More stories

  • in

    Evolution of woodcutting behaviour in Early Pliocene beaver driven by consumption of woody plants

    Dipoides sp. palaeoecology
    The Bayesian mixing model indicates that Dipoides sp. consumed both woody plants and freshwater macrophytes in approximately equal proportions (Figs. 6 and 7), although it relied slightly more on freshwater macrophytes. This suggests that Dipoides sp. spent a greater proportion of time feeding in the water than on land.
    Figure 7

    (a) Proportion versus Source Boxplot generated using SIAR, indicating the relative proportion that moss, woody vegetation, and aquatic macrophytes contributed to the diet of Dipoides sp. at the Beaver Pond site. Darker shaded areas indicate highest probability of source proportion. The Proportion versus Source Boxplots for (b) extant Castor canadensis and (c) late Pleistocene Castoroides have been included for comparison. Note the differences in dietary Source data used to distinguish C. canadensis and Castoroides diet (primarily the sub-division of aquatic plants into categories based on habitat within the water column). b and c from Plint et al.13.

    Full size image

    The distribution of Dipoides sp. δ13Ccol and δ15Ncol is not entirely enclosed within the three primary producer functional groups analyzed (Fig. 6). This is likely the result of the relatively small plant macrofossil sample size. Submerged aquatic macrophytes, for example, are under-represented in the plant macrofossils available for stable isotope analysis. Macrophytes have highly variable δ13C and may have contributed more to Dipoides sp. diet than the mixing model suggests. Submerged macrophytes can be highly enriched in 13C because of physiological differences (primarily the use of 13C-enriched dissolved bicarbonate) or environmental conditions in the water column (i.e. boundary-layer effect)54,55,56. In addition, tree bark is more enriched in 13C than tree foliage57 and may have been a key resource for Dipoides sp.
    The results of the dietary mixing model support the interpretation that woody plants were an important contributor to Dipoides sp. diet. It is likely that Dipoides sp. also used shrubs and trees as a source of construction material10,11, but more evidence is needed to confirm this. Similar to extant Castor, Dipoides sp. may have also demonstrated regional differences in diet, where northern and southern populations utilized different resources according to their availability.
    Nitrogen content and C/N as indicators of forage quality
    Plant macrofossil nitrogen content (N wt%) and C/N are indicators of forage quality and may be used to interpret the relative nutrition of dietary inputs. Plants with high N (wt%) contain more protein and energy—likewise, low N (wt%) correlates with low plant digestibility, high fiber and high lignin compound content58. Beaver Pond plant macrofossil N (wt%) and C/N are highly variable (Table 2, Fig. 5). Although there is considerable variability in C/N ratios depending upon which plant part was analyzed (i.e. seeds versus woody tissue), woody vegetation tends to have higher C/N ratios than macrophytes, and thus tends to be of lower food quality. However, the increased structural tissues in woody plants may have rendered them more effective winter cache foods.
    In extremely seasonal environments such as the High Arctic, herbivores must use plant resources in a highly efficient manner. Herbivores must consume the highest quality forage possible during the brief growing season to maximize nutrient and energy gain. High quality forage typically includes young leaves with high nitrogen content, minimal structural (fibrous) tissues, and low defense compound content59,60.
    Within the Beaver Pond macrofossil assemblage, pod grass (an emergent macrophyte) and birch have the highest nitrogen content and lowest C/N (Fig. 5). A larger sample set is necessary to confirm this observation; however, current data supports the conclusion that emergent macrophytes and deciduous broadleaf trees were among the more nutritious types of forage available to Dipoides sp. at the Beaver Pond site. It should be noted that forage quality is not the only factor that governs herbivore feeding behaviour. Animals may preferentially target plants with higher biomass to minimize energy expenditure traveling between forage sites or select plants that grow in locations that minimize the risk of predation.
    The C/N of high Arctic shrubs decreases over the course of the growing season58. As there is no time-constraint on macrofossil deposition at the Beaver Pond site, variation in C/N may also be due to differences in plant phenological stage at time of incorporation into the peat layer. The incorporation of senescent plants into the peat deposit at the end of each growing season may in part account for the lower than expected macrofossil N (wt%) values reported from this site.
    Beaver Pond site flora δ13C and δ15N
    The Beaver Pond macrofossil assemblage contains a diverse range of terrestrial and freshwater plant species. The identified plant species in this study concur with previous interpretations that this was an open-forest landscape interspersed with shallow wetlands. Larch trees and cool-climate woody shrubs dominated the forest community. The wetlands supported both vascular macrophytes and dense assemblages of bryophytes.
    The macrofossil δ13C are all within the range expected for primary producers utilizing the C3 photosynthetic pathway and accessing either ambient or dissolved atmospheric CO2 as their dominant carbon source. The δ15N of the macrofossils are also within the expected range for a riparian ecosystem in a cool climate biome.
    While Dipoides sp. most likely consumed leafy tree branches and woody tissues (cambium), it is worth noting that plant seeds and cone bracts were analyzed in this study due to ease of macrofossil taxonomic identification. Leaf δ13C is typically lower than that of other plant parts61, although there is no clear pattern in intra-plant variation of δ15N.
    Moss
    Samples of the dominant Beaver Pond site bryophyte, Scorpidium (hooked scorpion moss), have very low δ13C for a primary producer (− 36.6 to − 34.6‰). This pattern is consistent with modern mosses collected from freshwater habitats in Subarctic and Arctic regions13,62,63,64.
    Environmental conditions dictate moss δ13C rather than species-specific physiological differences. Peat mosses can grow partially or fully submerged in water. Given that mid-Pliocene atmospheric CO2 concentration levels were similar to modern (~ 400 ppm)15,65,66, moss exposed to the atmosphere would preferentially have used the abundant 12CO2, resulting in low δ13C. Alternatively, low δ13C in peat mosses can also indicate an underwater growing environment rich in 13C-depleted respired CO2 from surrounding plants62.
    Moss macrofossil δ15N is relatively high for a photosynthetic organism (mean =  + 4.8‰). This is indicative of either the presence of 15N-enriched sources of bioavailable N (i.e. dissolved nitrates, organic proteins such as urea or amino acids), or increased nutrient availability45,61. Unlike vascular plants, mosses do not uptake compounds through their roots. Rather, they obtain nutrients from wet or dry deposition through their leaves67,68. Today, beaver ponds are considered to be N sinks, with elevated rates of bacterially mediated denitrification69. These bacterial processes result in 15N-enriched products that are readily dissolved and used by plants (including moss) living in an aqueous environment. Decomposition processes also increase plant δ15N over time47 and remineralized organic debris decomposing in wetlands may be particularly 15N-enriched.
    Beaver Pond bulk peat samples and moss macrofossils show a similar isotopic pattern (low δ13C and high δ15N), which suggests that hooked scorpion moss contributed substantially to peat biomass accumulation at the Beaver Pond site.
    Macrophytes
    Beaver Pond macrophyte δ13C fall well within the albeit very wide known range for modern freshwater plants (− 50 to − 11‰, see Osmond et al.70, Keeley and Sandquist54, Mendonça et al.55, and Chappuis et al.56). It is reasonable to assume, however, that the very small sample size in this study hides the potential extent of the carbon isotope variability of macrophytes at the site.
    Pod grass and bogbean are classified as emergent macrophytes (they grow rooted in water-logged substrates, but their leaves are exposed to the atmosphere), while pondweed grows entirely submerged. Submerged macrophytes become more enriched in 13C as the dissolved CO2 pool (the dominant carbon source) becomes increasingly limited54.
    The Beaver Pond site pondweed δ13C is relatively low (− 26.5‰) for a submerged macrophyte. This indicates that it grew in an aquatic environment with adequate dissolved CO2. This is in keeping with the interpretation that the Beaver Pond was a fen (near neutral pH, cool water temperature) during the Pliocene. A low δ13C may also indicate high influx of terrestrial organic biomass or mosses (with low δ13C) into the water that subsequently remineralized and contributed to the dissolved inorganic carbon pool.
    Environmental conditions strongly influence aquatic plant δ15N. Beaver Pond macrophyte δ15N (range =  + 0.2 to + 2.7‰) indicate interspecific access and use of a variety of different sources of bioavailable N within the water column and substrate. The most likely N sources are microbial-fixed atmospheric N2 (which ranges from –2 to + 2‰), the products of nitrification/denitrification processes (15N-enriched NH4+ or NOx), and remineralized 15N-enriched organic material (either terrestrial or aquatic)71,72,73.
    Larch
    Larch (the extinct species Larix groenlandii) is the most common vascular plant species in this macrofossil assemblage.
    There is an offset of ~ 2‰ between the δ13C of (i) larch shoots/buds (which bear the needles) and cone bracts, and (ii) larch seeds. Larch shoots and cones (δ13C range = ‒25.4 to − 25.1‰; mean = − 25.3‰) are more depleted of 13C than larch seeds (δ13C range = − 23.3 to − 22.7‰, mean = − 23.1‰). This could be indicative of seasonal physiological or environmental conditions experienced by larch trees at the Beaver Pond site. The cones and shoots of extant larch trees begin growing in the early spring and have lower δ13C, whereas their seeds (higher δ13C) do not develop and mature until mid to late summer74.
    A number of physiological and environmental conditions could be responsible for this offset between needle/-bearing structures and seeds. Atmospheric vapor pressure deficit (aridity) induces stomatal closure in vascular plants75. This restricts not only the rate of water leaving the needle/tree, but also that of atmospheric CO2 entering it. Stomatal closure reduces CO2 entry and results in less discrimination against 13CO2. High levels of solar irradiance in the summer increase the rate of CO2 assimilation. Plants growing at very high latitudes experience 24-h of daylight during the summer. This creates a greater demand for CO2 to maintain photosynthesis and less discrimination against 13CO2. Both aridity and increased light levels could contribute to why Beaver Pond larch tissues grown late in the summer are more 13C-enriched than those grown in the early spring/the previous fall.
    Alternatively, trees can use water and carbon (in the form of sugars) stored during the previous year to promote new growth during the early spring when leaves are absent and light levels are low. Tissues that develop early in the growing season (i.e. needle-bearing buds and shoots) can therefore reflect the δ13C of photosynthetic conditions from the previous growing season76,77. In addition, differences in the macromolecular (lipid, protein, sugar) composition of larch buds/needles versus seeds could account for their offset in δ13C (i.e. lipids are typically more 13C-depleted than proteins).
    Larch δ15N (mean =  + 2.7‰) indicate that these conifer trees had access to N sources other than “light” fixed atmospheric N2. Given the proximity of wetlands, the root systems of larch trees may have had access to 15N-enriched dissolved nitrates in the surrounding water-logged soils. Increasing foliar N concentration due to atmospheric N deposition also drives up plant δ15N78,79.
    Aridity may also have influenced terrestrial plants growing at the Beaver Pond site. Higher rainfall is inversely correlated with δ15N, where rainier ecosystems tend to produce more 15N-depleted plants80.
    Similar to δ13C, there is an offset in δ15N (and N wt%) between larch needle-bearing structures (mean =  + 3.8‰; 0.9%) and larch seeds (mean =  + 2.1‰; 0.3%). This could indicate differences in the macromolecular composition of these different tissue types (where high N content typically indicates higher tissue protein content).
    Comparison of Dipoides within Castoridae
    The composition of Dipoides sp. diet differs from that of other members of Castoridae that lived in North America during the late Cenozoic. Pliocene High Arctic Dipoides sp. (n = 5), modern subarctic Castor canadensis (n = 4) (Table 1), and late Pleistocene Castoroides ohioensis (n = 11) (Table 1) δ13Ccol and δ15Ncol are compared in Fig. 3. A correction for the Suess effect was first necessary render the δ13C of all three genera comparable. The carbon isotope composition of atmospheric CO2 has changed over time with global climatic conditions. More recently, anthropogenic burning of fossil fuels that has rapidly released CO2 enriched in 12C into the atmosphere46,81. Hence, a correction is needed when comparing δ13C of organic samples from different time periods to account for this isotopic variation in the primary carbon source of photosynthetic organisms at the base of the food web.
    Suess effect corrections of + 2.02‰ and − 0.1‰ were applied to the δ13Ccol of modern C. canadensis (collected in 2013 and 2014) and Castoroides (late Pleistocene in age), respectively. These corrections were based on the average δ13C of atmospheric CO2 (δ13CCO2) calculated from Pliocene dual-benthic and planktonic foraminifera proxy records, spanning from ~ 4.1 to 3.8 Ma (average δ13CCO2 = − 6.55‰)76. These foraminifera proxy records are approximately contemporary with the Beaver Pond site. Average δ13CCO2 for 2014 (− 8.57‰) was compiled from the Scripps CO2 monitoring program. Average δ13CCO2 for the late Pleistocene (− 6.45‰) was compiled using ice core data from Schmitt et al.82.
    Plants growing during these three different time periods (Pliocene, late Pleistocene, and modern/2014) would reflect the δ13C of contemporary atmospheric CO2. Therefore, changes in δ13CCO2 help explain differences in δ13C between Dipoides sp. and modern C. canadensis. Additional factors, however, are important in explaining the wide range of δ13C and large enrichment in 13C measured for Castoroides.
    Dipoides sp. diet composition differs from that of Castoroides (the Pleistocene giant beaver) (Figs. 3 and 7). Castoroides’ high δ13Ccol and δ15Ncol (mean δ13Ccol = − 17.6‰ and mean δ15Ncol =  + 5.8‰) indicate a diet composed predominantly of aquatic (particularly submerged) macrophytes and minimal woody plant material (Table 1)13.
    In comparison with Castoroides, both Dipoides sp. and C. canadensis have a relatively small range of δ13Ccol and δ15Ncol (Table 1) (Fig. 3). Dipoides sp. mean δ13Ccol and δ15Ncol are higher than those of modern C. canadensis (Fig. 3). This is attributable to either variation in diet between the two species, or changes in global C and N baselines over geologic time.
    Previous mixing model studies predict that extant C. canadensis diet is composed of approximately equal proportions of woody terrestrial plants and aquatic macrophytes13. However, this can vary by latitude and season. For example, extant C. canadensis in the Canadian subarctic vary their winter diet significantly depending on habitat83. It is worth noting that extant C. canadensis does not occur north of 70° latitude and High Arctic Dipoides sp. living at 78° latitude may have employed different dietary strategies.
    Dipoides sp. may have relied more heavily than C. canadensis on underwater stores of tree branches to survive the long, dark polar winter. Tree bark is more 13C-enriched than leafy vegetation57 and increased consumption could account for the higher δ13Ccol seen in Dipoides sp. Variation in the quantity and type of macrophytes consumed by each beaver species could also account for this difference (i.e. emergent and floating macrophytes are, on average more 15N-enriched than submerged macrophytes).
    Changes in the isotopic composition of the C and N baseline between the Pliocene and the present could also account for the isotopic offset between beaver species. Further investigation of possible changes in the δ15N baseline of flora in terrestrial high latitude environments during the Pliocene would be a valuable avenue of future research.
    Dipoides sp. behaviour and evolutionary implications
    Evidence from the Beaver Pond site has implications for our understanding of Dipoides sp. ecology. These data also contribute to our understanding of the evolution of behavioural transitions within Castoridae. In particular, how Castor’s distinctive complex of behavioural traits (tree harvesting, underwater food caching, and construction behaviour) may have evolved. A new hypothesis of behavioural evolution in castorids based on evidence from the fossil record (i.e. fossil burrows, cut wood, and stable isotope measurements) and skeletal-dental morphology is mapped onto a simplified phylogenetic tree in Fig. 842,84,85,86,87.
    Figure 8

    Simplified Castoridae phylogeny showing behavioural reconstructions, including new evidence of woody plant consumption in Dipoides sp. Diagram based on phylogenetic analysis by Rybczynski9, which used a matrix of 88 morphological characters and 38 taxa. The origination of dam building is a minimum age (~ 7–8 Ma), corresponding to the time of divergence of Castor canadensis and C. fiber, inferred from molecular evidence96 and supported by fossil evidence97. Legend: CIRCLE—taxa that burrowed (Dipoides and Castoroides may have burrowed, but direct fossil evidence is currently lacking); WP—taxa with significant woody plant contribution to their diet; NWP—taxa that did not generally consume woody plants (the terrestrial burrowing clade is associated with open plains and unforested habitat, and therefore assumed to have not consumed significant amounts of woody plants); Plio—Pliocene; Q—Quaternary. Age range sources: Castor96,97,103, nowdatabase.org; Steneofiber eseri104; Fossorial clade84,86,94,105; Eutypomys94, Fossilworks.org, nowdatabase.org; Dipoides, including D. tanneri: Fossilworks.org, nowdatabase.org; Castoroides106, Fossilworks.org, nowdatabase.org. Fossil taxa behavioural evidence sources: Steneofiber eseri104; Castoroides13; Dipoides (this study); Fossorial clade90.

    Full size image

    Castoridae is a group of herbivorous rodents comprising roughly two dozen genera. Most fossil castorids fall within two major groups: a clade of fossorial specialists (Palaeocastorinae) and a semiaquatic clade42,84,86,88,89,90. The latter includes Castor and Dipoides. Members of the fossorial clade (~ 7 genera) possess striking specializations such large digging claws, extremely reduced tails, and broad, procumbent incisors for digging. In some cases, specimens have been found within fossil burrows (i.e. Palaeocastor, or “The devil’s corkscrew” burrows discovered in the plains of North America88). The semiaquatic group comprises two subfamilies, Castorinae (~ 6 genera, including Steneofiber and the extant Castor), and Castoroidinae (~ 7 genera, including Dipoides and the giant beaver, Castoroides). The oldest definitive Castorinae in the fossil record is Steneofiber eseri from the early Miocene (France, MN2, ~ 23 Ma). S. eseri shows evidence of living in family groups and swimming specializations91. This, in combination with aDNA evidence12, suggests Castorinae and Castoroidinae are derived from a semiaquatic ancestor in the early Miocene.
    Digging behaviour was not just characteristic of the fossorial group and appears within the semiaquatic clade as well. Castor, though not morphologically highly specialized for the task, digs bank burrows and creates extensive canal systems92. In addition, the extinct semiaquatic beaver Steneofiber eseri was found within a burrow91. Considering the phylogenetic distribution of burrowing behaviour within the Castorid tree (Fig. 8), it is likely that the common ancestor of the fossorial and semiaquatic clades also burrowed. Thus, the appearance of burrowing behaviour within Castor and Steneofiber are seen as a retention of a primitive trait9.
    If burrowing behaviour in semiaquatic castorids is the primitive condition, it is likely Dipoides burrowed as well, as seen in other semiaquatic rodents today such as Castor, but also Crossomys (earless water rat), Myocastor (nutria), and Ondatra (muskrat)92. It is also possible that Dipoides constructed lodges. Extant Castor and Ondatra are known to construct burrows and lodges, depending on the characteristics of the habitat. Bank burrows are associated with stream environments, whereas lodges are better suited to calmer waters92. Unlike Castor, extant Ondatra construct their push-up lodges using cattails and other fibrous vegetation rather than wood. The abundance of cut wood at the Beaver Pond site11 suggests that Dipoides sp. had the option to incorporate wood into their nesting structures, and possibly built lodges.
    Given the occurrence of woodcutting and woody plant consumption within both subfamilies of semiaquatic castorids (represented by Castor and Dipoides in Fig. 8), it seems likely these behaviours appeared in the common ancestor of the semiaquatic group. Woody plant consumption may have preadapted castorids to exploit colder environments that arose during and after the late Miocene. Castor canadensis does not hibernate, but builds and sink rafts of branches and foliage to use as a source of fresh food during the winter months1,93. Dipoides sp. may have also engaged in this behaviour and used underwater caches of branches as a primary food source to survive the consecutive months of darkness during the high latitude winter when plants become dormant. The use of woody plants in this way may have been key to allowing beavers to disperse between North American and Eurasia, which required crossing the Bering Isthmus94, a high latitude landmass. Curiously, given that a diet rich in woody plants appears to be the primitive condition of semiaquatic castorids, the absence of woody plant consumption seen in the Pleistocene giant beaver Castoroides13 must be interpreted here as an evolutionary loss and potentially a leading factor in their extinction (Fig. 8).
    Among living mammals, Castor’s dam construction is a unique and highly derived behaviour – an evolutionary puzzle, associated with a set of innate behavioural specializations95. For example, dam construction is well known to be triggered by the sound of running water alone95. The presence of such “hard-wired” behaviours may be associated with the ancient origins of this behaviour. Molecular and fossil occurrence records indicate that the split between Eurasian and North American Castor arose around 7.5 Ma ago96,97, implying that dam building behaviour itself is at least as old.
    Definitive fossil evidence for dam building by an extinct beaver is currently lacking. Consequently, dam building behaviour is shown as possibly arising only on the lineage leading to Castor. Hypothetically, dam building may have arisen from beavers collecting branches near their burrow/lodge for feeding purposes and the accumulations of sticks could have dammed streams by happenstance. The effects may have been multifold. A deeper pond is an effective defense mechanism and provides a safe refuge from predators. Raised water levels also create more favourable conditions for underwater food caching of branches in sub-freezing winter conditions because the deeper water would prevent an underwater food cache from being locked in ice. As such, natural selection would have favoured animals that maintained the dam, presumably as an extension of their pre-existing nesting behaviour such as lodge building. In this scenario, the climate cooling that started around 15 Ma ago and continued into the Pleistocene would have provided an interval where behaviours promoting over-wintering survival, such as underwater food caching branches and dam building, would have been increasingly reinforced by natural selection.
    It seems unlikely that the common ancestor of all semiaquatic beavers was a dam-builder. Extant Castor is a large powerful rodent weighing 12–25 kg, with some individuals as large as 40 kg92. Its body size is one factor that allows the animal to harvest branches and whole trees to build lodges and maintain dams over multiple years. The Beaver Pond site Dipoides sp. was also a large rodent and was roughly two-thirds the size of an average extant Castor. In contrast, the less-derived semiaquatic beavers, such as the Miocene Eucastor tortus (Castoroidinae) and Steneofiber eseri (Castorinae) were small (~ 1 kg, or less), suggesting that the common ancestor of the semiaquatic lineage was also small bodied. Although the common ancestor of the semiaquatic beaver lineage is inferred to have consumed woody plants (this study), and may have used branches in creating food piles and wood for lodge construction, it would have been too small to have had the capacity to build and maintain dams. As such, if Dipoides sp. did exhibit dam building behaviour, it would be the result of parallel evolution within the Castoroidinae and Castorinae lineages. More

  • in

    Contribution to the ecology of the Italian hare (Lepus corsicanus)

    1.
    Angelici, F. M. & Luiselli, L. Distribution and status of the Apennine hare Lepus corsicanus in continental Italy and Sicily. Oryx 35, 245–249 (2001).
    Google Scholar 
    2.
    Trocchi, V. & Riga, F. I lagomorfi in Italia. Linee guida per la conservazione e la gestione (ed. Trocchi, V. & Riga, F) 107–110 (Ministero delle Politiche Agricole e Forestali–Istituto Nazionale Fauna Selvatica, 2005).

    3.
    Lo Valvo, M. Conservazione di Lepus corsicanus De Winton, 1898 e stato delle conoscenze (ed. de Filippo et al.) 89–95 (IGF Publishing, 2007).

    4.
    Mori, E., Menchetti, M., Mazza, G. & Scalisi, M. A new area of occurrence of an endemic Italian hare inferred by camera trapping. Boll. Mus. Civ. Sci. Nat. Torino. 30, 123–130 (2014).
    Google Scholar 

    5.
    Amori, G. & Castiglia, R. Mammal endemism in Italy: A review. Biogeographia 33, 19–31 (2018).
    Google Scholar 

    6.
    De Winton, W. E. On the hares of Western Europe and North Africa. J. Nat. Hist. 1, 149–158 (1898).
    Google Scholar 

    7.
    Miller, G. S. Catalogue of the Mammals of Western Europe (exclusive of Russia) in the Collection of the British Museum 15–1019 (Trustees of the British Museum,1912).

    8.
    Palacios, F. Systematic of the indigenous hares of Italy traditionally identified as Lepus europaeus Pallas, 1778 (Mammalia: Leporidae). Bonn. Zool. Beitr. 46, 59–91 (1996).
    Google Scholar 

    9.
    Riga, F., Trocchi, V., Randi, E. & Toso, S. Morphometric differentiation between the Italian hare (Lepus corsicanus De Winton, 1898) and the European brown hare (Lepus europaeus Pallas, 1778). J. Zool. 253, 241–252 (2001).
    Google Scholar 

    10.
    Pierpaoli, M., Riga, F., Trocchi, V. & Randi, E. Species distinction and evolutionary relationships of the Italian hare (Lepus corsicanus) as described by mitochondrial DNA sequencing. Mol. Ecol. 8, 1805–1817 (1999).
    CAS  PubMed  Google Scholar 

    11.
    Alves, P. C., Ferrand, N., Suchentrunk, F. & Harris, D. J. Ancient introgression of Lepus timidus mtDNA into L. granatensis and L. europaeus in the Iberian Peninsula. Mol. Phylogenet. Evol. 27, 70–80 (2003).
    CAS  PubMed  Google Scholar 

    12.
    Wu, C. et al. Molecular phylogenetics and biogeography of Lepus in Eastern Asia based on mitochondrial DNA sequences. Mol. Phylogenet. Evol. 37, 45–61 (2005).
    CAS  PubMed  Google Scholar 

    13.
    Alves, P. C. et al. Evidence for genetic similarity of two allopatric European hares (Lepus corsicanus and L. castroviejoi) inferred from nuclear DNA sequences. Mol. Phylogenet. Evol. 46, 1191–1197 (2008).
    CAS  PubMed  Google Scholar 

    14.
    Melo-Ferreira, J. et al. Recurrent introgression of mitochondrial DNA among hares (Lepus spp.) revealed by species-tree inference and coalescent simulations. Syst. Biol. 61, 367–381 (2012).
    CAS  PubMed  Google Scholar 

    15.
    Acevedo, P., Melo-Ferreira, J., Real, R. & Alves, P. C. Evidence for niche similarities in the allopatric sister species Lepus castroviejoi and Lepus corsicanus. J. Biogeogr. 41, 977–986 (2014).
    Google Scholar 

    16.
    Palomo, L. J., Gisbert, J. & Blanco, J. C. Atlas y libro rojo de los mamíferos terrestres de España (ed. Palomo, L. J., Gisbert, J. & Blanco, J. C.) 479–483 (Organismo Autónomo de Parques Nacionales, 2007).

    17.
    Randi, E. & Riga, F. Lepus corsicanus. The IUCN Red List of Threatened Species e.T41305A2952954, https://doi.org/10.2305/IUCN.UK.2019-2.RLTS.T41305A2952954.en. (2019).

    18.
    Angelici, F. M., Petrozzi, F. & Galli, A. Conservazione di Lepus corsicanus De Winton, 1898 e stato delle conoscenze (ed. de Filippo et al.) 103–109 (IGF Publishing, 2007).

    19.
    Fulgione, D., Maselli, V., Pavarese, G., Rippa, D. & Rastogi, R. K. Landscape fragmentation and habitat suitability in endangered Italian hare (Lepus corsicanus) and European hare (Lepus europaeus) populations. Eur. J. Wildl. Res. 55, 385–396 (2009).
    Google Scholar 

    20.
    Pierpaoli, M., Riga, F., Trocchi, V. & Randi, E. Hare populations in Europe: intra and interspecific analysis of mtDNA variation. Cr. Biol. 326, 80–84 (2003).
    Google Scholar 

    21.
    Kasapidis, P., Suchentrunk, F., Magoulas, A. & Kotoulas, G. The shaping of mitochondrial DNA phylogeographic patterns of the brown hare (Lepus europaeus) under the combined influence of Late Pleistocene climatic fluctuations and anthropogenic translocations. Mol. Phylogenet. Evol. 34, 55–66 (2005).
    CAS  PubMed  Google Scholar 

    22.
    Angelici, F. M. & Luiselli, L. Body size and altitude partitioning of the hares Lepus europaeus and L. corsicanus living in sympatry and allopatry in Italy. Wildlife Biol. 13, 251–257 (2007).
    Google Scholar 

    23.
    de Filippo, G., Caliendo, M. F., Fulgione, D., Fusco, L. & Troisi, S. R. Conservazione di Lepus corsicanus De Winton, 1898 e stato delle conoscenze (ed. de Filippo et al.) 97–102 (IGF Publishing, 2007).

    24.
    Buglione, M. et al. A pilot study on the application of DNA metabarcoding for non-invasive diet analysis in the Italian hare. Mamm. Biol. 88, 31–42 (2018).
    Google Scholar 

    25.
    Buglione, M. et al. Who is who? High Resolution Melting analysis to discern between hare species using non-invasive sampling. Conserv. Genet. Resour https://doi.org/10.1007/s12686-020-01153-9 (2020).
    Article  Google Scholar 

    26.
    Mengoni, C., Mucci, N. & Randi, E. Genetic diversity and no evidences of recent hybridization in the endemic Italian hare (Lepus corsicanus). Conserv. Genet. 16, 477–489 (2015).
    Google Scholar 

    27.
    Ferreira, C. M. et al. Genetic non-invasive sampling (gNIS) as a cost-effective tool for monitoring elusive small mammals. E. J. Wild. Res. 4, 46 (2018).
    Google Scholar 

    28.
    Taberlet, P., Coissac, E., Pompanon, F., Brochmann, C. & Willerslev, E. Towards next-generation biodiversity assessment using DNA metabarcoding. Mol. Ecol. 21, 2045–2050 (2012).
    CAS  PubMed  Google Scholar 

    29.
    Angelici, F. M., Petrozzi, F. & Galli, A. The Apennine hare Lepus corsicanus in Latium, Central Italy: a habitat suitability model and comparison with its current range. Hystrix 21, 177–182 (2010).
    Google Scholar 

    30.
    Scarselli, D. et al. Coming home: Reintroduction of Italian hares (Lepus corsicanus) in the Elba Island. Mamm. Biol. 81, 15–15 (2016).
    Google Scholar 

    31.
    Mori, E. et al. Safety or satiety? Spatiotemporal behaviour of a threatened herbivore. Mamm. Biol. 100, 49–61 (2020).
    Google Scholar 

    32.
    de Filippo, G., Fulgione, D. Fusco, L. & Troisi, S.R. Conservazione di Lepus corsicanus De Winton, 1898 e stato delle conoscenze (ed. de Filippo et al.) 123–136 (IGF Publishing, 2007).

    33.
    Freschi, P. et al. Diet of the Italian hare (Lepus corsicanus) in a semi-natural landscape of southern Italy. Mammalia 79, 51–59 (2015).
    Google Scholar 

    34.
    Adhikari, D., Barik, S. K. & Upadhaya, K. Habitat distribution modelling for reintroduction of Ilex khasiana Purk., a critically endangered tree species of northeastern India. Ecol. Eng. 40, 37–43 (2012).
    Google Scholar 

    35.
    Gelviz-Gelvez, S. M., Pavón, N. P., Illoldi-Rangel, P. & Ballesteros-Barrera, C. Ecological niche modeling under climate change to select shrubs for ecological restoration in Central Mexico. Ecol. Eng. 74, 302–309 (2015).
    Google Scholar 

    36.
    Begon, M., Townsend, C. R. & Harper, J. L. Ecology: from individuals to ecosystems. (Blackwell Publishing, 2006).

    37.
    Swets, J. A. Measuring the accuracy of diagnostic systems. Science 240, 1285–1293 (1988).
    ADS  MathSciNet  CAS  PubMed  MATH  Google Scholar 

    38.
    Franklin, J. Mapping species distributions: spatial inference and prediction 1–318 (Cambridge University Press, Cambridge, 2010).
    Google Scholar 

    39.
    Pearson, R. G., Raxworthy, C. J., Nakamura, M. & Townsend, P. A. Predicting species distributions from small numbers of occurrence records: A test case using cryptic geckos in Madagascar. J. Biogeogr. 34, 102–117 (2007).
    Google Scholar 

    40.
    Liu, C., Newell, G. & White, M. On the selection of thresholds for predicting species occurrence with presence-only data. Ecol. Evol. 6, 337–348 (2016).
    PubMed  Google Scholar 

    41.
    Tao, Y., Börger, L. & Hastings, A. Dynamic range size analysis of territorial animals: An optimality approach. Am. Nat. 4, 460–474 (2016).
    Google Scholar 

    42.
    Fernando, P., Vidya, T. N. C., Rajapakse, C., Dangolla, A. & Melnick, D. J. Reliable noninvasive genotyping: fantasy or reality?. J. Hered. 94, 115–123 (2003).
    CAS  PubMed  Google Scholar 

    43.
    Valentini, A., Pompanon, F. & Taberlet, P. DNA barcoding for ecologists. Trends Ecol. Evol. 24, 110–117 (2009).
    PubMed  Google Scholar 

    44.
    Ball, M. C. et al. Characterization of target nuclear DNA from faeces reduces technical issues associated with the assumptions of low-quality and quantity template. Conserv. Genet. 8, 577–586 (2007).
    Google Scholar 

    45.
    Brinkman, T. J., Schwartz, M. K., Person, D. K., Pilgrim, K. L. & Hundertmark, K. J. Effects of time and rainfall on PCR success using DNA extracted from deer fecal pellets. Conserv. Genet. 11, 1547–1552 (2010).
    CAS  Google Scholar 

    46.
    Bisi, F. et al. A Space use patterns of mountain hare (Lepus timidus) on the Alps. Eur. J. Wildl. Res. 57, 305–312 (2011).
    Google Scholar 

    47.
    Toschi, A. Mammalia: lagomorpha, rodentia, carnivora, artiodactyla, cetacea 1–648 (Calderini, 1965).

    48.
    Zuccagni-Orlandini, A. Corografia fisica, storica e statistica dell’Italia e delle sue isole, corredata di un atlante, di mappe geografiche e topografiche, e di altre tavole illustrative 1–755 (Firenze, presso gli editori, 1835).

    49.
    Trocchi, V. & Riga, F. Piano d’azione nazionale per la Lepre italica (Lepus corsicanus) (ed. Trocchi, V. & Riga, F.) 1–97 (Ministero delle Politiche Agricole e Forestali–Istituto Nazionale Fauna Selvatica, 2001).

    50.
    Pietri, C. Range and status of the Italian hare Lepus corsicanus in Corsica. Hystrix 26, 1–3 (2015).
    Google Scholar 

    51.
    Deagle, B. E., Kirkwood, R. & Jarman, S. N. Analysis of Australian fur seal diet by pyrosequencing prey DNA in faeces. Mol. Ecol. 18, 2022–2038 (2009).
    CAS  PubMed  Google Scholar 

    52.
    Ando, H. et al. Diet analysis by next-generation sequencing indicates the frequent consumption of introduced plants by the critically endangered red-headed wood pigeon (Columba janthina nitens) in oceanic island habitats. Ecol. Evol. 3, 4057–4069 (2013).
    PubMed  PubMed Central  Google Scholar 

    53.
    Ingerson-Mahar, J. M. Relating diet and morphology of the head, mandibles and proventriculus in adult carabid beetles. PhD. Thesis, Rutgers University Graduate School New Brunswick (2014).

    54.
    Holechek, J. L., Vavra, M. & Pieper, R. D. Botanical composition of determination of range herbivore diets. A review Grazing animals, forage resources. Rangel. Ecol. Manag. 35, 309–315 (1982).
    Google Scholar 

    55.
    Hirakawa, H. Coprophagy in leporids and other mammalian herbivores. Mamm. Rev. 31, 61–80 (2001).
    Google Scholar 

    56.
    Naumova, E. I., Zharova, G. K., Chistova, T. Y. & Kuznetsova, T. A. The effect of coprophagy on the size of plant fibers in the digestive tract of hares Lepus europaeus and L. timidus (Lagomorpha, Leporidae). Biol. Bull. 42, 426–431 (2015).
    Google Scholar 

    57.
    Sokos, C. et al. Conservation considerations for a management measure: An integrated approach to hare rearing and release. Environ. Manag. 55, 19–30 (2015).
    ADS  Google Scholar 

    58.
    León-Ruiz, E., Alcázar, P., Domínguez-Vilches, E. & Galán, C. Study of Poaceae phenology in a Mediterranean climate. Which species contribute most to airborne pollen counts?. Aerobiologia 27, 37–50 (2011).
    Google Scholar 

    59.
    Rizzardini, G. et al. Feeding preferences in dry season of the Italian hare (Lepus corsicanus) in two sites of Corsica. Eur. J. Wild. Res. 65, 43 (2019).
    Google Scholar 

    60.
    Hanski, I. Patch-occupancy dynamics in fragmented landscapes. Trends Ecol. Evol. 9, 131–135 (1994).
    CAS  PubMed  Google Scholar 

    61.
    McCullough, D. R. Metapopulation and wildlife conservation (Island Press, 1996).

    62.
    Margules, C. R. & Pressey, R. L. Systematic conservation planning. Nature 405, 243 (2000).
    CAS  PubMed  PubMed Central  Google Scholar 

    63.
    Mech, S. G. & Hallett, J. G. Evaluating the effectiveness of corridors: a genetic approach. Conserv. Biol. 15, 467–474 (2001).
    Google Scholar 

    64.
    Fortebraccio, M., Riga, F., Bottaro, R. & Donnini, S. Take it easy: scat detection dog and conservation of Italian hare (Lepus corsicanus) Hystrix 29, 66–67 (2018).

    65.
    Wasser, S. K., Houston, C. S., Koehler, G. M., Cadd, G. G. & Fain, S. R. Techniques for application of faecal DNA methods to field studies of Ursids. Mol. Ecol. 6, 1091–1097 (1997).
    CAS  PubMed  Google Scholar 

    66.
    QGIS Development Team. QGIS Geographic Information System (Open Source Geospatial Foundation Project 2019), https://www.qgis.org.

    67.
    Wittwer, C. T., Reed, G. H., Gundry, C. N., Vandersteen, J. G. & Pryor, R. J. High-resolution genotyping by amplicon melting analysis using LCGreen. Clin. Chem. 49, 853–860 (2003).
    CAS  PubMed  Google Scholar 

    68.
    Lo Valvo, M., Barera, A. & Seminara, S. Biometria e status della Lepre appenninica (Lepus corsicaus, de Winton) in Sicilia. Nat. Sicil. 21, 67–74 (1997).
    Google Scholar 

    69.
    Rondinini, C., Battistoni, A., Peronace, V. & Teofili, C. Lista rossa IUCN dei vertebrati italiani 50 (Comitato Italiano IUCN e Ministero dell’Ambiente e della Tutela del Territorio e del Mare, 2013).

    70.
    Elith, J. et al. Novel methods improve prediction of species’ distributions from occurrence data. Ecography 29, 129–151 (2006).
    Google Scholar 

    71.
    Baldwin, R. A. Use of maximum entropy modeling in wildlife research. Entropy 11, 854–866 (2009).
    ADS  Google Scholar 

    72.
    Monterroso, P., Brito, J. C., Ferreras, P. & Alves, P. C. Spatial ecology of the European wildcat in a Mediterranean ecosystem: Dealing with small radio-tracking datasets in species conservation. J. Zool. 279, 27–35 (2009).
    Google Scholar 

    73.
    Tinoco, B. A., Astudillo, P. X., Latta, S. C. & Graham, C. H. Distribution, ecology and conservation of an endangered Andean hummingbird: The Violet-throated Metaltail (Metallura baroni). Bird. Conserv. Int. 19, 63–76 (2009).
    Google Scholar 

    74.
    Linnell, M. A., Davis, R. J., Lesmeister, D. B. & Swingle, J. K. Conservation and relative habitat suitability for an arboreal mammal associated with old forest. Forest. Ecol. Manag. 402, 1–11 (2017).
    Google Scholar 

    75.
    Ortega-Huerta, M. A. & Peterson, A. T. Modeling ecological niches and predicting geographic distributions: a test of six presence-only methods. Rev. Mex. Biodivers. 79, 205–216 (2008).
    Google Scholar 

    76.
    Phillips, S. J. & Dudík, M. Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography 31, 161–175 (2008).
    Google Scholar 

    77.
    Papeş, M. & Gaubert, P. Modelling ecological niches from low numbers of occurrences: assessment of the conservation status of poorly known viverrids (Mammalia, Carnivora) across two continents. Divers. Distrib. 13, 890–902 (2007).
    Google Scholar 

    78.
    Dormann, C. F. et al. Collinearity: a review of methods to deal with it and a simulation study evaluating their performance. Ecography 36, 27–46 (2013).
    Google Scholar 

    79.
    Elith, J. & Leathwick, J. Predicting species distributions from museum and herbarium records using multiresponse models fitted with multivariate adaptive regression splines. Divers. Distrib. 13, 265–275 (2007).
    Google Scholar 

    80.
    Elith, J., Kearney, M. & Phillips, S. The art of modelling range-shifting species. Methods Ecol. Evol. 1, 330–342 (2010).
    Google Scholar 

    81.
    Doyle, J. J. & Doyle, J. L. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 19, 11–15 (1987).
    Google Scholar 

    82.
    Yokota, Y., Kawata, T., Iida, Y., Kato, A. & Tanifuji, S. Nucleotide sequences of the 5.8 S rRNA gene and internal transcribed spacer regions in carrot and broad bean ribosomal DNA. J. Mol. Evol. 29, 294–301 (1989).
    ADS  CAS  PubMed  Google Scholar 

    83.
    Aceto, S., Caputo, P., Cozzolino, S., Gaudio, L. & Moretti, A. Phylogeny and evolution of orchis and allied genera based on ITS DNA variation: Morphological gaps and molecular continuity. Mol. Phylogenet. Evol. 13, 67–76 (1999).
    CAS  PubMed  Google Scholar 

    84.
    Frediani, M. & Caputo, P. Phylogenetic relationships among annual and perennial species of the genus Cicer as inferred from ITS sequences of nuclear ribosomal DNA. Biol. Plant. 49, 47–52 (2005).
    CAS  Google Scholar 

    85.
    Coissac, E. OligoTag: a program for designing sets of tags for next-generation sequencing of multiplexed samples. Methods Mol. Biol. 888, 13–31 (2012).
    PubMed  Google Scholar 

    86.
    Harris, J. K. et al. Comparison of normalization methods for construction of large, multiplex amplicon pools for next-generation sequencing. Appl. Environ. Microbiol. 76, 3863–3868 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    87.
    D’Amore, R. et al. A comprehensive benchmarking study of protocols and sequencing platforms for 16S rRNA community profiling. BMC Genom. 17, 55 (2016).
    Google Scholar 

    88.
    Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    89.
    Luo, R. et al. SOAPdenovo2: An empirically improved memory-efficient short-read de novo assembler. Gigascience 1, 2047–2217 (2012).
    Google Scholar 

    90.
    Camacho, C. et al. BLAST+: Architecture and applications. BMC Bioinform. 10, 421 (2009).
    Google Scholar 

    91.
    Freitas, T. A. K., Li, P. E., Scholz, M. B. & Chain, P. S. Accurate read-based metagenome characterization using a hierarchical suite of unique signatures. Nucleic Acids Res. 43, 69 (2015).
    Google Scholar 

    92.
    Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760 (2009).
    CAS  PubMed  PubMed Central  Google Scholar 

    93.
    Li, H. et al. The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
    PubMed  PubMed Central  Google Scholar 

    94.
    Lassmann, T., Hayashizaki, Y. & Daub, C. O. SAMStat: monitoring biases in next generation sequencing data. Bioinformatics 27, 130–131 (2011).
    CAS  PubMed  Google Scholar 

    95.
    Mollot, G. et al. Cover cropping alters the diet of arthropods in a banana plantation: a metabarcoding approach. PLoS ONE 9, e93740 (2014).
    ADS  PubMed  PubMed Central  Google Scholar 

    96.
    Hammer, Ø, Harper, D. A. & Ryan, P. D. PAST: Paleontological statistics software package for education and data analysis. Palaeontol. Electron. 4, 9 (2001).
    Google Scholar 

    97.
    Shapiro, S. S. & Wilk, M. B. An analysis of variance test for normality (complete samples). Biometrika 52, 591–611 (1965).
    MathSciNet  MATH  Google Scholar 

    98.
    Kruskal, H. W. & Wallis, A. W. Use of ranks in one-criterion variance analysis. J. Am. Statist. Ass. 47, 583–621 (1952).
    MATH  Google Scholar  More

  • in

    Bi-layered architecture facilitates high strength and ventilation in nest mounds of fungus-farming termites

    Study species and site
    Odontotermes obesus is a fungus-farming termite species21 which makes cathedral-shaped, buttressed mounds9. It is widely distributed in India21 with mounds of several meters in height9. The study was conducted at the Indian Institute of Science Campus in Bangalore, India, which has a residual red soil formed from weathering of gneissic bedrock22. The soil is classified as inorganic clay of low plasticity and contains kaolinite and montmorillonite as dominant clay minerals, and quartz, mica and feldspar as non-clay mineral fractions22. It contains 43%, 34% and 23% of sand, silt and clay-sized fractions respectively22. For all analysis presented in this paper, the outermost region of termite mounds with surface conduits directly in contact with the atmosphere was considered as the buttress and the innermost region farthest away from the mound exterior was considered as the mound core. This was determined visually on a case to case basis based on the architecture of individual termite mounds (see Supplementary Fig. S1 online).
    Strength of termite mound regions
    To understand the scaling of strength with dimensions of mound samples, termite mound slices used by Kandasami et al.11 were obtained and samples were cored out with diameters 2 cm, 2.5 cm and 3.5 cm and standard aspect ratio of 223. These were tested under unconfined compression in a micro Universal Testing Machine (micro UTM) at a displacement of 1 mm/min. The unconfined compressive stress (UCS) for these samples were not significantly different (see Supplementary Fig. S2 online) and were similar to values in Kandasami et al.11 with samples of 6 cm × 3 cm (height: diameter) suggesting no effect of specimen dimension in strength testing; samples of small dimensions could therefore be used for further experiments. This validation was essential since it was not possible to get samples of 6 cm × 3 cm (height: diameter) dimensions from the buttress of the mound due to presence of pits in the mound walls (see Supplementary Fig. S1 online). Samples of 2 cm × 1 cm (height:diameter) were cored out from the core and buttress of the horizontal slices mentioned above. Depending on the availability of samples without channels/tunnels made by termites, 3 to 8 samples were cored out from each location, weighed and their densities were calculated. These samples acted as technical replicates. Samples were oven dried at 50 °C overnight since the mound slices from which they were derived had been stored under laboratory conditions. Samples were then tested under unconfined compression at 1 mm/min displacement and peak compressive stress recorded.
    In order to obtain biological replicates, a drill was attached to a sampling tube (see Supplementary Fig. S3 online), and soil samples were obtained from the core and buttress of occupied mounds (N = 6 mounds). Drilling was carried out at 90 cm and 120 cm from the base (see Supplementary Fig. S3 online). Termites repaired the drilled section within 24 h. This method of sample collection, therefore, ensured minimal damage to the mounds. Samples were carefully transported in zip lock bags to minimise moisture loss, were cored to the dimensions 2 cm × 1 cm (height:diameter), were tested under unconfined compression at 1 mm/min displacement and the peak compressive stress was recorded. The in situ moisture content of soil from the core of the occupied mounds was 6–10% and that for the buttress was 0–4%. Some moisture loss was observed during sample testing, which was attributable to moisture loss occurring during sampling and testing.
    Brazilian test for tensile strength of mound soil
    To determine the tensile strength of termite mound soil, we performed a set of Brazilian or diametral compression tests wherein a disc of diameter 13.70 mm and thickness 6.60 mm24 was subjected to compression (displacement rate = 1 mm/min) under displacement–controlled loading along its diametral plane. Due to the compression load, a tensile stress state develops in the specimen normal to the compressed diameter with peak values near the centre of the specimen (see details in Supplementary, see Supplementary Fig. S4 online). To avoid local failure at compressed ends due to stress concentration, a cushion arc subtending an angle 2α (12°) at the centre of the disc is used to distribute the load uniformly25. With increase in axial displacement, the axial load increases to a peak where a crack initiates near the centre of the specimen and propagates towards the compressed ends instantly. The tensile strength corresponding to this peak load is calculated using σt = 2P/πDt where P is the peak compression load at failure or first drop in the load displacement curve, t is the thickness of the disc and D is the diameter of the disc26.
    The tensile strength was estimated for samples extracted from different cross-sections at varying heights. For each cross section of the mound, several tests were performed (slice A2: n for buttress = 3, n for core = 3; slice A4: n for buttress = 2, n for core = 3; slice A7: n for buttress = 1, n for core = 3). The strength among these tests did not vary significantly (see “Results”).
    Stability analysis of termite mounds
    Slope stability analysis was performed on termite mounds to examine the effect of varying soil density and strength along the radial direction. Two geometrical models, triangular and trapezoidal, of the slope were used in this analysis (see Supplementary Fig. S5 online). The finite element method was used to perform slope stability analysis using a strength reduction factor. The advantage of using finite element-based slope stability analysis is that it does not require any à priori assumption of the failure surface27,28. The termite mound slope was modelled as an axisymmetric domain with an isotropic, homogeneous, linear elastic perfectly plastic Mohr–Coulomb material. The axisymmetric domains were discretized with six noded triangular elements with reduced integration to obtain the global stiffness matrix (see Supplementary Fig. S5 online). Discretization is a prerequisite for performing slope stability analysis using the finite element method. The termite mound was discretized into triangular elements, force balance was performed on each element and the results obtained from individual elements were integrated to obtain the overall slope stability of the mound. The bases of these domains were kept fixed for finite element analysis (soil below the termite mound was not considered). The finite element simulations were performed in Plaxis 2D software. As observed from uniaxial compression test data and density calculations, the strength and density of the mound soil varied in radial directions; to accommodate this variation four sets of model parameters were used (Table S1) for outer buttress, inner buttress, outer core, and inner core (see Supplementary Fig. S5 online). The parameters for inner core and outer buttress were the average of their values along the height of the specimen. For outer core and inner buttress, density and cohesion were linearly interpolated between the inner core and outer buttress. The tensile strength was considered to be constant throughout the domain as obtained in our Brazilian test results using samples from the abandoned mound. Since the soil density is comparable between occupied and abandoned mounds and the cementation is also expected to be the same, the tensile strength is expected to be similar between occupied and abandoned mounds.
    Cohesive strength (c) for slope stability analysis is half of the average uniaxial compressive strength. Friction (ϕ) and dilation (ψ) angle were set to zero as the termite soil is predominantly clayey. The parameters used for slope stability analysis are provided in Table S1 online.
    In the strength reduction factor method, strength parameters were continuously reduced until slope failure occurred. This method involves the reduction of strength by a strength reduction factor in a step-by-step procedure. The factor of safety corresponds to a stable strength reduction factor over a number of successive steps given that the slope failure is achieved in these steps28. A slope failure is identified by a contiguous surface/curve at the plastic limit (or pre-identified failure shear strain) whose end points lie at the boundary of the slope. The strength reduction factor at failure is approximately equal to the factor of safety as defined in limit equilibrium methods29,30.
    Porosity distribution from computed tomography
    X-ray computed tomography (XCT) was performed on samples (of diameter ~ 1.3 mm, aspect ratio of one) extracted from buttress and core at different heights for analysing the distribution of pores within mound soil. From the reconstructed XCT data, the scanned volume was segmented into two phases, air voids and termite mound soil, using thresholds corresponding to air–soil gray-level intensity cutoff. A typical slice of scanned volume data is presented in Fig. S6 (see Supplementary Fig. S6 online) along with a binarized image corresponding to air–soil gray level intensity cutoff. In order to obtain the distribution of porosity from the binarized volume data, a probing cube of 101 voxels (~ 1.3 mm) was traversed along all the interior voxels within the specimen with the cube residing completely in the specimen. The size of the pores within the cube is estimated as

    $$Pore;Size = sqrt[3]{Total ;number;of;voxels;in;the;cube – voxels;occupied;by;soil;in;the;cube}$$

    Porosity of all interior voxels was determined and frequencies of pore sizes were plotted for core and buttress of slices A2, A4 and A7 (Fig. 4). A total of 18 samples were scanned for this analysis (3 samples each for core and buttress within each slice). The pore sizes were divided into 1,000 bins between 0 and 1 mm for plotting. Any attempt towards reduction in the number of bins (say 500, 250, 200, 125, 100, 50, … bins) led to loss of information and statistical significance between core and buttress (see “Statistical analysis”).
    We also calculate the porosity of the whole specimen by the following relation

    $${text{Air}};{text{space}};{text{ratio}};left( {{text{porosity}}} right) = frac{number;of;voxels;in ;pores}{{total ;number;of;voxels;in ;a;sample}}$$

    The porosity measurements for buttress and core at different cross sections are listed in Table 1.
    Air permeability of mound soil
    To understand the functional significance of the differences in density and strength on the gaseous permeability of termite mound samples, one sample each from the core and buttress at different heights from the abandoned mound was examined (see Supplementary Fig. S1 online). Samples were also obtained from the core and buttress of six occupied mounds by drilling at 0.9 and 1.2 m heights from the base of the mounds. Samples of dimensions 2 cm × 1 cm (height:diameter) were cored and inserted inside custom-made glass T-tubes. The samples were sealed inside the tubes with a commercial adhesive. The adhesive was allowed to dry and harden for 24 h before permeability testing. To ensure that all air flow can be attributed to the permeability of the mound samples alone, it was confirmed that the adhesive itself is impermeable to air in the range of air pressures tested. The set-up used for testing the permeability of termite mound soil was modified from King et al.4. The glass T-tubes with the samples were attached to a source of synthetic air (80% N2, 20% O2, 0% RH) and the flow rate was regulated using mass flow controllers Alicat MFC-100 and Alicat MFC-500 in the range 10–100 sccm (standard cubic centimetres per minute) and 100–500 sccm, respectively. The corresponding pressure was measured using a custom-made 14,000 Pa MEMS (micro-electronic measurement sensor) pressure transducer (0.28% full scale error) (see Supplementary Fig. S7 online). Air flow velocity vs. pressure graphs were plotted for samples from occupied and abandoned mounds. The pressures recorded in our experiment fell beyond the full scale error suggesting that they are not due to measurement error and thus reflect a real phenomenon.
    Statistical analysis
    We analysed the data using the software package R version 3.3.3 (2017-03-06). Data were tested for normality using the Shapiro–Wilk test. For the data on the scaling of strength in termite mound soil, Mann–Whitney U tests were performed followed by Bonferroni corrections. For the unconfined compressive strength data obtained from the abandoned mound, no significant interactions were found; therefore, a type II analysis of variance (ANOVA)31 was performed using the model: Compressive Strength ~ Height + Region  by employing the Anova function in the car package where Compressive Strength denotes peak compressive strength for each sample, Height refers to distance of each slice of the termite mound from the base (A2–A7; Fig. 1 and see Supplementary Fig. S1 online), and Region denotes the region within a slice (core vs. buttress; see Supplementary Fig. S1 online). For compressive strength data from the occupied mound, unpaired t tests were performed to check for differences between core and buttress at 90 cm and 120 cm from the base of the mound. Type II analysis of variance (ANOVA) was performed using the model: Tensile Strength ~ Height + Region by employing the Anova function in the car package where Tensile Strength denotes the tensile strength for each sample (see details in “Methods” and Supplementary; see Supplementary Fig. S4 online), Height refers to distance of each slice of the termite mound from the base (A2, A4, A7; see Supplementary Fig. S1 online), and Region denotes the region within a slice (core vs. buttress; see Supplementary Fig. S1 online). Data for porosity distribution in termite mound wall were analysed using a Kolmogorov–Smirnov (KS) test. Pore size distribution of core and buttress were compared for slices A2, A4 and A7 individually. The actual pore size values were compared using unpaired Mann–Whitney U tests for slices A2, A4 and A7 individually. Since the sample sizes in all these cases were very large, random subsamples were also taken and were subjected to unpaired Mann–Whitney U tests; results showed that the difference between core and buttress remained significant even when the sample was reduced to 1/128th of its original size (only results from original sample size and reduction to 1/128th of sample sizes are shown). Any further reduction would not have provided a representative sample. More

  • in

    An early Pangaean vicariance model for synapsid evolution

    1.
    Futuyma, D. J. Evolutionary Biology (Sinauer Associates, Sunderland, 1986).
    Google Scholar 
    2.
    Coyne, J. A. & Orr, H. A. Speciation (Sinauer Associates, Sunderland, 2004).
    Google Scholar 

    3.
    Lieberman, B. S. Paleobiogeography: Using Fossil to Study Global Change, Plate Tectonics, and Evolution (Kluwer Academic/Plenum Publishers, New York, 2000).
    Google Scholar 

    4.
    Abe, F. R. & Lieberman, B. S. The nature of evolutionary radiations: A case study involving Devonian trilobites. Evol. Biol. 36, 225–234 (2009).
    Google Scholar 

    5.
    Nelson, G. & Platnick, N. Systematics and Biogeography (Columbia University Press, New York, 1981).
    Google Scholar 

    6.
    Torsvik, T. H. & Cocks, R. M. Earth History and Palaeogeography (Cambridge University Press, Cambridge, 2017).
    Google Scholar 

    7.
    Golonka, J. Phanerozoic paleoenvironment and paleolithofacies maps. Late Paleozoic. Geologia 33, 145–209 (2007).
    Google Scholar 

    8.
    Torsvik, T. H. et al. Phanerozoic polar wander, palaeogeography and dynamics. Earth Sci. Rev. 114, 325–368 (2012).
    ADS  Google Scholar 

    9.
    Davydov, V. I. & Cózar, P. The formation of the Alleghenian Isthmus triggered the Bashkirian glaciation: Constraints from warm-water benthic foraminifera. Palaeogeogr. Palaeoclimatol. Palaeoecol. 531B, 108403 (2019).
    ADS  Google Scholar 

    10.
    Grachev, A. F. & Nikolaev, V. A. East European platform development in the Late Precambrian and Paleozoic: Structure and sedimentation. Russ. J. Earth Sci. 8, ES4001 (2006).
    Google Scholar 

    11.
    Davydov, V. I. Precaspian Isthmus emergence triggered the Early Sakmarian glaciation: Evidence from the Lower Permian of the Urals, Russia. Palaeogeogr. Palaeoclimatol. Palaeoecol. 511, 403–418 (2018).
    Google Scholar 

    12.
    Global paleogeography and tectonics in Deep Time Series; https://deeptimemaps.com/ (2019).

    13.
    Golonka, J. Phanerozoic palaeoenvironment and palaeolithofacies maps of the Arctic region. Geol. Soc. Lond. Mem. 35, 79–129 (2011).
    Google Scholar 

    14.
    Ziegler, A. M., Hulver, M. L. & Rowley, D. B. Permian world topography and climate. In Late Glacial and Postglacial Environmental Changes: Quaternary, Carboniferous-Permian and Proterozoic (ed. Martini, I. P.) 111–146 (Oxford University Press, Oxford, 1997).
    Google Scholar 

    15.
    Molostovsky, E. A., Molostovskaya, I. I. & Crasquin-Soleau, S. Depositional conditions in the southern Cis-Urals basin during Late Permian (biostratigraphic, lithofacies and petromagnetic data). Geodiversitas 19, 261–278 (1997).
    Google Scholar 

    16.
    Haq, B. U. & Schutter, S. R. A chronology of Paleozoic sea-level changes. Science 322, 64–68 (2008).
    ADS  CAS  PubMed  Google Scholar 

    17.
    Cohen, K. M., Finney, S. C., Gibbard, P. L. & Fan, J.-X. The ICS international chronostratigraphic chart. Episodes 36, 199–204 (2013).
    Google Scholar 

    18.
    Ogg, J. G., Ogg, G. M. & Gradstein, F. M. A Concise Geologic Time Scale 2016. Permian (Elsevier, Amsterdam, 2016).
    Google Scholar 

    19.
    Beauchamp, B. & Grasby, S. E. Permian lysocline shoaling and ocean acidification along NW Pangea led to carbonate eradication and chert expansion. Palaeogeogr. Palaeoclimatol. Palaeoecol. 350–352, 73–90 (2012).
    Google Scholar 

    20.
    Metcalfe, I., Crowley, J. L., Nicoll, R. S. & Schmitz, M. High-precision U-Pb CA-TIMS calibration of Middle Permian to Lower Triassic sequences, mass extinction and extreme climate-change in eastern Australian Gondwana. Gondwana Res. 28, 61–81 (2015).
    ADS  CAS  Google Scholar 

    21.
    Davydov, V. I., Biakov, A. S., Schmitz, M. D. & Silantiev, V. V. Radioisotopic calibration of the Guadalupian (middle Permian) series: Review and updates. Earth Sci. Rev. 176, 222–240 (2018).
    ADS  CAS  Google Scholar 

    22.
    Davydov, V. I. et al. Middle Permian U-Pb zircon ages of the “glacial” deposits of the Atkan Formation, Ayan-Yuryakh anticlinorium, Magadan province, NE Russia: Their significance for global climatic interpretations. Gondwana Res. 38, 74–85 (2016).
    ADS  CAS  Google Scholar 

    23.
    Kossovaya, O. L. Artinskian-Wordian antitropical rugose coral associations: A palaeogeographical approach. Palaeoworld 18, 136–151 (2009).
    Google Scholar 

    24.
    Haig, D. W. et al. Late Artinskian-Early Kungurian (Early Permian) warming and maximum marine flooding in the East Gondwana interior rift, Timor and Western Australia, and comparisons across East Gondwana. Palaeogeogr. Palaeoclimatol. Palaeoecol. 468, 88–121 (2017).
    Google Scholar 

    25.
    Henderson, C. M. Permian conodont biostratigraphy. Geol. Soc. Lond. Spec. Publ. 450, 119–142 (2018).
    ADS  Google Scholar 

    26.
    Kukhtinov, D. A. & Crasquin-Soleau, S. Upper Permian and Triassic of the Precaspian Depression: Stratigraphy and palaeogeography. Geodiversitas 21, 325–346 (1999).
    Google Scholar 

    27.
    Zhuravlev, A. V., Kotlyar, G. V. & Shishlov, S. B. Paleobiogeoraphical and biostratigraphical analysis of the Kazanian (Middle Permian) conodonts of the east Russian Platform—Preliminary results. Permophiles 48, 15–20 (2006).
    Google Scholar 

    28.
    Dercourt, J. et al. Atlas Peri-Tethys, Palaeogeographical Maps (CCGM/CGMW, 2000).

    29.
    Gorsky, V. P., Gusseva, E. A., Crasquin-Soleau, S. & Broutin, J. Stratigraphic data of the Middle–Late Permian on Russian platform. Geobios 36, 533–558 (2003).
    Google Scholar 

    30.
    Molostovskaya, I. I. Stratigraphic correlation of the Upper Permian deposits from the south of the Cis-Ural marginal Trough and the adjacent areas of the Russian Plate. Geodiversitas 19, 247–259 (1997).
    Google Scholar 

    31.
    Forel, M.-B. Heterochronic growth of ostracods (Crustacea) from microbial deposits in the aftermath of the end-Permian extinction. J. Syst. Palaeontol. 13, 315–349 (2015).
    Google Scholar 

    32.
    Golonka, J. et al. Paleogeographic reconstructions and basins development of the Arctic. Mar. Petrol. Geol. 20, 211–248 (2003).
    Google Scholar 

    33.
    Legler, B. & Schneider, J. W. Marine ingressions into the Middle/Late Permian saline lake of the Southern Permian Basin (Rotliegend, Northern Germany) possibly linked to sea level highstands in the Arctic rift system. Palaeogeogr. Palaeoclimatol. Palaeoecol. 267, 102–114 (2008).
    Google Scholar 

    34.
    Chen, B. et al. Permian ice volume and palaeoclimate history; oxygen isotope proxies revisited. Gondwana Res. 24, 77–89 (2013).
    ADS  CAS  Google Scholar 

    35.
    Peryt, T. M. & Peryt, D. Zechstein foraminifera from the Fore-Sudetic monocline area (West Poland) and their paleoecology. Rocznik Polskiego Towarzystwa Geologicznego (Annales de la Societe Geologique de Pologne) 47, 301–326 (1977) ((In Polish with English summary)).
    Google Scholar 

    36.
    Paleobiology Database 1. Search for Bakevellia ceratophaga. https://paleobiodb.org/classic/basicTaxonInfo?taxon_no=131181. Accessed 10 May 2020.

    37.
    Hounslow, M. W. & Balabanov, Y. P. A. A geomagnetic polarity timescale for the Permian, calibrated to stage boundaries. Geol. Soc. Lond. Spec. Publ. 450, 61–103 (2017).
    Google Scholar 

    38.
    Chen, Z. Q., Jin, Y. G. & Shi, G. R. Permian transgression–regression sequences and sea-level changes of South China. Proc. R. Soc. Victoria 110, 345–367 (1998).
    Google Scholar 

    39.
    Wignall, P. B. et al. Facies analysis and sea-level change at the Guadalupian-Lopingian global stratotype (Laibin, South China), and its bearing on the end-Guadalupian mass extinction. J. Geol. Soc. Lond. 166, 655–666 (2009).
    Google Scholar 

    40.
    Kofukuda, D., Isozaki, Y. & Igo, H. A remarkable sea-level drop and relevant biotic responses across the Guadalupian-Lopingian (Permian) boundary in low-latitude mid-Panthalassa: Irreversible changes recorded in accreted paleo-atoll limestones in Akasaka and Ishiyama, Japan. J. Asian Earth Sci. 82, 47–65 (2014).
    ADS  Google Scholar 

    41.
    Stemmerik, L. Sequence stratigraphy of a low productivity carbonate platform succession: The Upper Permian Wegener Halvù Formation, Karstryggen Area, East Greenland. Sedimentology 48, 79–97 (2001).
    ADS  Google Scholar 

    42.
    Brikiatis, L. Arido-eustasy: A new example of non-glacial eustatic sea level change. Gondwana Res. 70, 25–35 (2019).
    ADS  Google Scholar 

    43.
    Arefiev, M. P. et al. Type and Reference Sections of The Permian–Triassic Continental Sequences of the East European Platform: Main Isotope, Magnetic, and Biotic Events (PIN RAS, 2015).

    44.
    Kotlyar, G. V., Golubev, V. K. & Silantiev, V. V. General stratigraphic scale of the Permian marine-continental and continental formations of the East European Platform. In Carboniferous and Permian Earth Systems, Stratigraphic Events, Biotic Evolution, Sedimentary Basins and Resources (ed. Nurgaliev, D. K.) (Kazan Federal University, Kazan, 2014).
    Google Scholar 

    45.
    Gradstein, F. M., Ogg, J. G., Schmitz, M. D. & Ogg, G. M. The Geologic Time Scale 2012 (Elsevier, Amsterdam, 2012).
    Google Scholar 

    46.
    Menning, M. et al. Global time scale and regional stratigraphic reference scales of Central and West Europe, East Europe, Tethys, South China, and North America as used in the Devonian–Carboniferous–Permian Correlation Chart 2003 (DCP 2003). Palaeogeogr. Palaeoclimatolol. Palaeoecol. 240, 318–372 (2006).
    ADS  Google Scholar 

    47.
    van der Walt, M., Day, M., Rubidge, B., Cooper, A. K. & Netterberg, I. A new GIS-based biozone map of the Beaufort Group (Karoo Supergroup), South Africa. Palaeontologia Africana 45, 1–5 (2010).
    Google Scholar 

    48.
    Rubidge, B. S. et al. Advances in nonmarine Karoo biostratigraphy: Significance for understanding basin development. In Origin and Evolution of the Cape Mountains and Karoo Basin (eds Linol, B. & de Wit, M. J.) 141–149 (Springer, Berlin, 2016).
    Google Scholar 

    49.
    Barbolini, N., Rubidge, B. & Bamford, M. K. A new approach to biostratigraphy in the Karoo retroarc foreland system: Utilising restricted-range palynomorphs and their first appearance datums for correlation. J. Afr. Earth Sci. 140, 114–133 (2018).
    ADS  Google Scholar 

    50.
    Day, M. O. Middle Permian continental biodiversity changes as reflected in the Beaufort Group of South Africa: a bio-and lithostratigraphic review of the Eodicynodon, Tapinocephalus, and Pristerognathus assemblage zones (Unpublished PhD Thesis, University of the Witwatersrand, South Africa, 2013); https://hdl.handle.net10539/14014

    51.
    Day, M. O. et al. When and how did the terrestrial mid-Permian mass extinction occur? Evidence from the tetrapod record of the Karoo Basin, South Africa. Proc. R. Soc. Lond. Ser. B 282, 20150834 (2015).
    Google Scholar 

    52.
    Rubidge, B. S., Erwin, D. H., Ramezani, J., Bowring, S. A. & de Klerk, W. J. High-precision temporal calibration of Late Permian vertebrate biostratigraphy: U-Pb zircon constraints from the Karoo Supergroup, South Africa. Geology 41, 363–366 (2013).
    ADS  CAS  Google Scholar 

    53.
    Viglietti, P. A. et al. The Daptocephalus Assemblage Zone (Lopingian), South Africa: A proposed biostratigraphy based on a new compilation of stratigraphic ranges. J. Afr. Earth Sci. 113, 153–164 (2016).
    ADS  Google Scholar 

    54.
    Lucas, S. G. Permian tetrapod biochronology, correlation and evolutionary events. Geol. Soc. Lond. Spec. Publ. 450, 405–444 (2017).
    Google Scholar 

    55.
    Sennikov, A. G. & Golubev, V. K. Sequence of Permian tetrapod faunas of Eastern Europe and the Permian-Triassic ecological crisis. Paleontolog J. 51, 600–611 (2017).
    Google Scholar 

    56.
    Golubev, V. K. Permian tetrapod stratigraphy. New Mexico Mus. Nat. Hist. Sci. Bull. 30, 95–99 (2005).
    Google Scholar 

    57.
    Opluštil, S., Schmitz, M., Cleal, C. J. & Martínek, K. A review of the Middle-Late Pennsylvanian west European regional substages and floral biozones, and their correlation to the Geological Time Scale based on new U-Pb ages. Earth Sci. Rev. 154, 301–335 (2016).
    ADS  Google Scholar 

    58.
    Davydov, V. I., Korn, D. & Schmitz, M. D. The Carboniferous Period in The Geologic Time Scale 2012 (eds. Gradstein, F. M., Ogg, J. G., Schmitz, M. D. & Ogg, G. M.) 603–651 (Elsevier, Amsterdam, 2012).

    59.
    Lucas, S. G. Defining North American Pennsylvanian stages. Newsl. Carbonif. Stratigr. 32, 42–46 (2016).
    Google Scholar 

    60.
    Zhang, Y.-C. & Wang, Y. Permian fusuline biostratigraphy. Geol. Soc. Lond. Spec. Publ. 450, 253–288 (2017).
    Google Scholar 

    61.
    Fritsch, A. Fauna der Gaskohle und der Kalksteine der Permformation Böhmens 1/4 (Selbstverlag, 1883).

    62.
    Stamberg, S. & Zajic, J. Carboniferous and Permian Faunas and their Occurrence in the Limnic Basins of the Czech Republic (Museum of Eastern Bohemia, 2008).

    63.
    Opluštil, S., Martínek, K. & Tasáryová, Z. Facies and architectural analysis of fluvial deposits of the Nýřany Member and the Týnec Formation (Westphalian D–Barruelian) in the Kladno-Rakovník and Pilsen basins. Bull. Geosci. 80, 45–66 (2005).
    Google Scholar 

    64.
    Falcon-Lang, H. J. A history of research at the Joggins Fossil Cliffs of Nova Scotia, Canada, the world’s finest Pennsylvanian section. Proc. Geol. Assoc. 117, 377–392 (2006).
    Google Scholar 

    65.
    Gray, M. & Finkel, Z. V. The Joggins Fossil Cliffs UNESCO World Heritage site: A review of recent research. Atlantic Geol. 47, 185–200 (2011).
    Google Scholar 

    66.
    Falcon-Lang, H. J. Late Carboniferous tropical dryland vegetation in an alluvial-plain setting, Joggins, Nova Scotia, Canada. Palaios 18, 197–211 (2003).
    ADS  Google Scholar 

    67.
    Davies, S. J., Gibling, M. R., Rygel, M. C., Calder, J. H. & Skilliter, D. M. The Pennsylvanian Joggins Formation of Nova Scotia: Sedimentological log and stratigraphic framework of the historic fossil cliffs. Atlantic Geol. 41, 115–142 (2005).
    Google Scholar 

    68.
    Gibling, M. R., Culshaw, N., Rygel, M. C. & Pascucci, V. The Maritimes Basin of Atlantic Canada: Basin creation and destruction in the collisional zone of Pangea. In Sedimentary Basins of the World, Vol .5 (ed. Miall, A. D.) 211–244 (Elsevier, Amsterdam, 2008).
    Google Scholar 

    69.
    Baird, D. New records of Paleozoic diplopod Myriapoda. J. Paleontol. 32, 239–241 (1958).
    Google Scholar 

    70.
    Carroll, R. L. A limnoscelid reptile from the Middle Pennsylvanian. J. Paleontol. 41, 1256–1261 (1967).
    Google Scholar 

    71.
    Gibling, M. R., Saunders, K. I., Tibert, N. E. & White, J. A. Sequence sets, high-accommodation events, and the coal window in the Carboniferous Sydney coalfield, Atlantic Canada. AAPG Stud. Geol. 51, 169–197 (2004).
    Google Scholar 

    72.
    Paleobiology Database 2. Search for Danville Locality. https://paleobiodb.org/classic/displayCollResults?collection_no=80714. Accessed 10 May 2020.

    73.
    Jacobson, R. A summary of references with changes in Pennsylvanian stratigraphic nomenclature at the ISGS since Bulletin 95 (2002). http://www.isgs.illinois.edu/sites/isgs/files/files/coal-maps/Penn_Strat_column_ISGS.pdf. Accessed 10 May 2020.

    74.
    McDowell, R. C. The Geology of Kentucky—A text to accompany the geologic map of Kentucky. U.S. Geological Survey Paper 1151-H (US Government Printing Office, 1986).

    75.
    Cunningham, C. R. et al. The Upper Carboniferous Hamilton Fossil-Lagerstatte in Kansas: A valley-till, tidally influenced deposit. Lethaia 26, 225–236 (1993).
    Google Scholar 

    76.
    Gentile, R. J., Thompson, T. L. & Mulvany, P. S. The Pennsylvanian System of Missouri (A comparison of former and present classifications) (Missouri Department of Natural Resources, Division of Geology and Land Survey, 2004).

    77.
    Modesto, S. P. et al. The oldest parareptile and the early diversification of reptiles. Proc. R. Soc. Lond. Ser. B 282, 20141912 (2015).
    Google Scholar 

    78.
    Sahney, S., Benton, M. J. & Falcon-Lang, H.-J. Rainforest collapse triggered Carboniferous tetrapod diversification in Euramerica. Geology 38, 1079–1082 (2010).
    ADS  Google Scholar 

    79.
    Reisz, R. R., Heaton, M. J. & Pynn, B. R. Vertebrate fauna of Late Pennsylvanian Rock Lake Shale near Garnett, Kansas: Pelycosauria. J. Paleontol. 56, 741–750 (1982).
    Google Scholar 

    80.
    Feldman, H. R. et al. Stratigraphic architecture of the Tonganoxie Paleovalley Fill (Lower Virgilian) in Northeastern Kansas. AAPG Bull. 79, 1019–1043 (1995).
    Google Scholar 

    81.
    Olson, E. C. Early Permian Vertebrates of Oklahoma. Oklahoma Geol. Surv. Circ. 74, 1–111 (1967).
    Google Scholar 

    82.
    Heran, W. D., Green, G. N. & Stoeser, D. B. A Digital Geologic Map Database for the State of Oklahoma (Geological Survey Open-File Report 03-247, 2003).

    83.
    Olmsted, R. W. Geochemical Studies of Uranium in South-Central Oklahoma (Master Thesis, Oklahoma State University, 1975); https://hdl.handle.net/11244/20105

    84.
    Bunn, J. R. Jefferson County, in Oil and Gas in Oklahoma (Oklahoma Geological Survey Bulletin, 40-PP., 1930).

    85.
    Lucas, S. G., Spielmann, J. A. & Krainer, K. Summary of geology of Cañon Del Cobre, Rio Arriba County, New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 49, 15–24 (2010).
    Google Scholar 

    86.
    Tanner, L. H. & Lucas, S. G. Pedogenic record of climate change across the Pennsylvanian-Permian boundary in red-bed strata of the Cutler Group, northern New Mexico, USA. Sediment. Geol. 373, 98–110 (2018).
    ADS  CAS  Google Scholar 

    87.
    Eberth, D. A. & Miall, A. D. Stratigraphy, sedimentology and evolution of a vertebrate-bearing, braided to anastomosed fluvial system, Cutler Formation (Permian-Pennsylvanian), north-central New Mexico. Sediment. Geol. 72, 225–252 (1991).
    ADS  Google Scholar 

    88.
    Lucas, S. G., Harris, S. K., Spielmann, J. A., Berman, D. S. & Henrici, A. C. Vertebrate biostratigraphy and biochronology of the Pennsylvanian-Permian Cutler Group, El Cobre Canyon, Northern New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 31, 128–139 (2005).
    Google Scholar 

    89.
    Huffman, A. C. & Condon, S. M. Stratigraphy, Structure, and Paleogeography of Pennsylvanian and Permian rocks, San Juan Basin and Adjacent Areas, Utah, Colorado, Arizona, and New Mexico (U.S. Geological Survey, Bulletin 1808, 1993).

    90.
    Catuneanu, O. Principles of Sequence Stratigraphy (Elsevier, Amsterdam, 2006).
    Google Scholar 

    91.
    Gibling, M. R., Nanson, G. C. & Maroulis, J. C. Anastomosing river sedimentation in the Channel Country of central Australia. Sedimentology 45, 595–619 (1998).
    ADS  CAS  Google Scholar 

    92.
    Sames, B., Wagreich, M., Conrad, C. P. & Iqbal, S. Aquifer-eustasy as the main driver of short-term sea-level fluctuations during Cretaceous hothouse climate phases. Geol. Soc. Lond. Spec. Publ. 498, 9–38 (2020).
    ADS  Google Scholar 

    93.
    Wendler, J. E., Wendler, I., Vogt, C. & Kuss, J. Link between cyclic eustatic sea-level change and continental weathering: Evidence for aquifer-eustasy in the Cretaceous. Palaeogeogr. Palaeoclimatolol. Palaeoecol. 441, 430–437 (2016).
    ADS  Google Scholar 

    94.
    Giles, J. M., Soreghan, M. J., Benison, K. C., Soreghan, G. S. & Hasiotis, S. T. Lakes, loess, and paleosols in the Permian Wellington Formation of Oklahoma, U.S.A.: Implications for paleoclimate and paleogeography of the midcontinent. J. Sediment. Res. 83, 825–846 (2013).
    ADS  CAS  Google Scholar 

    95.
    Berman, D. S., Eberth, D. A. & Brinkman, D. B. Stegotretus agyrus a new genus and species of microsaur (amphibian) from the Permo-Pennsylvanian of New Mexico. Ann. Carnegie Mus. 57, 293–323 (1988).
    Google Scholar 

    96.
    Harris, S. K., Lucas, S. G., Berman, D. S. & Henrici, A. C. Vertebrate fossil assemblage from the Upper Pennsylvanian Red Tanks Member of the Bursum Formation, Lucero Uplift, Central New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 25, 267–284 (2004).
    Google Scholar 

    97.
    Lucas, S. G., Barrick, J. E., Krainer, K. & Schneider, J. W. The Carboniferous-Permian boundary at Carrizo Arroyo, Central New Mexico, USA. Stratigraphy 10, 153–170 (2013).
    Google Scholar 

    98.
    Lucas, S. G. & Krainer, K. The Red Tanks Member of the Bursum Formation in the Lucero Uplift and regional stratigraphy of the Bursum Formation in New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 25, 43–51 (2004).
    Google Scholar 

    99.
    Hodnett, J.-P.M. & Lucas, S. G. Paleoichthyological assemblages of the Upper Carboniferous-Lower Permian of Socorro County, New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 77, 133–138 (2017).
    Google Scholar 

    100.
    Krainer, K. & Lucas, S. G. The Pennsylvanian-Permian Bursum Formation in Central New Mexico. New Mexico Mus. Nat. Hist. Sci. Bull. 59, 143–160 (2013).
    Google Scholar 

    101.
    Lucas, S. G. et al. Progress report on correlation of nonmarine and marine Lower Permian strata, New Mexico, USA. Permophiles 61, 10–17 (2015).
    Google Scholar 

    102.
    Hentz, T. F., Ambrose, W. A. & Hamlin, H. S. Upper Pennsylvanian and Lower Permian Shelf-to-Basin Facies Architecture and Trends, Eastern Shelf of the Southern Midland Basin, West Texas (University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 282, 2017).

    103.
    Hentz, T. F. & Brown, L. F. Jr. Geologic Atlas of Texas, Wichita Falls-Lawton Sheet, map scale 1:250,000 (University of Texas at Austin, Bureau of Economic Geology, 1987).
    Google Scholar 

    104.
    Sander, P. M. Early Permian depositional environments and pond bonebeds in central archer County, Texas. Palaeogeogr. Palaeoclimatol. Palaeoecol. 69, 1–21 (1989).
    Google Scholar 

    105.
    Holterhoff, P. F., Walsh, T. R. & Barrick, J. A. Artinskian (Early Permian) conodonts from the Elm Creek Limestone, a heterozoan carbonate sequence on the eastern shelf of the Midland Basin, West Texas, U.S.A. New Mexico Mus. Nat. Hist. Sci. Bull. 60, 109–119 (2013).
    Google Scholar 

    106.
    Lucas, S. G. A global hiatus in the Middle Permian tetrapod fossil record. Stratigraphy 1, 47–64 (2004).
    Google Scholar 

    107.
    Johnson, K. S. Gypsum caves of North Texas and Western Oklahoma in Hypogene Karst of Texas (eds. Stafford, K. W. & Veni, G.) 111–122 (Texas Speleological Survey, TSS Monograph 3, 2018).

    108.
    Olson, E. C. Late Permian Terrestrial Vertebrates, U.S.A. and U.S.S.R. Vol. 52, no. 2 (Transactions of the American Philosophical Society, New Series, 1962).

    109.
    Olson, E. C. New Permian Vertebrates from the Chickasha Formation in Oklahoma. Oklahoma Geol. Sur. Circ. 70, 1–70 (1965).
    Google Scholar 

    110.
    Silantiev, V. V., Kotlyar, G. V., Zorina, S. O., Golubev, V. K. & Liberman, V. B. The geological setting and Permian stratigraphy of the Volga and Kama River regions. In Type and Reference Sections of the Middle and Upper Permian of the Volga and Kama River Regions (eds Nurgaliev, D. K. et al.) 10–23 (Kazan University Press, Kazan, 2015).
    Google Scholar 

    111.
    Lozovsky, V. R. Olson’s gap or Olson’s bridge, that is the question. New Mexico Mus. Nat. Hist. Sci. Bull. 30, 179–184 (2005).
    Google Scholar 

    112.
    Mullakaev, A. & Khasanov, R. Aeolian factor in the formation of the Sheshmian Horizon sands and sandstones in the Permian of the South-Tatar Arch. In Kazan Golovkinsky Stratigraphic Meeting, 2017 (eds. Barclay, M., Nikolaeva, S. & Silantiev, V.) 381–385 (Filodiritto International Proceedings, 2018).

    113.
    Güven, S., Rubidge, B. S. & Abdala, F. Taxonomy of tapinocephalid Dinocephalia from the South African Karoo Basin. Research abstract in the 16th Biennial Conference of the Palaeontological Association of Southern Africa, Cape Town. Palaeontologia Africana 47, 29–59 (2012).
    Google Scholar 

    114.
    Falcon-Lang, H. J. Latest Mid-Pennsylvanian tree-fern forests in retrograding coastal plain deposits, Sydney Mines Formation, Nova Scotia, Canada. J. Geol. Soc. Lond. 163, 81–93 (2006).
    Google Scholar 

    115.
    Cascales-Miñana, B. & Cleal, C. J. The plant fossil record reflects just two great extinction events. Terra Nova 26, 195–200 (2013).
    ADS  Google Scholar 

    116.
    Lynch-Stieglitz, J., Stocker, T. F., Broecker, W. S. & Fairbanks, R. G. The influence of air-sea exchange on the isotopic composition of oceanic carbon: Observations and modelling. Glob. Biogeochem. Cycles 9, 653–665 (1995).
    ADS  CAS  Google Scholar 

    117.
    Bond, D. P. G. et al. An abrupt extinction in the Middle Permian (Capitanian) of the Boreal Realm (Spitsbergen) and its link to anoxia and acidification. GSA Bull. 127, 1411–1421 (2015).
    CAS  Google Scholar 

    118.
    Cao, W., Flament, N., Zahirovic, S., Williams, S. & Müller, R. D. The interplay of dynamic topography and eustasy on continental flooding in the late Paleozoic. Tectonophysics 761, 108–121 (2019).
    ADS  Google Scholar 

    119.
    Davydov, V. Warmwater benthic foraminifera document the Pennsylvanian-Permian warming and cooling events—The record from the Western Pangea tropical shelve. Palaeogeogr. Palaeoclimatol. Palaeoecol. 414, 284–295 (2014).
    Google Scholar 

    120.
    Laurin, M. & de Buffrénil, V. Microstructural features of the femur in early ophiacodontids: A reappraisal of ancestral habitat use and lifestyle of amniotes. C. R. Palevol. 15, 115–127 (2015).
    Google Scholar 

    121.
    Ruta, M., Coates, M. I. & Quicke, D. L. J. Early tetrapod relationships revisited. Biol. Rev. 78, 251–345 (2003).
    PubMed  Google Scholar 

    122.
    Haubold, H. et al. Interpretation of the tetrapod footprints from the Early Pennsylvanian of Alabama in Pennsylvanian Footprints in the Black Warrior Basin of Alabama (eds. Buta, R. J., Rindsberg, A. K. & Kopaska-Merkel, D. C.) 75–112 (Alabama Paleontological Society Monograph 1, 2005).

    123.
    Brocklehurst, N. & Fröbisch, J. The definition of bioregions in palaeontological studies of diversity and biogeography affects interpretations: Palaeozoic tetrapods as a case study. Front. Earth Sci. 6, 200 (2018).
    ADS  Google Scholar 

    124.
    Dunne, E. M. et al. Diversity change during the rise of tetrapods and the impact of the ‘Carboniferous rainforest collapse’. Proc. R. Soc. Lond. B Biol. Sci. 285, 20172730 (2018).
    Google Scholar 

    125.
    Brocklehurst, N., Dunne, E. M., Cashmore, D. D. & Frӧbisch, J. Physical and environmental drivers of Paleozoic tetrapod dispersal across Pangaea. Nat. Commun. 9, 5216. https://doi.org/10.1038/s41467-018-07623-x (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    126.
    de Queiroz, A. The resurrection of oceanic dispersal in historical biogeography. Trends Ecol. Evol. 20, 68–73 (2005).
    PubMed  Google Scholar 

    127.
    Gardner, J. D., Surya, K. & Organ, C. L. Early tetrapodomorph biogeography: Controlling for fossil record bias in macroevolutionary analyses. C. R. Palevol. 18, 699–709 (2019).
    Google Scholar 

    128.
    Lucas, S. G., Lozovsky, V. R. & Shishkin, M. A. Tetrapod footprints from Early Permian redbeds of the Northern Caucasus, Russia. Ichnos 6, 277–281 (1999).
    Google Scholar 

    129.
    Lucas, S. G. Chinese Fossil Vertebrates (Columbia University Press, New York, 2001).
    Google Scholar 

    130.
    Pardo, J. D., Small, B. J., Milner, A. R. & Huttenlocker, A. K. Carboniferous-Permian climate change constrained early land vertebrate radiations. Nat. Ecol. Evol. 3, 200–206 (2019).
    PubMed  Google Scholar 

    131.
    Benson, R. B. J. Interrelationships of basal synapsids: Cranial and postcranial morphological partitions suggest different topologies. J. Syst. Paleontol. 10, 601–624 (2012).
    Google Scholar 

    132.
    Spindler, F. et al. First arboreal ’pelycosaurs’ (Synapsida: Varanopidae) from the early Permian Chemnitz Fossil Lagerstätte, SE Germany, with a review of varanopid phylogeny. PalZ 92, 315–364 (2018).
    Google Scholar 

    133.
    Romer, A. S. A large ophiacodont pelycosaur from the Pennsylvanian of the Pittsburgh region. Breviora 144, 1–7 (1961).
    Google Scholar 

    134.
    Reisz, R. R. Pelycosaurian reptiles from the middle Pennsylvanian of North America. Bull. Mus. Comp. Zool. 144, 27–61 (1972).
    Google Scholar 

    135.
    Lucas, S. G. & Golubev, V. K. Age and duration of Olson’s Gap, a global hiatus in the Permian tetrapod fossil record. Permophiles 67, 20–23 (2019).
    Google Scholar 

    136.
    Lantink, M. L., Davies, J. H. F. L., Mason, P. R. D., Schaltegger, U. & Hilgen, F. J. Climate control on banded iron formations linked to orbital eccentricity. Nat. Geosci. 12, 369–374 (2019).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    137.
    Bond, D. P. G. et al. The mid-Capitanian (Middle Permian) mass extinction and carbon isotope record of South China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 292, 282–294 (2010).
    Google Scholar 

    138.
    Rey, K. et al. Stable isotope record implicates aridification without warming during the late Capitanian mass extinction. Gondwana Res. 59, 1–8 (2018).
    ADS  CAS  Google Scholar 

    139.
    Stevens, L. G., Hilton, J., Bond, D. P., Glasspool, I. J. & Jardine, P. E. Radiation and extinction patterns in Permian floras from North China as indicators for environmental and climate change. J. Geol. Soc. Lond. 168, 607–619 (2011).
    Google Scholar 

    140.
    Tierney, K. E. Carbon and strontium isotope stratigraphy of the Permian from Nevada and China: Implications from an icehouse to greenhouse transition (PhD Thesis, Ohio University, 2010); https://rave.ohiolink.edu/etdc/view?acc_num=osu1269625662 More

  • in

    Photocatalytic degradation of dissolved organic matter under ZnO-catalyzed artificial sunlight irradiation system

    DOC changes under photocatalytic degradation
    The changes in DOC during photocatalysis are illustrated in Figs. 1–2 and Table S1. DOC removal after 180-min irradiation varied from 16.76 to 60.88% depending on the experimental conditions. All observed degradation trends followed a pseudo-first-order kinetic model (R2 = 0.96–1.00), which has also been reported in other studies to describe the photodegradation of DOM15,16,17. The effects of ZnO dosage, pH level, and presence of inorganic anions on the degradation of DOM are described in detail in the following sub-sections.
    Figure 1

    Effect of ZnO dosage and pH on DOC removal for photocatalytic degradation of DOM: (a) degradation curves, (b) removal %, and (c) degradation rates.

    Full size image

    Figure 2

    Effect of inorganic anions on DOC removal for photocatalytic degradation of DOM with 0.2 g/L ZnO at pH 7: (a) degradation curves, (b) removal %, and (c) degradation rates.

    Full size image

    Effect of ZnO dosage
    The relationship between ZnO dosage and DOC degradation is presented in Fig. 1 and Table S1. When ZnO dosage increased from 0.1 to 0.3 g/L, DOC removal and apparent degradation rate (kapp) were ∆ 21.42% and 2.49-fold higher at pH 4, ∆ 29.64% and 2.53-fold higher at pH 7, and ∆ 27.62% and 2.28-fold higher at pH 10, respectively. It is assumed that more active sites become available with increasing ZnO dosage, thus facilitating the generation of OH and consequently greater DOC removal and a higher degradation rate.
    Effect of pH
    As illustrated in Fig. 1 and Table S1, DOC removal and kapp were highest at a pH of 7 for all ZnO dosage levels. This observation could be explained by the ionization of DOM and the zeta potential (ZP) of ZnO at different pH levels. The acidic functional groups (e.g., –COOH and –OHphenolic) in HA molecules would become more ionized as the aqueous pH increases because pKa,-COOH and pKa,-OH have been reported to be 4.7 and 12.5, respectively (Eqs. 2)20:

    $${text{HOOC}} cdots -{text{HA}}- cdots {text{OH}} to^{ – } {text{OOC}} cdots- {text{HA}} – cdots {text{O}}^{ – } , + , 2{text{H}}^{ + }$$
    (2)

    HAs are negatively charged over a wide pH range (2.0–10.7) 21, while the ZP of ZnO is positive at a pH range of 6.7–9.3 and negative otherwise (the pHZPC of ZnO = 9.0 ± 0.3) 22. Thus, at pH 7, where the dominant charges of ZnO and HA oppose each other, the electrostatic attraction between HA molecules and the ZnO surface would lead to the more rapid exposure of the HAs to reactive species (especially ·OH), resulting in the maximum DOC removal and photodegradation rate. On the other hand, at a pH of 4 and 10, both the HAs and ZnO are negatively charged, thus the repulsive force between the HA molecules and the ZnO surface would be strong. Therefore, there would be limited opportunities for the HA molecules to contact with reactive species near the ZnO surface, reducing removal and kapp.
    In addition, it was found in the present study that DOC removal was always higher at pH 10 than at pH 4 for the same ZnO dosage. It has been reported that ZnO aggregation occurs at pH 4 (pHZPC of ZnO = 9.0 ± 0.3)22, which would slow the mass transport rate and consequently reduce the active surface area of ZnO. In addition, acidic conditions (i.e., less OH−) are less favorable for the formation of ·OH via the hole oxidation of OH−, lowering the efficiency of the attack of ·OH on DOM and the photocatalytic oxidation rate15, 16. These two reasons lead to lower total DOC removal and a lower photodegradation rate at a pH of 4.
    Effect of inorganic anions
    Figure 2 and Table S1 show the effects of inorganic ions on DOC removal. When fitting a pseudo-first-order kinetics model (R2 = 0.96–0.99), the presence of Cl−, SO42−, and HCO3− anions inhibited DOC removal. This occurred possibly because of two reasons. First, the ZnO surface is positively charged at pH 7, and the anions can be easily adsorbed onto the positively charged surface of the catalyst by electrostatic attraction, leading to the competitive adsorption. Second, the anions acted as free radical scavengers by reducing the availability of positive holes and by competitively reacting with ·OH6, 23,24,25, as given by the following reactions (Eq. 3–10) (Table S2):

    $${text{Cl}}^{ – } + {text{h}}_{{{text{VB}}}}^{ + } to {text{ Cl}}^{ cdot }$$
    (3)

    $${text{Cl}}^{ – } + { }^{ cdot } {text{OH}} to {text{ HOCl}}^{ cdot – } ;;;;;;;;;{ }left( {k = 4.3 times 10^{9} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (4)

    $${text{HOCl}}^{ cdot – } to {text{Cl}}^{ – } + { }^{ cdot } {text{OH }};;;;;;;; left( {k = 6.1 times 10^{9} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (5)

    $${text{SO}}_{4}^{2 – } + {text{ h}}_{{{text{VB}}}}^{ + } { } to {text{ SO}}_{4}^{ cdot – }$$
    (6)

    $${text{SO}}_{4}^{2 – } +^{ cdot } {text{OH}} to {text{ SO}}_{4}^{ cdot – } + {text{ OH}}^{ – } { }left( {k = 1.18 times 10^{6} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (7)

    $${text{HCO}}_{3}^{ – } + {text{h}}_{{{text{VB}}}}^{ + } to {text{CO}}_{3}^{ cdot – } + {text{ H}}_{2} {text{O}}$$
    (8)

    $${text{HCO}}_{3}^{ – } +^{ cdot } {text{OH}} to {text{ CO}}_{3}^{ cdot – } + {text{ H}}_{2} {text{O }}left( {k = 8.5 times 10^{6} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (9)

    The strength of the inhibition effect followed the order of HCO3−  > SO42−  > Cl−  > no ions, possibly because HCO3− had the strongest capturing effect on ·OH (k = 8.5 × 106)25. The HCO3− quenched the ({h}_{VB}^{+}), which prevented the generation of ·OH (i.e., it inhibited the ({h}_{VB}^{+}) + H2O → ·OH + H+ reaction) and may, in turn, have led to the formation of ({CO}_{3}^{cdot-}) via the oxidation of ({HCO}_{3}^{cdot-}) by ({h}_{VB}^{+}) (Eq. 8), with a lower reactivity (E° = 1.78 V) than ·OH26. ({CO}_{3}^{cdot-}) has a weaker oxidative ability than ·OH and rarely reacts with organic matter, thus decreasing the reaction rate significantly6. In addition, HCO3- anions form a strong combination on the surface of the catalyst and can significantly inhibit the adsorption of HAs on the catalyst due to the weak absorption competition between HCO3− and HAs6.
    Change in UV254
    The degradation process as measured using UV254 is summarized in Figs. S2, S3 and Table S1. All photodegradation rates fit a pseudo-first-order kinetics model (R2 = 0.93–1.00), and there was a strong correlation between UV254 and DOC (R2 = 0.92–0.98), suggesting that chromophoric DOM accounted for the most significant proportion of DOC removal16. The effects of ZnO dosage, pH level (Fig. S2), and the presence of inorganic anions (Fig. S3) on UV254 removal and photodegradation rate were analogous to those described in the previous section. The highest in UV254 removal was 96.54% after 180 min of irradiation with a ZnO dosage of 0.3 g/L, a pH of 7, and no additional inorganic anions.
    Total removal and the photodegradation rate calculated based on UV254 were much higher than those calculated using DOC concentration under all experimental conditions, which may be because the terminal functional groups of the aromatic compounds (e.g., hydroxyl and carboxyl) reinforced the adoption affinity of the surface of the catalyst particles15,16,17 and/or some of the DOM chromophores were partially transformed into non-UV-absorbing compounds (e.g., low-molecular-weight organic acids, alcohols, etc.) in the photochemical reaction13, 24.
    The rapid reduction in UV254 with irradiation time (Fig. S2) suggests that the DOM chromophores, which mostly consisted of large aromatic rings, might have been rapidly broken down into smaller non-aromatic structures12, 25. The UV/Vis absorption spectra of DOM showed, as expected, rapid decrease with reaction time, and the remained absorption in UV range, even after 180-min irradiation implies the necessity of experimental optimizations (such as reaction time, power of light source, dosage of catalyst, etc.) for complete mineralization.
    Change in SUVA254
    Figure 3 presents the changes in SUVA254 during irradiation. Initially, the SUVA254 values were all higher than 4, ranging from 4.37 to 4.98, indicating that the organic matter was primarily composed of hydrophobic compounds with high molecular weights (HMWs)9, 26. There was a substantial reduction in SUVA254 (over 90% of initial values) after 180 min of irradiation in most of the samples except for two (pH 4 and pH 10 with 0.1 g/L ZnO). This was because of the preferential removal of aromatic chromophores over aliphatic moieties, followed by the transition of the DOM to non- or less-UV-absorbing substances28. This reduction in SUVA254 also indicates that HMW DOM was rapidly decomposed into organic compounds of lower molecular weight (LMW), which is supported by the lower DOC removal values compared to UV254-measured removal for the same reaction time27. A strong linear correlation (R2 = 0.92–0.98) was found between SUVA254 and DOC, and similar effects of ZnO dosage, pH level, and the presence of inorganic anions were observed.
    Figure 3

    Changes of SUVA254 of DOM during photocatalysis under different ZnO dosages and pHs: (a) SUVA254 and (b) a total reduction (%).

    Full size image

    Change in EEMs
    The changes in the EEMs of the DOM over 180 min of irradiation under optimal conditions (0.2 g/L ZnO and pH 7) are presented in Fig. S4. It was observed that the broad and strong peak at emission wavelengths above ~ 350 nm, commonly referred to as the humic-like peak, decreased significantly with increasing irradiation time. After 180 min, the fluorescence intensity for the measured wavelengths was almost zero, with no clear peaks.
    The lower fluorescence intensity in the EEM plots of the DOM was likely due to the preferential photocatalytic degradation of the HMW fraction27, which led to an increase in the LMW fraction. This was supported by size-exclusion chromatography using DOC and UV254 detection, which also observed a reduction in fluorescence intensity with lower molecular weights based on the synchronous scan spectra of Aldrich HA fractions obtained with ultrafiltration after photocatalysis27. Moreover, the photocatalytic degradation of the HMW compounds in the DOM was similar to the previously reported photocatalytic degradation of NOM from a bog lake29. The photocatalytic degradation of DOM followed a similar sequence to other oxidation processes, such as the chlorination of NOM30 and the photocatalytic degradation of commercial HA using TiO2 and a solar UV-light simulator27.
    Behavior of the components during photocatalysis
    EEM-PARAFAC components
    Using 125 EEM samples from 25 experiments, two components (C1 and C2) were identified using PARAFAC modeling (Fig. S5 and Table S3). It was considered reasonable to extract two fluorophores from the samples because Sigma-Aldrich HA is known to be pedogenic with quite uniform sources31. C1 produced a maximum peak at an Ex/Em of 261 nm/ ≥ 500 nm, exhibiting a broad excitation spectrum and gradual emission above 350 nm, while C2 peaked at an Ex/Em of  pH 10  > pH 4 with 0.1 g/L ZnO and the order pH 10  > pH 7  > pH 4 with 0.2 g/L ZnO. With a ZnO dosage of 0.3 g/L, total removal followed the order pH 10  > pH 7  > pH 4, while the photodegradation rate followed the order pH 7  > pH 4  > pH 10 (Fig. 4). A difference from previous results was also observed for the addition of inorganic anions. In the presence of inorganic anions, total Fmax removal and the photodegradation rate followed the order Cl−  > no anions  > SO42−  > HCO3−. The highest total Fmax removal was 99.36% after 180 min irradiation at a ZnO dosage of 0.3 g/L and a pH of 10, with no additional inorganic anions (Fig. 5).
    Because the degradation behavior of both PARAFAC components followed a first-order exponential decay process, their photocatalytic degradation and kinetic rates could be directly compared. Total Fmax removal and the photodegradation rate of C1 were higher than those of C2, which can be explained by the excitation and emission wavelengths of each component. Although both C1 and C2 were both identified as terrestrial humic-like organic matter, C1 represents a combination of peak A and peak C, exhibiting longer excitation and emission wavelengths than C2. With peaks at longer wavelengths, C1 may be associated with the structural condensation and polymerization of DOM15, 32. Indeed, more pronounced fluorescence at longer emission wavelengths in the EEMs of larger sized and/or more hydrophobic DOM fractions has been previously reported33. Therefore, the results indicate the preferential adsorption of more hydrophobic and larger DOM molecules onto minerals and/or nanoparticles, which has also been reported in previous studies15, 34. In addition, because C2 has shorter excitation wavelengths than C1, it would be less excited by visible light than C1.
    Total DOM removal and photodegradation rates calculated using DOC, UV254, and the PARAFAC components were also compared (Fig. 6). It was interesting to observe that the total removal and photodegradation rates calculated using the Fmax of the two PARAFAC components were higher than those calculated using DOC and UV254. In particular, under optimal conditions (0.2 g/L ZnO, pH 7, and no inorganic anions), the total removal of C1 (100%) and C2 (98.97%) was observed to be higher than total UV254 removal (95.54%) and much higher than total DOC removal (43.04%), while the photodegradation rate of C1 was 11.27-fold and 8.55-fold higher than the photodegradation rates calculated with DOC and UV254, respectively. Similarly, the photodegradation rate of C2 was 1.90-fold and 1.44-fold higher than the photodegradation rates calculated with DOC and UV254, respectively. The more rapid degradation of fluorescence components compared to UV-absorbing moieties (i.e., UV254) could be explained by the fluorescence arising from the π*–π transitions in DOM molecules and its rapid extinction under UV irradiation15, 16.
    Figure 6

    Changes in DOC, UV254, and two EEM-PARAFAC components during photocatalytic degradation of DOM with 0.2 g/L ZnO at pH 7 under artificial sunlight: (a) degradation curves and (b) removal % and degradation rates.

    Full size image

    The proposed reaction mechanism for the ZnO-assisted photocatalytic degradation of DOM under artificial sunlight is presented in Fig. S6. When ZnO is irradiated with artificial light containing photonic energy (hv), valence band hole (({h}_{VB}^{+})) and conduction band electron (({e}_{CB}^{-})) pairs are produced, as given in Eq. (10)11. The ({h}_{VB}^{+}) reacts with H2O and hydroxide ions to yield ·OH (Eqs. 12 and 13)7,8,9. The reduction of dissolved or adsorbed O2 to ({O}_{2}^{cdot-}) by ({e}_{CB}^{_}) is depicted in Eq. (14)7,8,9. The ({O}_{2}^{cdot-}) is converted to H2O2 via disproportionation with protons (Eq. 15) or forms ({HO}_{2}^{cdot}) via protonation, which has a short lifetime due to the rapid reaction with ({O}_{2}^{cdot-}) or ({HO}_{2}^{cdot}) to form the more stable H2O2 (Eqs. 16 and 17)7,8,9, 35. The one-electron reduction of H2O2 produces ·OH (Eq. 19), while H2O2 can also react with ({O}_{2}^{cdot-}) to form ·OH (Eqs. 20 and 21)7,8,9. The generated ·OH is a powerful oxidizing agent that can attack DOM at or near the ZnO surface (Eq. 23). The reaction of ·OH with HAs (as a representative form of DOM) results in the release of LMW acids, amino acids, and ammonia36. The ({O}_{2}^{cdot-}) can also oxidize the DOM molecules (Eq. 24)16. Moreover, upon absorbing light, DOM can act as a photosensitizer in the generation of reactive species such as singlet oxygen (1O2), ·OH, and triplet DOM states (3DOM*), as given in Eqs. (25–28)37. 3DOM* is a potent oxidant of many aquatic contaminants that react with target organic substances directly through electron and energy transfer mechanisms to generate reactive oxygen species such as 1O2, ·OH, and H2O237, thus significantly influencing on the degradation of various fluorophores.

    $$ZnO+hv to {h}_{VB}^{+}+{e}_{CB}^{-}$$
    (10)

    $${H}_{2}Oto {OH}^{-}+{H}^{+}$$
    (11)

    $${text{h}}_{{{text{VB}}}}^{ + } + {text{ H}}_{2} {text{O}} to { }^{ cdot } {text{OH }} + {text{ H}}^{ + }$$
    (12)

    $${text{h}}_{{{text{VB}}}}^{ + } + {text{ OH}}^{ – } to { }^{ cdot } {text{OH}}$$
    (13)

    $${text{e}}_{{{text{CB}}}}^{_} + {text{ O}}_{2} { } to {text{ O}}_{2}^{ cdot – }$$
    (14)

    $${text{O}}_{2}^{ cdot – } + 2{text{H}}^{ + } + {text{e}}_{{{text{CB}}}}^{_} { } to {text{H}}_{2} {text{O}}_{2}$$
    (15)

    $${text{O}}_{2}^{ cdot – } + {text{H}}^{ + } to {text{HO}}_{2}^{ cdot } ;;;;;;left( {k = 2.1 times 10^{10} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (16)

    $${text{HO}}_{2}^{ cdot } + {text{HO}}_{2}^{ cdot } to {text{H}}_{2} {text{O}}_{2} + {text{O}}_{2} ;;;;;;;;left( {k = 8.3 times 10^{5} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (17)

    $$^{ cdot } {text{OH }} +^{ cdot } {text{OH}} to {text{H}}_{2} {text{O}}_{2} ;;;;;;;;;;;;left( {k = 5.5 times 10^{9} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (18)

    $${text{H}}_{2} {text{O}}_{2} + {text{H}}^{ + } + {text{e}}_{{{text{CB}}}}^{_} to { }^{ cdot } {text{OH }} + {text{ H}}_{{2}} {text{O}}$$
    (19)

    $${text{H}}_{2} {text{O}}_{2} + {text{O}}_{2}^{ cdot – } to {text{OH}}^{ – } + {text{O}}_{2} + ^{ cdot } {text{OH}} ;;;;;;;left( {k = 0.13{text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (20)

    $${text{H}}_{2} {text{O}}_{2} + {text{hv}} to {2}^{ cdot } {text{OH}}$$
    (21)

    $${text{H}}_{2} {text{O}}_{2} +^{ cdot } {text{OH}} to {text{H}}_{{2}} {text{O}};;;;;;;left( {k = 2.7 times 10^{7} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (22)

    $${text{DOM }} +^{ cdot } {text{OH}} to {text{CO}}_{{2}} + {text{ H}}_{{2}} {text{O }} + {text{ Products}};;;;;;;;left( {k = 1.7 times 10^{8} {text{ Mc}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (23)

    $${text{DOM }} + {text{O}}_{2}^{ cdot – } to {text{CO}}_{2} + {text{ H}}_{{2}} {text{O }} + {text{ Products}}$$
    (24)

    $${text{DOM }} + {text{ hv}} to^{{1}} {text{DOM}}^{*} to^{{3}} {text{DOM}}^{*}$$
    (25)

    $$^{{3}} {text{DOM}}^{*} + {text{ O}}_{{2}} to^{{1}} {text{DOM }} +^{{1}} {text{O}}_{{2}}$$
    (26)

    $${text{DOM}} + {text{ O}}_{2}^{ cdot – } to {text{DOM}}^{ cdot – } + {text{O}}_{2}$$
    (27)

    $$2{text{O}}_{2}^{ cdot – } + 2{text{H}}^{ + } { } to {text{ H}}_{2} {text{O}}_{2} + {text{ O}}_{2} ;;;;;;;;;;left( {k = 4.0 times 10^{4} {text{ M}}^{ – 1} {text{s}}^{ – 1} } right)$$
    (28)

    The effects of photolysis, adsorption, and photocatalysis on the degradation of DOM was also assessed. Figure 7 compares the results for total removal and the degradation rate for these three processes. The total DOM removal and degradation rate calculated with DOC are illustrated in Fig. 7a. After 180-min irradiation, total DOM removal was 2.92% for photolysis, 10.15% for adsorption, and 43.04% for photocatalysis. The photocatalytic rate was 18.6-fold and 5.5-fold higher than that for photolysis and adsorption. The DOM removal and degradation rate calculated with UV254 (Fig. 7b) also revealed that photocatalysis was more effective than the other two processes. Specifically, total DOM removal was 4.03% for photolysis, 19.46% for adsorption, and 95.45% for photocatalysis, while the degradation rate was 93.6-fold and 17.0-fold higher than that of photolysis and adsorption, respectively. Based on these results, we can conclude that adsorption by ZnO only and photolysis only play a minor role in DOM removal, while the synergistic effects of photocatalysis are vital to this process.
    Figure 7

    Comparison of photolysis, photocatalysis and adsorption of DOM with 0.2 g/L ZnO at pH 7: (a) DOC concentrations and (b) UV254 values.

    Full size image More

  • in

    Climate change, tropical fisheries and prospects for sustainable development

    1.
    Dyck, A. J. & Sumaila, U. R. Economic impact of ocean fish populations in the global fishery. J. Bioeconomics 12, 227–243 (2010).
    Google Scholar 
    2.
    Golden, C. D. et al. Nutrition: fall in fish catch threatens human health. Nature 534, 317–320 (2016).
    Google Scholar 

    3.
    Teh, L. C. L. & Pauly, D. Who brings in the fish? The relative contribution of small-scale and industrial fisheries to food security in Southeast Asia. Front. Mar. Sci. 5, 44 (2018).
    Google Scholar 

    4.
    Gillett, R. Fisheries in the Economies of Pacific Island Countries and Territories (Pacific Community, 2016).

    5.
    Kawarazuka, N. & Béné, C. Linking small-scale fisheries and aquaculture to household nutritional security: an overview. Food Secur. 2, 343–357 (2010).
    Google Scholar 

    6.
    Sale, P. F. et al. Transforming management of tropical coastal seas to cope with challenges of the 21st century. Mar. Pollut. Bull. 85, 8–23 (2014).
    Google Scholar 

    7.
    Bell, J. D. et al. Planning the use of fish for food security in the Pacific. Mar. Policy 33, 64–76 (2009).
    Google Scholar 

    8.
    Hicks, C. C. et al. Harnessing global fisheries to tackle micronutrient deficiencies. Nature 574, 95–98 (2019).
    Google Scholar 

    9.
    Kennedy, G., Nantel, G. & Shetty, P. The scourge of “hidden hunger”: global dimensions of micronutrient deficiencies. Food Nutr. Agric. 32, 8–16 (2003).
    Google Scholar 

    10.
    Teh, L. S. L., Teh, L. C. L. & Sumaila, U. R. Quantifying the overlooked socio-economic contribution of small-scale fisheries in Sabah, Malaysia. Fish. Res. 110, 450–458 (2011).
    Google Scholar 

    11.
    Pauly, D. & Zeller, D. Sea Around Us Concepts, Design and Data. Sea Around Us http://www.seaaroundus.org (2015).

    12.
    Béné, C. Small-scale fisheries: assessing their contribution to rural livelihoods in developing countries. FAO Fish. Circ. 1008, 46 (2006).
    Google Scholar 

    13.
    Williams, P. & Reid, C. Overview of tuna fisheries in the western and central Pacific Ocean, including economic conditions-2017. WCPFC Sci. Comm. SC14-2018/GN-WP-01 66pp (2018).

    14.
    Pacific Islands Forum Fisheries Agency (FFA). Tuna Development Indicators 2016. https://ffa.int/system/files/FFA Tuna Development Indicators Brochure.pdf (Pacific Islands Forum Fisheries Agency, 2017).

    15.
    Teh, L. C. L. & Sumaila, U. R. Contribution of marine fisheries to worldwide employment. Fish Fish. 14, 77–88 (2013).
    Google Scholar 

    16.
    Jentoft, S. Life above water—essays on human experiences of small-scale fisheries. TBTI Global Book Series 1 (2019).

    17.
    Kurien, J. SSF guidelines: the beauty of the small. Samudra Rep. 72, 30–36 (2016).
    Google Scholar 

    18.
    Teh, L. S. L., Teh, L. C. L. & Sumaila, U. R. A global estimate of the number of coral reef fishers. PLoS One 8, e65397 (2013).
    Google Scholar 

    19.
    Alberti, M. et al. Research on coupled human and natural systems (CHANS): approach, challenges, and strategies. Bull. Ecol. Soc. Am. 92, 218–228 (2011).
    Google Scholar 

    20.
    Liu, J., Hull, V., Luo, J., Yang, W. & Liu, W. Multiple telecouplings and their complex interrelationships. Ecol. Soc. 20, 44 (2015).
    Google Scholar 

    21.
    Cinner, J. & McClanahan, T. R. Socioeconomic factors that lead to overfishing in small-scale coral reef fisheries of Papua New Guinea. Environ. Conserv. 33, 73–80 (2006).
    Google Scholar 

    22.
    McClanahan, T. R., Hicks, C. C. & Darling, E. S. Malthusian overfishing and efforts to overcome it on Kenyan coral reefs. Ecol. Appl. 18, 1516–1529 (2008).
    Google Scholar 

    23.
    Gardner, T. A., Côté, I. M., Gill, J. A., Grant, A. & Watkinson, A. R. Long-term region-wide declines in Caribbean corals. Science 301, 958–960 (2003).
    Google Scholar 

    24.
    Islam, M. S. & Tanaka, M. Impacts of pollution on coastal and marine ecosystems including coastal and marine fisheries and approach for management: a review and synthesis. Mar. Pollut. Bull. 48, 624–649 (2004).
    Google Scholar 

    25.
    Hodgson, G. & Dixon, J. A. Logging Versus Fisheries and Tourism in Palawan: An Environmental and Economic Analysis (East-West Environment and Policy Institute, 1988).

    26.
    Hodgson, G. & Dixon, J. A. in Resources & Environment in Asia’s Marine Sector (ed. Marsh, J. B.) 421–446 (CRC, 1992).

    27.
    Côté, I. M., Green, S. J. & Hixon, M. A. Predatory fish invaders: insights from Indo-Pacific lionfish in the western Atlantic and Caribbean. Biol. Conserv. 164, 50–61 (2013).
    Google Scholar 

    28.
    Lehodey, P., Senina, I., Calmettes, B., Hampton, J. & Nicol, S. Modelling the impact of climate change on Pacific skipjack tuna population and fisheries. Clim. Change 119, 95–109 (2013).
    Google Scholar 

    29.
    Asch, R. G., Cheung, W. W. L. & Reygondeau, G. Future marine ecosystem drivers, biodiversity, and fisheries maximum catch potential in Pacific Island countries and territories under climate change. Mar. Policy 88, 285–294 (2018).
    Google Scholar 

    30.
    Jones, M. C. & Cheung, W. W. L. Multi-model ensemble projections of climate change effects on global marine biodiversity. ICES J. Mar. Sci. 72, 741–752 (2015).
    Google Scholar 

    31.
    Lam, V. W. Y., Cheung, W. W. L., Reygondeau, G. & Sumaila, U. R. Projected change in global fisheries revenues under climate change. Sci. Rep. 6, 32607 (2016).
    Google Scholar 

    32.
    Cinner, J. E. et al. Building adaptive capacity to climate change in tropical coastal communities. Nat. Clim. Change 8, 117–123 (2018).
    Google Scholar 

    33.
    Barange, M. et al. Impacts of Climate Change on Fisheries and Aquaculture. Synthesis of Current Knowledge, Adaptation and Mitigation Options (Food and Agriculture Organization of the United Nations, 2018).

    34.
    Pörtner, H.-O. et al. IPCC Special Report on the Ocean and Cryosphere in a Changing Climate (Intergovernmental Panel on Climate Change (IPCC), 2019).

    35.
    Bindoff, N. L. et al. Detection and Attribution of Climate Change: From Global to Regional (Intergovernmental Panel on Climate Change (IPCC), 2013).

    36.
    Bindoff, N. L., Cheung, W. W. L. & Kairo, J. G. in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate Ch. 5 (eds Pörtner, H.-O. et al.) (Intergovernmental Panel on Climate Change (IPCC), 2019).

    37.
    Rodgers, K. B., Lin, J. & Frölicher, T. L. Emergence of multiple ocean ecosystem drivers in a large ensemble suite with an Earth system model. Biogeosciences 12, 3301–3320 (2015).
    Google Scholar 

    38.
    Frölicher, T. L., Rodgers, K. B., Stock, C. A. & Cheung, W. W. L. Sources of uncertainties in 21st century projections of potential ocean ecosystem stressors. Glob. Biogeochem. Cycles 30, 1224–1243 (2016).
    Google Scholar 

    39.
    Pörtner, H.-O. et al. in Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change Ch. 6 (eds Field, C. B. et al.) 411–484 (Cambridge Univ. Press, 2014).

    40.
    Abram, N. et al. in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate Ch. 1 (Intergovernmental Panel on Climate Change (IPCC), 2019).

    41.
    Cheng, L. et al. Record-setting ocean warmth continued in 2019. Adv. Atmos. Sci. 37, 137–142 (2020).
    Google Scholar 

    42.
    Huang, B. et al. Extended reconstructed sea surface temperature version 4 (ERSST. v4). Part I: upgrades and intercomparisons. J. Clim. 28, 911–930 (2015).
    Google Scholar 

    43.
    Cheng, L., Abraham, J., Hausfather, Z. & Trenberth, K. E. How fast are the oceans warming. Science 363, 128–129 (2019).
    Google Scholar 

    44.
    Intergovernmental Panel on Climate Change (IPCC). in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate (ed. Pörtner, H.-O. et al) (Intergovernmental Panel on Climate Change (IPCC), 2019).

    45.
    Xie, S.-P. et al. Global warming pattern formation: Sea surface temperature and rainfall. J. Clim. 23, 966–986 (2010).
    Google Scholar 

    46.
    Hobday, A. J. et al. A hierarchical approach to defining marine heatwaves. Prog. Oceanogr. 141, 227–238 (2016).
    Google Scholar 

    47.
    Frölicher, T. L. & Laufkötter, C. Emerging risks from marine heat waves. Nat. Commun. 9, 650 (2018).
    Google Scholar 

    48.
    Smale, D. A. et al. Marine heatwaves threaten global biodiversity and the provision of ecosystem services. Nat. Clim. Change 9, 306–312 (2019).
    Google Scholar 

    49.
    Hughes, T. P. et al. Global warming and recurrent mass bleaching of corals. Nature 543, 373–377 (2017).
    Google Scholar 

    50.
    Oliver, E. C. J. et al. Longer and more frequent marine heatwaves over the past century. Nat. Commun. 9, 1324 (2018).
    Google Scholar 

    51.
    Holbrook, N. J. et al. Keeping pace with marine heatwaves. Nat. Rev. Earth Environ. https://doi.org/10.1038/s43017-020-0068-4 (2020).
    Article  Google Scholar 

    52.
    Frölicher, T. L., Fischer, E. M. & Gruber, N. Marine heatwaves under global warming. Nature 560, 360–364 (2018).
    Google Scholar 

    53.
    Collins, M. et al. in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate Ch. 6 (eds Pörtner, H.-O. et al.) (Intergovernmental Panel on Climate Change (IPCC), 2019).

    54.
    Cai, W. et al. Increasing frequency of extreme El Niño events due to greenhouse warming. Nat. Clim. Change 4, 111–116 (2014).
    Google Scholar 

    55.
    Capotondi, A., Alexander, M. A., Bond, N. A., Curchitser, E. N. & Scott, J. D. Enhanced upper ocean stratification with climate change in the CMIP3 models. J. Geophys. Res. Oceans 117, C04031 (2012).
    Google Scholar 

    56.
    Ganachaud, A. S. et al. in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change (eds Bell, J. D., Johnson, J. E. & Hobday, A. J.) 101–187 (Secretariat of the Pacific Community, 2011).

    57.
    Stramma, L., Johnson, G. C., Sprintall, J. & Mohrholz, V. Expanding oxygen-minimum zones in the tropical oceans. Science 320, 655–658 (2008).
    Google Scholar 

    58.
    Ito, T., Minobe, S., Long, M. C. & Deutsch, C. Upper ocean O2 trends: 1958–2015. Geophys. Res. Lett. 44, 4214–4223 (2017).
    Google Scholar 

    59.
    Schmidtko, S., Stramma, L. & Visbeck, M. Decline in global oceanic oxygen content during the past five decades. Nature 542, 335–339 (2017).
    Google Scholar 

    60.
    Helm, K. P., Bindoff, N. L. & Church, J. A. Observed decreases in oxygen content of the global ocean. Geophys. Res. Lett. 38, L23602 (2011).
    Google Scholar 

    61.
    Bopp, L. et al. Multiple stressors of ocean ecosystems in the 21st century: projections with CMIP5 models. Biogeosciences 10, 6225–6245 (2013).
    Google Scholar 

    62.
    Cocco, V. et al. Oxygen and indicators of stress for marine life in multi-model global warming projections. Biogeosciences 10, 1849–1868 (2013).
    Google Scholar 

    63.
    Doney, S. C., Fabry, V. J., Feely, R. A. & Kleypas, J. A. Ocean acidification: the other CO2 problem. Annu. Rev. Mar. Sci. 1, 169–192 (2009).
    Google Scholar 

    64.
    Gattuso, J.-P. et al. Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios. Science 349, aac4722 (2015).
    Google Scholar 

    65.
    Burger, F. A., Frölicher, T. L. & John, J. G. Increase in ocean acidity variability and extremes under increasing atmospheric CO2. Biogeosci. Discuss. https://doi.org/10.5194/bg-2020-22 (2020).

    66.
    Oppenheimer, M. et al. in IPCC Special Report on the Ocean and Cryosphere in a Changing Climate Ch. 4 (eds Pörtner, H.-O. et al.) (Intergovernmental Panel on Climate Change (IPCC), 2019).

    67.
    Moon, J. H., Song, Y. T., Bromirski, P. D. & Miller, A. J. Multidecadal regional sea level shifts in the Pacific over 1958–2008. J. Geophys. Res. Oceans 118, 7024–7035 (2013).
    Google Scholar 

    68.
    Han, W. et al. Intensification of decadal and multi-decadal sea level variability in the western tropical Pacific during recent decades. Clim. Dyn. 43, 1357–1379 (2014).
    Google Scholar 

    69.
    Thompson, P. R. & Mitchum, G. T. Coherent sea level variability on the North Atlantic western boundary. J. Geophys. Res. Oceans 119, 5676–5689 (2014).
    Google Scholar 

    70.
    England, M. H. et al. Recent intensification of wind-driven circulation in the Pacific and the ongoing warming hiatus. Nat. Clim. Change 4, 222–227 (2014).
    Google Scholar 

    71.
    Hamlington, B. D. et al. Uncovering an anthropogenic sea-level rise signal in the Pacific Ocean. Nat. Clim. Change 4, 782–785 (2014).
    Google Scholar 

    72.
    McGregor, S. et al. Recent Walker circulation strengthening and Pacific cooling amplified by Atlantic warming. Nat. Clim. Change 4, 888–892 (2014).
    Google Scholar 

    73.
    Le Borgne, R. et al. in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change (eds Bell, J. D., Johnson, J. E. & Hobday, A. J.) 189–249 (Secretariat of the Pacific Community, 2011).

    74.
    Steinacher, M. et al. Projected 21st century decrease in marine productivity: a multi-model analysis. Biogeosciences 7, 979–1005 (2010).
    Google Scholar 

    75.
    Laufkötter, C. et al. Drivers and uncertainties of future global marine primary production in marine ecosystem models. Biogeosciences 12, 6955–6984 (2015).
    Google Scholar 

    76.
    Stock, C. A., Dunne, J. P. & John, J. G. Drivers of trophic amplification of ocean productivity trends in a changing climate. Biogeosciences 11, 7125–7135 (2014).
    Google Scholar 

    77.
    Cheung, W. W. L. & Pauly, D. in Explaining Ocean Warming: Causes, Scale, Effects and Consequences (eds Laffoley D. & Baxter J. M.) 239–253 (IUCN, 2016).

    78.
    Hoegh-Guldberg, O. et al. The Coral Triangle and Climate Change: Ecosystems, People and Societies at Risk (WWF Australia, 2009).

    79.
    Hughes, T. P. et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science 359, 80–83 (2018).
    Google Scholar 

    80.
    Hoegh-Guldberg, O. et al. in Global Warming of 1.5°C (eds Masson-Delmotte, V. et al.) (Intergovernmental Panel on Climate Change (IPCC), 2018).

    81.
    Li, X., Bellerby, R., Craft, C. & Widney, S. E. Coastal wetland loss, consequences, and challenges for restoration. Anthropocene Coasts 1, 1–15 (2018).
    Google Scholar 

    82.
    Pörtner, H.-O. et al. Climate induced temperature effects on growth performance, fecundity and recruitment in marine fish: developing a hypothesis for cause and effect relationships in Atlantic cod (Gadus morhua) and common eelpout (Zoarces viviparus). Cont. Shelf Res. 21, 1975–1997 (2001).
    Google Scholar 

    83.
    Pörtner, H. O. & Farrell, A. P. Physiology and climate change. Science 322, 690–692 (2008).
    Google Scholar 

    84.
    Pauly, D. & Cheung, W. W. L. Sound physiological knowledge and principles in modeling shrinking of fishes under climate change. Glob. Change Biol. 24, e15–e26 (2018).
    Google Scholar 

    85.
    Pinsky, M. L., Worm, B., Fogarty, M. J., Sarmiento, J. L. & Levin, S. A. Marine taxa track local climate velocities. Science 341, 1239–1242 (2013).
    Google Scholar 

    86.
    Poloczanska, E. S. et al. Global imprint of climate change on marine life. Nat. Clim. Change 3, 919–925 (2013).
    Google Scholar 

    87.
    Cheung, W. W. L., Dunne, J., Sarmiento, J. L. & Pauly, D. Integrating ecophysiology and plankton dynamics into projected maximum fisheries catch potential under climate change in the Northeast Atlantic. ICES J. Mar. Sci. 68, 1008–1018 (2011).
    Google Scholar 

    88.
    Perry, A. L., Low, P. J., Ellis, J. R. & Reynolds, J. D. Climate change and distribution shifts in marine fishes. Science 308, 1912–1915 (2005).
    Google Scholar 

    89.
    Cheung, W. W. L., Lam, V. W. Y. & Pauly, D. in Modelling Present and Climate-shifted Distribution of Marine Fishes and Invertebrates 5–50 (Fisheries Centre, 2008).

    90.
    Cheung, W. W. L. et al. Projecting global marine biodiversity impacts under climate change scenarios. Fish Fish. 10, 235–251 (2009).
    Google Scholar 

    91.
    Mueter, F. J. & Litzow, M. A. Sea ice retreat alters the biogeography of the Bering Sea continental shelf. Ecol. Appl. 18, 309–320 (2008).
    Google Scholar 

    92.
    Dulvy, N. K. et al. Climate change and deepening of the North Sea fish assemblage: a biotic indicator of warming seas. J. Appl. Ecol. 45, 1029–1039 (2008).
    Google Scholar 

    93.
    Pörtner, H.-O. Oxygen- and capacity-limitation of thermal tolerance: a matrix for integrating climate-related stressor effects in marine ecosystems. J. Exp. Biol. 213, 881–893 (2010).
    Google Scholar 

    94.
    Pauly, D. & Kinne, O. Gasping Fish and Panting Squids: Oxygen, Temperature and the Growth of Water-Breathing Animals Vol. 22 (International Ecology Institute, 2010).

    95.
    Mackenzie, C. L. et al. Ocean warming, more than acidification, reduces shell strength in a commercial shellfish species during food limitation. PLoS One 9, e86764 (2014).
    Google Scholar 

    96.
    Rosas-Navarro, A., Langer, G. & Ziveri, P. Temperature affects the morphology and calcification of Emiliania huxleyi strains. Biogeosciences 13, 2913–2926 (2016).
    Google Scholar 

    97.
    Pörtner, H.-O., Bock, C. & Mark, F. C. Oxygen- and capacity-limited thermal tolerance: bridging ecology and physiology. J. Exp. Biol. 220, 2685–2696 (2017).
    Google Scholar 

    98.
    Daufresne, M., Lengfellner, K. & Sommer, U. Global warming benefits the small in aquatic ecosystems. Proc. Natl Acad. Sci. USA 106, 12788–12793 (2009).
    Google Scholar 

    99.
    Baudron, A. R., Needle, C. L. & Marshall, C. T. Implications of a warming North Sea for the growth of haddock Melanogrammus aeglefinus. J. Fish. Biol. 78, 1874–1889 (2011).
    Google Scholar 

    100.
    Sheridan, J. A. & Bickford, D. Shrinking body size as an ecological response to climate change. Nat. Clim. Change 1, 401–406 (2011).
    Google Scholar 

    101.
    Cheung, W. W. L. et al. Shrinking of fishes exacerbates impacts of global ocean changes on marine ecosystems. Nat. Clim. Change 3, 254–258 (2013).
    Google Scholar 

    102.
    Lotze, H. K. et al. Global ensemble projections reveal trophic amplification of ocean biomass declines with climate change. Proc. Natl Acad. Sci. USA 116, 12907–12912 (2019).
    Google Scholar 

    103.
    Cheung, W. W. L. et al. Structural uncertainty in projecting global fisheries catches under climate change. Ecol. Model. 325, 57–66 (2016).
    Google Scholar 

    104.
    Food and Agriculture Organization of the United Nations (FAO) The State of World Fisheries and Aquaculture 2018. Meeting the Sustainable Development Goals (Food and Agriculture Organization of the United Nations (FAO), 2018).

    105.
    Pauly, D. & Zeller, D. Catch reconstructions reveal that global marine fisheries catches are higher than reported and declining. Nat. Commun. 7, 10244 (2016).
    Google Scholar 

    106.
    Swartz, W., Sumaila, R. & Watson, R. Global ex-vessel fish price database revisited: a new approach for estimating ‘missing’ prices. Environ. Resour. Econ. 56, 467–480 (2013).
    Google Scholar 

    107.
    Pauly, D., Christensen, V., Dalsgaard, J., Froese, R. & Torres, F. Fishing down marine food webs. Science 279, 860–863 (1998).
    Google Scholar 

    108.
    Worm, B. et al. Rebuilding global fisheries. Science 325, 578–585 (2009).
    Google Scholar 

    109.
    Costello, C. et al. Status and solutions for the world’s unassessed fisheries. Science 338, 517–520 (2012).
    Google Scholar 

    110.
    Garcia, S. M. & Rosenberg, A. A. Food security and marine capture fisheries: characteristics, trends, drivers and future perspectives. Philos. Trans. R. Soc. B Biol. Sci. 365, 2869–2880 (2010).
    Google Scholar 

    111.
    Anderson, S. C., Branch, T. A., Ricard, D. & Lotze, H. K. Assessing global marine fishery status with a revised dynamic catch-based method and stock-assessment reference points. ICES J. Mar. Sci. 69, 1491–1500 (2012).
    Google Scholar 

    112.
    Costello, C. et al. Global fishery prospects under contrasting management regimes. Proc. Natl Acad. Sci. USA 113, 5125–5129 (2016).
    Google Scholar 

    113.
    Thorson, J. T., Branch, T. A. & Jensen, O. P. Using model-based inference to evaluate global fisheries status from landings, location, and life history data. Can. J. Fish. Aquat. Sci. 69, 645–655 (2012).
    Google Scholar 

    114.
    McOwen, C. J., Cheung, W. W. L., Rykaczewski, R. R., Watson, R. A. & Wood, L. J. Is fisheries production within large marine ecosystems determined by bottom-up or top-down forcing? Fish Fish. 16, 623–632 (2015).
    Google Scholar 

    115.
    Stock, C. A. et al. Reconciling fisheries catch and ocean productivity. Proc. Natl Acad. Sci. USA 114, E1441–E1449 (2017).
    Google Scholar 

    116.
    Free, C. M. et al. Impacts of historical warming on marine fisheries production. Science 363, 979–983 (2019).
    Google Scholar 

    117.
    Cheung, W. W. L., Watson, R. & Pauly, D. Signature of ocean warming in global fisheries catch. Nature 497, 365–368 (2013).
    Google Scholar 

    118.
    Cheung, W. W. L., Reygondeau, G. & Frölicher, T. L. Large benefits to marine fisheries of meeting the 1.5 C global warming target. Science 354, 1591–1594 (2016).
    Google Scholar 

    119.
    Cheung, W. W. L., Jones, M. C., Reygondeau, G. & Frölicher, T. L. Opportunities for climate-risk reduction through effective fisheries management. Glob. Change Biol. 24, 5149–5163 (2018).
    Google Scholar 

    120.
    Sumaila, U. R., Cheung, W. W. L., Lam, V. W. Y., Pauly, D. & Herrick, S. Climate change impacts on the biophysics and economics of world fisheries. Nat. Clim. Change 1, 449–456 (2011).
    Google Scholar 

    121.
    Allison, E. H. et al. Vulnerability of national economies to the impacts of climate change on fisheries. Fish Fish. 10, 173–196 (2009).
    Google Scholar 

    122.
    Bell, J. D. et al. Adaptations to maintain the contributions of small-scale fisheries to food security in the Pacific Islands. Mar. Policy 88, 303–314 (2018).
    Google Scholar 

    123.
    The Pacific Community (SPC) Implications of Climate-driven Redistribution of Tuna for Pacific Island Economies (The Pacific Community (SPC), 2019).

    124.
    Blasiak, R. et al. Climate change and marine fisheries: least developed countries top global index of vulnerability. PLoS One 12, e0179632 (2017).
    Google Scholar 

    125.
    Srinivasan, U., Cheung, W., Watson, R. & Sumaila, U. Food security implications of global marine catch losses due to overfishing. J. Bioeconomics 12, 183–200 (2010).
    Google Scholar 

    126.
    Oyinlola, M. A., Reygondeau, G., Wabnitz, C. C. C. & Cheung, W. W. L. Projecting global mariculture diversity under climate change. Glob. Change Biol. 26, 2134–2148 (2020).
    Google Scholar 

    127.
    Froehlich, H. E., Gentry, R. R. & Halpern, B. S. Global change in marine aquaculture production potential under climate change. Nat. Ecol. Evol. 2, 1745–1750 (2018).
    Google Scholar 

    128.
    Porter, J. R. et al. in Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change Ch. 7 (eds Field, C. B. et al.) 485–533 (Cambridge Univ. Press, 2014).

    129.
    Gaines, S. D. et al. Improved fisheries management could offset many negative effects of climate change. Sci. Adv. 4, eaao1378 (2018).
    Google Scholar 

    130.
    Smith, M. D. et al. Sustainability and global seafood. Science 327, 784–786 (2010).
    Google Scholar 

    131.
    Asche, F., Bellemare, M. F., Roheim, C., Smith, M. D. & Tveteras, S. Fair enough? Food security and the international trade of seafood. World Dev. 67, 151–160 (2015).
    Google Scholar 

    132.
    Kurien, J. Responsible Fish Trade and Food Security (Food and Agriculture Organization of the United Nations (FAO), 2005).

    133.
    Watson, R. A., Nichols, R., Lam, V. W. Y. & Sumaila, U. R. Global seafood trade flows and developing economies: Insights from linking trade and production. Mar. Policy 82, 41–49 (2017).
    Google Scholar 

    134.
    Food and Agriculture Organization of the United Nations (FAO) FAO Yearbook of Fishery and Aquaculture Statistics (Food and Agriculture Organization of the United Nations (FAO), 2017).

    135.
    Gorez, B. West Africa: fishmeal, mealy deal. Samudra Rep. 78, 33–35 (2018).
    Google Scholar 

    136.
    Corten, A., Braham, C.-B. & Sadegh, A. S. The development of a fishmeal industry in Mauritania and its impact on the regional stocks of sardinella and other small pelagics in Northwest Africa. Fish. Res. 186, 328–336 (2017).
    Google Scholar 

    137.
    Merino, G., Barange, M., Mullon, C. & Rodwell, L. Impacts of global environmental change and aquaculture expansion on marine ecosystems. Glob. Environ. Change 20, 586–596 (2010).
    Google Scholar 

    138.
    Naylor, R. L. et al. Feeding aquaculture in an era of finite resources. Proc. Natl Acad. Sci. USA 106, 15103–15110 (2009).
    Google Scholar 

    139.
    Cashion, T., Le Manach, F., Zeller, D. & Pauly, D. Most fish destined for fishmeal production are food-grade fish. Fish Fish. 18, 837–844 (2017).
    Google Scholar 

    140.
    New, M. B. & Wijkström, U. N. Use of fishmeal and fish oil in aquafeeds: further thoughts on the fishmeal trap. FAO Fisheries Circular No. 975 (2002).

    141.
    Jackson, A. & Shepherd, J. in Advancing the Aquaculture Agenda: Workshop Proceedings 331–343 (OECD, 2010).

    142.
    Kristofersson, D. & Anderson, J. L. Is there a relationship between fisheries and farming? Interdependence of fisheries, animal production and aquaculture. Mar. Policy 30, 721–725 (2006).
    Google Scholar 

    143.
    Deutsch, L. et al. Feeding aquaculture growth through globalization: Exploitation of marine ecosystems for fishmeal. Glob. Environ. Change 17, 238–249 (2007).
    Google Scholar 

    144.
    Mullon, C. et al. Modeling the global fishmeal and fish oil markets. Nat. Resour. Model. 22, 564–609 (2009).
    Google Scholar 

    145.
    Merino, G. et al. Can marine fisheries and aquaculture meet fish demand from a growing human population in a changing climate? Glob. Environ. Change 22, 795–806 (2012).
    Google Scholar 

    146.
    Liu, Y. & Sumaila, U. R. Can farmed salmon production keep growing? Mar. Policy 32, 497–501 (2008).
    Google Scholar 

    147.
    Pinsky, M. L. et al. Preparing ocean governance for species on the move. Science 360, 1189–1191 (2018).
    Google Scholar 

    148.
    Jacobs, A. China’s appetite pushes fisheries to the brink. New York Times (30 Apr 2017).

    149.
    Tickler, D., Meeuwig, J. J., Palomares, M.-L., Pauly, D. & Zeller, D. Far from home: Distance patterns of global fishing fleets. Sci. Adv. 4, eaar3279 (2018).
    Google Scholar 

    150.
    Campling, L. Trade politics and the global production of canned tuna. Mar. Policy 69, 220–228 (2016).
    Google Scholar 

    151.
    Eurofish. Overview of the Spanish fisheries and aquaculture sector. https://www.eurofish.dk/spain (2019).

    152.
    Arrizabalaga, H. et al. Global habitat preferences of commercially valuable tuna. Deep Sea Res. Part II Top. Stud. Oceanogr. 113, 102–112 (2015).
    Google Scholar 

    153.
    Erauskin-Extramiana, M. et al. Large-scale distribution of tuna species in a warming ocean. Glob. Change Biol. 25, 2043–2060 (2019).
    Google Scholar 

    154.
    FFA and SPC. Future of Fisheries: A Regional Roadmap for Sustainable Pacific Fisheries (FFA and SPC, 2015).

    155.
    Heltberg, R., Siegel, P. B. & Jorgensen, S. L. Addressing human vulnerability to climate change: toward a ‘no-regrets’ approach. Glob. Environ. Change 19, 89–99 (2009).
    Google Scholar 

    156.
    Brouwer, S. et al. The Western and Central Pacific Tuna Fishery: 2018 Overview and Status of Stocks (Pacific Community, 2019).

    157.
    Bell, J. D. et al. Diversifying the use of tuna to improve food security and public health in Pacific Island countries and territories. Mar. Policy 51, 584–591 (2015).
    Google Scholar 

    158.
    Senina, I. et al. Predicting skipjack tuna dynamics and effects of climate change using SEAPODYM with fishing and tagging data. Scientific Committee Twelfth Regular Session, Western and Central Pacific Fisheries Commission 1–70 (2016).

    159.
    Robinson, M. Climate Justice: Hope, Resilience, and the Fight for a Sustainable Future (Bloomsbury Publishing, 2018).

    160.
    United Nations. Transforming Our World: the 2030 agenda for sustainable development https://doi.org/10.1891/9780826190123.ap02 (2015).

    161.
    Pecl, G. T. et al. Biodiversity redistribution under climate change: Impacts on ecosystems and human well-being. Science 355, eaai9214 (2017).
    Google Scholar 

    162.
    Singh, G. G. et al. Climate impacts on the ocean are making the Sustainable Development Goals a moving target travelling away from us. People Nat. 1, 317–330 (2019).
    Google Scholar 

    163.
    Guillaumont, P. Assessing the economic vulnerability of small island developing states and the least developed countries. J. Dev. Stud. 46, 828–854 (2010).
    Google Scholar 

    164.
    Narayan, S. et al. The effectiveness, costs and coastal protection benefits of natural and nature-based defences. PLoS One 11, e0154735 (2016).
    Google Scholar 

    165.
    Moosavi, S. Ecological coastal protection: pathways to living shorelines. Procedia Eng. 196, 930–938 (2017).
    Google Scholar 

    166.
    Mutombo, K. & Ölçer, A. A three-tier framework for port infrastructure adaptation to climate change: balancing breadth and depth of knowledge. Ocean Yearb. Online 30, 564–577 (2016).
    Google Scholar 

    167.
    Forzieri, G. et al. Escalating impacts of climate extremes on critical infrastructures in Europe. Glob. Environ. Change 48, 97–107 (2018).
    Google Scholar 

    168.
    Beiler, M. O., Marroquin, L. & McNeil, S. State-of-the-practice assessment of climate change adaptation practices across metropolitan planning organizations pre-and post-Hurricane Sandy. Transp. Res. Part A Policy Pract. 88, 163–174 (2016).
    Google Scholar 

    169.
    Thorne, J. H. et al. The impact of climate change uncertainty on California’s vegetation and adaptation management. Ecosphere 8, e02021 (2017).
    Google Scholar 

    170.
    Ziervogel, G. & Ericksen, P. J. Adapting to climate change to sustain food security. Wiley Interdiscip. Rev. Clim. Change 1, 525–540 (2010).
    Google Scholar 

    171.
    Heenan, A. et al. A climate-informed, ecosystem approach to fisheries management. Mar. Policy 57, 182–192 (2015).
    Google Scholar 

    172.
    Poulain, F., Himes-Cornell, A. & Shelton, C. in Impacts of Climate Change on Fisheries and Aquaculture. Synthesis of Current Knowledge, Adaptation and Mitigation Options FAO Fisheries and Aquaculture Technical Paper 627 Ch. 25 535–566 (Food and Agriculture Organization of the United Nations (FAO), 2018).

    173.
    Bell, J. et al. Impacts and effects of ocean warming on the contributions of fisheries and aquaculture to food security (IUCN, 2016).

    174.
    Cochrane, K. L., Andrew, N. L. & Parma, A. M. Primary fisheries management: a minimum requirement for provision of sustainable human benefits in small-scale fisheries. Fish Fish. 12, 275–288 (2011).
    Google Scholar 

    175.
    Free, C. M. et al. Realistic fisheries management reforms could mitigate the impacts of climate change in most countries. PLoS One 15, e0224347 (2020).
    Google Scholar 

    176.
    Armitage, D. Adaptive capacity and community-based natural resource management. Environ. Manage. 35, 703–715 (2005).
    Google Scholar 

    177.
    MECM/MFMR. Solomon Islands National Plan of Action. Coral Triangle Initiative on Coral Reefs, Fisheries and Food Security (Solomon Islands Government, 2010).

    178.
    Bell, J. D. et al. Optimising the use of nearshore fish aggregating devices for food security in the Pacific Islands. Mar. Policy 56, 98–105 (2015).
    Google Scholar 

    179.
    Tilley, A. et al. Nearshore fish aggregating devices show positive outcomes for sustainable fisheries development in Timor-Leste. Front. Mar. Sci. 6, 487 (2019).
    Google Scholar 

    180.
    Mills, D. J. et al. Developing Timor-Leste’s Coastal Economy: Assessing Potential Climate Change Impacts and Adaptation Options. Final Report to the Australian Government Coral Triangle Initiative on Coral Reefs, Fisheries and Food Security National Initiative (WorldFish, 2013).

    181.
    Pomeroy, R. S. Community-based and co-management institutions for sustainable coastal fisheries management in Southeast Asia. Ocean Coast. Manag. 27, 143–162 (1995).
    Google Scholar 

    182.
    Foale, S., Cohen, P., Januchowski-Hartley, S., Wenger, A. & Macintyre, M. Tenure and taboos: origins and implications for fisheries in the Pacific. Fish Fish. 12, 357–369 (2011).
    Google Scholar 

    183.
    Tompkins, E. L. & Adger, W. N. Does adaptive management of natural resources enhance resilience to climate change? Ecol. Soc. 9, 10 (2004).
    Google Scholar 

    184.
    Biggs, R. et al. Toward principles for enhancing the resilience of ecosystem services. Annu. Rev. Environ. Resour. 37, 421–448 (2012).
    Google Scholar 

    185.
    Cohen, P. J. & Foale, S. J. Sustaining small-scale fisheries with periodically harvested marine reserves. Mar. Policy 37, 278–287 (2013).
    Google Scholar 

    186.
    Carvalho, P. G. et al. Optimized fishing through periodically harvested closures. J. Appl. Ecol. 56, 1927–1936 (2019).
    Google Scholar 

    187.
    Cinner, J. E. et al. Evaluating social and ecological vulnerability of coral reef fisheries to climate change. PLoS One 8, e74321 (2013).
    Google Scholar 

    188.
    Ford, J. D. et al. Including indigenous knowledge and experience in IPCC assessment reports. Nat. Clim. Change 6, 349–353 (2016).
    Google Scholar 

    189.
    McNamara, K. E. & Westoby, R. Local knowledge and climate change adaptation on Erub Island, Torres Strait. Local Environ. 16, 887–901 (2011).
    Google Scholar 

    190.
    Miller, D. D., Ota, Y., Sumaila, U. R., Cisneros-Montemayor, A. M. & Cheung, W. W. L. Adaptation strategies to climate change in marine systems. Glob. Change Biol. 24, e1–e14 (2018).
    Google Scholar 

    191.
    Weeks, R. & Jupiter, S. D. Adaptive comanagement of a marine protected area network in Fiji. Conserv. Biol. 27, 1234–1244 (2013).
    Google Scholar 

    192.
    Ogier, E. M. et al. Fisheries management approaches as platforms for climate change adaptation: comparing theory and practice in Australian fisheries. Mar. Policy 71, 82–93 (2016).
    Google Scholar 

    193.
    Bruno, J. F., Côté, I. M. & Toth, L. T. Climate change, coral loss, and the curious case of the parrotfish paradigm: why don’t marine protected areas improve reef resilience? Annu. Rev. Mar. Sci. 11, 307–334 (2019).
    Google Scholar 

    194.
    Oremus, K. L. et al. Governance challenges for tropical nations losing fish species due to climate change. Nat. Sustain. 3, 277–280 (2020).
    Google Scholar 

    195.
    Mendenhall, E. et al. Climate change increases the risk of fisheries conflict. Mar. Policy 117, 103954 (2020).
    Google Scholar 

    196.
    Moore, B. R. et al. Defining the stock structures of key commercial tunas in the Pacific Ocean I: current knowledge and main uncertainties. Fish. Res. https://doi.org/10.1016/j.fishres.2020.105525 (2020).

    197.
    Moore, B. R. et al. Defining the stock structures of key commercial tunas in the Pacific Ocean II: Sampling considerations and future directions. Fish. Res. https://doi.org/10.1016/j.fishres.2020.105524 (2020).

    198.
    Gattuso, J.-P. et al. Ocean solutions to address climate change and its effects on marine ecosystems. Front. Mar. Sci. 5, 337 (2018).
    Google Scholar 

    199.
    Sumaila, U. R. et al. Benefits of the Paris Agreement to ocean life, economies, and people. Sci. Adv. 5, eaau3855 (2019).
    Google Scholar 

    200.
    Gallo, N. D., Victor, D. G. & Levin, L. A. Ocean commitments under the Paris Agreement. Nat. Clim. Change 7, 833–838 (2017).
    Google Scholar 

    201.
    Mcleod, E. et al. A blueprint for blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2. Front. Ecol. Environ. 9, 552–560 (2011).
    Google Scholar 

    202.
    Herr, D. & Landis, E. Coastal Blue Carbon Ecosystems. Opportunities for Nationally Determined Contributions. Policy Brief (IUCN, 2016).

    203.
    Goldstein, A. et al. Protecting irrecoverable carbon in Earth’s ecosystems. Nat. Clim. Change 10, 287–295 (2020).
    Google Scholar 

    204.
    Pendleton, L. et al. Estimating global “blue carbon” emissions from conversion and degradation of vegetated coastal ecosystems. PLoS One 7, e43542 (2012).
    Google Scholar 

    205.
    Zarate-Barrera, T. G. & Maldonado, J. H. Valuing blue carbon: carbon sequestration benefits provided by the marine protected areas in Colombia. PLoS One 10, e0126627 (2015).
    Google Scholar 

    206.
    Wylie, L., Sutton-Grier, A. E. & Moore, A. Keys to successful blue carbon projects: lessons learned from global case studies. Mar. Policy 65, 76–84 (2016).
    Google Scholar 

    207.
    Locatelli, T. et al. Turning the tide: how blue carbon and payments for ecosystem services (PES) might help save mangrove forests. Ambio 43, 981–995 (2014).
    Google Scholar 

    208.
    Barbesgaard, M. C. Blue carbon: ocean grabbing in disguise? Transnational Institute https://www.tni.org/en/publication/blue-carbon-ocean-grabbing-in-disguise (2016).

    209.
    Sharp, M. The benefits of fish aggregating devices in the Pacific. SPC Fish. Newsl. 135, 28–36 (2011).
    Google Scholar 

    210.
    Grafton, R. Q. Adaptation to climate change in marine capture fisheries. Mar. Policy 34, 606–615 (2010).
    Google Scholar 

    211.
    Kurien, J. Voluntary guidelines for securing sustainable small-scale fisheries in the context of food security and poverty eradication: summary (Food and Agriculture Organization of the United Nations (FAO), 2015).

    212.
    Castree, N. et al. Changing the intellectual climate. Nat. Clim. Change 4, 763–768 (2014).
    Google Scholar 

    213.
    Allison, E. H. & Bassett, H. R. Climate change in the oceans: Human impacts and responses. Science 350, 778–782 (2015).
    Google Scholar 

    214.
    Bobrowsky, P., Cronin, V. S., Di Capua, G., Kieffer, S. W. & Peppoloni, S. 11. The emerging field of geoethics. Sci. Integr. Ethics Geosci. 73, 175 (2017).
    Google Scholar 

    215.
    Bohle, M. One realm: thinking geoethically and guiding small-scale fisheries? Eur. J. Dev. Res. 31, 253–270 (2019).
    Google Scholar 

    216.
    UNEP-WCMC, WorldFish Centre, WRI & TNC. Global distribution of warm-water coral reefs, compiled from multiple sources including the Millennium Coral Reef Mapping Project. Version 4.0. Includes contributions from IMaRS-USF and IRD (2005), IMaRS-USF (2005) and Spalding et al. (2001) (UN Environment World Conservation Monitoring Centre, 2018).

    217.
    UNEP-WCMC & Short, F. T. Global distribution of seagrasses (version 5.0). Fourth update to the data layer used in Green and Short (2003) (UNEP World Conservation Monitoring Centre, 2017).

    218.
    Giri, C. et al. Status and distribution of mangrove forests of the world using earth observation satellite data (version 1.3, updated by UNEP-WCMC). Glob. Ecol. Biogeogr. 20, 154–159 (2011).
    Google Scholar 

    219.
    Mcowen, C. J. et al. A global map of saltmarshes. Biodivers. Data J. 5, e11764 (2017).
    Google Scholar 

    220.
    Lehodey, P. et al. in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change (eds Bell, J. D., Johnson, J. E. & Hobday, A. J.) 433–492 (Secretariat of the Pacific Community, 2011).

    221.
    Lehodey, P. et al. Modelling the impact of climate change including ocean acidification on Pacific yellowfin tuna. Scientific Committee Thirteenth Regular Session, Western and Central Pacific Fisheries Commission 1–25 (2017).

    222.
    Senina, I. et al. Impact of climate change on tropical Pacific tuna and their fisheries in Pacific Islands waters and high seas areas. Scientific Committee Fourteenth Regular Session, Western and Central Pacific Fisheries Commission 1–43 (2018).

    223.
    Bell, J. D. et al. Mixed responses of tropical Pacific fisheries and aquaculture to climate change. Nat. Clim. Change 3, 591–599 (2013).
    Google Scholar 

    224.
    Bell, J. D. et al. in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change (eds Bell, J. D., Johnson, J. E. & Hobday, A. J.) 733–801 (Secretariat of the Pacific Community, 2011).

    225.
    Bell, J. D. et al. in Impacts of Climate Change on Fisheries and Aquaculture. Synthesis of Current Knowledge, Adaptation and Mitigation Options FAO Fisheries and Aquaculture Technical Paper 627 Ch. 14 305–324 (Food and Agriculture Organization of the United Nations (FAO), 2018).

    226.
    Scott, F., Scott, R., Yao, N., Pilling, G. M. & Hampton, J. Considering uncertainty when testing and monitoring WCPFC harvest strategies. Scientific Committee Fifteenth Regular Session, Western Central Pacific Fisheries Commission 1–23 (2019).

    227.
    Pratchett, M. S. et al. Vulnerabilty of coastal fisheries in the tropical Pacific to climate change (eds Bell, J. D., Johnson, J. E. & Hobday, A. J.) in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change 493–576 (Secretariat of the Pacific Community, 2011).

    228.
    Gasalla, M. A., Abdallah, P. R. & Lemos, D. in Climate Change Impacts on Fisheries and Aquaculture. A Global Analysis. Vol. 1 (eds Philips, B. F. & Pérez-Ramírez, M.) 455–477 (Wiley, 2017).

    229.
    Popova, E. et al. From global to regional and back again: common climate stressors of marine ecosystems relevant for adaptation across five ocean warming hotspots. Glob. Change Biol. 22, 2038–2053 (2016).
    Google Scholar 

    230.
    Schulz, C. et al. Physical, ecological and human dimensions of environmental change in Brazil’s Pantanal wetland: synthesis and research agenda. Sci. Total Environ. 687, 1011–1027 (2019).
    Google Scholar 

    231.
    Barros, D. F. & Albernaz, A. L. M. Possible impacts of climate change on wetlands and its biota in the Brazilian Amazon. Braz. J. Biol. 74, 810–820 (2014).
    Google Scholar 

    232.
    Johnson, J. E. et al. Climate change adaptation: vulnerability and challenges facing small-scale fisheries on small islands. FAO Fish. Aquacult. Proc. 61, 65–80 (2019).
    Google Scholar 

    233.
    Martins, I. M. & Gasalla, M. A. Perceptions of climate and ocean change impacting the resources and livelihood of small-scale fishers in the South Brazil Bight. Clim. Change 147, 441–456 (2018).
    Google Scholar 

    234.
    Martins, I. M., Gammage, L. C., Jarre, A. & Gasalla, M. A. Different but similar? Exploring vulnerability to climate change in Brazilian and South African small-scale fishing communities. Hum. Ecol. 47, 515–526 (2019).
    Google Scholar 

    235.
    Gasalla, M. A. Six decades of change in the South Brazil Bight Ecosystem in Proc. 3rd GLOBEC Open Science Meeting: From Ecosystem Function to Ecosystem Prediction (2008).

    236.
    Dahlet, L. I., Downey-Breedt, N., Arce, G., Sauer, W. H. H. & Gasalla, M. A. Comparative study of skipjack tuna Katsuwonus pelamis (Scombridae) fishery stocks from the South Atlantic and western Indian oceans. Sci. Mar. 83, 19–30 (2019).
    Google Scholar 

    237.
    Araújo, F. G., Teixeira, T. P., Guedes, A. P. P., de Azevedo, M. C. C. & Pessanha, A. L. M. Shifts in the abundance and distribution of shallow water fish fauna on the southeastern Brazilian coast: a response to climate change. Hydrobiologia 814, 205–218 (2018).
    Google Scholar 

    238.
    Gasalla, M. A. An overview of climate change effects in South Brazil Bight fisheries in Proc. 6th World Fisheries Congress (2012).

    239.
    Santos, L. C. M., Gasalla, M. A., Dahdouh-Guebas, F. & Bitencourt, M. D. Socio-ecological assessment for environmental planning in coastal fishery areas: a case study in Brazilian mangroves. Ocean Coast. Manag. 138, 60–69 (2017).
    Google Scholar 

    240.
    Zou, D. & Gao, K. in Seaweeds and Their Role in Globally Changing Environments (eds Seckbach, J., Einav, R. & Israel, A.) 115–126 (Springer, 2010).

    241.
    Ramaglia, A. C., de Castro, L. M. & Augusto, A. Effects of ocean acidification and salinity variations on the physiology of osmoregulating and osmoconforming crustaceans. J. Comp. Physiol. B 188, 729–738 (2018).
    Google Scholar 

    242.
    Freduah, G., Fidelman, P. & Smith, T. F. The impacts of environmental and socio-economic stressors on small scale fisheries and livelihoods of fishers in Ghana. Appl. Geogr. 89, 1–11 (2017).
    Google Scholar 

    243.
    Bunce, M., Rosendo, S. & Brown, K. Perceptions of climate change, multiple stressors and livelihoods on marginal African coasts. Environ. Dev. Sustain. 12, 407–440 (2010).
    Google Scholar 

    244.
    Burke, L.M., Reytar, K., Spalding, M. & Perry, A. Reefs at risk revisited: World Resources Institute. (2017).

    245.
    Roberts, C. M. et al. Marine biodiversity hotspots and conservation priorities for tropical reefs. Science 295, 1280–1284 (2002).
    Google Scholar 

    246.
    Lam, V. W. Y., Cheung, W. W. L., Swartz, W. & Sumaila, U. R. Climate change impacts on fisheries in West Africa: implications for economic, food and nutritional security. Afr. J. Mar. Sci. 34, 103–117 (2012).
    Google Scholar 

    247.
    Barange, M. et al. Impacts of climate change on marine ecosystem production in societies dependent on fisheries. Nat. Clim. Change 4, 211–216 (2014).
    Google Scholar 

    248.
    Blanchard, J. L. et al. Potential consequences of climate change for primary production and fish production in large marine ecosystems. Philos. Trans. R. Soc. B Biol. Sci. 367, 2979–2989 (2012).
    Google Scholar 

    249.
    Belhabib, D., Lam, V. W. Y. & Cheung, W. W. L. Overview of West African fisheries under climate change: impacts, vulnerabilities and adaptive responses of the artisanal and industrial sectors. Mar. Policy 71, 15–28 (2016).
    Google Scholar  More

  • in

    Extreme hyperthermia tolerance in the world’s most abundant wild bird

    1.
    Sears, M. W., Raskin, E. & Angilletta, M. J. Jr. The world is not flat: defining relevant thermal landscapes in the context of climate change. Integr. Comp. Biol. 51, 666–675 (2011).
    PubMed  Google Scholar 
    2.
    du Plessis, K. L., Martin, R. O., Hockey, P. A. R., Cunningham, S. J. & Ridley, A. R. The costs of keeping cool in a warming world: implications of high temperatures for foraging, thermoregulation and body condition of an arid-zone bird. Glob. Change Biol. 18, 3063–3070 (2012).
    ADS  Google Scholar 

    3.
    Araújo, M. B. et al. Heat freezes niche evolution. Ecol. Lett. 16, 1206–1219 (2013).
    PubMed  Google Scholar 

    4.
    Speakman, J. R. & Król, E. Maximal heat dissipation capacity and hyperthermia risk: neglected key factors in the ecology of endotherms. J. Anim. Ecol. 79, 726–746 (2010).
    PubMed  Google Scholar 

    5.
    Daghir, N. J. Poultry production in hot climates 2nd edn. (CAB International, Wallingford, 2008).
    Google Scholar 

    6.
    Nyoni, N. M. B., Grab, S. & Archer, E. R. M. Heat stress and chickens: climate risk effects on rural poultry farming in low-income countries. Clim. Dev. 11, 83–90. https://doi.org/10.1080/17565529.2018.1442792 (2018).
    Article  Google Scholar 

    7.
    Laszlo, A. The effects of hyperthermia on mammalian cell structure and function. Cell Prolif. 25, 59–87 (1992).
    CAS  PubMed  Google Scholar 

    8.
    Roti Roti, J. L. Cellular responses to hyperthermia (40–46 C): Cell killing and molecular events. Int. J. Hyperthermia 24, 3–15 (2008).
    ADS  PubMed  Google Scholar 

    9.
    Feder, M. E. & Hofmann, G. E. Heat-shock proteins, molecular chaperones, and the stress response: evolutionary and ecological physiology. Annu. Rev. Physiol. 61, 243–282 (1999).
    CAS  PubMed  Google Scholar 

    10.
    Hochachka, P. W. & Somero, G. N. Biochemical Adaptation (Princeton University Press, Princeton, 1984).
    Google Scholar 

    11.
    Pörtner, H. Climate change and temperature-dependent biogeography: oxygen limitation of thermal tolerance in animals. Naturwissenschaften 88, 137–146 (2001).
    ADS  PubMed  Google Scholar 

    12.
    Pörtner, H.-O. Oxygen-and capacity-limitation of thermal tolerance: a matrix for integrating climate-related stressor effects in marine ecosystems. J. Exp. Biol. 213, 881–893 (2010).
    PubMed  Google Scholar 

    13.
    Clusella-Trullas, S., Blackburn, T. M. & Chown, S. L. Climatic predictors of temperature performance curve parameters in ectotherms imply complex responses to climate change. Am. Nat. 177, 738–751 (2011).
    PubMed  Google Scholar 

    14.
    McKechnie, A. E. & Wolf, B. O. The physiology of heat tolerance in small endotherms. Physiology 34, 302–313 (2019).
    CAS  PubMed  Google Scholar 

    15.
    Arad, Z. & Marder, J. Strain differences in heat resistance to acute heat stress, between the bedouin desert fowl, the white leghorn and their crossbreeds. Comp. Biochem. Physiol. A 72, 191–193 (1982).
    Google Scholar 

    16.
    Randall, W. C. Factors influencing the temperature regulation of birds. Am. J. Physiol. 139, 56–63 (1943).
    Google Scholar 

    17.
    Tieleman, B. I., Williams, J. B., LaCroix, F. & Paillat, P. Physiological responses of Houbara bustards to high ambient temperatures. J. Exp. Biol. 205, 503–511 (2002).
    PubMed  Google Scholar 

    18.
    Chappell, M. A. & Bartholomew, G. A. Activity and thermoregulation of the antelope ground squirrel Ammospermophilus leucurus in winter and summer. Physiol. Zool. 54, 215–223 (1981).
    Google Scholar 

    19.
    Lovegrove, B. G., Heldmaier, G. & Ruf, T. Perspectives of endothermy revisited: the endothermic temperature range. J. Therm. Biol 16, 185–197 (1991).
    Google Scholar 

    20.
    Cory Toussaint, D. & McKechnie, A. E. Interspecific variation in thermoregulation among three sympatric bats inhabiting a hot, semi-arid environment. J. Comp. Physiol. B 182, 1129–1140 (2012).
    PubMed  Google Scholar 

    21.
    Dawson, W. R. In University of California Publications in Zoology Vol. 59 (eds Bartholomew, G. A. et al.) 81–123 (University of California Press, California, 1954).
    Google Scholar 

    22.
    Paulissen, M. A. Ontogenetic comparison of body temperature selection and thermal tolerance of Cnemidophorus sexlineatus. J. Herpetol. 22, 473–476 (1988).
    Google Scholar 

    23.
    Weathers, W. W. Energetics and thermoregulation by small passerines of the humid, lowland tropics. Auk 114, 341–353 (1997).
    Google Scholar 

    24.
    Southwick, E. E. Remote sensing of body temperature in a captive 25-g bird. Condor 75, 464–466 (1973).
    Google Scholar 

    25.
    Elliott, C. C. H. In Quelea quelea: Africa’s bird pest (eds Bruggers, R. L. & Elliott, C. C. H.) (Oxford University Press, Oxford, 1989).
    Google Scholar 

    26.
    Craig, A. J. F. K. In Roberts birds of southern Africa (eds Hockey, P. A. R. et al.) 1025–1026 (The Trustees of the John Voelcker Bird Book Fund, Cape Town, 2005).
    Google Scholar 

    27.
    Whitfield, M. C., Smit, B., McKechnie, A. E. & Wolf, B. O. Avian thermoregulation in the heat: scaling of heat tolerance and evaporative cooling capacity in three southern African arid-zone passerines. J. Exp. Biol. 218, 1705–1714 (2015).
    PubMed  Google Scholar 

    28.
    McKechnie, A. E. et al. Avian thermoregulation in the heat: efficient evaporative cooling allows for extreme heat tolerance in four southern Hemisphere columbids. J. Exp. Biol. 219, 2145–2155 (2016).
    PubMed  Google Scholar 

    29.
    Smith, E. K., O’Neill, J. J., Gerson, A. R., McKechnie, A. E. & Wolf, B. O. Avian thermoregulation in the heat: resting metabolism, evaporative cooling and heat tolerance in Sonoran Desert songbirds. J. Exp. Biol. 220, 3290–3300 (2017).
    PubMed  Google Scholar 

    30.
    Smit, B. et al. Avian thermoregulation in the heat: phylogenetic variation among avian orders in evaporative cooling capacity and heat tolerance. J. Exp. Biol. 221, jeb174870 (2018).
    PubMed  Google Scholar 

    31.
    Karasov, W. H. In Studies in Avian Biology (eds Morrison, M. L. et al.) 391–415 (Cooper Ornithological Society, California, 1990).
    Google Scholar 

    32.
    Swanson, D. L., Drymalski, M. W. & Brown, J. R. Sliding vs static cold exposure and the measurement of summit metabolism in birds. J. Therm. Biol 21, 221–226 (1996).
    Google Scholar 

    33.
    Kemp, R. & McKechnie, A. E. Thermal physiology of a range-restricted desert lark. J. Comp. Physiol. B 189, 131–141. https://doi.org/10.1007/s00360-018-1190-1 (2019).
    Article  PubMed  Google Scholar 

    34.
    Lighton, J. R. B. Measuring Metabolic Rates: A Manual for Scientists (Oxford University Press, Oxford, 2008).
    Google Scholar 

    35.
    Walsberg, G. E. & Wolf, B. O. Variation in the respirometry quotient of birds and implications for indirect calorimetry using measurements of carbon dioxide production. J. Exp. Biol. 198, 213–219 (1995).
    CAS  PubMed  Google Scholar 

    36.
    Tracy, C. R., Welch, W. R., Pinshow, B. & Porter, W. P. Properties of air: a manual for use in biophysical ecology. 4th Ed. The University of Wisconsin Laboratory for Biophysical Ecology: Technical Report (2010).

    37.
    R Core Team R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, Austria, 2019).

    38.
    Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D. & R Core Team. nlme: Linear and Nonlinear Mixed Effects Models. R Package Version 3. 57. (2009).

    39.
    Muggeo, V. M. R. Segmented: an R package to fit regression models with broken-line relationships. R News 8(1), 20–25 (2008).
    Google Scholar 

    40.
    McKechnie, A. E. et al. Avian thermoregulation in the heat: evaporative cooling in five Australian passerines reveals within-order biogeographic variation in heat tolerance. J. Exp. Biol. 220, 2436–2444 (2017).
    PubMed  Google Scholar 

    41.
    O’Connor, R. S., Wolf, B. O., Brigham, R. M. & McKechnie, A. E. Avian thermoregulation in the heat: efficient evaporative cooling in two southern African nightjars. J Comp Physiol B 187, 477–491 (2017).
    PubMed  Google Scholar 

    42.
    Hoffmann, A. A., Chown, S. L. & Clusella-Trullas, S. Upper thermal limits in terrestrial ectotherms: how constrained are they? Funct. Ecol. 27, 934–949 (2013).
    Google Scholar 

    43.
    Tieleman, B. I., Williams, J. B. & Bloomer, P. Adaptation of metabolic rate and evaporative water loss along an aridity gradient. Proc. R. Soc. Lond. 270, 207–214 (2003).
    Google Scholar 

    44.
    Xie, S., Tearle, R. & McWhorter, T. J. Heat shock protein expression is upregulated after acute heat exposure in three species of Australian desert birds. Avian Biol. Res. 11, 263–273 (2018).
    Google Scholar 

    45.
    Czenze, Z. J. et al. Regularly-drinking desert birds have greater evaporative cooling capacity and higher heat tolerance limits than non-drinking species. Funct. Ecol. https://doi.org/10.1111/1365-2435.13573 (2020).
    Article  Google Scholar 

    46.
    Midtgård, U. Scaling of the brain and the eye cooling system in birds: a morphometric analysis of the rete ophthalmicum. J. Exp. Zool. 225, 197–207 (1983).
    PubMed  Google Scholar 

    47.
    Kilgore, D. L., Bernstein, M. H. & Hudson, D. M. Brain temperatures in birds. J Comp Physiol 110, 209–215 (1976).
    Google Scholar 

    48.
    Bernstein, M. H., Curtis, M. B. & Hudson, D. M. Independence of brain and body temperatures in flying American kestrels, Falco sparverius. Am. J. Physiol. 237, R58–R62 (1979).
    CAS  PubMed  Google Scholar 

    49.
    Kregel, K. C. Invited review: heat shock proteins: modifying factors in physiological stress responses and acquired thermotolerance. J. Appl. Physiol. 92, 2177–2186 (2002).
    CAS  PubMed  Google Scholar 

    50.
    McKechnie, A. E. et al. Avian thermoregulation in the heat: evaporative cooling capacity in an archetypal desert specialist, Burchell’s sandgrouse (Pterocles burchelli). J. Exp. Biol. 219, 2137–2144 (2016).
    PubMed  Google Scholar 

    51.
    Talbot, W. A., McWhorter, T. J., Gerson, A. R., McKechnie, A. E. & Wolf, B. O. Avian thermoregulation in the heat: evaporative cooling capacity of arid-zone Caprimulgiformes from two continents. J. Exp. Biol. 220, 3488–3498 (2017).
    PubMed  Google Scholar 

    52.
    McWhorter, T. J. et al. Avian thermoregulation in the heat: evaporative cooling capacity and thermal tolerance in two Australian parrots. J. Exp. Biol. 221, jeb168930 (2018).
    PubMed  Google Scholar 

    53.
    Talbot, W. A., Gerson, A. R., Smith, E. K., McKechnie, A. E. & Wolf, B. O. Avian thermoregulation in the heat: metabolism, evaporative cooling and gular flutter in two small owls. J. Exp. Biol. 221, jeb171108 (2018).
    PubMed  Google Scholar 

    54.
    Smith, E. K., O’Neill, J., Gerson, A. R. & Wolf, B. O. Avian thermoregulation in the heat: resting metabolism, evaporative cooling and heat tolerance in Sonoran Desert doves and quail. J. Exp. Biol. 218, 3636–3646 (2015).
    PubMed  Google Scholar  More

  • in

    Sediment microbial fuel cells as a barrier to sulfide accumulation and their potential for sediment remediation beneath aquaculture pens

    1.
    El-Naggar, M. Y. et al. Electrical transport along bacterial nanowires from Shewanella oneidensis MR-1. Proc. Natl. Acad. Sci. USA 107, 18127–18131 (2010).
    ADS  CAS  PubMed  Google Scholar 
    2.
    Du Toit, A. Exporting electrons. Nat. Rev. Microbiol. 16, 657 (2018).
    PubMed  Google Scholar 

    3.
    Lovley, D. R. Microbial fuel cells: novel microbial physiologies and engineering approaches. Curr. Opin. Biotechnol. 17, 327–332 (2006).
    CAS  PubMed  Google Scholar 

    4.
    Myers, J. M. & Myers, C. R. Role for outer membrane cytochromes OmcA and OmcB of Shewanella putrefaciens MR-1 in reduction of manganese dioxide. Appl. Environ. Microbiol. 67, 260–269 (2001).
    CAS  PubMed  PubMed Central  Google Scholar 

    5.
    Bond, D. R. & Lovley, D. R. Electricity production by Geobacter sulfurreducens attached to electrodes. Appl. Environ. Microbiol. 69, 1548–1555 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    6.
    Rabaey, K., Boon, N., Siciliano, S. D., Verhaege, M. & Verstraete, W. Biofuel cells select for microbial consortia that self-mediate electron transfer. Appl. Environ. Microbiol. 70, 5373–5382 (2004).
    CAS  PubMed  PubMed Central  Google Scholar 

    7.
    Rabaey, K., Boon, N., Höfte, M. & Verstraete, W. Microbial phenazine production enhances electron transfer in biofuel cells. Environ. Sci. Technol. 39, 3401–3408 (2005).
    ADS  CAS  PubMed  Google Scholar 

    8.
    Potter, M. C. Electrical effects accompanying the decomposition of organic compounds. Proc. R. Soc. B Biol. Sci. 84, 260–276 (1911).
    ADS  Google Scholar 

    9.
    Logan, B. E. & Regan, J. M. Microbial fuel cells—challenges and applications. Environ. Sci. Technol. 40, 5172–5180 (2006).
    ADS  CAS  PubMed  Google Scholar 

    10.
    Trapero, J. R., Horcajada, L., Linares, J. J. & Lobato, J. Is microbial fuel cell technology ready? An economic answer towards industrial commercialization. Appl. Energy 185, 698–707 (2017).
    CAS  Google Scholar 

    11.
    Reimers, C. E., Tender, L. M., Fertig, S. & Wang, W. Harvesting energy from the marine sediment−water interface. Environ. Sci. Technol. 35, 192–195 (2001).
    ADS  CAS  PubMed  Google Scholar 

    12.
    Tender, L. M. et al. Harnessing microbially generated power on the seafloor. Nat. Biotechnol. 20, 821–825 (2002).
    CAS  PubMed  Google Scholar 

    13.
    Kubota, K. et al. Operation of sediment microbial fuel cells in Tokyo Bay, an extremely eutrophicated coastal sea. Bioresour. Technol. Rep. 6, 39–45 (2019).
    Google Scholar 

    14.
    Chun, C. L., Payne, R. B., Sowers, K. R. & May, H. D. Electrical stimulation of microbial PCB degradation in sediment. Water Res. 47, 141–152 (2013).
    CAS  PubMed  Google Scholar 

    15.
    Gajda, I., Greenman, J. & Ieropoulos, I. A. Recent advancements in real-world microbial fuel cell applications. Curr. Opin. Electrochem. 11, 78–83 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    16.
    Bond, D. R., Holmes, D. E., Tender, L. M. & Lovley, D. R. Electrode-reducing microorganisms that harvest energy from marine sediments. Science 295, 483–485 (2002).
    ADS  CAS  PubMed  Google Scholar 

    17.
    Froelich, P. N. et al. Early oxidation of organic matter in pelagic sediments of the eastern equatorial Atlantic: suboxic diagenesis. Geochim. Cosmochim. Acta 43, 1075–1090 (1979).
    ADS  CAS  Google Scholar 

    18.
    Hasvold, Ø et al. Sea-water battery for subsea control systems. J. Power Sources 65, 253–261 (1997).
    ADS  CAS  Google Scholar 

    19.
    Li, H. et al. Pilot-scale benthic microbial electrochemical system (BMES) for the bioremediation of polluted river sediment. J. Power Sources 356, 430–437 (2017).
    ADS  CAS  Google Scholar 

    20.
    Sherafatmand, M. & Ng, H. Y. Using sediment microbial fuel cells (SMFCs) for bioremediation of polycyclic aromatic hydrocarbons (PAHs). Bioresour. Technol. 195, 122–130 (2015).
    CAS  PubMed  Google Scholar 

    21.
    Sajana, T. K., Ghangrekar, M. M. & Mitra, A. Application of sediment microbial fuel cell for in situ reclamation of aquaculture pond water quality. Aquac. Eng. 57, 101–107 (2013).
    Google Scholar 

    22.
    Sajana, T. K., Ghangrekar, M. M. & Mitra, A. Effect of operating parameters on the performance of sediment microbial fuel cell treating aquaculture water. Aquac. Eng. 61, 17–26 (2014).
    Google Scholar 

    23.
    Giles, H. Using Bayesian networks to examine consistent trends in fish farm benthic impact studies. Aquaculture 274, 181–195 (2008).
    Google Scholar 

    24.
    Karakassis, I., Tsapakis, M., Hatziyanni, E., Papadopoulou, K. N. & Plaiti, W. Impact of cage farming of fish on the seabed in three Mediterranean coastal areas. ICES J. Mar. Sci. 57, 1462–1471 (2000).
    Google Scholar 

    25.
    Nøhr Glud, R., Gundersen, J. K., Barker Jørgensen, B., Revsbech, N. P. & Schulz, H. D. Diffusive and total oxygen uptake of deep-sea sediments in the eastern South Atlantic Ocean: in situ and laboratory measurements. Deep. Res. I 41, 1767–1788 (1994).
    Google Scholar 

    26.
    Van Duyl, F. C., Kop, A. J., Kok, A. & Sandee, A. J. J. The impact of organic matter and macrozoobenthos on bacterial and oxygen variables in marine sediment boxcosms. Neth. J. Sea Res. 29, 343–355 (1992).
    Google Scholar 

    27.
    Brooks, K. M. & Mahnken, C. V. Interactions of Atlantic salmon in the Pacific northwest environment II. Organic wastes. Fish. Res. 62, 255–293 (2003).
    Google Scholar 

    28.
    Mackin, J. E. & Swider, K. T. Organic matter decomposition pathways and oxygen consumption in coastal marine sediments. J. Mar. Res. 47, 681–716 (1989).
    CAS  Google Scholar 

    29.
    Holmer, M. & Kristensen, E. Impact of marine fish cage farming on metabolism and sulfate reduction of underlying sediments. Mar. Ecol. Prog. Ser. 80, 191–201 (1992).
    ADS  CAS  Google Scholar 

    30.
    Carroll, M. L., Cochrane, S., Fieler, R., Velvin, R. & White, P. Organic enrichment of sediments from salmon farming in Norway: Environmental factors, management practices, and monitoring techniques. Aquaculture https://doi.org/10.1016/S0044-8486(03)00475-7 (2003).
    Article  Google Scholar 

    31.
    Hargrave, B. T. Empirical relationships describing benthic impacts of salmon aquaculture. Aquac. Environ. Interact. 1, 33–46 (2010).
    Google Scholar 

    32.
    Bagarinao, T. Sulfide as an environmental factor and toxicant: tolerance and adaptations in aquatic organisms. Aquat. Toxicol. 24, 21–62 (1992).
    CAS  Google Scholar 

    33.
    Hargrave, B. T., Holmer, M. & Newcombe, C. P. Towards a classification of organic enrichment in marine sediments based on biogeochemical indicators. Mar. Pollut. Bull. 56, 810–824 (2008).
    CAS  PubMed  Google Scholar 

    34.
    Ryckelynck, N., Stecher, H. A. & Reimers, C. E. Understanding the anodic mechanism of a seafloor fuel cell: Interactions between geochemistry and microbial activity. Biogeochemistry 76, 113–139 (2005).
    Google Scholar 

    35.
    Ishii, S. et al. Identifying the microbial communities and operational conditions for optimized wastewater treatment in microbial fuel cells. Water Res. 47, 7120–7130 (2013).
    CAS  PubMed  Google Scholar 

    36.
    Fader, G.B.J. & Miller, R.O. Surficial Geology, Halifax Harbour, Nova Scotia. Bulletin of the Geological Survey of Canada (2008).

    37.
    Grant, J., Emerson, C. W., Hargrave, B. T. & Shortle, J. L. Benthic oxygen consumption on continental shelves off eastern Canada. Cont. Shelf Res. 11, 1083–1097 (1991).
    ADS  Google Scholar 

    38.
    Logan, B. E. Microbial fuel cells. In Treatise on Water Science, Vol. 4 (ed. Wilderer, P.) 641–665 (Wiley, New York, 2010).
    Google Scholar 

    39.
    Taillefert, M. et al. Early diagenesis in the sediments of the Congo deep-sea fan dominated by massive terrigenous deposits: part II—Iron—sulfur coupling. Deep. Res. II Top. Stud. Oceanogr. 142, 151–166 (2017).
    ADS  CAS  Google Scholar 

    40.
    Canfield, D. E., Raiswell, R. & Bottrell, S. The reactivity of sedimentary iron minerals toward sulfide. Am. J. Sci. 292, 659–683 (1992).
    ADS  CAS  Google Scholar 

    41.
    Boudreau, B. P. Diagenetic models and their implementation: modelling transport and reactions in aquatic sediments (Springer, Berlin Heidelberg, 1996).
    Google Scholar 

    42.
    Glud, R. N. Oxygen dynamics of marine sediments. Mar. Biol. Res. 4, 243–289 (2008).
    Google Scholar 

    43.
    Berg, P., Risgaard-petersen, N. & Silkeborg, D. Interpretation of measured concentration profiles in sediment pore water. Limnol. Oceanogr. 43, 1500–1510 (1998).
    ADS  CAS  Google Scholar 

    44.
    Hargrave, B. T. Seasonal changes in oxygen uptake by settled particulate matter and sediments in a marine bay. J. Fish. Res. Board Can. 35, 1621–1628 (1978).
    CAS  Google Scholar 

    45.
    Viggi, C. C. et al. Bridging spatially segregated redox zones with a microbial electrochemical snorkel triggers biogeochemical cycles in oil-contaminated River Tyne (UK) sediments. Water Res. 127, 11–21 (2017).
    CAS  PubMed  Google Scholar 

    46.
    Brüchert, V. & Arnosti, C. Anaerobic carbon transformation: Experimental studies with flow-through cells. Mar. Chem. 80, 171–183 (2003).
    Google Scholar 

    47.
    Arnosti, C. Microbial extracellular enzymes and their role in dissolved organic matter cycling. Aquat. Ecosyst. https://doi.org/10.1016/b978-012256371-3/50014-7 (2003).
    Article  Google Scholar 

    48.
    Lehman, R. M. & O’Connell, S. P. Comparison of extracellular enzyme activities and community composition of attached and free-living bacteria in porous medium columns. Appl. Environ. Microbiol. 68, 1569–1575 (2002).
    CAS  PubMed  PubMed Central  Google Scholar 

    49.
    Reimers, C. E. et al. Microbial fuel cell energy from an ocean cold seep. Geobiology 4, 123–136 (2006).
    CAS  Google Scholar 

    50.
    Jørgensen, B. B., Findlay, A. J. & Pellerin, A. The biogeochemical sulfur cycle of marine sediments. Front. Microbiol. 10, 849 (2019).
    PubMed  PubMed Central  Google Scholar 

    51.
    Lovley, D. R. Happy together: Microbial communities that hook up to swap electrons. ISME J. 11, 327–336 (2017).
    CAS  PubMed  Google Scholar 

    52.
    Finster, K., Liesack, W. & Thamdrup, B. Elemental sulfur and thiosulfate disproportionation by Desulfocapsa sulfoexigens sp. nov., a new anaerobic bacterium isolated from marine surface sediment. Appl. Environ. Microbiol. 64, 119–125 (1998).
    CAS  PubMed  PubMed Central  Google Scholar 

    53.
    Kelly, D. P., Shergill, J. K., Lu, W. P. & Wood, A. P. Oxidative metabolism of inorganic sulfur compounds by bacteria. Int. J. Gen. Mol. Microbiol. 71, 95–107 (1997).
    CAS  Google Scholar 

    54.
    Keeley, N. B., Forrest, B. M. & Macleod, C. K. Novel observations of benthic enrichment in contrasting flow regimes with implications for marine farm monitoring and management. Mar. Pollut. Bull. 66, 105–116 (2013).
    CAS  PubMed  Google Scholar 

    55.
    Cranford, P., Brager, L., Elvines, D., Wong, D. & Law, B. A revised classification system describing the ecological quality status of organically enriched marine sediments based on total dissolved sulfides. Mar. Pollut. Bull. 154, 111088 (2020).
    CAS  PubMed  Google Scholar 

    56.
    Soetaert, K., Hofmann, A. F., Middelburg, J. J., Meysman, F. J. R. & Greenwood, J. The effect of biogeochemical processes on pH. Mar. Chem. 105, 30–51 (2007).
    CAS  Google Scholar 

    57.
    Seitaj, D. et al. Cable bacteria generate a firewall against euxinia in seasonally hypoxic basins. Proc. Natl. Acad. Sci. USA 112, 13278–13283 (2015).
    ADS  CAS  PubMed  Google Scholar 

    58.
    Di Toro, D. M. et al. Acid volatile sulfide predicts the acute toxicity of cadmium and nickel in sediments. Environ. Sci. Technol. 26, 96–101 (1992).
    ADS  Google Scholar 

    59.
    Brooks, K. M. & Mahnken, C. V. W. Interactions of Atlantic salmon in the Pacific Northwest environment. III. Accumulation of zinc and copper. Fish. Res. 62, 295–305 (2003).
    Google Scholar 

    60.
    Fitridge, I., Dempster, T., Guenther, J. & de Nys, R. The impact and control of biofouling in marine aquaculture: A review. Biofouling 28, 649–669 (2012).
    PubMed  Google Scholar 

    61.
    FOA. The State of World Fisheries and Aquaculture 2016. Contributing to food security and nutrition for all (2016).

    62.
    Millero, F. J., Plese, T. & Fernandez, M. The dissociation of hydrogen sulfide in seawater. Limnol. Oceanogr. 33, 269–274 (1988).
    ADS  CAS  Google Scholar  More