Latitudinal gradients in species diversity

Last updated
Map latitudinal gradient of living terrestrial vertebrate species richness (Mannion 2014) Map latitudinal gradient of biodiversity mannion 2014.png
Map latitudinal gradient of living terrestrial vertebrate species richness (Mannion 2014)

Species richness, or biodiversity, increases from the poles to the tropics for a wide variety of terrestrial and marine organisms, often referred to as the latitudinal diversity gradient. [1] The latitudinal diversity gradient is one of the most widely recognized patterns in ecology. [1] It has been observed to varying degrees in Earth's past. [2] A parallel trend has been found with elevation (elevational diversity gradient), [3] though this is less well-studied. [4]

Contents

Explaining the latitudinal diversity gradient has been called one of the great contemporary challenges of biogeography and macroecology (Willig et al. 2003, Pimm and Brown 2004, Cardillo et al. 2005). [5] The question "What determines patterns of species diversity?" was among the 25 key research themes for the future identified in 125th Anniversary issue of Science (July 2005). There is a lack of consensus among ecologists about the mechanisms underlying the pattern, and many hypotheses have been proposed and debated. A recent review [6] noted that among the many conundrums associated with the latitudinal diversity gradient (or latitudinal biodiversity gradient) the causal relationship between rates of molecular evolution and speciation has yet to be demonstrated.

Understanding the global distribution of biodiversity is one of the most significant objectives for ecologists and biogeographers. Beyond purely scientific goals and satisfying curiosity, this understanding is essential for applied issues of major concern to humankind, such as the spread of invasive species, the control of diseases and their vectors, and the likely effects of global climate change on the maintenance of biodiversity (Gaston 2000). Tropical areas play prominent roles in the understanding of the distribution of biodiversity, as their rates of habitat degradation and biodiversity loss are exceptionally high. [7]

Patterns in the past

The latitudinal diversity gradient is a noticeable pattern among modern organisms that has been described qualitatively and quantitatively. It has been studied at various taxonomic levels, through different time periods and across many geographic regions (Crame 2001). The latitudinal diversity gradient has been observed to varying degrees in Earth's past, possibly due to differences in climate during various phases of Earth's history. Some studies indicate that the gradient was strong, particularly among marine taxa, while other studies of terrestrial taxa indicate it had little effect on the distribution of animals. [2]

Hypotheses for pattern

Although many of the hypotheses exploring the latitudinal diversity gradient are closely related and interdependent, most of the major hypotheses can be split into three general hypotheses.

Spatial/Area hypotheses

There are five major hypotheses that depend solely on the spatial and areal characteristics of the tropics.

Mid-domain effect

Using computer simulations, Colwell and Hurt (1994) and Willing and Lyons (1998) first pointed out that if species’ latitudinal ranges were randomly shuffled within the geometric constraints of a bounded biogeographical domain (e.g. the continents of the New World, for terrestrial species), species' ranges would tend to overlap more toward the center of the domain than towards its limits, forcing a mid-domain peak in species richness. Colwell and Lees (2000) called this stochastic phenomenon the mid-domain effect (MDE), presented several alternative analytical formulations for one-dimensional MDE (expanded by Connolly 2005), and suggested the hypothesis that MDE might contribute to the latitudinal gradient in species richness, together with other explanatory factors considered here, including climatic and historical ones. Because "pure" mid-domain models attempt to exclude any direct environmental or evolutionary influences on species richness, they have been claimed to be null models (Colwell et al. 2004, 2005). On this view, if latitudinal gradients of species richness were determined solely by MDE, observed richness patterns at the biogeographic level would not be distinguishable from patterns produced by random placement of observed ranges called dinosures(Colwell and Lees 2000). Others object that MDE models so far fail to exclude the role of the environment at the population level and in setting domain boundaries, and therefore cannot be considered null models (Hawkins and Diniz-Filho 2002; Hawkins et al. 2005; Zapata et al. 2003, 2005). Mid-domain effects have proven controversial (e.g. Jetz and Rahbek 2001, Koleff and Gaston 2001, Lees and Colwell, 2007, Romdal et al. 2005, Rahbek et al. 2007, Storch et al. 2006; Bokma and Monkkonen 2001, Diniz-Filho et al. 2002, Hawkins and Diniz-Filho 2002, Kerr et al. 2006, Currie and Kerr, 2007). While some studies have found evidence of a potential role for MDE in latitudinal gradients of species richness, particularly for wide-ranging species (e.g. Jetz and Rahbek 2001, Koleff and Gaston 2001, Lees and Colwell, 2007, Romdal et al. 2005, Rahbek et al. 2007, Storch et al. 2006; Dunn et al. 2007) [5] [8] others report little correspondence between predicted and observed latitudinal diversity patterns (Bokma and Monkkonen 2001, Currie and Kerr, 2007, Diniz-Filho et al. 2002, Hawkins and Diniz-Filho 2002, Kerr et al. 2006).

Geographical area hypothesis

Another spatial hypothesis is the geographical area hypothesis (Terborgh 1973). It asserts that the tropics are the largest biome and that large tropical areas can support more species. More area in the tropics allows species to have larger ranges and consequently larger population sizes. Thus, species with larger ranges are likely to have lower extinction rates (Rosenzweig 2003). Additionally, species with larger ranges may be more likely to undergo allopatric speciation, which would increase rates of speciation (Rosenzweig 2003). The combination of lower extinction rates and high rates of speciation leads to the high levels of species richness in the tropics.

A critique of the geographical area hypothesis is that even if the tropics is the most extensive of the biomes, successive biomes north of the tropics all have about the same area. Thus, if the geographical area hypothesis is correct these regions should all have approximately the same species richness, which is not true, as is referenced by the fact that polar regions contain fewer species than temperate regions (Gaston and Blackburn 2000). To explain this, Rosenzweig (1992) suggested that if species with partly tropical distributions were excluded, the richness gradient north of the tropics should disappear. Blackburn and Gaston 1997 tested the effect of removing tropical species on latitudinal patterns in avian species richness in the New World and found there is indeed a relationship between the land area and the species richness of a biome once predominantly tropical species are excluded. Perhaps a more serious flaw in this hypothesis is some biogeographers suggest that the terrestrial tropics are not, in fact, the largest biome, and thus this hypothesis is not a valid explanation for the latitudinal species diversity gradient (Rohde 1997, Hawkins and Porter 2001). In any event, it would be difficult to defend the tropics as a "biome" rather than the geographically diverse and disjunct regions that they truly include.

The effect of area on biodiversity patterns has been shown to be scale-dependent, having the strongest effect among species with small geographical ranges compared to those species with large ranges who are affected more so by other factors such as the mid-domain and/or temperature. [5]

Species-energy hypothesis

The species energy hypothesis suggests the amount of available energy sets limits to the richness of the system. Thus, increased solar energy (with an abundance of water) at low latitudes causes increased net primary productivity (or photosynthesis). This hypothesis proposes the higher the net primary productivity the more individuals can be supported, and the more species there will be in an area. Put another way, this hypothesis suggests that extinction rates are reduced towards the equator as a result of the higher populations sustainable by the greater amount of available energy in the tropics. Lower extinction rates lead to more species in the tropics.

One critique of this hypothesis has been that increased species richness over broad spatial scales is not necessarily linked to an increased number of individuals, which in turn is not necessarily related to increased productivity. [9] Additionally, the observed changes in the number of individuals in an area with latitude or productivity are either too small (or in the wrong direction) to account for the observed changes in species richness. [9] The potential mechanisms underlying the species-energy hypothesis, their unique predictions and empirical support have been assessed in a major review by Currie et al. (2004). [10]

The effect of energy has been supported by several studies in terrestrial and marine taxa. [7]

Climate harshness hypothesis

Another climate-related hypothesis is the climate harshness hypothesis, which states the latitudinal diversity gradient may exist simply because fewer species can physiologically tolerate conditions at higher latitudes than at low latitudes because higher latitudes are often colder and drier than tropical latitudes. Currie et al. (2004) [10] found fault with this hypothesis by stating that, although it is clear that climatic tolerance can limit species distributions, it appears that species are often absent from areas whose climate they can tolerate.

Climate stability hypothesis

Similarly to the climate harshness hypothesis, climate stability is suggested to be the reason for the latitudinal diversity gradient. The mechanism for this hypothesis is that while a fluctuating environment may increase the extinction rate or preclude specialization, a constant environment can allow species to specialize on predictable resources, allowing them to have narrower niches and facilitating speciation. The fact that temperate regions are more variable both seasonally and over geological timescales (discussed in more detail below) suggests that temperate regions are thus expected to have less species diversity than the tropics.

Critiques for this hypothesis include the fact that there are many exceptions to the assumption that climate stability means higher species diversity. For example, low species diversity is known to occur often in stable environments such as tropical mountaintops. Additionally, many habitats with high species diversity do experience seasonal climates, including many tropical regions that have highly seasonal rainfall (Brown and Lomolino 1998).

Historical/Evolutionary hypotheses

There are four main hypotheses that are related to historical and evolutionary explanations for the increase of species diversity towards the equator.

The historical perturbation hypothesis

The historical perturbation hypothesis proposes the low species richness of higher latitudes is a consequence of an insufficient time period available for species to colonize or recolonize areas because of historical perturbations such as glaciation (Brown and Lomolino 1998, Gaston and Blackburn 2000). This hypothesis suggests that diversity in the temperate regions has not yet reached equilibrium and that the number of species in temperate areas will continue to increase until saturated (Clarke and Crame 2003). However, in the marine environment, where there is also a latitudinal diversity gradient, there is no evidence of a latitudinal gradient in perturbation.

The evolutionary speed hypothesis

The evolutionary speed hypothesis [11] argues higher evolutionary rates due to shorter generation times in the tropics have caused higher speciation rates and thus increased diversity at low latitudes. [12] Higher evolutionary rates in the tropics have been attributed to higher ambient temperatures, higher mutation rates, shorter generation time and/or faster physiological processes, [13] [12] and increased selection pressure from other species that are themselves evolving. [14] Faster rates of microevolution in warm climates (i.e. low latitudes and altitudes) have been shown for plants, [15] mammals, [16] birds, [17] fish [18] and amphibians. [19] Bumblebee species inhabiting lower, warmer elevations have faster rates of both nuclear and mitochondrial genome-wide evolution. [20] Based on the expectation that faster rates of microevolution result in faster rates of speciation, these results suggest that faster evolutionary rates in warm climates almost certainly have a strong influence on the latitudinal diversity gradient. However, recent evidence from marine fish [21] and flowering plants [22] have shown that rates of speciation actually decrease from the poles towards the equator at a global scale. Understanding whether extinction rate varies with latitude will also be important to whether or not this hypothesis is supported. [23]

The hypothesis of effective evolutionary time

The hypothesis of effective evolutionary time assumes that diversity is determined by the evolutionary time under which ecosystems have existed under relatively unchanged conditions, and by evolutionary speed directly determined by effects of environmental energy (temperature) on mutation rates, generation times, and speed of selection. [12] It differs from most other hypotheses in not postulating an upper limit to species richness set by various abiotic and biotic factors, i.e., it is a nonequilibrium hypothesis assuming a largely non-saturated niche space. It does accept that many other factors may play a role in causing latitudinal gradients in species richness as well. The hypothesis is supported by much recent evidence, in particular, the studies of Allen et al. [13] and Wright et al. [24]

The integrated evolutionary speed hypothesis

The integrated evolutionary speed hypothesis argues that species diversity increases due to faster rates of genetic evolution and speciation at lower latitudes where ecosystem productivity is generally greater. [25] It differs from the effective evolutionary time hypothesis by recognizing that species richness generally increases with increasing ecosystem productivity [26] [27] [28] and declines where high environmental energy (temperature) causes water deficits. [29] It also proposes that evolutionary rate increases with population size, abiotic environmental heterogeneity, environmental change and via positive feedback with biotic heterogeneity. There is considerable support for faster rates of genetic evolution in warmer environments, [25] some support for a slower rate among plant species where water availability is limited [30] and for a slower rate among bird species with small population sizes. [31] Many aspects of the hypothesis, however, remain untested.

Biotic hypotheses

Biotic hypotheses claim ecological species interactions such as competition, predation, mutualism, and parasitism are stronger in the tropics and these interactions promote species coexistence and specialization of species, leading to greater speciation in the tropics. These hypotheses are problematic because they cannot be the ultimate cause of the latitudinal diversity gradient as they fail to explain why species interactions might be stronger in the tropics. An example of one such hypothesis is the greater intensity of predation and more specialized predators in the tropics has contributed to the increase of diversity in the tropics (Pianka 1966). This intense predation could reduce the importance of competition (see competitive exclusion) and permit greater niche overlap and promote higher richness of prey. Some recent large-scale experiments suggest predation may indeed be more intense in the tropics, [32] [33] although this cannot be the ultimate cause of high tropical diversity because it fails to explain what gives rise to the richness of the predators in the tropics. Interestingly, the largest test of whether biotic interactions are strongest in the tropics, which focused on predation exerted by large fish predators in the world's open oceans, found predation to peak at mid-latitudes. Moreover, this test further revealed a negative association of predation intensity and species richness, thus contrasting the idea that strong predation near the equator drives or maintains high diversity. [34] Other studies have failed to observe consistent changes in ecological interactions with latitude altogether (Lambers et al. 2002), [1] suggesting that the intensity of species interactions is not correlated with the change in species richness with latitude. Overall, these results highlight the need for more studies on the importance of species interactions in driving global patterns of diversity.

Synthesis and conclusions

There are many other hypotheses related to the latitudinal diversity gradient, but the above hypotheses are a good overview of the major ones still cited today. It is important to note that many of these hypotheses are similar to and dependent on one another. For example, the evolutionary hypotheses are closely dependent on the historical climate characteristics of the tropics.

The generality of the latitudinal diversity gradient

An extensive meta-analysis of nearly 600 latitudinal gradients from published literature tested the generality of the latitudinal diversity gradient across different organismal, habitat and regional characteristics. [1] The results showed that the latitudinal gradient occurs in marine, terrestrial, and freshwater ecosystems, in both hemispheres. The gradient is steeper and more pronounced in richer taxa (i.e. taxa with more species), larger organisms, in marine and terrestrial versus freshwater ecosystems, and at regional versus local scales. The gradient steepness (the amount of change in species richness with latitude) is not influenced by dispersal, animal physiology (homeothermic or ectothermic) trophic level, hemisphere, or the latitudinal range of study. The study could not directly falsify or support any of the above hypotheses, however, results do suggest a combination of energy/climate and area processes likely contribute to the latitudinal species gradient. Notable exceptions to the trend include the ichneumonidae, shorebirds, penguins, and freshwater zooplankton. Also, in terrestrial ecosystems the soil bacterial diversity peaks in temperate climatic zones, [35] [36] and has been linked to carbon inputs and the microscale distribution of aqueous habitats. [37]

Data robustness

One of the main assumptions about latitudinal diversity gradients and patterns in species richness is that the underlying data (i.e., the lists of species at specific locations) are complete. However, this assumption is not met in most cases. For instance, diversity patterns for blood parasites of birds suggest higher diversity in tropical regions, however, the data may be skewed by undersampling in rich faunal areas such as Southeast Asia and South America. [38] For marine fishes, which are among the most studied taxonomic groups, current lists of species are considerably incomplete for most of the world's oceans. At a 3° (about 350 km2) spatial resolution, less than 1.8% of the world's oceans have above 80% of their fish fauna currently described. [39]

Conclusion

The fundamental macroecological question that the latitudinal diversity gradient depends on is "What causes patterns in species richness?". Species richness ultimately depends on whatever proximate factors are found to affect processes of speciation, extinction, immigration, and emigration. While some ecologists continue to search for the ultimate primary mechanism that causes the latitudinal richness gradient, many ecologists suggest instead this ecological pattern is likely to be generated by several contributory mechanisms (Gaston and Blackburn 2000, Willig et al. 2003, Rahbek et al. 2007). For now, the debate over the cause of the latitudinal diversity gradient will continue until a groundbreaking study provides conclusive evidence, or there is general consensus that multiple factors contribute to the pattern.

See also

Related Research Articles

Phylogeography is the study of the historical processes that may be responsible for the past to present geographic distributions of genealogical lineages. This is accomplished by considering the geographic distribution of individuals in light of genetics, particularly population genetics.

Allopatric speciation – also referred to as geographic speciation, vicariant speciation, or its earlier name the dumbbell model – is a mode of speciation that occurs when biological populations become geographically isolated from each other to an extent that prevents or interferes with gene flow.

The metabolic theory of ecology (MTE) is the ecological component of the more general Metabolic Scaling Theory and Kleiber's law. It posits that the metabolic rate of organisms is the fundamental biological rate that governs most observed patterns in ecology. MTE is part of a larger set of theory known as metabolic scaling theory that attempts to provide a unified theory for the importance of metabolism in driving pattern and process in biology from the level of cells all the way to the biosphere.

Macroecology is a subfield in ecology that uses a methodological approach that investigates the empirical patterns and mechanistic processes by which the particulate components of complex ecological systems generate emergent structures and dynamics Unlike traditional ecology, which focuses on local and small-scale interactions, macroecology seeks to identify general emergent patterns within and across spatial and temporal scales.

An evolutionary radiation is an increase in taxonomic diversity that is caused by elevated rates of speciation, that may or may not be associated with an increase in morphological disparity. A significantly large and diverse radiation within a relatively short geologic time scale is often referred to as an explosion. Radiations may affect one clade or many, and be rapid or gradual; where they are rapid, and driven by a single lineage's adaptation to their environment, they are termed adaptive radiations.

<span class="mw-page-title-main">Species richness</span> Variety of species in an ecological community, landscape or region

Species richness is the number of different species represented in an ecological community, landscape or region. Species richness is simply a count of species, and it does not take into account the abundances of the species or their relative abundance distributions. Species richness is sometimes considered synonymous with species diversity, but the formal metric species diversity takes into account both species richness and species evenness.

Rapoport's rule is an ecogeographical rule that states that latitudinal ranges of plants and animals are generally smaller at lower latitudes than at higher latitudes.

<span class="mw-page-title-main">Intermediate disturbance hypothesis</span> Model proposing regional biodiversity is increased by a moderate level of ecological disturbance

The intermediate disturbance hypothesis (IDH) suggests that local species diversity is maximized when ecological disturbance is neither too rare nor too frequent. At low levels of disturbance, more competitive organisms will push subordinate species to extinction and dominate the ecosystem. At high levels of disturbance, due to frequent forest fires or human impacts like deforestation, all species are at risk of going extinct. According to IDH theory, at intermediate levels of disturbance, diversity is thus maximized because species that thrive at both early and late successional stages can coexist. IDH is a nonequilibrium model used to describe the relationship between disturbance and species diversity. IDH is based on the following premises: First, ecological disturbances have major effects on species richness within the area of disturbance. Second, interspecific competition results in one species driving a competitor to extinction and becoming dominant in the ecosystem. Third, moderate ecological scale disturbances prevent interspecific competition.

The hypothesis of effective evolutionary time attempts to explain gradients, in particular latitudinal gradients, in species diversity. It was originally named "time hypothesis".

Tropical ecology is the study of the relationships between the biotic and abiotic components of the tropics, or the area of the Earth that lies between the Tropic of Cancer and the Tropic of Capricorn. The tropical climate experiences hot, humid weather and rainfall year-round. While many might associate the region solely with the rainforests, the tropics are home to a wide variety of ecosystems that boast a great wealth of biodiversity, from exotic animal species to seldom-found flora. Tropical ecology began with the work of early English naturalists and eventually saw the establishment of research stations throughout the tropics devoted to exploring and documenting these exotic landscapes. The burgeoning ecological study of the tropics has led to increased conservation education and programs devoted to the climate.

<span class="mw-page-title-main">Klaus Rohde</span> German biologist

Klaus Rohde is a German biologist and parasitologist at the University of New England (UNE), Australia. He is known particularly for his work on marine parasitology, evolutionary ecology, zoogeography, phylogeny, and ultrastructure of lower invertebrates.

The Janzen–Connell hypothesis is a well-known hypothesis for the maintenance of high species biodiversity in the tropics. It was published independently in the early 1970s by Daniel Janzen, who focused on tropical trees, and Joseph Connell who discussed trees and marine invertebrates. According to their hypothesis, host-specific herbivores, pathogens, or other natural enemies make the areas near a parent tree inhospitable for the survival of seeds or seedlings. These natural enemies are referred to as 'distance-responsive predators' if they kill seeds or seedlings near the parent tree, or 'density-dependent predators' if they kill seeds or seedlings where they are most abundant. Such predators can prevent any one species from dominating the landscape, because if that species is too common, there will be few safe places for its seedlings to survive. Both Janzen and Connell originally proposed that for natural enemies to increase local diversity, they must be host-specific and relatively immobile, such that they disproportionately reduce the density of the more locally common tree species. This prevents any one species from becoming dominant and excluding other species through competition, allowing more species to coexist in small areas. This can be classified as a stabilizing mechanism.

<span class="mw-page-title-main">Elevational diversity gradient</span> Ecological pattern in which biodiversity changes with elevation

Elevational diversity gradient (EDG) is an ecological pattern where biodiversity changes with elevation. The EDG states that species richness tends to decrease as elevation increases, up to a certain point, creating a "diversity bulge" at middle elevations. There have been multiple hypotheses proposed for explaining the EDG, none of which accurately describe the phenomenon in full.

<span class="mw-page-title-main">Avian clutch size</span>

Clutch size refers to the number of eggs laid in a single brood by a nesting pair of birds. The numbers laid by a particular species in a given location are usually well defined by evolutionary trade-offs with many factors involved, including resource availability and energetic constraints. Several patterns of variation have been noted and the relationship between latitude and clutch size has been a topic of interest in avian reproduction and evolution. David Lack and R.E. Moreau were among the first to investigate the effect of latitude on the number of eggs per nest. Since Lack's first paper in the mid-1940s there has been extensive research on the pattern of increasing clutch size with increasing latitude. The proximate and ultimate causes for this pattern have been a subject of intense debate involving the development of ideas on group, individual, and gene-centric views of selection.

The term phylogenetic niche conservatism has seen increasing use in recent years in the scientific literature, though the exact definition has been a matter of some contention. Fundamentally, phylogenetic niche conservatism refers to the tendency of species to retain their ancestral traits. When defined as such, phylogenetic niche conservatism is therefore nearly synonymous with phylogenetic signal. The point of contention is whether or not "conservatism" refers simply to the tendency of species to resemble their ancestors, or implies that "closely related species are more similar than expected based on phylogenetic relationships". If the latter interpretation is employed, then phylogenetic niche conservatism can be seen as an extreme case of phylogenetic signal, and implies that the processes which prevent divergence are in operation in the lineage under consideration. Despite efforts by Jonathan Losos to end this habit, however, the former interpretation appears to frequently motivate scientific research. In this case, phylogenetic niche conservatism might best be considered a form of phylogenetic signal reserved for traits with broad-scale ecological ramifications. Thus, phylogenetic niche conservatism is usually invoked with regards to closely related species occurring in similar environments.

Microbial biogeography is a subset of biogeography, a field that concerns the distribution of organisms across space and time. Although biogeography traditionally focused on plants and larger animals, recent studies have broadened this field to include distribution patterns of microorganisms. This extension of biogeography to smaller scales—known as "microbial biogeography"—is enabled by ongoing advances in genetic technologies.

<span class="mw-page-title-main">Altitudinal migration</span>

Altitudinal migration is a short-distance animal migration from lower altitudes to higher altitudes and back. Altitudinal migrants change their elevation with the seasons making this form of animal migration seasonal. Altitudinal migration can be most commonly observed in species inhabiting temperate or tropical ecosystems. This behavior is commonly seen among avian species but can also be observed within other vertebrates and some invertebrates. It is commonly thought to happen in response to climate and food availability changes as well as increasingly due to anthropogenic influence. These migrations can occur both during reproductive and non-reproductive seasons.

<span class="mw-page-title-main">Reinforcement (speciation)</span> Process of increasing reproductive isolation

Reinforcement is a process of speciation where natural selection increases the reproductive isolation between two populations of species. This occurs as a result of selection acting against the production of hybrid individuals of low fitness. The idea was originally developed by Alfred Russel Wallace and is sometimes referred to as the Wallace effect. The modern concept of reinforcement originates from Theodosius Dobzhansky. He envisioned a species separated allopatrically, where during secondary contact the two populations mate, producing hybrids with lower fitness. Natural selection results from the hybrid's inability to produce viable offspring; thus members of one species who do not mate with members of the other have greater reproductive success. This favors the evolution of greater prezygotic isolation. Reinforcement is one of the few cases in which selection can favor an increase in prezygotic isolation, influencing the process of speciation directly. This aspect has been particularly appealing among evolutionary biologists.

Allochronic speciation is a form of speciation arising from reproductive isolation that occurs due to a change in breeding time that reduces or eliminates gene flow between two populations of a species. The term allochrony is used to describe the general ecological phenomenon of the differences in phenology that arise between two or more species—speciation caused by allochrony is effectively allochronic speciation.

Douglas Schemske is an evolutionary ecologist who made major contributions to research on pollination, the latitudinal gradient in species diversity, the evolution of polyploidy, and plant mating systems.

References

References
  1. 1 2 3 4 Hillebrand, H. (February 2004). "On the generality of the latitudinal diversity gradient" (PDF). The American Naturalist . 163 (2): 192–211. doi:10.1086/381004. PMID   14970922. S2CID   9886026.
  2. 1 2 Sahney, Sarda; Benton, Michael J (7 April 2008). "Recovery from the most profound mass extinction of all time". Proceedings of the Royal Society B: Biological Sciences. 275 (1636): 759–765. doi:10.1098/rspb.2007.1370. PMC   2596898 . PMID   18198148.
  3. McCain, Christy M. (February 2005). "Elevational Gradients in Diversity of Small Mammals" (PDF). Ecology . 86 (2): 366–372. doi:10.1890/03-3147. S2CID   37759984. Archived from the original (PDF) on 2020-02-08.
  4. Rahbek, Carsten (June 1995). "The elevational gradient of species richness: a uniform pattern?". Ecography . 18 (2): 200–205. doi: 10.1111/j.1600-0587.1995.tb00341.x .
  5. 1 2 3 Mora, Camilo; Robertson, D. Ross (July 2005). "Causes of latitudinal gradients in species richness: a test with fishes of the Tropical Eastern Pacific" (PDF). Ecology. 86 (7): 1771–1782. doi:10.1890/04-0883. S2CID   35273509. Archived from the original (PDF) on 2019-03-07.
  6. Dowle, E. J.; Morgan-Richards, M.; Trewick, S. A. (2013). "Molecular evolution and the latitudinal biodiversity gradient". Heredity. 110 (6): 501–510. doi:10.1038/hdy.2013.4. PMC   3656639 . PMID   23486082.
  7. 1 2 Tittensor, Derek P.; Mora, Camilo; Jetz, Walter; Lotze, Heike K.; Ricard, Daniel; Berghe, Edward Vanden; Worm, Boris (August 2010). "Global patterns and predictors of marine biodiversity across taxa". Nature. 466 (7310): 1098–1101. Bibcode:2010Natur.466.1098T. doi:10.1038/nature09329. PMID   20668450. S2CID   4424240.
  8. Mora, Camilo; Chittaro, Paul M.; Sale, Peter F.; Kritzer, Jacob P.; Ludsin, Stuart A. (February 2003). "Patterns and processes in reef fish diversity". Nature. 421 (6926): 933–936. Bibcode:2003Natur.421..933M. doi:10.1038/nature01393. PMID   12606998. S2CID   4311686.
  9. 1 2 Cardillo, M.; Orme, C. D. L.; Owens, I. P. F. (2005). "Testing for latitudinal bias in diversification rates: An example using New World birds". Ecology. 86 (9): 2278–2287. doi:10.1890/05-0112.
  10. 1 2 Currie, D. J.; Mittelbach, G. G.; Cornell, H. V.; Kaufman, D. M.; Kerr, J. T.; Oberdorff, T. (2004). "Predictions and tests of climate-based hypotheses of broad-scale variation in taxonomic richness". Ecology Letters. 7 (11): 1121–1134. doi:10.1111/j.1461-0248.2004.00671.x.
  11. Rensch, Bernhard (1959-03-02). Evolution Above the Species Level. Columbia University Press. doi:10.7312/rens91062. ISBN   978-0-231-88186-9.
  12. 1 2 3 Rohde, Klaus (1992). "Latitudinal Gradients in Species Diversity: The Search for the Primary Cause". Oikos. 65 (3): 514–527. doi:10.2307/3545569. ISSN   0030-1299. JSTOR   3545569.
  13. 1 2 Allen, Andrew P.; Gillooly, James F. (August 2006). "Assessing latitudinal gradients in speciation rates and biodiversity at the global scale". Ecology Letters. 9 (8): 947–954. doi: 10.1111/j.1461-0248.2006.00946.x . ISSN   1461-023X. PMID   16913938.
  14. Schemske, Douglas W.; Mittelbach, Gary G.; Cornell, Howard V.; Sobel, James M.; Roy, Kaustuv (December 2009). "Is There a Latitudinal Gradient in the Importance of Biotic Interactions?" (PDF). Annual Review of Ecology, Evolution, and Systematics. 40 (1): 245–269. doi:10.1146/annurev.ecolsys.39.110707.173430. S2CID   53470632. Archived from the original (PDF) on 2020-10-30.
  15. Wright, Shane D.; Keeling, Jeannette; Gillman, Len N. (May 2006). "The road from Santa Rosalia: a faster tempo of evolution in tropical climates". Proceedings of the National Academy of Sciences. 103 (20): 7718–7722. doi: 10.1073/pnas.0510383103 . PMC   1472511 . PMID   16672371.
  16. Gillman, Len N.; Keeling, D. Jeanette; Ross, Howard A.; Wright, Shane D. (2009-09-22). "Latitude, elevation and the tempo of molecular evolution in mammals". Proceedings of the Royal Society B: Biological Sciences. 276 (1671): 3353–3359. doi:10.1098/rspb.2009.0674. PMC   2817169 . PMID   19556254.
  17. Gillman, Len N.; McCowan, Luke S. C.; Wright, Shane D. (September 2012). "The tempo of genetic evolution in birds: body mass and climate effects: The tempo of evolution in birds". Journal of Biogeography. 39 (9): 1567–1572. doi:10.1111/j.1365-2699.2012.02730.x. S2CID   83114083.
  18. Wright, Shane D.; Ross, Howard A.; Jeanette Keeling, D.; McBride, Paul; Gillman, Len N. (2011-03-01). "Thermal energy and the rate of genetic evolution in marine fishes". Evolutionary Ecology. 25 (2): 525–530. doi:10.1007/s10682-010-9416-z. ISSN   1573-8477. S2CID   38535658.
  19. Wright, Shane D.; Gillman, Len N.; Ross, Howard A.; Keeling, D. Jeanette (June 2010). "Energy and the tempo of evolution in amphibians: Energy and the tempo of evolution in amphibians". Global Ecology and Biogeography: no. doi:10.1111/j.1466-8238.2010.00549.x.
  20. Lin, Gonghua; Huang, Zuhao; Wang, Lei; Chen, Zhenhua; Zhang, Tongzuo; Gillman, Lennard N; Zhao, Fang (2019-06-01). "Evolutionary Rates of Bumblebee Genomes Are Faster at Lower Elevations". Molecular Biology and Evolution. 36 (6): 1215–1219. doi:10.1093/molbev/msz057. ISSN   0737-4038. PMC   6526908 . PMID   30865278.
  21. Rabosky, Daniel L.; Chang, Jonathan; Title, Pascal O.; Cowman, Peter F.; Sallan, Lauren; Friedman, Matt; Kaschner, Kristin; Garilao, Cristina; Near, Thomas J.; Coll, Marta; Alfaro, Michael E. (July 2018). "An inverse latitudinal gradient in speciation rate for marine fishes". Nature. 559 (7714): 392–395. doi:10.1038/s41586-018-0273-1. ISSN   1476-4687. PMID   29973726. S2CID   49574382.
  22. Igea, Javier; Tanentzap, Andrew J. (April 2020). Davies, Jonathan (ed.). "Angiosperm speciation cools down in the tropics". Ecology Letters. 23 (4): 692–700. doi:10.1111/ele.13476. ISSN   1461-023X. PMC   7078993 . PMID   32043734.
  23. Rolland, Jonathan; Condamine, Fabien L.; Jiguet, Frederic; Morlon, Hélène (2014-01-28). "Faster Speciation and Reduced Extinction in the Tropics Contribute to the Mammalian Latitudinal Diversity Gradient". PLOS Biology. 12 (1): e1001775. doi: 10.1371/journal.pbio.1001775 . ISSN   1545-7885. PMC   3904837 . PMID   24492316.
  24. Wright, Shane; Keeling, Jeannette; Gillman, Len (2006-05-16). "The road from Santa Rosalia: A faster tempo of evolution in tropical climates". Proceedings of the National Academy of Sciences. 103 (20): 7718–7722. doi: 10.1073/pnas.0510383103 . ISSN   0027-8424. PMC   1472511 . PMID   16672371.
  25. 1 2 Gillman, Len N.; Wright, Shane D. (January 2014). Ladle, Richard (ed.). "Species richness and evolutionary speed: the influence of temperature, water and area". Journal of Biogeography. 41 (1): 39–51. doi: 10.1111/jbi.12173 . ISSN   0305-0270. S2CID   84888131.
  26. Gillman, Len N.; Wright, Shane D. (May 2006). "The Influence of Productivity on the Species Richness of Plants: A Critical Assessment". Ecology. 87 (5): 1234–1243. doi:10.1890/0012-9658(2006)87[1234:TIOPOT]2.0.CO;2. ISSN   0012-9658. PMID   16761602.
  27. Cusens, Jarrod; Wright, Shane D.; McBride, Paul D.; Gillman, Len N. (October 2012). "What is the form of the productivity–animal-species-richness relationship? A critical review and meta-analysis". Ecology. 93 (10): 2241–2252. doi:10.1890/11-1861.1. ISSN   0012-9658. PMID   23185885.
  28. Gillman, Len N.; Wright, Shane D.; Cusens, Jarrod; McBride, Paul D.; Malhi, Yadvinder; Whittaker, Robert J. (January 2015). "Latitude, productivity and species richness". Global Ecology and Biogeography. 24 (1): 107–117. doi: 10.1111/geb.12245 . ISSN   1466-822X.
  29. Hawkins, Bradford A.; Field, Richard; Cornell, Howard V.; Currie, David J.; Guégan, Jean-François; Kaufman, Dawn M.; Kerr, Jeremy T.; Mittelbach, Gary G.; Oberdorff, Thierry; O'Brien, Eileen M.; Porter, Eric E. (December 2003). "Energy, Water, and Broad-Scale Geographic Patterns of Species Richness". Ecology. 84 (12): 3105–3117. doi: 10.1890/03-8006 . ISSN   0012-9658. S2CID   15859214.
  30. Goldie, Xavier; Gillman, Len; Crisp, Mike; Wright, Shane (2010-09-07). "Evolutionary speed limited by water in arid Australia". Proceedings of the Royal Society B: Biological Sciences. 277 (1694): 2645–2653. doi:10.1098/rspb.2010.0439. PMC   2982047 . PMID   20410038.
  31. Wright, Shane D.; Gillman, Len N.; Ross, Howard A.; Keeling, D. Jeanette (September 2009). "Slower Tempo of Microevolution in Island Birds: Implications for Conservation Biology". Evolution. 63 (9): 2275–2287. doi: 10.1111/j.1558-5646.2009.00717.x . PMID   19473390. S2CID   23035757.
  32. Roslin, Tomas; Hardwick, Bess; Novotny, Vojtech; Petry, William K.; Andrew, Nigel R.; Asmus, Ashley; Barrio, Isabel C.; Basset, Yves; Boesing, Andrea Larissa; Bonebrake, Timothy C.; Cameron, Erin K.; Dáttilo, Wesley; Donoso, David A.; Drozd, Pavel; Gray, Claudia L.; Hik, David S.; Hill, Sarah J.; Hopkins, Tapani; Huang, Shuyin; Koane, Bonny; Laird-Hopkins, Benita; Laukkanen, Liisa; Lewis, Owen T.; Milne, Sol; Mwesige, Isaiah; Nakamura, Akihiro; Nell, Colleen S.; Nichols, Elizabeth; Prokurat, Alena; Sam, Katerina; Schmidt, Niels M.; Slade, Alison; Slade, Victor; Suchanková, Alžběta; Teder, Tiit; Nouhuys, Saskya van; Vandvik, Vigdis; Weissflog, Anita; Zhukovich, Vital; Slade, Eleanor M. (19 May 2017). "Higher predation risk for insect prey at low latitudes and elevations". Science. 356 (6339): 742–744. Bibcode:2017Sci...356..742R. doi:10.1126/science.aaj1631. PMID   28522532. S2CID   206653702.
  33. Hargreaves, A. L.; Suárez, Esteban; Mehltreter, Klaus; Myers-Smith, Isla; Vanderplank, Sula E.; Slinn, Heather L.; Vargas-Rodriguez, Yalma L.; Haeussler, Sybille; David, Santiago; Muñoz, Jenny; Almazán-Núñez, R. Carlos; Loughnan, Deirdre; Benning, John W.; Moeller, David A.; Brodie, Jedediah F.; Thomas, Haydn J. D.; M, P. A. Morales (1 February 2019). "Seed predation increases from the Arctic to the Equator and from high to low elevations". Science Advances. 5 (2): eaau4403. Bibcode:2019SciA....5.4403H. doi:10.1126/sciadv.aau4403. PMC   6382403 . PMID   30801010.
  34. Roesti, Marius; Anstett, Daniel N.; Freeman, Benjamin G.; Lee-Yaw, Julie A.; Schluter, Dolph; Chavarie, Louise; Rolland, Jonathan; Holzman, Roi (31 March 2020). "Pelagic fish predation is stronger at temperate latitudes than near the equator". Nature Communications. 11 (1): 1527. Bibcode:2020NatCo..11.1527R. doi:10.1038/s41467-020-15335-4. PMC   7109113 . PMID   32235853.
  35. Bahram, Mohammad; Hildebrand, Falk; Forslund, Sofia K.; Anderson, Jennifer L.; Soudzilovskaia, Nadejda A.; Bodegom, Peter M.; Bengtsson-Palme, Johan; Anslan, Sten; Coelho, Luis Pedro; Harend, Helery; Huerta-Cepas, Jaime; Medema, Marnix H.; Maltz, Mia R.; Mundra, Sunil; Olsson, Pål Axel (August 2018). "Structure and function of the global topsoil microbiome". Nature. 560 (7717): 233–237. doi:10.1038/s41586-018-0386-6. hdl: 1887/73861 . ISSN   1476-4687. PMID   30069051. S2CID   256768771.
  36. Bickel, Samuel; Chen, Xi; Papritz, Andreas; Or, Dani (2019-08-20). "A hierarchy of environmental covariates control the global biogeography of soil bacterial richness". Scientific Reports. 9 (1): 12129. doi:10.1038/s41598-019-48571-w. ISSN   2045-2322. PMC   6702155 . PMID   31431661.
  37. Bickel, Samuel; Or, Dani (2020-01-08). "Soil bacterial diversity mediated by microscale aqueous-phase processes across biomes". Nature Communications. 11 (1): 116. doi:10.1038/s41467-019-13966-w. ISSN   2041-1723. PMC   6949233 . PMID   31913270.
  38. Clark, Nicholas; Clegg, S.; Lima, M. (2014). "A review of global diversity in avian haemosporidians (Plasmodium and Haemoproteus: Haemosporida): new insights from molecular data". International Journal for Parasitology. 44 (5): 329–338. doi:10.1016/j.ijpara.2014.01.004. hdl: 10072/61114 . PMID   24556563.
  39. Mora, Camilo; Tittensor, Derek P; Myers, Ransom A (22 January 2008). "The completeness of taxonomic inventories for describing the global diversity and distribution of marine fishes". Proceedings of the Royal Society B: Biological Sciences. 275 (1631): 149–155. doi:10.1098/rspb.2007.1315. PMC   2596190 . PMID   17999950.
Bibliography