Open access

Introductory Chapter: Factors That Affect Biodiversity and Species Richness of Ecosystems - A Review

Written By

Levente Hufnagel and Ferenc Mics

Published: 08 July 2022

DOI: 10.5772/intechopen.105890

From the Edited Volume

Biodiversity of Ecosystems

Edited by Levente Hufnagel

Chapter metrics overview

253 Chapter Downloads

View Full Metrics

1. Introduction

The general latitudinal trend is that species diversity declines as latitude increases [1]. This appears to be the case for almost all terrestrial plants and animals [2, 3]. It is also usually used for the distribution of marine species [4, 5]. This latitudinal pattern in species richness is detectable in different spatial scales, habitats, and taxonomic groups [6]. However, because of the lack of agreement on the dynamics of site-to-site heterogeneity in species composition (b-diversity) through latitudinal gradients, latitudinal variations in species co-occurrence remain a central issue in ecology [7, 8]. Living species colonized and changed nearly all aquatic and terrestrial ecosystems on Earth, and thereby developed in their many shapes, physiology, and life history from origins more than three billion years ago. Ecosystems are not uniformly rich in species but this richness shows pattern across the world, and observers have long been puzzled about the origins of striking diversity trends such as the latitudinal diversity gradient, elevational gradients in terrestrial ecosystems, and bathymetric gradients in the sea. Since Darwin and Wallace, biologists have been focused on elucidating the mechanisms that lead to diversity. To explain why some regions of the world host greater numbers of taxa than other areas, studies on lineage differentiation and longevity (evolutionary biology) and organismal survivorship and coexistence (ecology, etc.) must be combined. In the last decade or so, new data on the pattern of biodiversity, fine-scale maps of climate and environmental factors, and huge developments in the reconstruction of the tree of life have been ongoing. Biogeographical and phylogenetic data integrations also restructured the patterns of the global distribution of species diversity as well as phylogenetic connections of organisms on various taxonomic scales. The processes that underlie species diversity trends are crucial to study, particularly when people influence important environmental changes [9]. Humans are diminishing habitat space for the majority of the world’s organisms, while global temperature and precipitation regimes are shifting, and these factors are in the focus of many hypotheses describing the nature and maintenance of species diversity. To complicate things, some of these assumptions involve the same factors, namely geographic area, energy climate stability, and biotic interactions, and generate similar predictions about diversity and the rate of speciation because most of the hypothetical factors coincide with the latitude. However, the intense theoretical and methodological attention focused on species richness gradients has not been applied to consider the broadscale distribution of other essential components of biodiversity, such as intraspecific diversity. Intraspecific diversity—whether defined as functional diversity, phylogenetic diversity, population richness, or genetic diversity within and among populations—can influence species’ geographic distributions and responses to environmental change [10, 11, 12, 13], community structure, and ecosystem functioning [14, 15]. Given the alarming pace of species loss and human modification of natural ecosystems, the importance to quantify biodiversity is crucial (e.g., [16, 17, 18, 19, 20]). Species richness, namely the number of species per unit area, is perhaps the simplest and most often used indicator of biological diversity. A substantial amount of biological research has been done using species richness as a metric to explain what causes, and is caused by, biodiversity. Species richness is often used as a measure of diversity within a single biological community, ecosystem, or habitat (e.g., [21]). Pianka [22] wrote the first review paper on large-scale diversity gradients and reviewed major hypotheses to explain the latitudinal diversity gradient. Latitudinal gradients have been known for almost a century in species diversity, and nowadays, some of these polar-equatorial patterns have been explored in depth and several authors added new hypotheses to explain latitudinal gradients [8, 23, 24]. The gradient-forming drivers are differentiated on the basis of significance to different stages of lineage divarication, survival, and allopatric; parapatric and sympatric speciation can be integrated into this framework [25].

Advertisement

2. Abiotic factors

When a population of a certain taxon extends, its range to an extra-tropical area has to face evolutionary trade-offs that influence speciation. Cold tolerance is the most important trait to avoid frost damage poleward from frost line, in the cold season species without frost tolerance can be easily eliminated, even if freezing does not occur every year. Significant energy must be invested in frost resistance at the expanse of growth and reproduction [26]. Thus, they cannot grow and produce offsprings as fast as frost-intolerant species and are out-competed in frost-free areas [27]. The breadth of tropical climate zone and physiological pressures combined with biotic factors hinder northward wandering and this phenomenon is true for plants and animals (termed Dobzhansky-MacArthur phenomenon) [28]. Animals have to struggle with frost, but they are more mobile and can migrate or become dormant in winter, so this trade-off is not pronounced, but vertebrate and invertebrate fauna almost completely turns over northward and southward moving away from the equator [29]. In addition, many animals’ distribution coincides with that of feeding plants [30].

Many taxonomic groups are of tropical origins, and the main source of diversity is the tropical region (such as birds, amphibians, angiosperms, and many marine invertebrates) [23, 31, 32, 33]. Terrestrial phylogenetical lineages’ expansion is restricted by physiological tolerance toward cold or arid environment, evolution of adaptation is particularly difficult, and these traits have evolved fairly infrequently [34]. Hence, most speciation events occur within the border of biomes to lineages that do not cross boundaries of bioregion. Bacteria do not show particulate biome adherence [35], and criteria that link biodiversity to bioregions are not confirmed; hence, the prevalence of biome crossing lineages, dispersal rates of lineages, and the size of transition zones between biomes should be determined [36]. On the regional scale, speciation is believed to be affected by the population size with the presumption that bigger areas support a larger population [37]. Population size, in turn, is positively related to the physical extent of the bioregion and genetic diversity. Size of distribution area is believed positively correlated with allopatric speciation, and small-ranged species have no opportunity to be divided by a barrier within the range. If ranges are large enough to surround a barrier, the species has also less chance to be divided into two new species. Leading to conclusion that medium-ranged species are predicted to have the highest speciation rate [38]. Bioregions with tectonic activity such as mountain lifting have improved opportunities for speciation and environmental gradient by emerging new mountain ranges. Finally, wide population ranges result in great genetic and phenotypic variability increasing the possibility of survival in a highly heterogenic environment, which is a prerequisite for parapatric speciation [39]. There is a clear correlation between the extent of the area and the likelihood of extinction. In smaller bioregions, the average range of species is also smaller. This is associated with a higher likelihood of extinction due to catastrophic events [40]. In addition, in small populations, both genetic and phenotypic variations are small, which easily leads to inbreeding, and are therefore less able to adapt to changing environmental conditions [41].

The extent and position of the temperate belt have changed many times in the history of the Earth, to which glaciations have contributed [42]. Of course, temperate bioregions have existed for at least 50 million years, where mammals, birds, reptiles, amphibians, and seed plants also lived [43].

The size of the population also positively correlated with the amount of available energy. Of course, in a larger bioregion and in a larger population, this is even more pronounced [44]. This is analogous to the idea of population abundances from metabolic theory, and this assumes that more available energy in one region allows for a larger population size, affecting speciation and extinction rate [45]. High-productivity areas tend to be more heterogeneous, with several sources available to populations, than low-productivity bioregions, increasing the likelihood of speciation [46]. It should be mentioned that energy as a factor influencing speciation rate is not equal to the theory that the total energy available limits the number of species in a given region. Global trends of annual net primary production (NPP) of natural biomes are critical to understanding global (natural and anthropogenic) carbon budgets, and they are essential to understanding the adaptive relationships and the evolution of ecosystems and biomes. The competitive exclusion hypothesis is one of many hypotheses that seek to understand biodiversity dynamics by focusing on primary productivity [47], the energy-richness (or more individuals) hypothesis [48], the integrated evolutionary speed hypothesis [49], and the biological relativity to water-energy dynamics hypothesis [50, 51]. The first of these theories is based on the premise that the prevailing relationship between productivity and species diversity is unimodal (or hump-shaped), while the other three are based on the assumption that the predominant relationship is positive. Another different perspective for explaining the factors that ensure species coexistence is the hypothesis that there is a carrying capacity determined by the energy available. The ecoregions attain equilibrium over time in terms of the number of species and the speciation and extinction are balanced. The addition of a new species to the area causes a decrease in the population size of the previous species, increasing the chance of extinction because there is a limit on the number of individuals for each species. New arrival may lead to extinction of a local species [52]. The ecological needs of different species differ, which concludes that there is more competition between closely related species because they have similar ecological needs. Over time, diversification slows down resulting in a pattern of diversity-dependent diversification [53]. The assumption is that nearby species compete for available resources in a zero-sum game, and the ecological limit determines the maximum number of species within a clade in a region. During the Cenozoic period, when the climate became cooler, speciation also slowed down, as lower average temperatures also lowered net primary production. In the case of a lower net primary production, the number of new species that could theoretically be formed is also smaller [54]. In the case of higher net primary production, more species may develop but a slowdown can be observed here as well, reaching a limit.

A number of phylogenetic studies have reported diversity-dependent species formation, but in many cases due to methodological and sampling problems, in the practice, it has failed to demonstrate that the evolution of new species would slow down after reaching a theoretical limit [55].

Machac and Graham [56], on the other hand, found in their study that the formation of new species does not slow down in the tropical region and that the slowdown may be an artifact and there is no real limit to species formation.

This does not mean that co-occurring species’ biological traits do not impact their diversification rates. Several examples have been reported that how the evolution of new species is influenced by species interactions [57]. However, it is very difficult to detect the persisting and speciation-promoting effect of the area and the source on clade-level properties in lineages and to distinguish it from the diversity-limiting effect of area and productivity [58].

In addition, the theory of ecological limitations did not provide a strong conceptual relation between how the volume of resources is associated with carrying capacity for all organisms, and how resources could then be specifically related to the diversity of species [59]. Large-scale dynamics of species diversity did not correlate well with numbers of organisms, according to previous reviews that found little evidence for the species-energy hypothesis [60]. While new immigration to an area (or new species emerging in situ) would inevitably consume some of the region’s available resources, it is unclear how a new species would affect the community abundances of all other native species unless only one other species exists in the area. The relative abundances of any of these species will differ in an infinite number of ways given a certain amount of resources and a certain degree of species diversity [61]. Furthermore, observational experiments have shown that common organisms in an area may have extraordinarily high relative abundances [62]. If newcomer species decrease the abundance of common species while not reducing the abundance of rarer species, so more rare species can be easily accommodated within the field. Our knowledge of how the relative abundance of species changes over time is complicated by the Milankovitch cycle, which also changes the extent and shape of bioregions in 14, 21, and 100,000-year cycles. These periods are often shorter than we would expect for species formation or extinction [63, 64]. Furthermore, with their recent histories, certain bioregions such as tundra and boreal bioregions are unlikely to have reached equilibrium diversity for trees or vertebrates [65].

Finally, it is unclear which scientific evidence would be required to assess whether bioregions have attained their carrying potential for species diversity. Rather than thinking about whether or not a bioregion represents an equilibrium or nonequilibrium structure, I believe it is more efficient to concentrate on how various drivers can affect speciation and extinction [66].

Species energy, time-integrated area, and tropical conservatism hypotheses are based on some mechanistic assumptions, taking into account population size, speciation, and extinction rates, these are worth combining. These concepts are based on the fact that by observing area and energy over a long period of time, diversity and diversification rates can be estimated on a large scale. Marin et al. [67] concluded that one of the most important factors in predicting species richness is a time-integrated area (area through time). Ecological and climatic stability influences species richness indirectly, altering the evolutionary time (i.e., persistence time) and rate. According to their discovery, global heterogeneity in species richness can be primarily explained by the duration of evolution. Colville et al. [68]. In South Africa, 4813 plant species from Cape Floristic Province and 21 molecularly dated endemic clade were added to the simulation study, and the age and area hypothesis was tested taking into account known climates and topography going back 140,000 years. The regression model showed that long-term stability is the deciding factor in explaining species richness. In general, fossil analyses of both aquatic and terrestrial clades have shown higher rates of speciation (origination) in tropical areas [69, 70]. In the sea, the gradient is less steep because temperatures change less and organisms do not have to adapt to such large fluctuations as on land [71]. The gradient is quite conservative, dispersing events occurring in 4–5% of temperate regions. In most cases, dispersion occurs in other tropical regions. Duchêne and Cardillo [72] came to this conclusion in their study by integrating the phylogeny of 9000 bird species. Dispersal events toward temperate areas are generally no older than the Eocene-Oligocene Climate Transition. Rivadeneira et al. [73] used data from 328 marine mollusk species and 159 genera to explore the evolutionary processes that lead to the emergence of the current observed distribution. To do this, the fossil record, nestedness analysis, and projection matrix are used as complementary approaches. Geographical distribution was nested irrespective of the region of origin of genera, and according to the distribution dynamics model, dispersion events were common from temperate areas to the tropics, where extinction events were much rarer. In conclusion, despite the difficulty of distinguishing the signs of speciation and extinction in phylogenetic studies of diversification, all large-scale phylogenic research papers provide findings that are consistent with the tropical conservatism hypothesis and the area/energy/time hypothesis, namely that species have been concentrated in tropical bioregions to a greater extent than extratropical bioregions. The climate stability hypothesis complements the area/energy/time hypothesis in the sense that it is based on a similar mechanistic basis: the genetic variability of population and the geographical extent and, in return, these factors influence speciation and extinction rates. Building on the projections that a larger area and greater genetic variability can influence species formation and extinction, climate stability postulates that these factors are highly mediated by orbitally enforced range dynamics. Therefore, the time-integrated area of temperate climate regions is likely to fluctuate strongly during cold periods and increase the extinction rate. As the chances of extinction increase due to minimum temperatures, a cumulative time-integrated area may not be the best way to test the relationship between species richness and total area available for populations within a region. Instead, if contraction has resulted in extinction events, it is worth working with the minimum area when determining the species richness. Paleoclimatic changes during the Paleo and Neogene allowed or hindered species wandering between areas, as well as regional extinctions, leaving a mark on species and genera distribution. Few species and genera are disjunct between eastern North America and East Asia among temperate plants, suggesting past connectivity over the Bering land bridge as well as climate-driven extinctions from Europe and western North America [74]. Quaternary glacial-interglacial oscillations have left legacies in existing species ranges, according to a wide body of evidence. Ordonez and Svenning [75] examined the distribution patterns of Europe’s contemporary vegetation using six independent lineages (Caryophyllales, Brassicales, Ranunculales, Saxifragales, Rosales, and Malpighiales). The Pleistocene climate, the former location and extent of refugium, the accessibility of areas after the Ice Age, and the contemporary environment play a major role in their current distribution. In concordance with a previous study, Costa et al. [76] concluded that climate stability played key role in Holocene biodiversity of the Amazonian-basin using 30 kyr pollen record and random forest classification. Decreases or fluctuations in temperature negatively correlate with species richness [77]. Deep-sea invertebrates are exception to the role because they do not exhibit equator-pole diversity gradient. Marine invertebrates can be much more diverse in deep waters, which are always roughly 4°C, than in shallow seas. But even in this case, stability is very important. So, constant temperature and access to resources have no variance over time. Deep sea is not productive at all, but resource availability is constant throughout the year [78]. A study of the last 500 million years shows that a steep gradient exists only in the last 30 million years due to cooling. During warm periods, the gradient was much shallower or it was turned around, and there were times when the temperate zone had maximum biodiversity. At that time, the climate was much more stable throughout the world, with few fluctuations [79]. High temperatures have direct positive effect on ectothermic animals, as physiological processes accelerate and evolutional processes can be faster. Faster metabolism leads to shorter generational times, which, together with a higher mutation rate, leads to the formation of new species [80].

Due to the impact of temperature on fostering reproductive isolation, the evolutionary pace hypothesis suggests that speciation rate may be higher in warmer bioregions. Orton et al. [49] conducted a comprehensive study involving multiple phyla (Arthropoda, Chordata, Mollusca, Annelida, Echinodermata, and Cnidaria), COI (Cytochrome c Oxidase subunit I) sequences, and 8037 lineages, but only 51.6% of pairs exhibit higher mutation rate near the equator. For the remainder, the mutation rate was higher in the higher latitudes. This is the most comprehensive study in the literature that has occurred in recent years, and although the result has proved significant, the results are not entirely conclusive to prove evolutionary speed hypothesis.

Advertisement

3. Biotic factors

Organisms also interact with one another, and the effect on one another other also influences evolution. Andresen et al. [81] proved relationships are stronger and more diverse than in the temperate zone. Four different hypotheses attempt to explain species richness: Two involve the speciation (enemy-mediated habitat and the geographic mosaic of coevolution) and two relate to the existence and retention of diversity (competition causing finer niches and predation promoting coexistence). If access to resources differs in two different habitats, the plant will find it more difficult to regenerate the damage caused by herbivores and parasites in places where nutrients are less accessible. Therefore, they need to produce several compounds that keep animals away. However, due to protection, they grow more slowly and are less competitive with specimens living in nutrient-rich areas [82]. Because there are more enemies in productive environments in the tropics, this effect is also stronger. During the formation of the new species, it specializes in a particular habitat with growth protection optimization [83]. Different selection and speciation probabilities between interdepending mutualist or antagonistic species, in tropical climates, should be higher. Organisms are rarely killed by abiotic stress in tropical climates, with most of the selection processes in the tropics controlled by biotic factors. These factors evolve themselves and give an impetus to the evolution of the other species. In tropical environments, where seasonal differences are small, there are several interacting species (plants, herbivores, pathogens, predators, mutualists, etc.) and their relationships persist throughout the year. In cold climates, seasonality disrupts relations, slowing down co-evolution. In stable warm environments, co-evolution is faster and easier to develop mutualistic interrelationships [84]. The novel protection mechanism protects the plant from enemies, which enables the plant to increase its range, stimulating allopatric specificity. But if this enemy is able to adapt to the defense mechanism, then the range of the plant may shrink. Similarly, specialization may lead to speciation through the development of the host race or the increase in patches of geographical specialists that lead to differentiation. These processes will also increase over time and generate a positive feedback loop [85]. An example is the relationship between butterflies in the Nymphalidae family and solanaceae feed plants. The Solanaceae family separated from its sister groups 49−68 million years ago and then diversified (29–47 MYA). Around this time, the Ithomiini subfamily of the Nymphalidae family was also formed and diversified [86]. But there are also examples where this diversification effect has not been confirmed. According to Kaczvinsky and Hardy [87], the emergence of a new plant-predator relationship could increase the chances of extinction. The development of a connection with the new plant can be due to the significant fitness reduction of ancient feed plant. This can be caused by several factors, such as increasing competition for diminishing resources, the development of a new defense mechanism, the emergence of a new invasive species, and the emergence of new natural enemies. If a change in the feeding plant of an insect is caused by such an effect, it will cause a decrease in the population size and chance of speciation. With decreased overall performance of ancestral host and presumably minimal overall performance of new host, the growth rates and adequate sizes of herbivorous insect populations should shrink, as well as their geographic and climatic area.

Two other hypotheses explain that in areas where interrelationships between organisms are stronger, the width of the niche is smaller, in the sense that it uses habitat and resources. As competition intensifies, niches can split up, resulting in a finer niche for evolving new species. Thus, the existence of more species can be imagined within the community [22]. Harmáčková et al. [88] tested the hypothesis by using 298 Australian songbirds’ (Passeriformes) data. It was expected that the species richness-specialization relationship is stronger in particularly species-rich communities, where annual precipitation is high and vegetation is complex. They also tested the extent of niche overlap. A positive relationship was found between specialization and species richness, but the direction and strength of relationship vary according to traits and area size. The specialization-species richness relationship is clear only in the forage stratum and has increased toward a smaller area only for habitat and diet. At the same time, local communities had a high overlap in habitat and diet. In a particularly species-rich community, no particularly strong link was found between species richness and specialization. However, they found a negative connection between specialization and overlap, meaning that species separate the ecological space on the basis of where food is found. These just weakly supported their expectations. The specialization in forage stratum has probably been significant in promoting species coexistence. On the other hand, while several species were habitat and diet specialists, high overlap in these traits did not rule out coexistence. The stronger predator (or herbivore) is able to drive selection in the antipredator traits in the prey (or plant). This is the basis of predator-victim specialization, the prey (or plant) can compete for a predator-free environment, leading to greater coexistence of the prey (or plant) species. Horst and Venable [89] studied the seed predation of rodents in the Sonora Desert. The rodents prefer the seeds of a certain species out of three different plants, which allows the two other plants to coexist. Regular predation on the more common species reduces interspecific competition between the three plants.

Phylogenetic studies of trees, birds, mammals, reptiles, and amphibians show that tropical regions are the source of biodiversity, composed of both ancient and recently evolved lineages. Moderate regions contain only a fraction of this. Over the past millions of years, the climate of tropical areas has been more stable, which has helped new species to evolve and reduced extinction rate. And niche conservatism has prevented lineages to wander to the temperate zone. It seems that the energy available is a secondary factor because deep sea life has become almost completely independent of it. But temperature and availability of resources are stable there in the long run. The time-integrated area is the most important controller of evolutionary processes. The seasonality of temperature and productivity affects the number of resources available on land, which determines the size of the populations. And lastly, speciation and coexistence are augmented by biotic interactions.

References

  1. 1. Briggs JC. Ocean Biogeography. International Encyclopedia of Geography: People, the Earth, Environment and Technology. Hoboken, New Jersey, U.S.: John Wiley & Sons Ltd; 2016. pp. 1-5
  2. 2. Cirtwill AR, Stouffer DB, Romanuk TN. Latitudinal gradients in biotic niche breadth vary across ecosystem types. Proceedings of the Royal Society B: Biological Sciences. 2015;282:20151589
  3. 3. Maestri R, Patterson BD. Patterns of species richness and turnover for the south american rodent fauna. PLoS One. 2016;11(3):e0151895
  4. 4. Kinlock NL, Prowant L, Herstoff EM, Foley CM, Akin-Fajiye M, Bender N, et al. Explaining global variation in the latitudinal diversity gradient: Meta-analysis confirms known patterns and uncovers new ones. Global Ecology and Biogeography. 2017;27(1):125-141
  5. 5. Saupe EE, Myers CE, Townsend Peterson A, Soberón J, Singarayer J, Valdes P, et al. Spatio-temporal climate change contributes to latitudinal diversity gradients. Nature Ecology & Evolution. 2019;3:1419-1429
  6. 6. Etienne RS, Cabral JS, Hagen O, Hartig F, Hurlbert Pellissier L, Pontarp M, et al. A minimal model for the latitudinal diversity gradient suggests a dominant role for ecological limits. The American Naturalist. 2019;194(5):122-133
  7. 7. Fordyce JA, DeVries PJ. A tale of two communities: Neotropical butterfly assemblages show higher beta diversity in the canopy compared to the understory. Oecologia. 2016;181:235-243
  8. 8. Willig MR, Presley SJ. Latitudinal gradients of biodiversity: Theory and empirical patterns. In: DellaSala DA, Goldstein MI, editors. The Encyclopedia of the Anthropocene. Oxford: Elsevier; 2018
  9. 9. Hudson L, Newbold T, Contu S, Hill S, Lysenko I, De Palma A, et al. The database of the PREDICTS (projecting responses of ecological diversity in changing terrestrial systems) project. Ecology and Evolution. 2017;7:145-188
  10. 10. Bernatchez L. On the maintenance of genetic variation and adaptation to environmental change: Considerations from population genomics in fishes. Journal of Fish Biology. 2016;89:2519-2556
  11. 11. Matuszewski S, Hermisson J, Kopp M. Catch me if you can: Adaptation from standing genetic variation to a moving phenotypic optimum. Genetics. 2015;200(4):1255-1274
  12. 12. Sollars ESA, Harper AL, Kelly LJ, Sambles CM, Ramirez-Gonzalez RH, Swarbreck D, et al. Genome sequence and genetic diversity of European ash trees. Nature. 2016;541:212-216
  13. 13. Matz MV, Treml EA, Aglyamova GV, Bay LK. Potential and limits for rapid genetic adaptation to warming in a Great Barrier Reef coral. PLoS Genetics. 2018;14(4):e1007220
  14. 14. Raffard A, Santoul F, Cucherousset J, Blanchet S. The community and ecosystem consequences of intraspecific diversity: A meta-analysis. Biological Reviews. 2019;94(2):648-661
  15. 15. Siefert A, Violle C, Chalmandrier L, Albert CH, Taudiere A, Fajardo Aarssen LW, et al. A global meta-analysis of the relative extent of intraspecific trait variation in plant communities. Ecology Letters. 2015;18(12):1406-1419
  16. 16. Suo AN, Lin Y, Sun YG. Impact of sea reclamation on zoobentic community in adjacent sea area: A case study in Caofeidian, North China. Applied Ecology and Environmental Research. 2017;15(3):871-880
  17. 17. Jahanbakhsh GM, Khorasani N, Morshedi J, Danehkar A, Naderi M. Factors influencing abundance and species richness of overwintered waterbirds in Parishan International Wetland in Iran. Applied Ecology and Environmental Research. 2017;15(4):1565-1579
  18. 18. Fornal-Pieniak B, Ollik M, Schwerk A. Impact of surroundings landscape structure on formation of plant species in aforestrated manor parks. Applied Ecology and Environmental Research. 2018;16(5):6483-6497
  19. 19. Rajamurugan J, Mohandass D, Campbell MC, Jayakrishnan P, Balachandran N, Shao S-C. Fragmentation causes woody plant composition decline in sacred grove patches in the Puducherry Region of Shoutheast. Applied Ecology and Environmental Research. 2021;19(3):1625-1643
  20. 20. Zhang GX, Yuan XZ, Wang KH, Zhang MJ, Zhou LL, Zhang QY, et al. Biodiversity conservation in agricultural landscapes: An ecological opportunity for coal mining subsidence areas. Applied Ecology and Environmental Research. 2020;18(3):4283-4308
  21. 21. Tufan-Çetin Ö. Determination of lichen diversity variations in habitat type of mediterranean maquis and arborescent matorral. Applied Ecology and Environmental Research. 2019;17(4):10173-10193
  22. 22. Pianka ER. Latitudinal gradients in species diversity: A review of concepts. The American Naturalist. 1966;100(910):33-46
  23. 23. Jablonski D, Huang S, Roy K, Valentine JW, Bronstein JL. Shaping the latitudinal diversity gradient: New perspectives from a synthesis of paleobiology and biogeography. The American Naturalist. 2017;189(1):1-12
  24. 24. Schluter D, Pennell MW. Speciation gradients and the distribution of biodiversity. Nature. 2017;546:48-55
  25. 25. Fine PVA. Ecological and evolutionary drivers of geographic variation in species diversity. Annual Review of Ecology, Evolution, and Systematics. 2015;46:369-392
  26. 26. Sebastian-Azcona J, Harmann A, Hacke UG, Rweyongeza D. Survival, growth and cold hardiness tradeoffs in white spruce populations: Implications for assisted migration. Forest Ecology and Management. 2019;433:544-552
  27. 27. Bucher SF, Feiler R, Buchner O, Neuner G, Rosbakh S, Leiterer M, et al. Temporal and spatial trade-offs between resistance and performance traits in herbaceous plant species. Environmental and Experimental Botany. 2019;157:187-196
  28. 28. Preisser W. Latitudinal gradients of parasite richness: A review and new insights from helminths of cricetid rodents. Ecography. 2019;42(7):1315-1330
  29. 29. Kolomiytsev N, Poddubnaya N. Temporal and spatial variability of environments drive the patterns of species richness along latitudinal, elevational, and depth gradients. Communications Biology. 2018;63(3):189-201
  30. 30. Du C, Chen J, Jiang L, Qiao G. High correlation of species diversity patterns between specialist herbivorous insects and their specific hosts. Journal of Biogeography. 2020;00:1-14
  31. 31. Marin J, Hedges SB. Time best explains global variation in species richness of amphibians, birds and mammals. Journal of Biogeography. 2016;43(6):1069-1079
  32. 32. Pyron RA, Costa GC, Patten MA, Burbrink FT. Phylogenetic niche conservatism and the evolutionary basis of ecological speciation. Biological Reviews. 2015;90(4):1248-1262
  33. 33. Qian H, Ricklefs RE. Out of the tropical lowlands: Latitude versus elevation. Trends in Ecology & Evolution. 2016;31(10):738-741
  34. 34. Schubert M, Grønvold L, Sandve SR, Hvidsten TR, Fjellheim S. Evolution of cold acclimation and its role in niche transition in the temperate grass subfamily pooideae. Plant Physiology. 2019;180:404-419
  35. 35. Moss JA, Henriksson NL, Pakulski JD, Snyder RA, Jeffrey WH. Oceanic microplankton do not adhere to the latitudinal diversity gradient. Microbial Ecology. 2020;79:511-515
  36. 36. Edwards EJ, Donoghue MJ. Is it easy to move and easy to evolve? Evolutionary accessibility and adaptation. Journal of Experimental Botany. 2013;64(13):4047-4052
  37. 37. Patiño J, Weigelt P, Guilhaumon F, Kreft H, Triantis KA, Naranjo-Cigala A, et al. Differences in species–area relationships among the major lineages of land plants: A macroecological perspective. Global Ecology and Biogeography. 2013;23(11):1275-1283
  38. 38. Castiglione S, Mondanaro A, Melchionna M, Serio C, Di Febbraro M, Carotenuto F, et al. Diversification rates and the evolution of species range size frequency distribution. Frontiers in Ecology and Evolution. 2017;5:147
  39. 39. Polechová J, Barton NH. Limits to adaptation along environmental gradients. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(20):6401-6406
  40. 40. Chichorro F, Juslén A, Cardoso P. A review of the relation between species traits and extinction risk. Biological Conservation. 2019;237:220-229
  41. 41. Weeks AR, Stoklosa J, Hoffmann AA. Conservation of genetic uniqueness of populations may increase extinction likelihood of endangered species: The case of Australian mammals. Frontiers in Zoology. 2016;13:31
  42. 42. Svenning J, Eiserhardt W, Normand S, Ordonez A, Sandel B. The influence of paleoclimate on present-day patterns in biodiversity and ecosystems. Annual Review of Ecology, Evolution, and Systematics. 2015;46(1):551-572
  43. 43. Anisha D, Akash G. Cenozoic era. In: Vonk J, Shackelford TK, editors. Encyclopedia of Animal Cognition and Behavior. Switzerland: Springer Nature; 2021
  44. 44. Woolley SNC, Tittensor DP, Dunstan PK, Guillera-Arroita G, Lahoz-Monfort JJ, Wintle BA, et al. Deep-sea diversity patterns are shaped by energy availability. Nature. 2016;533(7603):393-393
  45. 45. Belmaker J, Jetz W. Relative roles of ecological and energetic constraints, diversification rates and region history on global species richness gradients. Ecology Letters. 2015;18:563-571
  46. 46. Armitage DW. Experimental evidence for a time-integrated effect of productivity on diversity. Ecology Letters. 2015;18(11):1216-1225
  47. 47. Pocheville A. (2015): The ecological niche: History and Recent Controversies. In: Heams T, Huneman P, Lecointre G, Silberstein M. (eds). Handbook of Evolutionary Thinking in the Sciences. Dordrecht: Springer; 2015
  48. 48. Rabosky DL, Hurlbert AH, Price T. Species richness at continental scales is dominated by ecological limits. The American Naturalist. 2015;185(5):572-583
  49. 49. Orton MG, May JA, Ly W, Lee DJ, Adamowicz SJ. Is molecular evolution faster in the tropics? Heredity. 2019;122(5):513-524
  50. 50. Dong Y, Wu N, Li F, Chen X, Zhang D, Zhang Y, et al. Influence of monsoonal water-energy dynamics on terrestrial mollusk species-diversity gradients in northern China. Science of the Total Environment. 2019;676:206-2014
  51. 51. Xu X, Wang Z, Rahbek C, Sanders NJ, Fang J. Geographical variation in the importance of water and energy for oak diversity. Journal of Biogeography. 2015;43(2):279-288
  52. 52. Pontarp M, Wiens JJ. The origin of species richness patterns along environmental gradients: Uniting explanations based on time, diversification rate and carrying capacity. Journal of Biogeography. 2016;44(4):722-735
  53. 53. Gatti RC. A conceptual model of new hypothesis on the evolution of biodiversity. Biologia. 2016;71(3):343-351
  54. 54. Condamine FL, Rolland J, Morlon H. Assessing the causes of diversification slowdowns: Temperature-dependent and diversity-dependent models receive equivalent support. Ecology Letters. 2019;22:1900-1912
  55. 55. Pannetier T, Martinez C, Bunnefeld L, Etienne RS. Branching patterns in phylogenies cannot distinguish diversity-dependent diversification from time-dependent diversification. Evolution. 2021;75(1):25-38
  56. 56. Machac A, Graham CH. Regional diversity and diversification in mammals. The American Naturalist. 2017;189(1):1-13
  57. 57. Zeng Y, Wiens JJ. Species interactions have predictable impacts on diversification. Ecology Letters. 2021;24(2):239-248
  58. 58. Šímová I, Storch D. The enigma of terrestrial primary productivity: Measurements, models, scales and the diversity–productivity relationship. Ecography. 2017;40(2):239-252
  59. 59. Storch D, Okie JG. The carrying capacity for species richness. Global Ecology and Biogeography. 2019;28:1519-1532
  60. 60. Storch D, Bohdalkova E, Okie J. He more-individuals hypothesis revisited: The role of community abundance in species richness regulation and the productivity–diversity relationship. Ecology Letters. 2018;21(6):920-937
  61. 61. Mateo RG, Mokany K, Guisan A. Biodiversity models: What if unsaturation is the rule? Trends in Ecology & Evolution. 2017;32(8):556-566
  62. 62. Hopkins MJG. Are we close to knowing the plant diversity of the Amazon? Annals of the Brazilian Academy of Sciences. 2019;91(Suppl. 3):e20190396
  63. 63. Piek M. Milankovitch cycles: Precession discovered and explained from Hipparchus to Newton [bachelor thesis]. Utrecht: University of Utrecht; 2015
  64. 64. Simões M, Breitkreutz L, Alvarado M, Baca S, Cooper JC, Heins L, et al. The evolving theory of evolutionary radiations. Trends in Ecology and Evolution. 2016;31(1):27-34
  65. 65. Schluter D, Bronstein JL. Speciation, ecological opportunity, and latitude. The American Naturalist. 2016;187(1):1-18
  66. 66. Grace JB, Anderson TM, Seabloom EW, Borer ET, Adler PB, Harpole WS, et al. Integrative modelling reveals mechanisms linking productivity and plant species richness. Nature. 2016;529:390-393
  67. 67. Marin J, Rapacciuolo G, Costa GC, Graham CH, Brooks TM, Young BE, et al. Evolutionary time drives global tetrapod diversity. Proceedings of the Royal Society B: Biological Sciences. 2018;285:20172378
  68. 68. Colville JF, Beale CM, Forest Altwegg R, Huntley B, Cowling RM. Plant richness, turnover, and evolutionary diversity track gradients of stability and ecological opportunity in a megadiversity center. Proceedings of the National Academy of Sciences of the United States of America. 2020;117(33):20027-20037
  69. 69. Cowman PF, Parravicini V, Kulbicki M, Floeter SR. The biogeography of tropical reef fishes: Endemism and provinciality through time. Biological Reviews. 2017;92(4):2112-2130
  70. 70. Tamma K, Ramakrishnan U. Higher speciation and lower extinction rates influence mammal diversity gradients in Asia. BMC Evolutionary Biology. 2015;15:11
  71. 71. Yasuhara M, Danovaro R. Temperature impacts on deep-sea biodiversity. Biological Reviews. 2016;91(2):275-287
  72. 72. Duchêne DA, Cardillo M. Phylogenetic patterns in the geographic distributions of birds support the tropical conservatism hypothesis. Global Ecology and Biogeography. 2015;24(11):1261-1268
  73. 73. Rivadeneira MM, Alballay AH, Villafaña JA, Raimondi PT, Blanchette CA, Fenberg PB. Geographic patterns of diversification and the latitudinal gradient of richness of rocky intertidal gastropods: The ‘into the tropical museum’ hypothesis. Global Ecology and Biogeography. 2015;24:1149-1158
  74. 74. Xiang J-Y, Wen J, Peng H. Evolution of the eastern Asian–North American biogeographic disjunctions in ferns and lycophytes. Journal of Systematics and Evolution. 2015;53(1):2-32
  75. 75. Ordonez A, Svenning J-C. Consistent role of quaternary climate change in shaping current plant functional diversity patterns across European plant orders. Scientific Reports. 2017;7:42988
  76. 76. Costa GC, Hampe A, Ledru M-P, Martinez PA, Mazzochini GG, Shepard DB, et al. Biome stability in South America over the last 30 kyr: Inferences from long-term vegetation dynamics and habitat modelling. Global Ecology and Biogeography. 2018;27(3):285-297
  77. 77. Rasconi S, Winter K, Kaintz MJ. Temperature increase and fluctuation induce phytoplankton biodiversity loss—Evidence from a multi-seasonal mesocosm experiment. Ecology and Evolution. 2017;7(9):2936-2946
  78. 78. Chaudhary C. Global-scale distribution of marine species diversity: An analysis of latitudinal, longitudinal and depth gradients [PhD thesis]. New Zealand: University of Auckland; 2019
  79. 79. Mannion PB, Upchurch P, Benson RBJ, Goswami A. The latitudinal biodiversity gradient through deep time. Trends in Ecology & Evolution. 2014;29:42-50
  80. 80. Puurtinen M, Elo M, Jalasvuori M, Kahilainen A, Ketola T, Kotiaho JS, et al. Temperature-dependent mutational robustness can explain faster molecular evolution at warm temperatures, affecting speciation rate and global patterns of species diversity. Ecography. 2016;39(11):1025-1033
  81. 81. Andresen E, Arroyo-Rodríguez V, Escobar F. Tropical biodiversity: The importance of biotic interactions for its origin, maintenance, function, and conservation. In: Dáttilo W, Rico-Gray V. Ecological Networks in the Tropics. Cham:Springer; 2018
  82. 82. Jia S, Wang X, Yuan Z, Lin F, Ye J, Lin G, et al. Tree species traits affect which natural enemies drive the Janzen-Connell effect in a temperate forest. Nature Communications. 2020;11:286
  83. 83. Viswanathan A, Ghazoul J, Lewis OT, Honwad G, Bagchi R. Effects of forest fragment area on interactions between plants and their natural enemies: Consequences for plant diversity at multiple spatial scales. Frontiers in Forests and Global Change. 2020;2:88
  84. 84. Medeiros LP, Garcia G, Thompson JN, Guimaraes PR. The geographic mosaic of coevolution in mutualistic networks. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(47):12017-12022
  85. 85. Harmon LJ, Andreazzi CS, Débarre F, Drury J, Goldberg EE, Martins AB, et al. Detecting the macroevolutionary signal of species interactions. Journal of Evolutionary Biology. 2019;32(8):769-782
  86. 86. Khyade VB, Gaikwad PM, Vare PR. Explanation of nymphalidae butterflies. International Academic Journal of Science and Engineering. 2018;5(4):24-47
  87. 87. Kaczvinsky C, Hardy NB. Do major host shifts spark diversification in butterflies? Ecology and Evolution. 2020;0(0):1-11
  88. 88. Harmáčková L, Remešová E, Remeš V. Specialization and niche overlap across spatial scales: Revealing ecological factors shaping species richness and coexistence in Australian songbirds. Journal of Animal Ecology. 2019;88:1766-1776
  89. 89. Horst JL, Venable DL. Frequency-dependent seed predation by rodents on Sonoran Desert winter annual plants. Ecology. 2018;99(1):196-203

Written By

Levente Hufnagel and Ferenc Mics

Published: 08 July 2022