ArticlePDF Available

Hypothesis testing in biogeography

Authors:

Abstract and Figures

Often, biogeography is applied only as a narrative addition to phylogenetic studies and lacks scientific rigour. However, if research questions are framed as hypotheses, biogeographical scenarios become testable. In this review, we explain some problems with narrative biogeography and show how the use of explicit hypotheses is changing understanding of how organisms came to be distributed as they are. Developing synergies between biogeography, ecology, molecular dating and palaeontology are providing novel data and hypothesis-testing opportunities. New approaches are challenging the classic 'Gondwana' paradigm and a more complicated history of the Southern Hemisphere is emerging, involving not only general drivers such as continental drift and niche conservatism, but also drowning and re-emergence of landmasses, biotic turnover and long-distance colonization.
Content may be subject to copyright.
Hypothesis testing in biogeography
Michael D. Crisp
1
, Steven A. Trewick
2
and Lyn G. Cook
3
1
Research School of Biology, The Australian National University, Canberra, ACT 0200, Australia
2
Institute of Natural Resources, Massey University, Palmerston North, New Zealand
3
The University of Queensland, School of Biological Sciences, Brisbane, Qld 4072, Australia
Often, biogeography is applied only as a narrative addi-
tion to phylogenetic studies and lacks scientific rigour.
However, if research questions are framed as hypothe-
ses, biogeographical scenarios become testable. In this
review, we explain some problems with narrative bio-
geography and show how the use of explicit hypotheses
is changing understanding of how organisms came to be
distributed as they are. Developing synergies between
biogeography, ecology, molecular dating and palaeon-
tology are providing novel data and hypothesis-testing
opportunities. New approaches are challenging the clas-
sic ‘Gondwana’ paradigm and a more complicated his-
tory of the Southern Hemisphere is emerging, involving
not only general drivers such as continental drift and
niche conservatism, but also drowning and re-emer-
gence of landmasses, biotic turnover and long-distance
colonization.
What is biogeography?
Biogeography is the study of the distribution and evolution
of organisms through space and time [1]. New methods
have given impetus to the discipline: for example, geo-
graphic information systems (GIS) for spatial analysis
[2]; Bayesian molecular phylogenetics for dating diver-
gences between lineages [3]; and integrative models for
reconstructing distributional change through evolutionary
time, using either maximum likelihood [4] or Bayesian
inference [5]. Above all, renewed recognition that ecologi-
cal factors (e.g. climatic tolerance and dispersal limitation)
underlie deep historical events (i.e. speciation, extinction
and distributional change) [6,7] has rekindled interest in
old questions, such as ‘how do ecological factors influence
the processes of vicariance and long-distance dispersal and
establishment (LDDE)?’ (see Glossary) [68]. It has also
stimulated new questions, such as ‘what is the role of niche
conservatism in large-scale community assembly?’ [810].
In the beginning, with Wallace and Darwin, biogeography
was an exploration of evolution and it is popular today
because, with new methods, it can open windows on the
geographical dimensions of speciation. Although hypothe-
ses about ancient ecological processes are not testable by
direct observation or experiment, their predictions about
present-day biota can potentially be tested. These include
predictions about distributional patterns, fossils, likeli-
hoods of dispersal, and the shapes and timing of phyloge-
nies [11].
A purely inductive approach (‘pattern before process’) is
not science
Unfortunately, biogeography often lacks rigour when it is
presented as a geo-historical narrative for a single taxon,
commonly as an addendum to a phylogenetic analysis.
Biogeography deals with historical events that can neither
be observed directly nor manipulated experimentally, and
this limitation has been used to justify inductivism; that is,
the view that researchers should first observe and analyse
the present-day pattern and only then might explanations
emerge in terms of historical processes (‘pattern before
process’) [12,13]. In a commonly used inductivist approach,
Opinion
Glossary
Area cladogram: a phylogeny in which the names of the organisms at
the tips are replaced by those of the areas in which they occur (e.g.
[13,19]).
Ancestral area reconstruction (AAR): inference of hypothetical ances-
tral areas at the internal nodes (and root) of a phylogeny by ‘optimiz-
ing’ from known areas at the tips of an area cladogram. Several
methods are used for AAR, including parsimony and increasingly
complex models using maximum likelihood and Bayesian inference.
Biotic turnover: extinction and replacement of floras and faunas in
the fossil record, usually driven by global environmental change.
Crown age: the age of the most recent common ancestor shared by
the extant species of a monophyletic lineage. The crown age of a
lineage might be considerably younger than its stem age (Box 3,
Figure Ia). See also ‘Stem age’.
Stem age: the time when a lineage diverged from its sister group;
that is, from the lineage that includes its nearest living relatives. See
also ‘Crown age’.
Long-distance dispersal and establishment (LDDE)*: allopatric (geo-
graphical) speciation caused by an exceptional dispersal event,
establishing a new population on the far side of a barrier that
sufficiently limits subsequent gene flow between the parent and
daughter populations. See also ‘Vicariance’.
Niche conservatism: the notion that major ecological niches are
more conserved than expected through evolutionary time is based
on the observation from phylogenetic studies that major niche shifts
have been relatively rare [9].
Vicariance*: allopatric (geographical) speciation caused by the orig-
ination of a barrier within the range of the ancestral species, dis-
rupting gene flow between the now separated subpopulations. See
also ‘LDDE’.
West Wind Drift: the strongly asymmetrical flow of wind and ocean
currents from west to east in the temperate latitudes of the Southern
Hemisphere, thought to be responsible for directionally biased LDDE
in that hemisphere [19,20].
*Note that allopatric speciation requires processes in addition to
those that cause the disjunction and establishment of disjunct
populations. See examples in main text; for example, plant species
shared by Tasmania and New Zealand. However, the speciation
processes should be similar under either the vicariance or the LDDE
model.
Corresponding author: Crisp, M.D. (mike.crisp@anu.edu.au).
66 0169-5347/$ see front matter ß2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.tree.2010.11.005 Trends in Ecology and Evolution, February 2011, Vol. 26, No. 2
Box 1. The pattern-first and hypothesis-testing approaches can lead to different conclusions
A common question in biogeography asks ‘what is the geographical
origin of taxon A?’ Recent examples include Nilsson et al. [65] in
respect of marsupials and Brown et al. [48] in respect of Rhododen-
dron sect. Vireya. Here, we illustrate two different approaches to
formulating and testing biogeographical hypotheses, using the
Southern Hemisphere Callitroid clade of the cypress family (Cupres-
saceae) as a hypothetical example. Each approach results in a
different interpretation of the biogeographical history.
The first is a ‘pattern before process’ approach (Figure Ia), in which
the distributions of extant taxa are mapped at the tips of a phylogeny
and ancestral areas are reconstructed at internal nodes using any of
several methods [14]. Here, both parsimony (mapped in Figure Ia) and
maximum likelihood infer that the common ancestor of the Callitroid
clade probably originated in Australia and that its descendants
subsequently dispersed to New Zealand (green), New Caledonia
(red, twice), Patagonia (purple, three times) and South Africa (yellow).
However, this commonly used trait-mapping approach fails to
consider alternative hypotheses or data that are independent of the
tree.
The second approach (Figure Ib) uses the same molecular
phylogeny to illustrate how process-based hypothesis testing gives
a different conclusion. Fossils can be used to test vicariance versus
dispersal hypotheses by adding extinct lineages and their distribu-
tions to the phylogeny (Figure Ib), and by adding a time calibration
to the tree. Here, interpretation of fossils indicates that most
Callitroid genera were once more widespread across Gondwana
but have suffered extensive extinction [5658].Knowingthis
enables one to frame hypotheses of vicariance for some nodes
(those with daughter lineages distributed among continents) and
assess them using the tests detailed in Box 2. This approach leads,
in many cases, to the conclusions that vicariance (nodes labelled
V?) cannot be rejected as the cause of divergence. For example,
Fitzroya cupressoides in Patagonia (extant) probably diverged from
its sister Fitzroya tasmaniensis in Australia (now extinct) between
30 and 40 mya, about the time when these landmasses separated as
East Gondwana broke up (yellow bar). Not all labelled divergences
overlap the geological separation bars (e.g. Papuacedrus,Wid-
dringtonia and Austrocedrus) but their confidence intervals prob-
ably would, in which case vicariance hypotheses could not be ruled
out for these either. (This is a hypothetical example and the
relations of fossils have not yet been fully verified by experts. A
more rigorous approach to estimating divergence times would also
use the fossils to calibrate nodes in a molecular dating analysis; e.g.
using BEAST [3].)
[()TD$FIG]
Aust - SAm - Ant
Gondwana - NZ
0
102030
40
506070
80
90
100
Million years before present
Character: Areas
Key:
Parsimony reconstruction
Unordered
Northern Hemisphere
Patagonia
Australia - New Guinea
New Zealand
New Caledonia
Africa
Antarctica
Gondwana
Actinostrobus
Callitris
Neocallitropsis
Fitzroya
Diselma
Widdringtonia
Austrocedrus
Libocedrus plumosa
Libocedrus bidwillii
Pilgerodendron
Libocedrus yateensis
Papuacedrus papuana
Cupressoids
Taxodium
Glyptostrobus
Cryptomeria
(a)
Actinostrobus
Callitris leaensis X
Callitris
Neocallitropsis
Fitzroya
Fitzroya tasmanensis X
Diselma
Widdringtonia americana X
Widdringtonia
Austrocedrus tasmanica X
Austrocedrus
Libocedrus sp X
Libocedrus acutifolius X
Libocedrus spp X
Libocedrus plumosa
Libocedrus bidwillii
Pilgerodendron
Libocedrus yateensis
Papuacedrus shenii X
Papuacedrus sp X
Papuacedrus papuana
Papuacedrus australis X
Cupressoids
Taxodium
Glyptostrobus
Cryptomeria
Papuacedrus prechilensis X
(b)
V?
V?
V?
V?
CC CC
TRENDS in Ecology & Evolution
Figure I.(a) Ancestral area reconstruction using parsimony to map distributions of cypresses on a phylogeny redrawn from [66] infers an Australian origin for the
Callitroid clade followed by dispersal to other Southern Hemisphere landmasses. (b) Adding relationships of fossils (and thus a time calibration) to the phylogeny leads
to a failure to reject vicariance hypotheses based on the break up of Gondwana. The blue-shaded bar indicates separation of Zealandia from the remainder of Gondwana
and the geological origin of New Zealand [26,59]; the yellow-shaded bar indicates separation of Australia, Antarctica and Patagonia. Extinct taxa are labelled with cross
symbols. The most recent common ancestor of the Callitroid clade is labelled ‘CC’. Divergences for which vicariance is not rejected as the cause of the disjunction are
labelled ‘V?’.
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
67
ancestral areas are reconstructed at internal nodes of the
phylogeny; for example, using ancestral area reconstruc-
tion (AAR) methods (reviewed in [14,15]), which are some-
times combined with relaxed molecular-clock dating of
nodes (Box 1).
Conceptually, AAR does not differ from mapping pheno-
typic traits (‘standard’ or ‘morphological’ characters) onto
phylogenies. Geographical distribution is also a trait that
can be modelled and, similar to any trait, it can change
through time. Thus, standard ancestral trait reconstruction
models, based on parsimony, maximum likelihood and
Bayesian inference [5,16], have been used in biogeography.
Subsequently, complex biogeographical models have been
developed to take account of: (i) geological, ecological and
geographical factors relevant to distributional change; and
(ii) causal links between distributional change (e.g. vicari-
ance or LDDE), speciation and extinction ([4,17,18] and
references therein). AAR models differ in their (sometimes
unspecified) assumptions about processes, such as whether
speciation is constrained to accompany LDDE or whether
vicariance is favoured over LDDE [14]. Methods also differ
in whether, and to what extent, they enable ancestors to
occupy multiple areas (a required assumption for vicariance
[14]), and whether they model directional bias in LDDE
[19,20]. Ancestral area reconstructions are then typically
used to describe a sequence of distributional change through
time. Correlates between the inferred distributional
changes and ‘events’, such as continental break up or climate
change, are sought and often inferred as causative.
A logical problem with this type of approach, which is
not exclusive to biogeography, is that a finite set of obser-
vations can be consistent with an almost unlimited set of
alternative explanations [1,2123]. Moreover, ‘observa-
tions’ could be subjective or biased if the observer filters
the data through an explanatory theory, even if this pro-
cess is subconscious [1,21]. Proponents of inductive (‘pat-
tern before process’) biogeography commonly work from
implicit process assumptions, usually of vicariance [22,23].
The inductive approach has been criticised as storytelling
and unscientific: alone, it cannot progress beyond being a
speculative first attempt to understand the biogeography
of a group, because it tends to generate, rather than test,
hypotheses [1,13].
Biogeography becomes a science in the Popperian sense
when it frames and tests hypotheses [1,13]. Biogeographi-
cal knowledge can progress beyond the inductive hypothe-
sis-creation stage by framing restrictive propositions and
testing specific predictions that can rule out many of the
alternatives [1]. Thus, an untestable question, such as
‘where did Nothofagus (southern beeches) first evolve?’
does not express a specific prediction and could be replaced
by a testable hypothesis, such as ‘the disjunction between
Nothofagus sister taxa in Australia and South America
was caused by vicariance’. A hypothesis about an unob-
served process can be tested if it predicts an observable
outcome (e.g. pattern or timing) that contrasts with that
from an alternative hypothesis [24]. How biogeographical
questions are phrased dictates how they are addressed,
and can affect the interpretation of past events (Box 1).
To avoid circularity, it is important to test a hypothesis
using data that are independent of those used to frame it in
the first place [23,24]. For example, the hypothesis that the
entire terrestrial biota of New Zealand established and
diversified after the Oligocene was proposed on the basis of
multiple lines of geological evidence that indicate total
marine inundation of the landscape before 23 million years
ago (mya) [23,25]. This hypothesis can be tested by its
prediction that no terrestrial lineage occupied New Zeal-
and continuously through the Oligocene. The drowning
hypothesis would be falsified by the existence in New
Zealand of an endemic radiation with a crown age reliably
dated to the Oligocene (2334 mya) or older [23,26]. Phy-
logenies used for the test should be calibrated using inde-
pendent data (e.g. from fossils or stratigraphy), rather than
the non-independent geological data used to erect the
drowning hypothesis.
Here, we discuss some specific approaches to testing
hypotheses, using as examples the well-known models of
vicariance and dispersal that have been used to explain
disjunct distributions. Examples of biogeographical hy-
potheses and their testable predictions are detailed in
Table S1 (supplementary material online).
Testing alternative hypotheses to explain current
disjunct distributions
Vicariance and LDDE are both geographical (allopatry-
based) explanations for the process of speciation and,
although both probably had a role in the diversification
of lineages [8,27], many biogeographers treat them as
exclusive alternative models. Vicariance, by definition,
results from processes that restrict the dispersal of indi-
viduals within the range of a species [6] and this can occur
only after the range of a species has already expanded via
dispersal. Long-distance dispersal and establishment
requires that organisms overcome some barrier to gene
flow, but infrequently enough that populations on either
side of the barrier (or filter) speciate [19].
The relative contributions of dispersal and vicariance to
distributions of organisms in the Southern Hemisphere,
where closely related terrestrial species are disjunct across
wide oceanic gaps, have been debated extensively. Follow-
ing the recognition of plate tectonics [28], these distribu-
tions have often been explained as arising by vicariance
through continental drift [13,29]. Vicariance biogeography
under this scenario postulates that, as Gondwana broke
up, populations were sundered, isolated on the newly
formed landmasses and subsequently diverged to become
different species [13,29]. This scenario requires that each
species was widespread across much of Gondwana before
the break up of the supercontinent. Vicariance has mean-
ing in the evolutionary sense only when it is tied to a
divergence event. Thus, continental drift leading to the
separation of lineages across oceans is not a cause of
vicariance if the lineages were already diverged by the
time continental drift separated the landmasses.
The alternative LDDE model for transoceanic disjunc-
tions posits that, driven by rare events (such as storms or
tsunamis), organisms have been carried across gaps, such
as oceans, that are not normally traversed. The model also
allows for cases where propagule dispersal is more fre-
quent but survival and establishment is rare (possibly
linked with ecological and genetic factors) [30,31]. With
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
68
no (or minimal) gene flow, the separated (allopatric) popu-
lations evolve independently and, ultimately, speciate.
Tests of vicariance
If the pattern and timing of the origin of potential vicari-
ance events are known from geological data, vicariance
hypotheses are testable because they make several pre-
dictions (Table S1 in Supplementary Material Online). The
advent of molecular dating has led to the ability to test the
timing of divergences and thus test hypotheses of vicari-
ance (Box 2). Surprisingly, most transoceanic plant dis-
junctions [8] and many of those in animal taxa [26,32,33]
have been determined to be asynchronous or too young to
be fully explained by the break up of Gondwana. This
applies even in the case of iconic taxa, such as Nothofagus
[34] and kauri pines (Agathis)[35] in New Zealand,
ostriches in Africa [36] and primates and rodents in South
America [37].
Importantly, divergences can be too old to have been
caused by a particular geological event [26,31]: the predic-
tion of timing requires a two-tailed test (Box 2). By this
criterion, many of the cases of species-poor lineages that
are presented as evidence of long-term occupancy resulting
from vicariance, for example, tuatara in New Zealand and
Amborella in New Caledonia, fail the test of a vicariance
explanation [26].
Another important prediction from a hypothesis of vi-
cariance is that multiple lineages will probably be affected
by the origin of the putative barrier [7,29]. Thus, a further
prediction is that there should be divergences in multiple
taxa either side of that barrier dating to that time [7,38,39].
For example, alternative vicariance hypotheses have been
proposed for the middle of the Baja Peninsula, California,
putatively owing to either climate change during the Pleis-
tocene or marine incursion during the late Mioceneearly
Pliocene [38]. These were tested for coincident divergence
times across the barrier in multiple animal and plant taxa,
with some support found for vicariance at the earlier time
in nine taxa [40].
Are hypotheses of dispersal testable?
Commonly, dispersal is inferred as the default explanation
of a biogeographical disjunction following rejection of a
vicariance hypothesis, for example by molecular dating.
Therefore, it is important that LDDE hypotheses should be
testable using independent evidence. Despite claims that
hypotheses of dispersal are not testable [13], careful fram-
ing of hypotheses enables some to be tested. As illustrated
by the following examples, ecology has an increasing role in
testing dispersal hypotheses in historical biogeography.
Example 1. Model-fitting approaches can be used to test
dispersal-based hypotheses. For example, Sanmartı
´net al.
[20] used parsimony-based tree fitting to test the predic-
tion [19,41,42] that atmospheric and oceanic West Wind
Drift should cause an easterly bias in plant dispersals in
the Southern Hemisphere. Inferred LDDE events in 23
phylogenies were significantly asymmetrical in the pre-
dicted direction, rejecting the null hypothesis of equal rates
of inferred dispersal in both directions, as determined from
randomizations.
Example 2. Stepping-stone dispersal routes have often
been inferred to explain what, for some, might be seeming-
ly impossible LDDE events across extreme barriers. This
approach has been especially adopted for terrestrial taxa
that are disjunct across oceans, such as between Australia,
New Zealand and New Caledonia [43,44], Antarctica and
Africa via the Kerguelen Plateau [36] and between Africa
and Madagascar [32]. However, stepping-stone routes
might be even more problematic than a single jump across
a wider gap, because a stepping-stone hypothesis assumes
that an intermediate, reproducing population was large
enough and existed long enough to produce a ‘propagule (or
migrant) pressure’ [30] sufficient to colonize the next land-
mass along the chain. For example, it has been suggested
that a single extreme LDDE event could be more probable
than multiple shorter LDDE (stepping-stone model)
events. Long-distance seed ‘dispersal kernels’ (i.e. proba-
bility distributions of LDDE) appear to be ‘fat tailed’
[45,46]; that is, extreme LDDE is not much less probable
than LDDE over much shorter distances. This is partly
because of stochasticity and partly because of infrequent
atypical processes (e.g. cyclones and tsunamis) [45]. Given
that probabilities multiply in a chain of independent
Box 2. Tests of vicariance are two-tailed
Divergence times in molecular phylogenies can be used to test
hypotheses of vicariance [27]. Vicariance hypotheses predict that the
divergence time between taxa on either side of a barrier should
coincide with the timing of the origin of that barrier. The test is two
tailed. Vicariance is rejected if the divergence between the taxa is
too young (post-dates the origin of the barrier) or too old (pre-dates
origin of barrier) and, thus, the barrier could not have caused the
divergence (Figure I). The test of vicariance is explicit as it addresses
a specific divergence (node) in the phylogeny, which is hypothe-
sized to be caused by the origin of a particular barrier. A rejection of
one vicariance event does not equate to ‘vicariance does not explain
the distribution of this taxon’. It can reject only the hypothesis that
‘vicariance event X explains node Y’.
[()TD$FIG]
0
20
40
60
80
100
Geological timing
of vicariance
(Australia-South America)
Geological timing of vicaria
-nce (Australia/South
America-New Zealand)
Australia
South America
New Zealand
Australia
South America
New Zealand
Australia
South America
New Zealand
Australia
South America
New Zealand
Ma
(a) (b) (c) (d)
Current geographic
distribution of lineage
TRENDS in Ecology & Evolution
Figure I. Four scenarios showing different timing of divergences between
three lineages. Each phylogeny enables the testing of two hypotheses of
vicariance: one between South America and Australia (with red confidence
interval bars) and another between Australia + South America and New
Zealand (with dark-blue confidence interval bars). All vicariance hypotheses
cannot be rejected in (a) and (b) because the divergence-time error bar
overlaps the relevant geological time bar in each case. In (c) and (d), all
vicariance hypotheses are rejected because the respective error bars and
divergence time bars do not overlap.
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
69
events, a single, long LDDE is likely to be more probable
than are multiple, shorter steps. Using the hypothetical
dispersal kernel of Nathan ([45]: Figure 2, corrected ver-
sion, published 17 October 2006), the probability of a single
seed arrival over 500 km is P=10
16
and that of a single
seed arrival over 1000 km is P=10
18
. However, the
probability of two consecutive jumps over 500 km, with
the second contingent on the first, is P= (10
16
)
2
=10
32
;
that is, more improbable than the single jump over
1000 km.
Example 3. Ecological parameters, such as the above
dispersal probability kernels, can be included in model-
based tests of alternative dispersal hypotheses [4,18]. This
approach integrates ‘historical’ and ‘ecological’ biogeogra-
phy, two domains once thought to be independent because
of their differing time scales and treatment of evolution
(ancient and evolutionary versus recent and non-evolution-
ary, respectively) [47]. Webb and Ree [18] compared two
alternative hypotheses from [48] to explain the occurrence
of species of Rhododendron sect. Vireya on both sides of
Wallace’s Line, a putatively ancient division between the
biotas of South-east Asia and Australasia, caused by plate
tectonics [49]. Webb and Ree used SHIBA [18], a program
that simulates lineage movement on a changing historical
landscape, as determined from geological data, and makes
probabilistic estimates of ancestral ranges. Their model
also included ecological parameters from the theory of
island biogeography [50], namely survival versus area,
and dispersal versus distance. The authors then used this
model to test contrasting hypotheses about the age of the
radiation of Rhododendron sect Vireya in the island archi-
pelago of Malesia by comparing the likelihoods of bio-
geographical reconstructions using the alternative root
ages (55 mya vs 12 mya). Their test determined that a
single LDDE event at 55 mya was more probable than
shorter stepping-stone dispersals through islands that
came into existence more recently.
The problem of extinction
Extinction has long been acknowledged as a key determi-
nant of observable biogeographical patterns, but is often
considered intractable and ignored [13]. One reason is that
it is difficult to reconstruct (Box 3) unless the fossil record
provides compelling evidence of the former presence of taxa
in areas where they are no longer found [51]. The reverse,
lack of fossil evidence of a former presence of a taxon in a
given area, should not be accepted prima facie as evidence
that it was always absent, given the stochastic nature of
the fossil record (Box 3,Figure Ib).
Despite the difficulties, it is essential to consider extinc-
tion in testing biogeographical hypotheses because it can
result in false reconstructions that appear to be well
supported (Box 3). Biotic turnover has probably been over-
looked because fossils of extinct lineages have been mis-
assigned to younger, related lineages that have
immigrated more recently, giving a false impression of
long-term occupancy of a region by the original lineage
(Box 3, Figure Ic). For example, in New Zealand, evidence
is emerging of previously overlooked floristic turnover
through the Cenozoic [52], for example in Nothofagus
[34], Ericaceae [53] and Agathis [35].
Thus, the fossil record, and the probable biases and/or
uncertainty it implies, should be considered as far as
possible in biogeographical analysis [51]. For example,
some extant ‘Gondwanan’ groups have an unequivocal,
even extensive, ancient fossil record in the Northern Hemi-
sphere, where they have apparently gone extinct (cf. Box 3,
Figure Id); for example, marsupials [54], Rhynchocephalia
(tuatara) [55], southern conifers such as Araucariaceae and
many Podocarpaceae [56]. Similarly, there are fossils of
several genera of the cypress family (Cupressaceae) from
Southern Hemisphere areas where they are now extinct
[5658];Box 1 illustrates how incorporating this fossil
evidence into hypotheses can change how researchers
assess the biogeographical history of the family.
Geographically restricted taxa that are species poor and
sister to a species-rich lineage (often referred to as ‘relicts’)
Box 3. Extinction needs to be considered in hypothesis
formulation
No current method of AAR using phylogenies can reconstruct as
ancestral an area that has not been observed in present-day species
and, consequently, has not been included in the analysis, cf. [67].
Only the fossil record (if available) can provide evidence of former
occurrences of taxa in areas where they are now extinct (e.g.
[51,64]). Extinction can mislead by differentially erasing any kind of
biogeographical pattern, including dispersal pattern, and it can
remove evidence from either a particular time period or a particular
region.
[()TD$FIG]
Space
Time
(a) (c) (d)(b)
(a) (c) (d)
TRENDS in Ecology & Evolution
Figure I. Biogeographical histories of hypothetical lineages. Open circles
represent living taxa, filled circles indicate fossil taxa, black lines indicate actual
phylogenetic relationships and coloured bars indicate hypothetical locations
through time. As biogeographical history is ‘known’, one can intuit that: (a) the
age of a crown group does not equate to the age of the lineage in a particular
area; (b) current absence of a lineage does not equate to past absence; (c)
current presence of a lineage (pink area) does not equate to continuous past
presence; and (d) relict taxa or living fossils do not necessarily indicate long
occupation of an area, but might reflect high levels of extinction. Using a
phylogeny from sequence data for extant taxa (at bottom of figure), rate
modelling and AAR, a biogeographer could infer a common ancestor for each
of (c) and (d) but could not infer where the ancestor existed through time.
Inclusion of one available dated fossil (solid-black circle) could result in correct
time calibration and inference of ancestral locality. Alternatively, inclusion of a
different dated fossil (solid-red circle) could lead to incorrect inferences of both
the location and age of the common ancestor. Thus, uncertainty about the
placement of fossils yields uncertainty about biogeographical inference,
regardless of the sophistication of the phylogenetic tools.
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
70
tend to invite speculation about their origins and biogeog-
raphy. Examples include Ginkgo in China, tuatara and
Agathis in New Zealand, and the endemic shrub Amborella
in New Caledonia. However, extant taxa indicate persis-
tence in time only, not in space (Box 3, Figure Id), and
‘relict’ lineages cannot be assumed to have occupied the
present space throughout the existence of the lineage. Such
lineages have probably been subject to considerable extinc-
tion and, in the absence of additional data, are essentially
uninformative about biogeographical history, presenting
little scope for erecting testable hypotheses. Even though
the above taxa (except Amborella) have an excellent an-
cient fossil record and have been geographically wide-
spread in the past, their restriction to a single surviving
species in a localized area is shrouded in mystery.
Conclusions and the way ahead
Understanding of how lineages became distributed as they
are has changed dramatically because biogeographers are
taking a more focused, critical approach. Sweeping ques-
tions such as ‘where did cypresses evolve?’ are being
replaced with focused, testable hypotheses, such as ‘the
ancestor of the extant cypress species of Libocedrus in New
Zealand arrived by LDDE after the Oligocene drowning’.
Consequently, it has been learned that the geographical
evolution of biota has been driven by a greater diversity of
processes with a more complex history than under a simple
vicariance (or dispersal) paradigm. For example, tests of
predictions from geology and ecology have shown that, to a
large degree, New Zealand and New Caledonia resemble
‘oceanic’ islands with young, immigrant biota, rather than
‘continental’ islands with relictual ‘Gondwanan’ biota
[26,59,60].
Future biogeographical models will become more com-
plex, sophisticated and realistic, as they incorporate esti-
mates of ecological parameters, such as dispersal kernels
[7,61]. Models can be used to test hypotheses by varying
the parameter under question while holding others con-
stant, within a statistical framework [7,11,18]. However,
models require validation with independent empirical data
on crucially important parameter values [19,23] and these
are difficult to obtain, especially in an historical context.
Important parameters to include are the shape of the tail of
the LDDE distribution and the distances beyond which
reduced gene flow leads to divergence. Such parameters
are difficult to quantify and are likely to be species or
ecology specific. Current historical models use parameter
values that are either best guesses, or worse, are estimated
from phylogenies and, thus, not independent of them.
Increasingly, geo-referenced ecological and climatic
parameters are being integrated into tests of alternative
spatial models of community diversification and distribu-
tion [7]. Current climatic models are well validated and
implemented in GIS at fine geographical and seasonal
scales. Extending such models to ancient time periods is
challenging, partly because past climates are commonly
reconstructed using fossil evidence, so using the recon-
structions for testing biogeographical hypotheses could
be circular.
The fossil record is emerging again as being crucially
important in biogeography (e.g. [51]), and we have reiter-
ated here, with examples, that ignorance of the role of
extinction can lead to misinterpretation. Auspiciously, new
collaborations between palaeontologists and molecular
systematists [8,35,53,62,63] are leading to reinterpretion
of fossils, resulting in improvement of the phylogenetic
placement of calibration points and more reliable diver-
gence time estimates. In addition, ecological parameters
estimated from extant organisms can help explain distri-
butional changes when compared with the fossil record.
For example, ecophysiological tolerances were measured in
living conifer genera, some of which are extinct in
Australia but have a fossil record there [64]. It was found
that, unlike the extant Australian genera, those that are
extinct probably had moisture tolerances that fell outside
the current range of climates in Australia [64]. This type of
integrative approach is resulting in more critical tests of
biogeographical hypotheses and is changing the current
view of the history of the biota of the world.
Acknowledgements
We thank the ‘papers-in-the-pub’ systematics discussion group at The
University of Queensland, especially Nate Hardy, and the Biogeography
and Systematics group, Massey University, especially Ian Henderson and
Mary Morgan-Richards, for their input that helped improve the article.
We also acknowledge funding support from the Australian Biological
Resources Study and the Australian Research Council.
Appendix A. Supplementary data
Supplementary data associated with this article can be
found, in the online version, at doi:10.1016/j.tree.2010.
11.005.
References
1 Ball, I.R. (1976) Nature and formulation of biogeographical hypotheses.
Syst. Zool. 24, 407430
2 Rosauer, D. et al. (2009) Phylogenetic endemism: a new approach for
identifying geographical concentrations of evolutionary history. Mol.
Ecol. 18, 40614072
3 Drummond, A.J. et al. (2006) Relaxed phylogenetics and dating with
confidence. PLoS Biol. 4, e88
4 Ree, R.H. and Smith, S.A. (2008) Maximum likelihood inference of
geographic range evolution by dispersal, local extinction, and
cladogenesis. Syst. Biol. 57, 414
5 Lemey, P. et al. (2009) Bayesian phylogeography finds its roots. PLoS
Comput. Biol. 5, 16
6 Wiens, J.J. and Donoghue, M.J. (2004) Historical biogeography, ecology
and species richness. Trends Ecol. Evol. 19, 639644
7 Riddle, B.R. et al. (2008) The role of molecular genetics in sculpting the
future of integrative biogeography. Prog. Phys. Geogr. 32, 173202
8 Crisp, M.D. et al. (2009) Phylogenetic biome conservatism on a global
scale. Nature 458, 754756
9 Donoghue, M.J. (2008) A phylogenetic perspective on the distribution of
plant diversity. Proc. Natl. Acad. Sci. U. S. A. 105, 1154911555
10 Wiens, J.J. et al. (2010) Niche conservatism as an emerging principle in
ecology and conservation biology. Ecol. Lett. 13, 13101324
11 Pigot, A.L. et al. (2010) The shape and temporal dynamics of
phylogenetic trees arising from geographic speciation. Syst. Biol. 59,
660673
12 Andersson, L. (1996) An ontological dilemma: epistemology and
methodology of historical biogeography. J. Biogeogr. 23, 269277
13 Humphries, C.J. (2004) From dispersal to geographic congruence:
comments on cladistic biogeography in the twentieth century. In
Milestones in Systematics (Williams, D.M. and Forey, P.L., eds), pp.
225260, CRC Press
14 Lamm, K.S. and Redelings, B.D. (2009) Reconstructing ancestral
ranges in historical biogeography: properties and prospects. J. Syst.
Evol. 47, 369382
15 Kodandaramaiah, U. (2010) Use of dispersal-vicariance analysis in
biogeography; a critique. J. Biogeogr. 37, 311
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
71
16 Lewis, P.O. (2001) A likelihood approach to estimating phylogeny from
discrete morphological character data. Syst. Biol. 50, 913925
17 Sanmartı
´n, I. et al. (2008) Inferring dispersal: a Bayesian approach to
phylogeny-based island biogeography, with special reference to the
Canary Islands. J. Biogeogr. 35, 428449
18 Webb, C.O. and Ree, R.H. Historical biogeography inference in Malesia.
In Biotic evolution and environmental change in Southeast Asia (Gower,
D. and Ruber, L., eds.), Cambridge University Press (in press)
19 Cook, L.G. and Crisp, M.D. (2005) Directional asymmetry of long-
distance dispersal and colonisation could mislead reconstructions of
biogeography. J. Biogeogr. 32, 741754
20 Sanmartı
´n, I. et al. (2007) West Wind Drift revisited: testing for
directional dispersal in the Southern Hemisphere using event-based
tree fitting. J. Biogeogr. 34, 398416
21 Popper, K.R. (1959) The Logic of Scientific Discovery, Hutchinson
22 McDowall, R.M. (2004) What biogeography is: a place for process. J.
Biogeogr. 31, 345351
23 Waters, J.M. and Craw, D. (2006) Goodbye Gondwana? New Zealand
biogeography, geology, and the problem of circularity. Syst. Biol. 55,
351356
24 Penny, D. and Phillips, M.J. (2004) The rise of birds and mammals: are
microevolutionary processes sufficient for macroevolution? Trends
Ecol. Evol. 19, 516522
25 Landis, C.A. et al. (2008) The Waipounamu Erosion Surface:
questioning the antiquity of the New Zealand land surface and
terrestrial fauna and flora. Geol. Mag. 145, 173197
26 Goldberg, J. et al. (2008) Evolution of New Zealand’s terrestrial fauna:
a review of molecular evidence. Philos. Trans. R. Soc. B, Biol. Sci. 363,
33193334
27 Donoghue, M.J. and Moore, B.R. (2003) Toward an integrative
historical biogeography. Integr. Comp. Biol. 43, 261270
28 Hammond, A.L. (1971) Plate tectonics: the geophysics of the Earth’s
surface. Science 173, 4041
29 Rosen, D.E. (1978) Vicariant patterns and historical explanation in
biogeography. Syst. Zool. 27, 159188
30 Simberloff, D. (2009) The role of propagule pressure in biological
invasions. Annu. Rev. Ecol. Evol. Syst. 40, 81102
31 Trewick, S.A. and Gibb, G.C. (2010) Vicars, tramps and assemblyof the
New Zealand avifauna: a review of molecular phylogenetic evidence.
Ibis 152, 226253
32 Poux, C. et al. (2005) Asynchronous colonization of Madagascar by the
four endemic clades of primates, tenrecs, carnivores, and rodents as
inferred from nuclear genes. Syst. Biol. 54, 719730
33 Wallis, G.P. and Trewick, S.A. (2009) New Zealand phylogeography:
evolution on a small continent. Mol. Ecol. 18, 35483580
34 Cook, L.G. and Crisp, M.D. (2005) Not so ancient: the extant crown
group of Nothofagus represents a post-Gondwanan radiation. Proc. R.
Soc. B, Biol. Sci. 272, 25352544
35 Biffin, E. et al. (2010) Did kauri (Agathis: Araucariaceae) really survive
the Oligocene drowning of New Zealand? Syst. Biol. 59, 594610
36 Phillips, M.J. et al. (2010) Tinamous and moa flock together:
mitochondrial genome sequence analysis reveals independent losses
of flight among ratites. Syst. Biol. 59, 90107
37 Poux, C. et al. (2006) Arrival and diversification of caviomorph rodents
and platyrrhine primates in South America. Syst. Biol. 55, 228244
38 Lindell, J. et al. (2006) Deep genealogies and the mid-peninsular
seaway of Baja California. J. Biogeogr. 33, 13271331
39 Crisp, M.D. and Cook, L.G. (2007) A congruent molecular signature of
vicariance across multiple plant lineages. Mol. Phylogenet. Evol. 43,
11061117
40 Riddle, B.R. and Hafner, D.J. (2006) A step-wise approach to
integrating phylogeographic and phylogenetic biogeographic
perspectives on the history of a core North American warm deserts
biota. J. Arid Envir. 66, 435461
41 Waters, J.M. and Roy, M.S. (2004) Out of Africa: the slow train to
Australasia. Syst. Biol. 53, 1824
42 Mun
˜oz, J. et al. (2004) Wind as a long-distance dispersal vehicle in the
Southern Hemisphere. Science 304, 11441147
43 Ladiges, P.Y. and Cantrill, D. (2007) New CaledoniaAustralian
connections: biogeographic patterns and geology. Aust. Syst. Bot. 20,
383389
44 Heads, M. (2008) Panbiogeography of New Caledonia, south-west
Pacific: basal angiosperms on basement terranes, ultramafic
endemics inherited from volcanic island arcs and old taxa endemic
to young islands. J. Biogeogr. 35, 21532175
45 Nathan, R. (2006) Long-distance dispersal of plants. Science 313, 786
788
46 Hardy, O.J. (2009) How fat is the tail? Heredity 103, 437438
47 Rosen, B.R. (1988) Biogeographic patterns: a perceptual overview. In
Analytical Biogeography (Myers, A.A. and Giller, P.S., eds), pp. 2355,
Chapman and Hall
48 Brown, G.K. et al. (2006) Historical biogeography of Rhododendron
section Vireya and the Malesian Archipelago. J. Biogeogr. 33, 1929
1944
49 Whitmore, T.C. (1982) Wallace’s Line: a result of plate tectonics. Ann.
Mo. Bot. Gard. 69, 668675
50 McArthur, R.H. and Wilson, E.O. (1967) The Theory of Island
Biogeography, Princeton University Press
51 Quental, T.B. and Marshall, C.R. (2010) Diversity dynamics: molecular
phylogenies need the fossil record. Trends Ecol. Evol. 25, 434441
52 Pole, M.S. (2001) Can long-distance dispersal be inferred from the New
Zealand plant fossil record? Aust. J. Bot. 49, 357366
53 Jordan, G.J. et al. (2010) Fossil Ericaceae from New Zealand:
deconstructing the use of fossil evidence in historical biogeography.
Am. J. Bot. 97, 5970
54 Vullo, R. et al. (2009) The oldest modern therian mammal from Europe
and its bearing on stem marsupial paleobiogeography. Proc. Natl Acad.
Sci. U. S. A. 106, 1991019915
55 Jones, M.E. et al. (2009) A sphenodontine (Rhynchocephalia) from the
Miocene of New Zealand and palaeobiogeography of the tuatara
(Sphenodon). Proc. R. Soc. B.Biol. Sci. 276, 13851390
56 Hill, R.S. and Brodribb, T.J. (1999) Southern conifers in time and
space. Aust. J. Bot. 47, 639696
57 Wilf, P. et al. (2009) Papuacedrus (Cupressaceae) in Eocene
Patagonia: a new fossil link to Australasian rainforests. Am.J.Bot.
96, 20312047
58 Paull, R. and Hill, R.S. (2010) Early Oligocene Callitris and Fitzroya
(Cupressaceae) from Tasmania. Am. J. Bot. 97, 809820
59 Trewick, S.A. et al. (2007) Hello New Zealand. J. Biogeogr. 34, 16
60 Grandcolas, P. et al. (2008) New Caledonia: a very old Darwinian
island? Philos. Trans. R. Soc. B, Biol. Sci. 363, 33093317
61 Nathan, R. et al. (2008) Mechanisms of long-distance seed dispersal.
Trends Ecol. Evol. 23, 638647
62 Ho, S.Y.W. and Phillips, M.J. (2009) Accounting for calibration
uncertainty in phylogenetic estimation of evolutionary divergence
times. Syst. Biol. 58, 367380
63 Sauquet, H. et al. (2009) Contrasted patterns of hyperdiversification in
Mediterranean hotspots. Proc. Natl. Acad. Sci. U. S. A. 106, 221225
64 Hill, R.S. (2004) Origins of the southeastern Australian vegetation.
Philos. Trans. R. Soc. B, Biol. Sci. 359, 15371549
65 Nilsson, M.A. et al. (2010) Tracking marsupial evolution using archaic
genomic retroposon insertions. PLoS Biol. 8, e1000436
66 Gadek, P.A. et al. (2000) Relationships within Cupressaceae sensu lato:
a combined morphological and molecular approach. Am. J. Bot. 87,
10441057
67 Oakley, T.H. and Cunningham, C.W. (2000) Independent contrasts
succeed where ancestor reconstruction fails in a known bacteriophage
phylogeny. Evolution 54, 397405
Opinion Trends in Ecology and Evolution February 2011, Vol. 26, No. 2
72
... Australian Berberidopsidaceae species thrive in the rainforests of eastern Australia (Harden, 1990;POWO, 2022), whereas in Chile, the habitat of B. corallina is restricted to the temperate rainforests of the coastal mountain range between latitudes 35°58′S and 40°51′S, at an altitude range from sea level to 700 m elevation (Ministerio del Medio Ambiente, 2022) (Fig. 1). Time-calibrated phylogenies can provide information about divergence times and thus can be used to test these two main biogeographic hypotheses (Crisp & al., 2011). Chile hosts the "Chilean winter rainfall-Valdivian forests" (25°S-47°S), one of the 36 world biodiversity hotspots, given its high level of endemism and threat (Myers & al., 2000;. ...
... In contrast, our dates provide evidence for a recent longdistance dispersal. These findings further support the idea that long-distance dispersal events cannot be discarded as the primary causal explanation in shaping the current distribution patterns of Gondwanic elements, being more common and important than previously thought (Sanmartín & Ronquist, 2004;Crisp & al., 2011;Heads, 2013). According to our results, the divergence time between the South American B. corallina and B. granitica took place between the late Miocene and the early Pleistocene. ...
Article
Full-text available
Berberidopsidales comprises two families: monotypic Aextoxicaceae (Aextoxicon punctatum in Chile and Argentina) and Berberidopsidaceae. The latter includes Australian monotypic Streptothamnus (S. moorei) and Berberidopsis (B. beckleri in Australia and B. corallina in Chile). A new Berberidopsis species from the Central Chilean Andes is here described. Phylogenetic analyses based on nuclear and chloroplast data clarified the relationships within Berberidopsidales. The new species, Berberidopsis granitica, is sister to the Chilean endemic B. corallina, and this clade is sister to the Australian B. beckleri. The dated molecular phylogeny places the split between the South American B. corallina and B. granitica into the late Miocene/early Pleistocene and the split between South American and Australian Berberidopsis to the late Miocene and Pliocene, suggesting transoceanic dispersal rather than vicar-iance. Climatic niche analyses show two distinct and non-overlapping climatic niches for the Australian and southern South American species. Berberidopsis granitica and B. corallina also differ clearly in their habitat and morphology, in addition to their climatic niche. Berberidopsis granitica has a very restricted distribution area and grows in the Andes under montane climatic conditions, unique within Berberidopsidales.
... However, the suitability of current distributions to reconstruct evolutionary history varies [2] as the approach rests upon the assumption that current ranges represent past distributions. In particular, the role of extinction in determining observable biogeographic patterns is too often ignored [3]. Rates of extinction are likely to vary in time and space, for example rapid climate change is likely to increase rates of extinction and the restricted land area of islands is likely to elevate rates of extinction of terrestrial organisms compared with continents [4]. ...
... These island endemics (including four monotypic genera) either have living relatives on some other land masses that have remained unrecognized by taxonomy, or their ancestral populations have never been sampled due to lack of resources or their extinction. Extinction is a key component of evolution but one that can readily mislead biogeographic inference [3,22]. ...
Article
Full-text available
The biota of continents and islands are commonly considered to have a source–sink relationship, but small islands can harbour distinctive taxa. The distribution of four monotypic genera of Orthoptera on young subantarctic islands indicates a role for long-distance dispersal and extinction. Phylogenetic relationships were inferred from whole mtDNA genomes and nuclear sequences (45S cassette; four histones). We used a fossil and one palaeogeographic event to calibrate molecular clock analysis. We confirm that neither the Australian nor Aotearoa-New Zealand Rhaphidophoridae faunas are monophyletic. The radiation of Macropathinae may have begun in the late Jurassic, but trans-oceanic dispersal is required to explain the current distribution of some lineages within this subfamily. Dating the most recent common ancestor of seven island endemic species with their nearest mainland relative suggests that each existed long before their island home was available. Time estimates from our fossil-calibrated molecular clock analysis suggest several lineages have not been detected on mainland New Zealand, Australia, or elsewhere most probably due to their extinction, providing evidence that patterns of extinction, which are not consistently linked to range size or lineage age, confound biogeographic signal.
... To avoid post hoc rationalization, however, we advocate the need to anticipate expectations before performing analyses. In other words, to adopt a hypothetico-deductive approach, as has been done, for example, in biogeography (10). ...
... Theoretical and empirical studies suggest that the limited gene flow between populations, imposed by barriers, can lead to population divergence and a process of speciation through different modes: [1] geographical vicariance, when the emergence of a physical barrier geographically divides two populations; [2] ecological vicariance, when environmental changes create an area with unsuitable conditions that impede gene flow between populations persisting in suitable areas. This disruption of gene flow leads to both geographical isolation and ecological divergence; [3] ecogeographical vicariance, when a physical barrier initially emerges and subsequently leads to adaptation of populations by changes in local environmental conditions; or [4] long-distance dispersal and establishment, when geographic isolation and limited gene flow lead to divergent populations after an exceptional dispersal event to isolated regions, independent of the emergence of vicariant barriers (Crisp et al., 2011). It is noteworthy that considering the speciation models [1][2][3], some temporal congruence between the speciation process and the emergence of the barrier is expected (Rincón-Barrado et al., 2021). ...
Article
Aim: We studied the niche evolution and diversification modes in transisthmian Alpheus shrimps by examining the interplay between environmental niche divergence and conservatism in allopatric sister species. In a broader perspective, the current study analysed the evolution of climatic niche and the role of the environment in species diversification of Alpheus transisthmian shrimp. Location: Atlantic and Eastern-Pacific oceans. Taxon: Alpheus shrimps (Caridea: Alpheidae). Methods: We assembled georeferenced occurrences for 33 species of Alpheus (with 24 sister species) from a time-calibrated molecular phylogeny. We modelled their ecological niches and assessed niche overlap through pairwise comparisons. Additionally, we performed phylogenetic reconstructions of the ancestral environmental niche, for each niche axis. Results: Our results demonstrate that thermal tolerances, food availability and hydrodynamic forces were relevant environmental axes in evolutionary processes in transisthmian species of Alpheus. Among the 528 paired comparisons, we found that most niches were divergent, including in 12 clades formed by pairs of sister species (in only two of these clades were the niches fully equivalent). Phylogenetic reconstructions of ancestral niches showed an initial niche conservatism in all axes, with divergences intensifying in the last 12 million years. Main Conclusions: We found evidence that confirms the relevance of the environmental changes that occurred in the West Atlantic and East Pacific for niche evolution in transisthmian Alpheus species, as well as for the emergence of some lineages. Our findings provide evidence for different modes of Alpheus species speciation in a period consistent with the closure of the Isthmus of Panama.
... Conversely, many biologists still rely on those philosophers to tell them on what grounds a good or bad scientific endeavour is. For instance, this occurs for some scientists working with Biogeography (e.g., Crisp et al., 2011), phylogenetic inference (e.g., Lienau & DeSalle, 2010), Ecology (Murray, 2001) and evolutionary studies (Richter & Wirkner, 2014); highlighting the necessity to discuss these issues with the philosophically directed biologists. In relation to the topic of explanations in biology, we can find a plurality of explanations that can give us a causal explanation in various fields within biology (Brigandt, 2013;Braillard & Malaterre, 2015). ...
Article
Full-text available
Walter Bock was committed to developing a framework for evolutionary biology. Bock repeatedly discussed how evolutionary explanations should be considered within the realm of Hempel's deductive-nomological model of scientific explanations. Explanation in evolution would then consist of functional and evolutionary explanations, and within the latter, an explanation can be of nomological-deductive and historical narrative explanations. Thus, a complete evolutionary explanation should include, first, a deductive functional analysis, and then proceed through no-mological and historical evolutionary explanations. However, I will argue that his views on the deductive proprieties of functional analysis and the deductive-nomo-logical parts of evolution fail because of the nature of evolution, which contains a historical element that the logic of deduction and Hempel's converting law model do not compass. Conversely, Bock's historical approach gives a critical consideration of the historical narrative element of evolutionary explanation, which is fundamental to the methodology of the historical nature of evolutionary theory. Herein, I will expand and discuss a modern view of evolutionary explanations of traits that includes the currentacknowledgement of the differences between experimental and the historical sciences, including the token and type event dichotomy, that mutually illuminate each other in order to give us a well confirmed and coherent hypothesis for evolutionary explanations. Within this framework, I will argue that the duality of evolutionary explanations is related to two components of character evolution: origin, with its evolutionary pathways along with the history, and maintenance, the function (mainly a current function) for the character being selected.
Article
Background Squamata (lizards, snakes, and amphisbaenians) is a Triassic lineage with an extensive and complex biogeographic history, yet no large-scale study has reconstructed the ancestral range of early squamate lineages. The fossil record indicates a broadly Pangaean distribution by the end- Cretaceous, though many lineages (e.g., Paramacellodidae, Mosasauria, Polyglyphanodontia) subsequently went extinct. Thus, the origin and occupancy of extant radiations is unclear and may have been localized within Pangaea to specific plates, with potential regionalization to distinct Laurasian and Gondwanan landmasses during the Mesozoic in some groups. Methods We used recent tectonic models to code extant and fossil squamate distributions occurring on nine discrete plates for 9,755 species, with Jurassic and Cretaceous fossil constraints from three extinct lineages. We modeled ancestral ranges for crown Squamata from an extant-only molecular phylogeny using a suite of biogeographic models accommodating different evolutionary processes and fossil-based node constraints from known Jurassic and Cretaceous localities. We hypothesized that the best-fit models would not support a full Pangaean distribution (i.e., including all areas) for the origin of crown Squamata, but would instead show regionalization to specific areas within the fragmenting supercontinent, likely in the Northern Hemisphere where most early squamate fossils have been found. Results Incorporating fossil data reconstructs a localized origin within Pangaea, with early regionalization of extant lineages to Eurasia and Laurasia, while Gondwanan regionalization did not occur until the middle Cretaceous for Alethinophidia, Scolecophidia, and some crown Gekkotan lineages. While the Mesozoic history of extant squamate biogeography can be summarized as a Eurasian origin with dispersal out of Laurasia into Gondwana, their Cenozoic history is complex with multiple events (including secondary and tertiary recolonizations) in several directions. As noted by previous authors, squamates have likely utilized over-land range expansion, land-bridge colonization, and trans-oceanic dispersal. Tropical Gondwana and Eurasia hold more ancient lineages than the Holarctic (Rhineuridae being a major exception), and some asymmetries in colonization (e.g., to North America from Eurasia during the Cenozoic through Beringia) deserve additional study. Future studies that incorporate fossil branches, rather than as node constraints, into the reconstruction can be used to explore this history further.
Article
Full-text available
Aim While the evolution of seasonal migration and its association with biogeography have been the subject of numerous studies, its influence on species diversification has rarely been examined. The aim of this study is to explain the atypical latitudinal diversity gradient in shorebirds: did seasonal migration influence diversification and did the gradient evolve from higher in situ diversification or repeated transitions? Location Global. Time Period Palaeogene to present. Major Taxa Studied Order Charadriiformes (shorebirds and allies). Shorebirds include the species with the longest annual migrations in birds and exhibit an inverse latitudinal diversity gradient, with more species breeding toward higher latitudes. Methods We first assessed the temporal and geographical framework of the evolution of migration using a newly reconstructed time‐calibrated phylogeny, and using the fossil record as a complementary source of information. We then used hidden state speciation and extinction models to test whether diversification and thus the latitudinal diversity gradient has been shaped by migratory behaviour or species latitudinal distribution. Results We found that the ancestor of Charadriiformes was likely a tropical year‐round resident and that the current clade's distribution of diversity was not shaped by dispersal events out of the tropics but rather by higher in situ diversification in temperate biomes compared to tropical biomes. Seasonal migration did not affect diversification rate in our models. However, the evolution of seasonal migration seems to be a prerequisite to breed in temperate biomes, indirectly enabling the higher diversification observed in temperate biomes. Main Conclusions Our results bring original insights into how the migratory behaviour acted as a precursor in the diversification and biogeographical history of a large bird clade, ultimately shaping an atypical latitudinal diversity gradient.
Article
Full-text available
New Caledonia has generally been considered a continental island, the biota of which largely dates back to Gondwanan times owing to its geological origin and the presence of phylogenetic relicts. This view is contradicted by geological evidence indicating long Palaeocene and Eocene submersions and by recent biogeographic and phylogenetic studies, with molecular or geophysical dating placing the biota no older than the Oligocene. Phylogenetic relicts do not provide conclusive information in this respect, as their presence cannot be explained by simple hypotheses but requires assumption of many ad hoc extinction events. The implication of this new scenario is that all the New Caledonian biota colonized the island since 37 Ma Local richness can be explained by local radiation and adaptation after colonization but also by many dispersal events, often repeated within the same groups of organisms. Local microendemism is another remarkable feature of the biota. It seems to be related to recent speciation mediated by climate, orography, soil type and perhaps unbalanced biotic interactions created by colonization disharmonies. New Caledonia must be considered as a very old Darwinian island, a concept that offers many more fascinating opportunities of study.
Article
Full-text available
The three southern conifer families, Araucariaceae, Cupressaceae and Podocarpaceae, have a long history and continue to be an important part of the vegetation today. The Araucariaceae have the most extensive fossil record, occurring in both hemispheres, and with Araucaria in particular having an ancient origin. In the Southern Hemisphere Araucaria and Agathis have substantial macrofossil records, especially in Australasia, and Wollemia probably also has an important macrofossil record. At least one extinct genus of Araucariaceae is present as a macrofossil during the Cenozoic. Cupressaceae macrofossils are difficult to identify in older sediments, but the southern genera begin their record in the Cretaceous (Athrotaxis) and become more diverse and extensive during the Cenozoic. Several extinct genera of Cupressaceae also occur in Cretaceous and Cenozoic sediments in Australasia. The Podocarpaceae probably begin their macrofossil record in the Triassic, although the early history is still uncertain. Occasional Podocarpaceae macrofossils have been recorded in the Northern Hemisphere, but they are essentially a southern family. The Cenozoic macrofossil record of the Podocarpaceae is extensive, especially in south-eastern Australia, where the majority of the extant genera have been recorded. Some extinct genera have also been reported from across high southern latitudes, confirming an extremely diverse and widespread suite of Podocarpaceae during the Cenozoic in the region. In the Southern Hemisphere today conifers achieve greatest abundance in wet forests. Those which compete successfully with broad-leaved angiosperms in warmer forests produce broad, flat photosynthetic shoots. In the Araucariaceae this is achieved by the planation of multiveined leaves into large compound shoots. In the other two families leaves are now limited to a single vein (except Nageia), and to overcome this limitation many genera have resorted to re-orientation of leaves and two-dimensional flattening of shoots. The Podocarpaceae show greatest development of this strategy with 11 of 19 genera producing shoots analogous to compound leaves. The concentration of conifers in wet forest left them vulnerable to the climate change which occurred in the Cenozoic, and decreases in diversity have occurred since the Paleogene in all regions where fossil records are available. Information about the history of the dry forest conifers is extremely limited because of a lack of fossilisation in such environments. The southern conifers, past and present, demonstrate an ability to compete effectively with angiosperms in many habitats and should not be viewed as remnants which are ineffectual against angiosperm competitors.
Article
Full-text available
New Zealand is generally thought to have been physically isolated from the rest of the world for over 60 million years. But physical isolation may not mean biotic isolation, at least on the time scale of millions of years. Are New Zealand’s present complement of plants the direct descendants of what originally rafted from Gondwana? Or has there been total extinction of this initial flora with replacement through long-distance dispersal (a complete biotic turnover)? These are two possible extremes which have come under recent discussion. Can the fossil record be used to decide the relative importance of the two endpoints, or is it simply too incomplete and too dependent on factors of chance? This paper suggests two approaches to the problem—the use of statistics to apply levels of confidence to first appearances in the fossil record and the analysis of trends based on the entire palynorecord. Statistics can suggest that the first appearance of a taxon was after New Zealand broke away from Gondwana—as long as the first appearance in the record was not due to an increase in biomass from an initially rare state. Two observations can be drawn from the overall palynorecord that are independent of changes in biomass: (1) The first appearance of palynotaxa common to both Australia and New Zealand is decidedly non-random. Most taxa occur first in Australia. This suggests a bias in air or water transport from west to east. (2) The percentage of endemic palynospecies in New Zealand shows no simple correlation with the time New Zealand drifted into isolation. The conifer macrorecord also hints at complete turnover since the Cretaceous.
Article
To understand biogeographical patterns one must first ask questions about the nature of patterns themselves. This in turn requires a consideration of the aims, approaches and methods of biogeography. The need for such a discussion reflects the fact that biogeography lacks cohesion and consistency (Chapter 1). There is too little sound knowledge of biogeographical processes, reflected by disparity about what can be assumed or should be investigated, a lack of integrated models of the biogeographical system and little recognition of process levels within such a system. These problems have been present in biogeography for a long time but more recently the subject has been shaken by outright controversy, and even unpleasant disagreement and bitterness about approaches and methods (e.g. some of the contributions in the volume edited by Nelson and Rosen, 1981; and see Ferris, 1980 for an interesting summary of this meeting). This controversy may well testify to the healthiness of the subject (Brown and Gibson, 1983), but more promisingly, it may presage major changes: “The proliferation of competing articulations, the willingness to try anything, the expression of explicit discontent, the recourse to philosophy and to debate over fundamentals, all these are symptoms of a transition from normal to extraordinary research” (Kuhn, 1970, p. 91).
Article
All biogeographical studies seem to pass through three main phases: an empirical or descriptive phase, a narrative phase, and an analytical phase, the last two being the true subject matter of explanatory historical biogeography. Narrative explanations are more or less inductively based and differ from the usual covering law models of explanation in that it is the particular circumstances in the form of an historical narrative that bear the explanatory load. Much biogeographical thinking is of this nature and the hypotheses so formulated suffer from their lack of predictive power and their basic untestability—usually they cannot be falsified and a preference between competing narrative hypotheses cannot be established on logical grounds. Examples of such hypotheses are given. The best biogeographical hypotheses are those that follow the hypothetico-deductive model and an example of how this may be done is provided. A phylogenetic systematic background is essential. Rigorously formulated hypotheses must have explanatory power, internal consistency, predictive power, and potential for falsification. Our aim should be to make our biogeographical hypotheses fit these requirements for only then will we begin to understand the history underlying modern distribution patterns. [Biogeography; Inductive and Deductive Methods; Dispersal.]
Article
A review is presented of current knowledge of the tectonic rifting of eastern Gondwana in the context of possible land continuity between Australia and New Caledonia and the dating of events. Comment is made on the level of endemicity of the modern-day biota of NewCaledonia, with examples of phylogenetic analyses, biogeographic patterns and estimated divergence dates among particular taxa, including birds (kagu-sunbittern clade), Nothofagus and Myrtaceae. The geological history of the region is complex but there is evidence of land persisting into the Cenozoic (Paleocene/Eocene) that possibly allowed old biota to persist. Vicariance explanations of divergences cannot be justifiably dismissed in favour of long-distance, over-water dispersal for such taxa.
Article
Rosen, D. E. (Department of Ichthyology, American Museum of Natural History, New York, New York 10024) 1978. Vicariant patterns and historical explanation in biogeography. Syst. Zool. 27:159-188.-Geographic coincidence of animal and plant distributions to form recognizable patterns suggests that the separate components of the patterns are historically connected with each other and with geographic history. To seek evidence of these historical connections, cladograms of geographic areas, representing sequences of disruptive geologic, climatic, or geographic events, may be compared with biological cladograms, representing sequences of allopatric speciation events in relation to those geographic areas. Such comparisons, when they meet the minimum requirements of being among dichotomized three- taxon cladograms, can resolve similar or dissimilar historical factors; two-taxon statements do not distinguish between groups with different histories. Congruence of biological and geological area-cladograms at a high confidence level (such as congruence of a five-taxon clado- gram or four three-taxon cladograms with a geological cladogram, where the confidence level can be shown in cladistic theory to be 99%) means that specified events of paleogeography can be adopted as an explanation of the biological patterns. In such a cause and effect relationship, where the earth and its life are assumed to have evolved together, paleogeography is taken by logical necessity to be the independent variable and biological history, the dependent variable. Drawing a mathematical simile, the biological cladogram y (dependent variable), is a function of the geological cladogram x (independent variable), as in a simple regression of effect y on cause x where we are given no free choice as to which is the independent variable. Such a view implies that any specified sequence in earth history must coincide with some discoverable biological patterns; it does not imply a necessary converse that each biological pattern must coincide with some discoverable paleogeographic pattern, because some biological distributions might have resulted from stochastic processes (chance dispersal). Determining that all discoverable biological patterns conflict with a given corroborated or observed sequence of geologic, climatic, or geographic change (i.e., that y is not a function of x), in theory, therefore should falsify vicariance biogeography. Because dispersal biogeography presupposes stochastic processes, and any failure to meet the expectation of a postulated dispersal is explained by an additional dispersal, dispersal biogeography is immune to falsification. Without resort to paleontology or earth history, whether a given historical relationship implied by congruence of biological area-cladograms is the result of dispersal or vicariance can also be thought of in terms which minimize the number of necessary assumptions: Did the sedentary organisms disperse with the vagile ones or did the vagile organisms vicariate with the sedentary ones? Cladistic congruence of a group of sedentary organisms with a group of vagile ones rejects dispersal for both. Hence, distributions of sedentary organisms have the potential to falsify dispersal theories as applied to vagile organisms, but distributions of vagile organisms cannot falsify vicariance theories as applied to sedentary ones. The problems that arise in various kinds of historical explanation are exemplified by several specific distributions of fishes and other organisms in North and Middle America and in the larger context of Pangaean history, and are discussed in relation to current species concepts. [Vicariance; species concepts; biocladistics; biohistory; geocladistics; geohistory; Neotropics; Gondwanaland.].