to all oceanic islands and views on this important subject remain controversial (Abbott &
594 Grant 1976; Marshall & Quental 2016). We think that our GTIB may help to resolve this controversy
595 because it suggests that the interplay between the timing needed to reach the equilibrium and the
596 frequency of the perturbations or the timing to next environmental changes is critical; an equilibrium
597 might never be achieved if the environment changes before an equilibrium is reached.
598
599 Our GTIB is a generalisation of ETIB, the latter being a particular case when the short-term extinction
600 rate is nil and when the environmental regime is stable. Along the life cycle of an island,
601 environmental changes are likely to occur either through climate change or because of island
602 configuration (Whittaker et al. 2007; Whittaker et al. 2008; Whittaker et al. 2017). When this occurs,
603 the number of niches is altered, which affect species at equilibrium and a new dynamic occurs. Our
604 GTIB could therefore be used as part of the general dynamic theory of oceanic island biogeography
605 developed by Whittaker and colleagues (Whittaker et al. 2007; Whittaker et al. 2008; Whittaker et al.
606 2017). Taken together with other works (Lomolino et al. 2009; Rominger et al. 2016; Whittaker et al.
607 2017; Veron et al. 2019), we think that our findings may help to improve our understanding of island
608 biodiversity dynamics and to progress toward a new synthesis of island biogeography. Our results
609 have important implications for ecological restoration and our model could be applied (i) to
610 determine the degree of direct human disturbance on species richness and (ii) to examine how
611 climate change might affect island biodiversity, because Bs in our model is fixed by the number of
612 available climatic niches that will be altered as climate changes.
613
614
615 ACKNOWLEDGEMENTS

616

617 This work was supported by the ‘Centre National de la Recherche Scientifique’ (CNRS), the Research
618 Programme CPER CLIMIBIO (Feder, Nord-Pas de Calais), the regional programme INDICOP (Pas-de-
619 Calais) and the ANR project TROPHIK. The authors also thank the French Ministère de l'Enseignement
620 Supérieur et de la Recherche, the Hauts de France Region and the European Funds for Regional
621 Economical Development for their financial support to this project.

622

623 AUTHORS CONTRIBUTION

624

625 GB conceived the study and designed the models. GB, LK and EG prepared the data. GB and LK made
626 the data analyses. GB prepared the first draft. GB, RRK, LK, CL and EG discussed the results and
627 contributed to writing.

628

629 Competing financial interests


630
631 The authors declare no competing financial interests
632
633 Data accessibility statement
634

635 The biological data that support the findings in this study are available in Supplementary Table S2-S3.

636

637
638 REFERENCES

639 1.
640 Abbott, I. & Grant, P.R. (1976). Nonequilibrial bird faunas on islands. The American Naturalist, 110,
641 507-528.
642 2.
643 Arrhenius, O. (1921). Species and area. Journal of Ecology, 9, 95-99.
644 3.
645 Beaugrand, G. (2015). Marine biodiversity, climatic variability and global change. Routledge, London.
646 4.
647 Beaugrand, G., Edwards, M., Raybaud, V., Goberville, E. & Kirby, R.R. (2015). Future vulnerability of
648 marine biodiversity compared with contemporary and past changes. Nature Climate Change,
649 5, 695-701.
650 5.
651 Beaugrand, G., Kirby, R.R. & Goberville, E. (2020). The mathematical influence on global patterns of
652 biodiversity. Ecology and Evolution, 10, 6494-6511.
653 6.
654 Beaugrand, G., Luczak, C., Goberville, E. & Kirby, R.R. (2018). Marine biodiversity and the chessboard
655 of life Plos One, 13, e0194006.
656 7.
657 Beaugrand, G., Rombouts, I. & Kirby, R.R. (2013). Towards an understanding of the pattern of
658 biodiversity in the oceans. Global Ecology and Biogeography, 22, 440–449.
659 8.
660 Benton, M.J. & Emerson, B.C. (2007). How did life become so diverse? The dynamics of diversification
661 according to the fossil record and molecular phylogenetics. . Palaeontology, 50, 23-40.
662 9.
663 Blackburn, T.M., Delean, S., Pysek, P. & Cassey, P. (2016). On the island biogeography of aliens: a
664 global analysis of the richness of plant and bird species on oceanic islands. Global Ecology
665 and Biogeography, 25, 859-868.
666 10.
667 Bush, M.B. & Whittaker, R.H. (1993). Non-equilibrium in island theory of Krakatau. Journal of
668 Biogeography, 20, 453-457.
669 11.
670 Cadena, C.D., Ricklefs, R.E., Jiménez, I. & Bermingham, E. (2005). Is speciation driven by species
671 diversity? Nature, 438, E1-E2.
672 12.
673 Carlquist, S. (1965). Island life: a natural history of the islands of the world. The American Museum of
674 Natural History, New York.
675 13.
676 Chai, T. & Draxler, R.R. (2014). Root mean square error (RMSE) or mean absolute error (MAE)?
677 Arguments against avoiding RMSE in the literature. Geoscientific Model Development, 7,
678 1247-1250.
679 14.
680 Darlington, P.J. (1957). Zoogeography: the geographical distribution of animals. Wiley, New York.
681 15.
682 Diamond, J.M. (1969). Avifaunal equilibria and species turnover rates on the Channel islands of
683 California. Proceedings of the National Academy of Sciences of the United States of America,
684 64, 57-63.
685 16.
686 Elton, C. (1927). Animal ecology. Sidgwick and Jackson, London.
687 17.
688 Emerson, B.C. & Kolm, N. (2005a). Emerson & Kolm reply. Nature, 438, E2.
689 18.
690 Emerson, B.C. & Kolm, N. (2005b). Species diversity can drive speciation. Nature, 434, 1015-1017.
691 19.
692 Fick, S.E. & Hijmans, R.J. (2017). WorldClim 2: New 1-km spatial resolution climate surfaces for global
693 land areas. International Journal of Climatology, 37, 4302–4315
694 20.
695 Gause, G.F. (1934). The struggle for coexistence. MD: Williams and Wilkins, Baltimore.
696 21.
697 Gillespie, R. (2004). Community assembly through adaptive radiation in Hawaiian spiders. Science,
698 303, 356-359.
699 22.
700 Gillespie, R.G. & Roderick, G.K. (2002). Arthropods on islands: Colonization, Speciation, and
701 Conservation. Annual Review of Entomology, 47, 595-632.
702 23.
703 Gleason, H.A. (1922). On the relation between species and area. Ecology, 3, 158-162.
704 24.
705 Guo, J. ( 2008). Fire and life. Nature 454, 930-932.
706 25.
707 He, F. & Legendre, P. (1996). On species-area relations. The American Naturalist, 148, 719-737.
708 26.
709 Hijmans, R.J., Cameron, S.E., Parra, J.L., Jones, P.G. & Jarvis, A. (2005). Very high resolution
710 interpolated climate surfaces for global land areas. International Journal of Climatology, 25,
711 1965-1978.
712 27.
713 Hirzel, A.H., Hausser, J., Chessel, D. & Perrin, N. (2002). Ecological-niche factor analysis: how to
714 compute habitat-suitability maps without absence data? Ecology, 83, 2027-2036.
715 28.
716 Hutchinson, G.E. (1957). Concluding remarks. Cold Spring Harbor Symposium Quantitative Biology,
717 22, 415-427.
718 29.
719 Hutchinson, G.E. (1978). An introduction to population ecology. Yale University Press, New Haven.
720 30.
721 Klein, D.R. (1968). The introduction, increase and crash of reindeer on Saint Mattew Island. Journal of
722 Wildlife Management, 32, 350-367.
723 31.
724 Kueffer, C., Drake, D.R. & Fernandez-Palacios, J.M. (2014). Island biology: looking towards the future.
725 Biology Letters, 10, 20140719.
726 32.
727 Lack, D. (1970). Island birds. Biotropica, 2, 29-31.
728 33.
729 Lack, D. (1976). Island biology, illustrated by the land birds of Jamaica. Blackwell Scientific
730 Publications, Oxford.
731 34.
732 Liu, X., Li, X., Liu, Z., Tingley, R., Kraus, F., Guo, Z. et al. (2014). Congener diversity, topographic
733 heterogeneity and humanassisted dispersal predict spread rates of alien herpetofauna at a
734 global scale. Ecology Letters, 17, 821-829.
735 35.
736 Lomolino, M.V., Brown, J.H. & Sax, D.F. (2009). Island biogeography theory. Reticulations and
737 reintegration of "a biogeography of the species". In: The theory of island biogeography
738 revisited (eds. Losos, JB & Ricklefs, RE). Princeton University Press Princeton, pp. 13-51.
739 36.
740 Lomolino, M.V., Riddle, B.R. & Brown, J.H. (2006). Biogeography. 3 edn. Sinauer Associates, Inc.,
741 Sunderland.
742 37.
743 MacArthur, R.H. & Wilson, E.O. (1963). An equilibrium theory of island zoogeography. Evolution, 17,
744 373-387.
745 38.
746 MacArthur, R.H. & Wilson, E.O. (1967). The theory of island biogeography. Princeton University Press,
747 Princeton
748 39.
749 Marshall, C.R. & Quental, T.B. (2016). The uncertain role of diversity dependence in species
750 diversification and the need to incorporate time-varying carrying capacities. Philosophical
751 Transactions of the Royal Society B, 371, 20150217.
752 40.
753 Mayr, E. (1965). The nature of colonisation in birds. In: The genetics of colonizing species. (eds. Baker,
754 HG & Stebbins, GL). Academic Press New York.
755 41.
756 Rominger, A.J., Goodman, K.R., Lim, J.Y., Armstrong, E.E., Becking, L.E., Bennett, G.M. et al. (2016).
757 Community assembly on isolated islands: macroecology meets evolution. Global Ecology and
758 Biogeography, 25, 769-780.
759 42.
760 Santos, A.M.C., Field, R. & Ricklefs, R.E. (2016). New directions in island biogeography. Global Ecology
761 and Biogeography, 25, 751-768.
762 43.
763 Schoener, T.W. (1983). Rate of species turnover decreases from lower to higher organisms: areview
764 of the data. Oikos, 41, 372-377.
765 44.
766 Schoener, T.W. (2009). The MacArthur-Wilson equilibrium model. A chronicle of what it said and how
767 it was tested. In: The theory of island biogeography revisited (eds. Losos, JB & Ricklefs, RE).
768 Princeton University Press Princeton, pp. 52-87.
769 45.
770 Sepkoski Jr, J.J. (1978). A kinetic model of Phanerozoic taxonomic diversity I. Analysis of marine
771 orders. Paleobiology, 4, 223-251.
772 46.
773 Simberloff, D. (1976). Experimental zoogeography of islands: effects of island size. Ecology, 57, 629-
774 648.
775 47.
776 Simberloff, D.S. (2009). Equilibrium theory of island biogeography and ecology. In: The theory of
777 island biogeography revisited (eds. Losos, JB & Ricklefs, RE). Princeton University Press
778 Princeton, pp. 161-182.
779 48.
780 Sokal, R.R. & Rohlf, F.J. (1995). Biometry. W.H. Freeman and compagny, New York.
781 49.
782 Storch, D., Šímová, I., Smyčka, J., Bohdalková, E., Toszogyova, A. & Okie, J.G. (2021). Biodiversity
783 dynamics in the Anthropocene: how human activities change equilibria of species richness.
784 Ecography, 44, 1-19.
785 50.
786 Stracey, C.M. & Pimm, S.L. (2009). Testing island biogeography theory with visitation rates of birds to
787 British islands. Journal of Biogeography, 36, 1532-1539.
788 51.
789 Thornton, I.W.B., Zann, R.A., Rawlinson, P.A., Tidemann, C.R., Adikerana, A.S. & Widjoya, A.H.T.
790 (1988). Colonization of the Krakatau Islands by vertebrates: Equilibrium, succession, and
791 possible delayed extinction Proceedings of the National Academy of Sciences of the United
792 States of America, 85, 515-518.
793 52.
794 Thornton, I.W.B., Zann, R.A. & Stephenson, D.G. (1990). Colonization of the Krakatau islands by land
795 birds, and the approach to an equilibrium number of species. Philosophical Transactions of
796 the Royal Society of London B, 327, 55-93.
797 53.
798 Thornton, I.W.B., Zann, R.A. & van Balen, S. (1993). Colonization of Rakata (Krakatau Is.) by non-
799 migrant land birds from 1883 to 1992 and implications for the value of island equilibrium
800 theory. Journal of Biogeography, 20, 441-452.
801 54.
802 Tjørve, E. (2012). Arrhenius and Gleason revisited: new hybrid models resolve an old controversy.
803 Journal of Biogeography, 39, 629-639.
804 55.
805 Valente, L.M., Illera, J.C., Havenstein, K., Pallien, T., Etienne, R.S. & Tiedemann, R. (2017). Equilibrium
806 bird species diversity in Atlantic Islands. . Current Biology, 27, 1660-1666.
807 56.
808 Veron, S., Haevermans, T., Govaerts, R., Mouchet, M. & Pellens, R. (2019). Distribution and relative
809 age of endemism across islands worldwide. Scientific Reports, 9, 11693
810 57.
811 Warren, B.H., Simberloff, D., Ricklefs, R.E., Aguilée, R., Condamine, F.L., Gravel, D. et al. (2015).
812 Islands as model systems in ecology and evolution: prospects fifty years after MacArthur-
813 Wilson. Ecology Letters, 18, 200-217.
814 58.
815 Whittaker, R.H. (1975). Communities and ecosystems. 2 edn. Macmillan, New York.
816 59.
817 Whittaker, R.J. & Fernandez-Palacios, J.M. (2007). Island biogeography. 2 edn. Oxford University
818 Press, Oxford.
819 60.
820 Whittaker, R.J., Fernandez-Palacios, J.M., Matthews, T.J., Borregaard, M.K. & Triantis, K.A. (2017).
821 Island biogeography: Taking the long view of nature’s laboratories. Science, 357, eaam8326.
822 61.
823 Whittaker, R.J., Ladle, R.J., Araujo, M.B., Fernandez-Palacios, J.M., Delgado, J.D. & Arévalo, J.R.
824 (2007). The island immaturity - speciation pulse model of island evolution: an alternative to
825 the "diversity begets diversity" model. Ecography, 30, 321-327.
826 62.
827 Whittaker, R.J., Triantis, K.A. & Ladle, R.J. (2008). A general dynamic theory of oceanic island
828 biogeography. Journal of Biogeography, 35, 977-994.
829 63.
830 Witt, C.C. & Maliakal-Witt, S. (2007). Why are diversity and endemism linked on islands? Ecography,
831 30, 331-333.
832 64.
833 Wright, S.J. (1981). Extinction-mediated competition: the anolis lizards and insectivorous birds of the
834 west indies. The American Naturalist, 117, 181-192.
835
1

2 Supplementary Information
3

6 Supplementary text (S1)

7 Supplementary figures (S1-2)

8 Supplementary tables (S1-4)

10

11

12

13

14

15

1
16 Supplementary text

17

18 Supplementary text S1: Mathematical description of the biogeographical model

19
20 We applied a model developed to reconstruct and investigate large-scale biodiversity patterns in the
21 terrestrial and marine realms1,2. We calculated rectangular niches in this model with a 0 corresponding
22 to an absence and a 1 to a presence (Supplementary fig. S2)3. All potential thermal niches ranged from
23 ρmin=tmin=-1.8°C to ρmax=tmax=44°C and all potential precipitation niches ranged from ρmin=pmin=0 mm
24 to ρmax=pmax=3000 mm; these thresholds were investigated in previous works and best fit the data in the
25 oceanic and terrestrial realms1. The ecological amplitude α of a niche (αT for temperature or αP
26 precipitation) varied between 1°C and 45.5°C for temperature and from 100 mm to 3000 mm for
27 precipitation by step of µ (µT for temperature and µP for precipitation). The amplitude α of a niche with
28 respect to temperature or precipitation was calculated as follows:
29
30 αi=αi-1+ µ with 2 ≤ i ≤ p (1)
31
32 With µ, the increment between niche amplitudes. µT (temperature) was fixed to 0.5°C and µP
33 (precipitation) was 100 mm. α1=1°C for temperature and 100 mm for precipitation. p was calculated as
34 follows:
35
 − 
=  max  1  +1
36 (2)
p 
  
37
38 The maximum amplitude αmax (αT for temperature or αP precipitation) was calculated as follows:
39
40 αmax= ρmax - ρmin (3)
41
42 Where ρmax=tmax and ρmin=tmin for temperature and ρmax=pmax and ρmin=pmin for precipitation.
43
44 Therefore, p varied as a function of both the minimum (α1) and maximum (αmax) niche amplitude, as
45 well as the increment between niche amplitudes (temperature or precipitation) µ. Column vector Ap=[αi]
46 with p=90 for temperature (α1=1°C and α90=45.5°C) and p=30 for precipitation (α1=100 mm and
47 α30=3000 mm). When α is large, the niche corresponds to an euryoecious species having the potential to
48 colonise many terrestrial (temperature and/or precipitation) regions. The weight of these euryoecious
49 species in the modeled biodiversity was low, however.
50

2
51 For a given niche amplitude αi (1≤ i ≤ p), the starting point of a niche x was a function of ρmin and ρmax
52 and the degree of overlapping between niches k, which was fixed to kT=0.5°C for temperature and
53 kP=100 mm for precipitation. No species had exactly the same niche according to the principle of
54 competitive exclusion of Gause4. For each niche amplitude αi, the starting point of a niche was calculated
55 as follows:
56 𝑥𝑖,𝑗 = 𝑥𝑖,𝑗−1 + 𝑘 1 ≤ i ≤ p 2 ≤ j ≤ q𝑖 (4)
57
58 With x.1= ρmin; x.1= tmin=-1.8°C for temperature and x.1= pmin=100 mm for precipitation. qi was calculated
59 as follows:
60
𝛼𝑚𝑎𝑥 +𝑘−𝛼𝑖
61 𝑞𝑖 = ⌊ 𝑘
⌋ 1 ≤ i ≤ p (5)

62
63 With αTmax= αT90=45.5 for temperature and αPmax= αP30=3000 for precipitation. Column vector Qp=[qi]
64 (Q90 for temperature and Q30 for precipitation). The ending point of a niche (temperature or
65 precipitation) y was determined by adding the niche amplitude to the starting point:
66
67 yi,j=xi,j+αi 1≤i≤p 1 ≤ j ≤ qi (6)
68
69 A total of r niches was built for temperature (rT) or precipitation (rP):
70
p
71 r = q (7)
i
i =1

72
73 With p being calculated in Equation (2). To remain close to 1 million niches and limit calculation times
74 for some islands (e.g. Greenland), we only considered odd qi to build the pool of potential niches for
75 precipitation. This led to rP=255 instead of rP= 495. rT=4185 for temperature.
76

77 The total number of niches R was the result of the multiplication of rT by rP:
78
79 R= rT . rP (8)
80
81 With rT and rP the number of niches based on temperature and precipitation, respectively.
82
83 Therefore, we had a total of 1,067,175 potential niches (rP = 255 precipitation x rT = 4185 thermal
84 niches).
85

3
86 Supplementary figures

87

88 Supplementary fig. S1. Ecogeographical patterns (a) and latitudinal biodiversity gradients (b) in
89 simulated richness assessed by the biodiversity model. The value in (b) is the median of all longitudes
90 for a given latitude. The vertical dashed line denotes the equator. Simulated patterns of species
91 richness are highly correlated with observed patterns of biodiversity for a variety of taxonomic
92 groups 1. In this simulation, one climatic niche gives one simulated species.

93

94 Supplementary fig. S2. Sketch diagram that summarises how the number of climatic niches was
95 assessed for each island.

96

4
97 Supplementary fig. S1

98
99

100

5
101 Supplementary fig. S2

102

103
104

6
105 Supplementary tables
106

107 Supplementary table S1. List of main symbols and acronyms used in the text.
108 Supplementary table S2. Estimated and observed species richness of birds and plants in
109 some islands. Data on observed species richness, area, latitude and distance to land are from
110 Blackburn and colleagues 5. NE: no estimation.
111 Supplementary table S3. Abiotic and biotic characteristics of islands used for herpetofauna.
112 Islands were selected according to data availability. Herpetofauna species richness was
113 assessed, island by island, from different resources of the literature, found with Google
114 scholar. Distances from land were calculated using Google Earth. Areas of islands were found
115 on Wikipedia or in the publication referenced in the table. Latitude and longitude are in
116 decimal degree. It corresponds to the mean latitude and longitude of the rectangle used to
117 delimit the position of the island. For Corse, species richness from 1989 (20 species) was
118 confirmed by recent observation recorded on Inaturalist.org (18 species). The highest
119 expectation, from Castanet and Guyetant 1989, was kept. Tokara, Amami, Okinawa, Miyako,
120 Yaeyama and Senkaku archipelagos were grouped as in Ota H 2000B. Japan islands were
121 grouped as in Ota H 2000A., and Jersey, Guernsey archipelagos as in Edgar P 2010.
122 Supplementary table S4. Statistics of the linear model assessing species richness as a function
123 of the mean number of available climatic niches (M), island area (A) and distance to mainland
124 (d). All variables were log10-transformed. Multi-regression coefficients ϒ1, ϒ2, ϒ3 are related to
125 M, A and d, respectively. Coefficient β is the y-intercept.
126

7
127 Supplementary Table S1.

128

129

Symbol or Definition Unit


acronym
A Area of an island km²
d Distance to mainland km
M Maximum number of climatic niches Niche
Bs Species richness at saturation Species
B0 Species richness at year t=0 Species
I0 Immigration rate at year t=0 Species . yr-1
Is Immigration rate at saturation Species . yr-1
E0 Long-term extinction rate at t=0 Species . yr-1
Es Long-term extinction rate at saturation Species . yr-1
It Immigration rate at year t Species . yr-1
Ft Short-term extinction rate at year t Species . yr-1
Et Long-term extinction rates at year t Species . yr-1
Gt Total extinction rates at year t Species . yr-1
Bt Species richness at year t Species
Beq Species richness at equilibrium Species
Teq Year at which species richness is at equilibrium Year
ETIB Equilibrium Theory of Island Biogeography -
TEI Theory of Ecological Impoverishment -
METAL MacroEcological Theory on the Arrangement of Life -
RMSE Root Mean Square Error -
130

131

132

8
133

134 Supplementary table S2.


Number of
Location climatic niches Alien species Native species Area Latitude Distance to land
Birds
Annobon 86051.5 1 11 17.5 -1 358
Ascension 36488.7 8 0 88 -8 1548
Azores 92975.4 3 21 2333 38.5 1388
Bermuda 60848.4 10 14 54 32 1042
Canaries 46328.7 28 64 7493 28 102
CapeVerde 59238.7 4 20 4033 16 577
Principe 92487.6 4 43 136 1.5 219
SaoTome 146679.8 4 63 854 0 242
SouthGeorgia 24163.2 1 8 3528 -54 1748
StHelena NE 9 11 121 -16 1859
Bahamas 52520.5 15 94 13880 24 106
Cuba 132435.3 9 126 110860 21.5 203
Hispaniola 143615.5 10 108 76480 18.5 575
Jamaica 136314.8 16 111 10991 18 662
PuertoRico 118868 42 121 13790 18 720
Aldabra 42984.1 2 20 153.8 -9.5 632
Amsterdam 61035.3 2 1 55 -37.5 3383
Andaman/NicobarIslands 144919.5 4 104 8249 10 286
Chagos 81060.4 6 3 56.13 -6 1630
Christmas(IndianOc) 78271.5 3 11 135 -10.5 1322
Cocos 116157.5 1 4 14.2 -12.5 1000
Comoros 134634.3 11 59 2034 -12 294
Madagascar 184949.3 7 185 587040 -20 418
Mauritius 83140.2 19 25 2040 -20 1870
Reunion 112646 23 31 2511 -21 1668

9
Rodrigues 45677.8 8 15 108 -19.5 2451
ChathamIslands 56559.4 11 35 966 -44 2870
CookIslands 92538 3 14 92.7 -15.5 5036
Easter 57193.5 5 9 163.6 -27 3515
Galapagos 51317.5 4 42 8010 0 943
Guam 79994.7 7 40 541 13.5 2912
HawaiiArchipelago 275796.9 54 52 16624 21 3673
Henderson 60453 0 5 37.3 -24 5421
JuanFernandez 50990.1 2 12 99.6 -33 600
KermadecIsland 73118.3 4 7 33 -30 2761
LordHowe 64897.3 7 24 14.55 -31.5 575
Marquesas 113870 8 13 1049.3 -9 4812
Nauru 65689.3 1 3 21 -0.5 2878
New Zealand 241061.2 34 95 268021 -42 1724
NewCaledonia 156959.6 13 70 18576 -21.5 1230
NorfolkIsland 59881.2 12 22 35 -29 1397
Palau 109417.2 5 31 459 7 2220
Samoa 219186.1 6 32 2842 -14 3773
Societies 120909.6 8 22 1590 -17 5760
Taiwan 282327.9 35 160 36193 24 140
Tonga 118605.8 5 22 748 -21 3254
Tuamotu NE 3 10 850 -19 5970
Vanuatu 188040 10 57 12190 -17.5 1778
WakeIsland 37777.9 2 6 2.85 19 4270
Antipodes 25398.7 4 4 22 -49.5 2681
Auckland 50776.2 10 11 510 -51 1935
CampbellIsland 41343.8 10 4 113 -52.5 2233
Cochons 53910.8 0 2 67 -46 2379
Falklands 22412.5 3 35 12200 -51.5 507
Gough 88243.8 0 2 91 -40 2586

10
Heard 14605 0 1 368 -53 3885
Inaccessible 60958 0 4 14 -37 2816
Kerguelen 21112.5 0 3 7215 -49.5 3888
MacquarieIsland 22426.1 4 0 128 -54.5 1971
Marion 59252.3 0 1 335 -47 1720
McDonald 14605 0 1 2.5 -53 3885
Nightingale 62369.3 0 3 3.2 -37 2798
Pinguoins NE 0 2 3 -46.5 2412
Possession 51810 0 2 150 -46.5 2497
PrinceEdward 48190 0 1 45 -46.5 1720
Snares 49562.1 4 6 3.5 -48 1782
St.Paul 12403.7 1 0 6 -39 3367
TristandaCunha 88290.8 1 3 98 -37 2774
Plants
Ascension 36488.7 266 23 88 -8 1548
Azores 92975.4 660 245 2333 38.5 1388
Bermuda 60848.4 303 165 54 32 1042
Canaries 46328.7 662 1366 7493 28 102
CapeVerde 59238.7 445 309 4033 16 577
Greenland 66180 86 427 2166086 71 847
Jan Mayen 16912.5 4 57 377 71 928
Madeira Archipelago 53468.7 579 646 801 32.5 648
Salvage Islands(Portugal) NE 24 80 2.73 30 378
South Georgia 24163.2 37 26 3528 -54 1748
StHelena NE 260 70 121 -16 1859
Bahamas 52520.5 246 1104 13880 24 106
CaymanIslands 53731.2 65 536 264 19.5 635
Cuba 132435.3 376 5790 110860 21.5 203
GuadeloupeandMartinique NE 360 1668 2756 15 427
PuertoRico 118868 633 2333 13790 18 720

11
Amsterdam 61035.3 57 38 55 -37.5 3383
Christmas(IndianOc) 78271.5 151 201 135 -10.5 1322
Cocos 116157.5 53 61 14.2 -12.5 1000
Mauritius 83140.2 731 685 2040 -20 1870
Reunion 112646 628 675 2511 -21 1668
Rodrigues 45677.8 280 154 108 -19.5 2451
SeychelleIslands 90388 247 233 459 -4.5 1312
CookIslands 92538 286 278 92.7 -15.5 5036
DesventuradasArchipelago 32273.7 6 21 5.36 -26 874
Easter 57193.5 64 46 163.6 -27 3515
Fiji 158059.6 323 1302 18274 -18 2621
FrenchPolynesia NE 520 959 4167 -17 5497
Galapagos 51317.5 266 492 8010 0 943
Guam 79994.7 185 327 541 13.5 2912
HawaiiArchipelago 275796.9 841 1146 16624 21 3673
Heron Island NE 25 27 0.29 -23.5 64
JuanFernandez 50990.1 266 196 99.6 -33 600
KermadecIsland 73118.3 88 117 33 -30 2761
Lord Howe 64897.3 202 219 14.55 -31.5 575
Makatea (TuamotuIslands) 61224.3 102 60 24 -16 5970
Mangareva Island 67443.7 60 85 18 -23 5745
Nauru 65689.3 85 50 21 -0.5 2878
New Caledonia 156959.6 324 3001 18576 -21.5 1230
New Zealand 241061.2 2069 2065 268021 -42 1724
NorfolkIsland 59881.2 244 157 35 -29 1397
Northern Line Islands
(Kiribati) 124648.1 41 35 435.7 -2 5357
NukuHiva(Marquesas
Islands) 105526.1 215 254 339 -9 4802
Okinawa NE 186 115 2271.3 26.5 406
Pitcairn Island 60453 40 40 4.6 -25 5622

12
Rurutu (Austral Islands) 78429.6 157 126 29 -22.5 5554
SanNicolas (Channel
Islands,CA) 32308.7 131 114 58.93 33 98
Santa Cruz Calif 32746.2 157 462 250 34 30
Tahiti 117566.3 373 495 1045 -17.5 5910
Taiwan 282327.9 270 3875 36193 24 140
Antipodes 25398.7 4 66 22 -49.5 2681
Auckland 50776.2 33 187 510 -51 1935
CampbellIsland 41343.8 70 144 113 -52.5 2233
Falklands 22412.5 97 163 12200 -51.5 507
Heard 14605 1 11 368 -53 3885
Kerguelen 21112.5 67 22 7215 -49.5 3888
Macquarie Island 22426.1 3 44 128 -54.5 1971
Marion 59252.3 14 22 335 -47 1720
Possession 51810 9 17 150 -46.5 2497
Prince Edward 48190 10 21 45 -46.5 1720
Souhern Shetland Islands 9290 0 2 3687 -62 1172
Tristanda Cunha 88290.8 113 70 98 -37 2774
135
136

13
137 Supplementary table S3.
138
Number
Area Distance from Species of
Islands Latitude Longitude Sources
(km²) land richness climatic
niches
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Montserrat 16.7475 -62.195 102 670 13 62680,3 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Ascension Island -7.94 -14.36 91 1600 8 36488,8 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Isle of Man 54.2375 -4.55 572 530 3 47661,8 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Henderson island -24.38 -128.325 37.3 5500 7 60453 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Britain 54.45 -2.5 229850 37 25 97825,1 Arnold R.H. Atlas of amphibians and reptiles in Britain. Natural Environment Research Council. 1995. 40p.
King J.L. et al. Ireland Red List No. 5: Amphibians, Reptiles & Freshwater Fish. National Parks and Wildlife Service,
Ireland 53.375 -8.125 81638 420 10 75892,1
Department of Arts, Heritage and the Gaeltacht, Dublin, Ireland. 2011. 77p
Daugherty C.H. et al. Taxonomic and conservation review of the New Zealand herpetofauna. New Zealand Journal of
New Zeland -40.75 172.5 268680 1660 65 241061,3
Zoology. 1994. Vol 21(4): 317-323.
Brown J.L. et al. Spatial Biodiversity Patterns of Madagascar's Amphibians and Reptiles. PLoS ONE. 2016. Vol 11(1):
Madagascar -18.6 46.85 587041 430 745 184949,3
e0144076. doi:10.1371/journal.pone.0144076.
Cox N. Chanson J. and Stuart S. The Status and Distribution of Reptiles and
Chypre 35.1375 33.475 9251 75 27 52596,3 Amphibians of the Mediterranean Basin. International Union for Conservation of Nature and Natural Resources (IUCN).
2006. 42p.
Cox N. Chanson J. and Stuart S. The Status and Distribution of Reptiles and
Malte 35.9375 14.3825 316 250 11 35878,8 Amphibians of the Mediterranean Basin. International Union for Conservation of Nature and Natural Resources (IUCN).
2006. 42p.
Japan (mains islands) 37.365 137.915 377915 185 66 166750,7 Ota H. Current status of the threatened amphibians and reptiles of Japan. Population Ecology. 2000A. Vol 42: 5-9
Tasmania -42 146.5 67031 230 31 98679,3 Cogger H.G. Reptiles and amphibians of Australia. Seventh edition. CSIRO Publishing.2014. 1033p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Guernsey archipelago 49.5675 -2.4275 78 50 5 34993,3 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Jersey archipelago 49.215 -2.1375 116 22 9 34237,8 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.
Edgar P. The Amphibians and Reptiles of the UK Overseas Territories, Crown Dependencies and Sovereign Base Areas.
Bermudes 32.3175 -64.765 53.2 1030 13 60848,4 Species Inventory and Overview of Conservation and Research Priorities. Amphibian and reptiles conservation. 2010.
132p.

14
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Tokara group 29.3875 129.475 101.35 760 8 113798,7
Tropics. 2000B. Vol 10(1): 51-62.
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Amami group 27.775 129.15 1030 710 31 107168,9
Tropics. 2000B. Vol 10(1): 51-62.
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Okinawa group 26.5 127.475 1206.98 650 47 84989,6
Tropics. 2000B. Vol 10(1): 51-62.
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Miyako group 24.8 125.05 159.26 560 22 76819,3
Tropics. 2000B. Vol 10(1): 51-62.
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Yaeyama group 24.325 123.65 591.46 415 36 86988,1
Tropics. 2000B. Vol 10(1): 51-62.
Ota H. The current geographic faunal pattern of reptiles and amphibians of the Ryukyu archipelago and adjacent regions.
Senkaku group 25.82 124.01 7 340 6 92631,3
Tropics. 2000B. Vol 10(1): 51-62.
Castanet J. and Guyetant R. Atlas de répartition des amphibiens et reptiles de France. Société Herpétologique de France.
1989. 191p.
Corse 42.1875 9.05 8722 85 20 62315
Inaturalist. Observations Corse. The CaliforniaAcademy of Sciences. Consult in April 2018. On
ligne.https://www.inaturalist.org/observations?place_id=10574&view=species&iconic_taxa=Amphibia,Reptilia
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Jamaica 18.1 -77.25 11425 640 70 136314,8
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Redonda 16.9375 -62.345 1.25 690 4 48187,9
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Saba 17.635 -63.235 13 780 9 57248
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Puerto Rico 18.2 -66.425 8870 715 79 118868
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Hispaniola 18.8 -71.4 76480 580 366 143615,5
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Cuba 21.55 -79.5375 104945 160 305 132435,3
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Marie Galante 15.935 -61.265 158.01 572 11 52574,3
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
St Martin 18.065 -63.08 87 810 25 49786
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Virgin Gorda 18.4675 -64.385 21 870 18 50467,3
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Guana 18.4775 -64.57 3.5 875 17 47435,7
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Dominica 15.425 -61.35 754 500 27 97542,9
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Martinique 14.6375 -61.02 1128 425 44 96334,2
Natural History. 2012. Vol 51(2): 85-166.
Powell R. and Henderson R.W. Island lists of West Indian amphibians and reptiles. Bulletin of the Florida Museum of
Guadeloupe 16.25 -61.53 1438 580 54 83434,8
Natural History. 2012. Vol 51(2): 85-166.
139

15
140 Supplementary Table S4.

141

Constants of the multiple regression analyses


ϒ1 (M) ϒ2 (A) ϒ3 (d) β
Plants 0.7595 0.2026 -0.4333 -1.6859
Herpetofauna 1.0934 0.1340 0.0186 -4.3839
Birds 1.4662 0.1618 -0.5380 -3.5392
142

143

16
144 References

145

146 1 Beaugrand, G., Kirby, R. R. & Goberville, E. The mathematical influence on global patterns of
147 biodiversity. Ecology and Evolution 10, 6494-6511, doi: 10.1002/ece3.6385 (2020).
148 2 Beaugrand, G., Luczak, C., Goberville, E. & Kirby, R. R. Marine biodiversity and the chessboard
149 of life Plos One 13, e0194006, doi:https://doi.org/10.1371/journal.pone.0194006 (2018).
150 3 Beaugrand, G., Rombouts, I. & Kirby, R. R. Towards an understanding of the pattern of
151 biodiversity in the oceans. Global Ecology and Biogeography 22, 440–449 (2013).
152 4 Gause, G. F. The struggle for coexistence. (MD: Williams and Wilkins, 1934).
153 5 Blackburn, T. M., Delean, S., Pysek, P. & Cassey, P. On the island biogeography of aliens: a
154 global analysis of the richness of plant and bird species on oceanic islands. Global Ecology
155 and Biogeography 25, 859-868 (2016).
156

17

You might also like