More stories

  • in

    Old African fossils provide new evidence for the origin of the American crocodiles

    1.
    Uetz, P. & Hošek, J. The Reptile Database. https://www.reptile-database.org (2019). Accessed 5 Nov 2019.
    2.
    Oaks, J. R. A time-calibrated species tree of Crocodylia reveals a recent radiation of the true crocodiles. Evolution 65(11), 3285–3297 (2011).
    Article  Google Scholar 

    3.
    Nicolaï, M. P. J. & Matzke, N. J. Trait-based range expansion aided in the global radiation of Crocodylidae. Glob. Ecol. Biogeogr. 1, 2. https://doi.org/10.1111/geb.12929 (2019) (in press).
    Article  Google Scholar 

    4.
    Hekkala, E. et al. An ancient icon reveals new mysteries: mummy DNA resurrects a cryptic species within the Nile crocodile. Mol. Ecol. 20, 4199–4215 (2011).
    CAS  Article  Google Scholar 

    5.
    Meredith, R. W., Hekkala, E. R., Amato, G. & Gatesy, J. A phylogenetic hypothesis for Crocodylus (Crocodylia) based on mitochondrial DNA: evidence for a trans-Atlantic voyage from Africa to the New World. Mol. Phylogenetics Evol. 60, 183–191 (2011).
    Article  Google Scholar 

    6.
    Brochu, C. A. Phylogenetic relationships and divergence timing of Crocodylus based on morphology and the fossil record. Copeia 2000(3), 657–673 (2000).
    Article  Google Scholar 

    7.
    Brochu, C. A. Congruence between physiology, phylogenetics and the fossil record on crocodylian historical biogeography. In Crocodilian Biology and Evolution (eds Grigg, G. C. et al.) 9–28 (Surrey Betty & Sons, New South Wales, 2001).
    Google Scholar 

    8.
    Brochu, C. A. Phylogenetic approaches toward crocodylian history. Annu. Rev. Earth Planet. Sci. 31, 357–397 (2003).
    ADS  CAS  Article  Google Scholar 

    9.
    Brochu, C. A. & Storrs, G. W. A giant crocodile from the Plio-Pleistocene of Kenya, the phylogenetic relationships of Neogene African crocodylines, and the antiquity of Crocodylus in Africa. J. Vertebr. Paleontol. 32(3), 587–602 (2012).
    Article  Google Scholar 

    10.
    Fermino, B. R. et al. Shared species of crocodilian trypanosomes carried by tabanid flies in Africa and South America, including the description of a new species from caimans Trypanosoma kaiowa n. sp. Parasites Vectors 12(1), 225 (2019).
    Article  Google Scholar 

    11.
    Hekkala, E. R., Amato, G., DeSalle, R. & Blum, M. J. Molecular assessment of population differentiation and individual assignment potential of Nile crocodile (Crocodylus niloticus) populations. Conserv. Genet. 11, 1435–1443 (2010).
    Article  Google Scholar 

    12.
    Schmitz, A. et al. Molecular evidence for species level divergence in African Nile Crocodiles Crocodylus niloticus (Laurenti, 1786). C. R. Palevol. 2(8), 703–712 (2003).
    Article  Google Scholar 

    13.
    Hecht, M. K. Fossil snakes and crocodilians from the Sahabi Formation of Libya. In Neogene Palaeontology and Geology of Sahabi (eds Boaz, N. T. et al.) 101–106 (Alan R. Liss. Inc., New York, 1987).
    Google Scholar 

    14.
    Maccagno, A. M. Descrizione di una nuova specie di “Crocodilus” del Giacimento di Sahabi (Sirtica). Atti Accad. Naz. Lincei Mem. Classe Sci. Fis. Mat. Nat. Sez. II Fis. Chim. Geol. Paleontol. Mineral. Ser. VIII 1(2), 61–96 (1947).
    Google Scholar 

    15.
    Maccagno, A. M. I coccodrilli di Sahabi. Rend. Accad. Naz. XL. 4(3), 73–117 (1952).
    Google Scholar 

    16.
    Rook, L. The discovery of Sahabi site: Ardito Desio or Carlo Petrocchi? In Boaz, N.T. El-Arnauti, A., Pavlakis, P. & Salem, M. (eds) Circum-mediterranean neogene geology and biotic evolution: the perspective from Libya. Garyounis Sci. Bull. 5, 11–18 (2008).
    Google Scholar 

    17.
    Pandolfi, L. & Rook, L. The latest Miocene Rhinocerotidae from Sahabi (Lybia). C. R. Palevol. 18, 442–448 (2019).
    Article  Google Scholar 

    18.
    Delfino, M. New remains of Crocodylus checchiai Maccagno, 1947 (Crocodylia, Crocodylidae) from the Late Miocene of As Sahabi, Libya. In Boaz N. T., El-Arnauti A., Pavlakis P. & Salem M. (eds). Circum-Mediterranean geology and biotic evolution during the Neogene Period: the perspective from Libya. Garyounis Sci. Bull. 5, 111–118 (2008).
    Google Scholar 

    19.
    Scheyer, T. M. et al. Crocodylian diversity peak and extinction in the late Cenozoic of the northern Neotropics. Nat. Commun. 4, 1907. https://doi.org/10.1038/ncomms2940 (2013).
    ADS  CAS  Article  PubMed  Google Scholar 

    20.
    El-Shawaihdi, M. H. Stratigraphy of the Neogene Sahabi units in the Sirt Basin, northeast Libya. J. Afr. Earth Sci. 118, 87–106 (2016).
    ADS  Article  Google Scholar 

    21.
    Bernor, R. L. & Rook, L. A current view of Sahabi large mammal biogeographic relationships. In Boaz, N. T., El-Arnauti, A., Pavlakis, P. & Salem, M. (eds) Circum-mediterranean neogene geology and biotic evolution: the perspective from Libya. Garyounis Sci. Bull. 5, 285–292 (2008).
    Google Scholar 

    22.
    Bernor, R. L., Boaz, N. T. & Rook, L. Eurygnathohippus feibeli (Perissodacyla: Mammalia) from the Late Miocene of As Sahabi (Libya) and its evolutionary and biogeographic significance. Boll. Soc. Paleont. Ital. 51, 39–48 (2012).
    Google Scholar 

    23.
    Gradstein, F., Ogg, J. & Smith, A. A Geologic Time Scale 2012 , Vol. 1176 (Cambridge University Press, Cambridge, 2012).
    Google Scholar 

    24.
    Tchernov, E. Evolution of the Crocodiles in East and North Africa (CNRS, Paris, 1986).
    Google Scholar 

    25.
    Leakey, M. G. et al. Lothagam: a record of faunal change in the Late Miocene of East Africa. J. Vertebr. Paleontol. 16(3), 556–570 (1996).
    Article  Google Scholar 

    26.
    Brochu, C. A. Pliocene crocodiles from Kanapoi, Turkana Basin Kenya . J. Hum. Evol. 140, 102410 (2020).
    Article  Google Scholar 

    27.
    Moreno-Bernal, J. W., Head, J. & Jaramillo, C. A. Fossil crocodilians from the High Guajira Peninsula of Colombia: Neogene faunal change in northernmost South America. J. Vertebr. Paleontol. 36(3), e1110586 (2016).
    Article  Google Scholar 

    28.
    Delfino, M., Böhme, M. & Rook, L. First European evidence for transcontinental dispersal of Crocodylus (late Neogene of southern Italy). Zool. J. Linn. Soc. 149, 293–307 (2007).
    Article  Google Scholar 

    29.
    Delfino, M. & Rook, L. African crocodylians in the Late Neogene of Europe. A revision of Crocodylus bambolii Ristori, 1890. J. Paleontol. 82(2), 336–343 (2008).
    Article  Google Scholar 

    30.
    Delfino, M. & Rossi, M. A. Fossil crocodylid remains from Scontrone (Tortonian, Southern Italy) and the Late Neogene Mediterranean biogeography of crocodylians. Geobios 46, 25–31 (2013).
    Article  Google Scholar 

    31.
    Montoya, P. et al. Las nuevas excavaciones (1995–2006) en el yacimiento del Mioceno final de Venta del Moro Valencia. Estud. Geol. 62, 313–326 (2006).
    Article  Google Scholar 

    32.
    Taplin, L. E. & Grigg, G. C. Historical zoogeography of the eusuchian crocodilians: a physiological perspective. Am. Zool. 29, 885–901 (1989).
    Article  Google Scholar 

    33.
    Brochu, C. A., Nieves-Rivera, ÁM., Vélez-Juarbe, J., Daza-Vaca, J. D. & Santos, H. Tertiary crocodylians from Puerto Rico: evidence for Late Tertiary endemic crocodylians in the West Indies?. Geobios 40, 51–59 (2007).
    Article  Google Scholar 

    34.
    Mannion, P. D. et al. Climate constrains the evolutionary history and biodiversity of crocodylians. Nat. Commun. 6, 8438 (2015).
    ADS  CAS  Article  Google Scholar 

    35.
    Díaz Aráez, J. L. et al. New remains of Diplocynodon (Crocodylia: Diplocynodontidae) from the Early Miocene of the Iberian Peninsula. C. R. Palevol. 16, 12–26 (2017).
    Article  Google Scholar 

    36.
    Iurino, D. A., Danti, M., Della Sala, S. W. & Sardella, R. Modern techniques for ancient bones: vertebrate palaeontology and medical CT analysis. Boll. Soc. Paleontol. Ital. 52, 145–155 (2013).
    Google Scholar 

    37.
    Iurino, D. A. & Sardella, R. CT scanning analysis of Megantereon whitei (Carnivora, Machairodontinae) from Monte Argentario (Early Pleistocene, central Italy): evidence of atavistic teeth. Naturwissenschaften 101(12), 1099–1106 (2014).
    ADS  CAS  Article  Google Scholar 

    38.
    Lautenschlager, S. & Butler, R. J. Neural and endocranial anatomy of Triassic phytosaurian reptiles and convergence with fossil and modern crocodylians. PeerJ. 4, e2251 (2016).
    Article  Google Scholar 

    39.
    Cherin, M. et al. Synchrotron radiation reveals the identity of the large felid from Monte Argentario (Early Pleistocene, Italy). Sci. Rep. 8(1), 8338 (2018).
    ADS  Article  Google Scholar 

    40.
    Goloboff, P. A., Farris, J. S. & Nixon, K. C. TNT, a free program for phylogenetic analysis. Cladistics 24(5), 774–786 (2008).
    Article  Google Scholar 

    41.
    Goloboff, P. A. Parsimony and weighting: a reply to Turner and Zandee. Cladistics 11(1), 91–104 (1995).
    Article  Google Scholar  More

  • in

    Discovery of a new mode of oviparous reproduction in sharks and its evolutionary implications

    Two modes of oviparity, i.e. single and multiple oviparity, are currently recognized in chondrichthyans1,2,3,5,6,7,8,9,10,11,12. Single oviparity (Fig. 3a) is a mode where each oviduct in pregnant females contains one egg case, i.e. a pair of egg cases in a pregnant female. These egg cases are retained in the oviduct only for a short time and deposited immediately before the embryo has begun developing. The embryos are not recognizable at oviposition and become visible in a few weeks. Oviposition is repeatedly performed, and each mature female can deposit tens of egg cases over the course of a spawning season6 (“Short single” oviparity in Fig. 4). Multiple oviparity (Fig. 3b) is a mode where several egg cases accumulate in each oviduct and are retained for several months before oviposition, in which time embryos begin development in the oviduct and the egg cases are deposited later when the embryos grow large to a certain developmental stage (“Multiple” oviparity in Fig. 4).
    Figure 3

    Three modes of oviparity in catsharks, showing egg cases in oviduct. (a) Short single oviparity (Galeus sauteri from Taiwan, uncatalogued), (b) multiple oviparity (Halalelurus buergeri, 410 mm TL from Kagoshima, Japan, uncatalogued), (c) Sustained single oviparity (Cephaloscyllium sarawakensis), (c1) egg cases without developing embryo (NMMB-P30890, 80.0 mm ECL), (c2) egg cases with a developing embryo in each (NMMB-P 30888, 81.6 mm, 83.0 mm ECL). Top three photographs (a,b,c1) cover whole abdominal cavities, showing difference of relative sizes of egg cases in three modes of oviparity. Cephaloscyllium sarawakensis (c1) has huge egg cases occupying most of the abdominal cavity.

    Full size image

    Figure 4

    Schematic diagram of five modes of reproduction in catsharks, showing differences of succession, duration and condition of egg cases/ embryos per one oviduct/uterus. Numerals show order of egg case produced in the oviduct/uterus.

    Full size image

    Cephaloscyllium sarawakensis does not fit the classic single oviparity, nor multiple oviparity. Pregnant females of this species always have a single egg case, never two or more, in each oviduct (Fig. 3c), and keep it until embryo attains a certain developmental stage (“Sustained single” oviparity in Fig. 4). These facts indicate that reproduction of C. sarawakensis represents a new mode of oviparity, which is herein termed “sustained single oviparity”. A 450 mm TL female of Cephaloscyllium silasi from the Indian Ocean had one egg case with a well-developed embryo in each oviduct13. Although they reported only one female specimen, this species possibly also displays the sustained single oviparity.
    Various technical terms have been used for oviparity. Single oviparity has at least three alternative names, “extended” oviparity3,4,6,12,14, “external” oviparity5,15, and “simple” oviparity5. Multiple oviparity has been also termed “retained” oviparity3,4,5,6,12,14,16. These terms are used for the same reproductive mode, and this could lead to confusion and misunderstanding about the oviparity. Therefore, we herein summarized the terms of oviparity and proposed a new set of technical terms as follows: (1) “short single oviparity” for the single oviparity previously known; (2) “sustained single oviparity” for the new type of oviparity reported in this study; and (3) “multiple oviparity” instead of the former “retained” oviparity. The new definition of oviparity in the cartilaginous fishes was summarized, with additions of two modes of yolk-sac viviparity in catsharks (Table 1).
    Table 1 New definitions of oviparity and yolk-sac viviparity in lecithotrophic (yolk-dependent) cartilaginous fishes.
    Full size table

    Typically in the catsharks displaying short single oviparity, the shell of egg case is tough and thick, with two pairs of long strong tendrils on its anterior and posterior ends (Fig. 5a‒c,e,f), although tendrils are very short or replaced by fine silky materials (Fig. 5d) or absent in some species. The posterior pair of tendrils is used to attach the egg case to the substrate on the sea bottom, to pull it out from the oviduct, and coil it around the substrate, together with anterior pair of tendrils. The egg case is firmly secured on the substrate until juvenile hatches. The tendrils in the multiple oviparous species (Fig. 5f) tend to be thinner and shorter than those of the short single oviparous species. Cephaloscyllium sarawakensis of the sustained single oviparity has a thick shell and long tendrils (Fig. 1b,c,d1). The two egg cases (Fig. 2a) collected from fishery landings can be inferred to have been deposited intentionally on the tube of a tubeworm Paradiopatra sp. based on the fact that the posterior tendrils are firmly twined around it.
    Figure 5

    Egg cases of catsharks. (a) Cephaloscyllium laticeps from Australia, (b) Cephaloscyllium umbratile from Japan, (c) Poroderma africanum from Ibaraki Prefectural Oarai Aquarium, Japan, (d) Galeus sauteri from Taiwan, (e) Haploblepharus fuscus from Ibaraki Prefectural Oarai Aquarium, Japan, (f) Halaelurus buergeri from Japan. (a‒e) short single oviparous species, (f) multiple oviparous species. Scales 30 mm.

    Full size image

    The egg cases in the species of sustained single oviparity are very large, with its length (ECL) ranging 16.5‒20.1% TL in Cephaloscyllium sarawakensis and 18.9‒19.2% TL in C. silasi13, while the egg cases of short single oviparous C. umbratile are 10.6‒15.1% TL2,17,18. The egg cases of other short single oviparous species are also small, with lengths 8.6‒9.0% TL in Holohalaelurus regani19, 9.2% TL in Galeus sauteripresent study, 9.2‒14.8% TL in four species of Apristurus20,21,22,23, 10.6‒14.9% TL in Atelomycterus marmoratus24, 11.4% TL in Schroederichthys maculatus25, 11.5‒11.9% TL in Scyliorhinus torazame26,presentstudy, 11.8% TL in S. capensis27, 12.0% TL in Bythaelurus dawsoni28 and 13.5‒18.9% TL from Fig. 10d29 of Parmaturus xaniurus. The egg case lengths of multiple oviparous species are about 11% TL in Halaelurus buergeripresent study and 10.4‒10.9% TL in H. quagga30. As seen in Fig. 3, the egg cases of C. sarawakensis (Fig. 3c1) are far larger than those of G. sauteri (Fig. 3a) and H. buergeri (Fig. 3b), occupying most of the available space of oviduct and abdominal cavity. Thus, the two species of sustained single oviparity have much larger egg cases than short single oviparous and multiple oviparous species, suggesting larger neonates at hatching in C. sarawakensis and C. silasi.
    The embryos in the egg cases recorded from the oviduct indicate that Cephaloscyllium sarawakensis retains the egg case until the embryo attains about 102 mm TL (Fig. 1d; largest embryo in the egg inside the oviduct), and then oviposition occurs later. However, the two egg cases (both 80.0 mm ECL) from the seabed contained a 65 mm TL and a 94.5 mm TL embryo each (Fig. 2a). These facts suggest the timing of oviposition is rather wide in this species, and the egg cases are laid when the embryo grows roughly 6‒10 + cm TL, stimulated by some internal or external factors. The smallest free-swimming juvenile collected was 125 mm TL with a remnant of external yolk-sac (Fig. 2b), suggesting the hatching size from egg case being around 120 mm TL in this species. These evidences indicate C. sarawakensis keeps the egg case in the oviduct until embryo has developed to 50 ~ 80+ % of its hatching size. Therefore, the retention of egg case in the oviduct continues for an extended period, perhaps several months or more, in C. sarawakensis and probably also in C. silasi.
    The other remarkable characteristic of Cephaloscyllium sarawakensis is the glassy transparent egg cases (Fig. 1b‒d), and the transparency is completely maintained even after oviposition (Fig. 2a). The egg cases of oviparous cartilaginous fishes are opaque, usually yellowish to dark brownish (Fig. 5), and sometimes with longitudinal or transverse ridges (Fig. 5a). The functional role of egg case is to protect the embryo from physical, physiological and biological hazards from the environment. The colored egg cases can be also effective to conceal the embryo in it. However, the egg case of C. sarawakensis is transparent and never cryptic that the orange yellow yolk would clearly be recognizable through transparent egg case, if the egg case is deposited immediately after egg case being formed.
    As shown in this study, C. sarawakensis retains an egg case in each oviduct until the embryo is developed with a distinct dark polka-dot color pattern on light brownish body (Figs. 1d, 2a), typically seen in the juveniles (Fig. 2b). One of the reasons for transparent egg case may be related to their vivid body color patterns. Benthic and reef-dwelling sharks have complicated color patterns, which are effective to blend the body into their background or for camouflage31. The present egg cases of C. sarawakensis (Fig. 2a) were deposited around the tube of a tubeworm sticking out from the seabed. Their vivid polka-dots and light brownish body coloration could function as more effective camouflage against the complex and dark background through the transparent egg case. Thus, the long retention of transparent egg cases and vivid embryonic coloration could suggest a new method of reproductive tactics in cartilaginous fishes.
    The oviparity is advantageous as a method to increase the fecundity in small elasmobranchs that have limited space in body cavity for care and storage of the embryos6. The species of short single oviparity (Fig. 3a) repeatedly deposits two egg cases immediately after the cases are completed and laid, resulting in 20‒100 eggs per season6. The captive Cephaloscyllium umbratile was recorded as depositing two egg cases at intervals of 11‒38 days (20 days in average) for whole year32, which means a single female deposited about 36 egg cases a year. Similarly, the captive C. laticeps laid two egg cases at intervals of up to 28 days throughout whole year11, equating more than 26 egg cases being deposited annually.
    The multiple oviparous species retains a number of egg cases in each oviduct (Fig. 3b) for several months until the embryos have developed to a certain stage. All the species of Halaelurus are multiple oviparous, and H. buergeri has been recorded to deposit 8 egg cases one by one at a stage when the embryos inside have attained 70 mm TL33, or 10 egg cases at one time34. One specimen of H. buergeri we (KN) collected (Fig. 3b) had 10 egg cases in the oviducts with a developing embryo in each. A captive H. maculosus deposited 11 egg cases containing 50‒70 mm TL embryos in five days (personal communication with Mr. K. Tokunaga of Ibaraki Prefectural Oarai Aquarium). The fecundity of H. lineatus is up to 16 eggs at a time35.
    The maternal environment offers the best protection to the developing embryos and can shorten the exposed period of time on the substrate. Therefore, the survival rate could be expected to be much higher in the sustained single and multiple oviparous species than the short single oviparous species. The species of the sustained single oviparity (Fig. 3c) and those of the multiple oviparity (Fig. 3b) are the same in that the egg cases have long maternal protection, but the number of the egg cases deposited at a time is considerably less in the sustained single oviparous species, i.e. 2 eggs vs. 4‒16 eggs per mother, respectively. Hence, the fecundity of Cephaloscyllium sarawakensis could be very low, 1/8–1/2 of the multiple oviparous species (see “Oviparity” in Fig. 4).
    It is crucial to produce a certain number of offspring to maintain a sustainable population, and the very low fecundity in C. sarawakensis could be decisively disadvantageous for the species. Similar issues exist in the yolk-sac viviparous species. The yolk-sac viviparous catsharks, such as Bythaelurus clevai, B. hispidus, B. lutarius, B. stewarti and Cephalurus cephalus retain only one embryo per uterus or per mother1,28,36,37,38,KN pers.obs, which is hence termed here “single pregnancy”. These species of single pregnancy would have also lower fecundity than the species of “multiple pregnancy” seen in Galeus polli (see “Yolk-sac viviparity” in Fig. 4).
    Species of the genus Cephaloscyllium are generally large in body sizes, mostly growing to more than 70 cm TL and some species (C. isabellum, C. laticeps, and C. umbratile) attain more than 100 cm TL39, whereas C. sarawakensis and C. silasi are dwarf species within the genus. Cephaloscyllium sarawakensis attains a maximum of only 39.7 cm TL in males and 49.5 cm TL in females40,present study, and matures at the sizes less than 32.5 cm TL and 35.4 cm TL in males and females, respectively41. Similarly, Cephaloscyllium silasi attains only 50 cm TL13 and reaches its maturity at less than 36.8 cm TL in males1 and less than 45 cm TL in females13. Bythaelurus clevai, B. hispidus and B. lutarius attain 42 cm TL, 36 cm TL and 39 cm TL, respectively37,39 and these species are also the smallest species for the genus. Cephalurus cephalus reaches 30 cm TL36,39, and this is known as one of the smallest sharks1.
    The ratio of length at maturity to the largest total length was reported at around 0.73 for elasmobranchs42, and this indicates the smaller species could reach their maturity at smaller sizes than the larger species. Actually, the captive Cephaloscyllium umbratile which grows up to 118 cm TL hatches at lengths of 16–22 cm TL32 and attains its full maturity at 96‒104 cm TL17. In contrast, C. sarawakensis which produces very large egg cases relative to the mother size is expected to hatch at about 12 cm TL and mature at less than 35 cm TL. Therefore, C. sarawakensis grows only about 23 cm in length until maturity is attained, whereas C. umbratile needs to grow 75–90 cm to its maturity. Cephalurus cephalus produces 7–9 cm TL neonates and attains maturity at 18–22 cm TL43, thus 9–15 cm in length to grow to attain maturity. Eridacnis radcliffei gives birth 10.5‒12.8 cm TL neonates36, and attains maturity at 18.3 cm TL, only 5.5‒7.8 cm in length to the maturity.
    Hence, these dwarf sharks of the sustained single oviparity and the yolk-sac viviparity of single pregnancy likely attain their maturity in a shorter time frame than the larger species do, enabling them to reproduce at an earlier age and keep their life-time fecundity high. Other factors to increase their lifetime fecundity include quick repetition of reproduction, longer lifetime reproduction and higher proportion of females, but these will not be referred to here.
    Figure 6a shows five modes of lecithotrophic (yolk-dependent) reproduction in the cartilaginous fishes, and Fig. 6b1–6 denote six combinations of these reproductive modes in the catsharks. Figure 6b1 (short single oviparity) represents ten genera such as Apristurus, Asymbolus, Atelomycterus, Figaro, Haploblepharus, Holohalaelurus, Parmaturus, Poroderma, Scyliorhinus and Schroederichthys. Figure 6b3 (multiple oviparity) and Fig. 6b6 (yolk-sac viviparity of single pregnancy) are Halaelurus and Cephalurus, respectively. However, Fig. 6b2,b4,b5 include two or three modes of reproduction. Cephaloscyllium (Fig. 6b2) performs short single oviparity + sustained single oviparity, Bythaelurus (Fig. 6b4) involves short single oviparity + yolk-sac viviparity of single pregnancy, and Galeus (Fig. 6b5) performs short single oviparity + multiple oviparity + yolk-sac viviparity of multiple pregnancy. These facts could suggest rather facile diversification of reproductive mode in closely related species, or may suggest necessity of some taxonomic reconsideration of them, as indicated for Bythaelurus38.
    Figure 6

    Modes of reproduction (a) and combination of the modes at generic level in catsharks (b). (a) Five modes of lecithotrophy (yolk-dependent reproduction) in cartilaginous fishes, (b1) Apristurus, Asymbolus, Atelomycterus, Figaro, Haploblepharus, Holohalaelurus, Parmaturus, Poroderma, Scyliorhinus and Schroederichthys, (b2) Cephaloscyllium, (b3) Halaelurus, (b4) Bythaelurus, (b5) Galeus, and (b6) Cephalurus.

    Full size image

    Oviparity has been suggested to be the ancestral mode of reproduction for vertebrates7,44, and it has also traditionally been believed as the ancestral mode for chondrichthyan fishes3,14,45. However, recent studies5,6,12,46,47 suggest that viviparity is ancestral for all chondrichthyans, with many reversions to oviparity and secondary reversions to viviparity. Phylogenetic interrelationships for the Galeomorphi (orders Heterodontiformes, Orectolobiformes, Lamniformes and Carcharhiniformes)4,45,48‒50 show that short single oviparity is the ancestral mode for the Galeomorphi, and also for the orders Heterodontiformes, Orectolobiformes and Carcharhiniformes. Multiple oviparity is generally considered to have evolved from short single oviparity5, or evolved intermediately between the short single oviparity and the yolk-sac viviparity1,7,14,51.
    The catsharks (now Pentanchidae and Scyliorhinidae) in the Carcharhiniformes are separated into a few isolated groups, based on genetic works49,50,52. Mapping of the reproductive modes on their phylogenetic relationships suggests that the short single oviparity is ancestral for each group. According to the phylogenetic result50 which covers more catshark taxa than the other works, their Scyliorhinidae I50 (Fig. 7a) includes nine genera of the family Pentanchidae. Six genera of these, i.e. Apristurus, Asymbolus, Figaro, Haploblepharus, Holohalaelurus and Parmaturus, display short single oviparity. The multiple oviparous genus Halaelurus is sister to the groups of short single oviparous catsharks, and the relationships suggest the multiple oviparity has derived from the short single oviparity.
    Figure 7

    Reproductive modes mapped on simplified relationship of (a) Scyliorhinidae I50 and (b) Bythaelurus53, and suggested evolution.

    Full size image

    The genus Bythaelurus, which is also deeply merged in groups of short single oviparous species in the Scyliorhinidae I50 (Fig. 7a), is currently comprised of 14 species38,53,54, with five short single oviparous species (B. bachi, B. canescens, B. dawsoni, B. naylori and B. vivaldi) and four yolk-sac viviparous species of single pregnancy (B. clevai, B. hispidus, B. lutarius and B. stewarti) (Fig. 6b4). Interrelationships of five Bythaelurus species53 (Fig. 7b) show they are clearly separable in two groups, i.e. short single oviparous species (B. bachi, B. naylori, B. dawsoni and B. canescens) and yolk-sac viviparous species of single pregnancy (B. hispidus). The short single oviparous B. dawsoni and B. canescens have a sister relation with short single oviparous genera Asymbolus + Figaro in the Scyliorhinidae I50 (Fig. 7a). These facts suggest that the short single oviparity is ancestral for Bythaelurus and the yolk-sac viviparity of single pregnancy could have derived from the short single oviparity1, maybe via sustained single oviparity.
    The genus Galeus contains 18 species with three reproductive modes (Fig. 6b5), i.e. short single oviparity (G. antillensis and seven other species), multiple oviparity (G. atlanticus, G. melastomus and G. piperatus) and yolk-sac viviparity of multiple pregnancy (G. polli). The genus Galeus is deeply embedded within groups of short single oviparous species in the Scyliorhinidae I50 (Fig. 7a), and the short single oviparity is considered to be ancestral for the genus Galeus, and the yolk-sac viviparity of multiple pregnancy in G. polli could have derived from short single oviparity via multiple oviparity.
    Their Scyliorhinidae II50 includes three genera Atelomycterus, Schroederichthys of short single oviparity and oviparous Aulohalaelurus55. Scyliorhinidae III50 includes three genera Cephaloscyllium, Scyliorhinus and Poroderma, and they all display short single oviparity. Cephaloscyllium sarawakensis and C. silasi of sustained single oviparity were not treated in their analysis50, but the relationships of Cephaloscyllium in the Scyliorhinidae III50 that is composed of short single oviparous species could suggest derivation of the sustained single oviparity directly from short single oviparity by longer retention of one egg case in an oviduct.
    The modes of reproduction in the catsharks were summarized in Table 2, and the phylogenetic evidences mentioned above suggest: (1) short single oviparity is ancestral for the catsharks; (2) more diverse modes of reproduction evolved in the family Pentanchidae than family Scyliorhinidae; (3) sustained single oviparity in Cephaloscyllium was derived directly from short single oviparity; (4) multiple oviparity in Halaelurus was derived from short single oviparity; (5) multiple oviparity in Galeus was derived from short single oviparity, and originated yolk-sac viviparity of multiple pregnancy; (6) yolk-sac viviparity of single pregnancy in Bythaelurus was derived from short single oviparity, possibly via sustained single oviparity; and (7) yolk-sac viviparity of single pregnancy in Cephalurus was derived possibly from short single oviparity via sustained single oviparity.
    Table 2 Modes of reproduction and evolution in catsharks.
    Full size table

    Materials and methods
    All specimens of Cephaloscyllium sarawakensis examined were bycatch from commercial bottom trawlers operating in the South China Sea off southwest Taiwan, and were collected at Hsin-da port (HD) and Ke-tzu-liao (KTL) in Kaohsiung. Specimens were fixed in 4% formalin and then transferred to 70% Ethanol or 50% Isopropanol ethanol. All specimens were deposited at Pisces collection of the National Museum of Marine Biology & Aquarium, Pingtung, Taiwan (NMMB-P). Total length (TL) was measured using a ruler or digital caliper, to nearest 1 or 0.1 mm, respectively.
    Egg cases (Table 3): a total of 8 egg cases without visible embryo on the yolk, and 15 egg cases with a developed embryo each were collected from the oviduct of thirteen specimens of C. sarawakensis. Two egg cases with a developed embryo tied on the tube of a tubeworm Paradiopatra sp. (family Onuphidae, order Eunicida) were collected from the fishery landings by fishers, and were kept frozen. Egg cases were deposited and catalogued in NMMB-P collection. Length (excluding tendrils, ECL) and width (ECW) of the egg case were measured by a ruler or digital caliper.
    Table 3 Measurements of egg cases and embryos, and sampling data of Cephaloscyllium sarawakensis.
    Full size table

    Juveniles: NMMB-P22719, 125 mm TL female, KTL, 2 Apr. 2015; NMMB-P17143, 143 mm TL male; NMMB-P24872 (1 of 6 specimens), 134 mm TL male, KTL, 18 Mar. 2016. More

  • in

    Effects of prey trophic mode on the gross-growth efficiency of marine copepods: the case of mixoplankton

    1.
    Almeda, R. et al. Trophic role and carbon budget of metazoan microplankton in northwest Mediterranean coastal waters. Limnol. Oceanogr. 56, 415–430 (2011).
    ADS  CAS  Google Scholar 
    2.
    Hansen, B., Bjornsen, P. K. & Hansen, P. J. The size ratio between planktonic predators and their prey. Limnol. Oceanogr. 39, 395–403 (1994).
    ADS  Google Scholar 

    3.
    Saiz, E., Griffell, K., Calbet, A. & Isari, S. Feeding rates and prey: predator size ratios of the nauplii and adult females of the marine cyclopoid copepod Oithona davisae. Limnol. Oceanogr. 59, 2077–2088 (2014).
    ADS  Google Scholar 

    4.
    Møller, E. F. Production of dissolved organic carbon by sloppy feeding in the copepods Acartia tonsa, Centropages typicus, and Temora longicornis. Limnol. Oceanogr. 52, 79–84 (2007).
    ADS  Google Scholar 

    5.
    Saiz, E., Calbet, A., Irigoien, X. & Alcaraz, M. Copepod egg production in the western Mediterranean: response to food availability in oligotrophic environments. Mar. Ecol. Prog. Ser. 187, 179–189 (1999).
    ADS  Google Scholar 

    6.
    Kiørboe, T. & Saiz, E. Planktivorous feeding in calm and turbulent environments, with emphasis on copepods. Mar. Ecol. Prog. Ser. 122, 135–146 (1995).
    ADS  Google Scholar 

    7.
    Broglio, E., Jónasdóttir, S. H., Calbet, A., Jakobsen, H. H. & Saiz, E. Effect of heterotrophic versus autotrophic food on feeding and reproduction of the calanoid copepod Acartia tonsa: relationship with prey fatty acid composition. Aquat. Microb. Ecol. 31, 267–278 (2003).
    Google Scholar 

    8.
    Kleppel, G. S. & Burkart, C. A. Egg production and the nutritional environment of Acartia tonsa: the role of food quality in copepod nutrition. ICES J. Mar. Sci. 52, 297–304 (1995).
    Google Scholar 

    9.
    Tang, K. W. & Dam, H. G. Limitation of zooplankton production: beyond stoichiometry. Oikos 84, 537–542 (1999).
    Google Scholar 

    10.
    Twining, B. S., Baines, S. B. & Fisher, N. S. Element stoichiometries of individual plankton cells collected during the Southern Ocean Iron Experiment (SOFeX). Limnol. Oceanogr. 49, 2115–2128 (2004).
    ADS  CAS  Google Scholar 

    11.
    Saiz, E. & Calbet, A. Copepod feeding in the ocean: scaling patterns, composition of their diet and the bias of estimates due to microzooplankton grazing during incubations. Hydrobiologia 666, 181–196 (2011).
    CAS  Google Scholar 

    12.
    Broglio, E., Saiz, E., Calbet, A., Trepat, I. & Alcaraz, M. Trophic impact and prey selection by crustacean zooplankton on the microbial communities of an oligotrophic coastal area (NW Mediterranean Sea). Aquat. Microb. Ecol. 35, 65–78 (2004).
    Google Scholar 

    13.
    Irigoien, X. et al. Copepod hatching success in marine ecosystems with high diatom concentrations. Nature 419, 387–389 (2002).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Stoecker, D. K., Hansen, P. J., Caron, D. A. & Mitra, A. Mixotrophy in the marine plankton. Ann. Rev. Mar. Sci. 9, 311–335 (2017).
    PubMed  Google Scholar 

    15.
    Flynn, K. J. et al. Mixotrophic protists and a new paradigm for marine ecology: where does plankton research go now?. J. Plankton Res. 41, 375–391 (2019).
    Google Scholar 

    16.
    Mitra, A. et al. Defining planktonic protist functional groups on mechanisms for energy and nutrient acquisition: incorporation of diverse mixotrophic strategies. Protist 167, 106–120 (2016).
    CAS  PubMed  Google Scholar 

    17.
    Ptacnik, R., Sommer, U., Hansen, T. & Martens, V. Effects of microzooplankton and mixotrophy in an experimental planktonic food web. Limnol. Oceanogr. 49, 1435–1445 (2004).
    ADS  Google Scholar 

    18.
    Moorthi, S. D. et al. The functional role of planktonic mixotrophs in altering seston stoichiometry. Aquat. Microb. Ecol. 79, 235–245 (2017).
    Google Scholar 

    19.
    Ward, B. A. & Follows, M. J. Marine mixotrophy increases trophic transfer efficiency, mean organism size, and vertical carbon flux. Proc. Natl. Acad. Sci. 113, 2958–2963 (2016).
    ADS  CAS  PubMed  Google Scholar 

    20.
    Waggett, R. J., Tester, P. A. & Place, A. R. Anti-grazing properties of the toxic dinoflagellate Karlodinium veneficum during predator-prey interactions with the copepod Acartia tonsa. Mar. Ecol. Prog. Ser. 366, 31–42 (2008).
    ADS  Google Scholar 

    21.
    Xu, J., Nielsen, L. T. & Kiørboe, T. Foraging response and acclimation of ambush feeding and feeding-current feeding copepods to toxic dinoflagellates. Limnol. Oceanogr. 63, 1449–1461 (2018).
    ADS  Google Scholar 

    22.
    Berge, T., Poulsen, L. K., Moldrup, M., Daugbjerg, N. & Hansen, P. J. Marine microalgae attack and feed on metazoans. ISME J. 6, 1926–1936 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    23.
    Vaqué, D. et al. Effects of the toxic dinoflagellate Karlodinium sp. (cultured at different N/P ratios) on micro and mesozooplankton. Sci. Mar. 70, 59–65 (2006).
    Google Scholar 

    24.
    Delgado, M. & Alcaraz, M. Interactions between red tide microalgae and herbivorous zooplankton: the noxious effects of Gyrodinium corsicum (Dinophyceae) on Acartia grani (Copepoda: Calanoida). J. Plankton Res. 21, 2361–2371 (1999).
    Google Scholar 

    25.
    D’Alelio, D. et al. The green-blue swing: plasticity of plankton food-webs in response to coastal oceanographic dynamics. Mar. Ecol. 36, 1155–1170 (2014).
    ADS  Google Scholar 

    26.
    Mitra, A. et al. The role of mixotrophic protists in the biological carbon pump. Biogeosciences 11, 995–1005 (2014).
    ADS  Google Scholar 

    27.
    Ho, P. C. et al. Body size, light intensity, and nutrient supply determine plankton stoichiometry in mixotrophic plankton food webs. Am. Nat. 195, E100–E111 (2020).
    PubMed  Google Scholar 

    28.
    Helenius, L. K. & Saiz, E. Feeding behaviour of the nauplii of the marine calanoid copepod Paracartia grani Sars: functional response, prey size spectrum, and effects of the presence of alternative prey. PLoS ONE 12, 1–20 (2017).
    Google Scholar 

    29.
    Henriksen, C. I., Saiz, E., Calbet, A. & Hansen, B. W. Feeding activity and swimming patterns of Acartia grani and Oithona davisae nauplii in the presence of motile and non-motile prey. Mar. Ecol. Prog. Ser. 331, 119–129 (2007).
    ADS  Google Scholar 

    30.
    Calbet, A. et al. Adaptations to feast and famine in different strains of the marine heterotrophic dinoflagellates Gyrodinium dominans and Oxyrrhis marina. Mar. Ecol. Prog. Ser. 483, 67–84 (2013).
    ADS  Google Scholar 

    31.
    Isari, S. & Saiz, E. Feeding performance of the copepod Clausocalanus lividus (Frost and Fleminger, 1968). J. Plankton Res. 33, 715–728 (2011).
    Google Scholar 

    32.
    Tong, M., Smith, J. L., Kulis, D. M. & Anderson, D. M. Role of dissolved nitrate and phosphate in isolates of Mesodinium rubrum and toxin-producing Dinophysis acuminata. Aquat. Microb. Ecol. 75, 169–185 (2015).
    PubMed  PubMed Central  Google Scholar 

    33.
    Menden-Deuer, S. & Lessard, E. J. Carbon to volume relationships for dinoflagellates, diatoms, and other protist plankton. Limnol. Oceanogr. 45, 569–579 (2000).
    ADS  CAS  Google Scholar 

    34.
    Li, A., Stoecker, D. K. & Coats, D. W. Mixotrophy in Gyrodinium galatheanum (Dinophyceae): grazing responses to light intensity and inorganic nutrients. J. Phycol. 36, 33–45 (2000).
    CAS  Google Scholar 

    35.
    Adolf, J. E., Bachvaroff, T. R. & Place, A. R. Environmental modulation of karlotoxin levels in strains of the cosmopolitan dinoflagellate, Karlodinium veneficum (Dinophyceae). J. Phycol. 45, 176–192 (2009).
    CAS  PubMed  Google Scholar 

    36.
    Sheng, J., Malkiel, E., Katz, J., Adolf, J. E. & Place, A. R. A dinoflagellate exploits toxins to immobilize prey prior to ingestion. Proc. Natl. Acad. Sci. 107, 2082–2087 (2010).
    ADS  CAS  PubMed  Google Scholar 

    37.
    Olivares, M., Saiz, E. & Calbet, A. Ontogenetic changes in the feeding functional response of the marine copepod Paracartia grani. Mar. Ecol. Prog. Ser. 616, 25–35 (2019).
    ADS  CAS  Google Scholar 

    38.
    Isari, S., Antó, M. & Saiz, E. Copepod foraging on the basis of food nutritional quality: can copepods really choose?. PLoS ONE 8, 1–12 (2013).
    Google Scholar 

    39.
    Saiz, E., Alcaraz, M. & Paffenhöfer, G. A. Effects of small-scale turbulence on feeding rate and gross-growth efficiency of three Acartia species (Copepoda: Calanoida). J. Plankton Res. 14, 1085–1097 (1992).
    Google Scholar 

    40.
    Calbet, A., Saiz, E. & Barata, C. Lethal and sublethal effects of naphthalene and 1,2-dimethylnaphthalene on the marine copepod Paracartia grani. Mar. Biol. 151, 195–204 (2007).
    CAS  Google Scholar 

    41.
    Saiz, E. & Kiørboe, T. Predatory and suspension feeding of the copepod Acartia tonsa in turbulent environments. Mar. Ecol. Prog. Ser. 122, 147–158 (1995).
    ADS  Google Scholar 

    42.
    Besiktepe, S. & Dam, H. G. Coupling of ingestion and defecation as a function of diet in the calanoid copepod Acartia tonsa. Mar. Ecol. Prog. Ser. 229, 151–164 (2002).
    ADS  Google Scholar 

    43.
    Burian, A., Grosse, J., Winder, M. & Boschker, H. T. S. Nutrient deficiencies and the restriction of compensatory mechanisms in copepods. Funct. Ecol. 32, 636–647 (2018).
    Google Scholar 

    44.
    Berggreen, U., Hansen, B. & Kiørboe, T. Food size spectra, ingestion and growth of the copepod Acartia tonsa during development: implications for determination of copepod production. Mar. Biol. 99, 341–352 (1988).
    Google Scholar 

    45.
    Broglio, E., Johansson, M. & Jonsson, P. R. Trophic interaction between copepods and ciliates: effects of prey swimming behavior on predation risk. Mar. Ecol. Prog. Ser. 220, 179–186 (2001).
    ADS  Google Scholar 

    46.
    Feinberg, L. R. & Dam, H. G. Effects of diet on dimensions, density and sinking rates of fecal pellets of the copepod Acartia tonsa. Mar. Ecol. Prog. Ser. 175, 87–96 (1998).
    ADS  Google Scholar 

    47.
    Dagg, M. J. & Walser, W. E. The effect of food concentration on fecal pellet size in marine copepods. Limnol. Oceanogr. 31, 1066–1071 (1986).
    ADS  CAS  Google Scholar 

    48.
    Liu, H., Chen, M., Zhu, F. & Harrison, P. J. Effect of diatom silica content on copepod grazing, growth and reproduction. Front. Mar. Sci. 3, 1–7 (2016).
    Google Scholar 

    49.
    Hansen, B., Fotel, F. L., Jensen, N. J. & Madsen, S. D. Bacteria associated with a marine planktonic copepod in culture. II. Degradation of fecal pellets produced on a diatom, a nanoflagellate or a dinoflagellate diet. J. Plankton Res. 18, 275–288 (1996).
    Google Scholar 

    50.
    Li, A., Stoecker, D. K. & Adolf, J. E. Feeding, pigmentation, photosynthesis and growth of the mixotrophic dinoflagellate Gyrodinium galatheanum. Aquat. Microb. Ecol. 19, 163–176 (1999).
    Google Scholar 

    51.
    Thor, P. & Wendt, I. Functional response of carbon absorption efficiency in the pelagic calanoid copepod Acartia tonsa Dana. Limnol. Oceanogr. 55, 1779–1789 (2010).
    ADS  Google Scholar 

    52.
    Saiz, E. et al. Ageing and caloric restriction in a marine planktonic copepod. Sci. Rep. 5, 1–10 (2015).
    Google Scholar 

    53.
    Calbet, A. et al. Intraspecific variability in Karlodinium veneficum: growth rates, mixotrophy, and lipid composition. Harmful Algae 10, 654–667 (2011).
    Google Scholar 

    54.
    Arias, A., Saiz, E. & Calbet, A. Towards an understanding of diel feeding rhythms in marine protists: consequences of light manipulation. Microb. Ecol. 79, 64–72 (2020).
    PubMed  Google Scholar 

    55.
    Hansen, P. J., Bjørnsen, P. K. & Hansen, B. W. Zooplankton grazing and growth: scaling within the 2–2,000-μm body size range. Limnol. Oceanogr. 42, 687–704 (1997).
    ADS  Google Scholar 

    56.
    Kleppel, G. S., Burkart, C. A. & Houchin, L. Nutrition and the regulation of egg production in the calanoid copepod Acartia tonsa. Limnol. Oceanogr. 43, 1000–1007 (1998).
    ADS  Google Scholar 

    57.
    Isari, S. et al. Lack of evidence for elevated CO2-induced bottom-up effects on marine copepods: a dinoflagellate–calanoid prey–predator pair. ICES J. Mar. Sci. 73, 650–658 (2016).
    Google Scholar 

    58.
    Miralto, A. et al. The insidious effect of diatoms on copepod reproduction. Nature 402, 173–176 (1999).
    ADS  CAS  Google Scholar 

    59.
    Rasmussen, S. A. et al. Karmitoxin: an amine-containing polyhydroxy-polyene toxin from the marine dinoflagellate Karlodinium armiger. J. Nat. Prod. 80, 1287–1293 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    60.
    Rodriguez, V., Guerreo, F. & Bautista, B. Egg production of individual copepods of Acartia grani Sars from coastal waters: seasonal and diel variability. J. Plankton Res. 17, 2233–2250 (1995).
    Google Scholar 

    61.
    Thingstad, T. F. et al. Nature of phosphorus limitation in the ultraoligotrophic eastern Mediterranean. Science (80-) 309, 1068–1071 (2005).
    ADS  CAS  Google Scholar 

    62.
    Millette, N. C., Pierson, J. J., Aceves, A. & Stoecker, D. K. Mixotrophy in Heterocapsa rotundata: a mechanism for dominating the winter phytoplankton. Limnol. Oceanogr. 62, 836–845 (2016).
    ADS  Google Scholar 

    63.
    Ballen-Segura, M., Felip, M. & Catalan, J. Some mixotrophic flagellate species selectively graze on archaea. Appl. Environ. Microbiol. 83, 1–11 (2017).
    Google Scholar 

    64.
    Berge, T. & Hansen, P. J. Role of the chloroplasts in the predatory dinoflagellate Karlodinium armiger. Mar. Ecol. Prog. Ser. 549, 41–54 (2016).
    ADS  CAS  Google Scholar 

    65.
    Smith, M. & Hansen, P. J. Interaction between Mesodinium rubrum and its prey: importance of prey concentration, irradiance and pH. Mar. Ecol. Prog. Ser. 338, 61–70 (2007).
    ADS  Google Scholar 

    66.
    Saiz, E., Calbet, A., Trepat, I., Irigoien, X. & Alcaraz, M. Food availability as a potential source of bias in the egg production method for copepods. J. Plankton Res. 19, 1–14 (1997).
    Google Scholar 

    67.
    Frost, B. W. Effects of size and concentration of food particles on the feeding behaviour of the marine planktonic copepod Calanus pacificus. Limnol. Oceanogr. 17, 805–815 (1972).
    ADS  Google Scholar 

    68.
    Saiz, E., Griffell, K. & Calbet, A. Ontogenetic changes in the elemental composition and stoichiometry of marine copepods with different life history strategies. J. Plankton Res. 42, 320–333 (2020).
    Google Scholar  More

  • in

    Impact of 10-Myr scale monsoon dynamics on Mesozoic climate and ecosystems

    1.
    Hinnov, L. A. Cyclostratigraphy and its revolutionizing applications in the earth and planetary sciences. Geol. Soc. Am. Bull. 125, 1703–1734 (2013).
    ADS  Google Scholar 
    2.
    Hays, J. D., Imbrie, J. & Shackleton, N. J. Variations in the Earth’s orbit: pacemaker of the ice ages. Science 194, 1121–1132 (1976).
    ADS  CAS  PubMed  Google Scholar 

    3.
    Laskar, J. et al. A long-term numerical solution for the insolation quantities of the Earth. Astron. Astrophys. 428, 261–285 (2004).
    ADS  Google Scholar 

    4.
    Laskar, J., Fienga, A., Gastineau, M. & Manche, H. La2010: a new orbital solution for the long-term motion of the Earth. Astron. Astrophys. 532, A89 (2011).
    ADS  MATH  Google Scholar 

    5.
    Boulila, S., Galbrun, B., Laskar, J. & Pälike, H. A ~9 Myr cycle in Cenozoic δ13C record and long-term orbital eccentricity modulation: Is there a link?. Earth Planet. Sci. Lett. 317, 273–281 (2012).
    ADS  Google Scholar 

    6.
    Ikeda, M. & Tada, R. Long period astronomical cycles from the Triassic to Jurassic bedded chert sequence (Inuyama, Japan); geologic evidences for the chaotic behavior of solar planets. Earth, Planets Space 65, 351–360 (2013).
    ADS  Google Scholar 

    7.
    Martinez, M. & Dera, G. Orbital pacing of carbon fluxes by a ∼9-My eccentricity cycle during the Mesozoic. Proc. Natl. Acad. Sci. 112, 12604–12609 (2015).
    ADS  CAS  PubMed  Google Scholar 

    8.
    Sprovieri, M. et al. Late Cretaceous orbitally-paced carbon isotope stratigraphy from the Bottaccione Gorge (Italy). Palaeogeogr. Palaeoclimatol. Palaeoecol. 379, 81–94 (2013).
    Google Scholar 

    9.
    Ikeda, M. & Tada, R. Reconstruction of the chaotic behavior of the solar system from geologic records. Earth Planet. Sci. Lett. 537, 116168 (2020).
    CAS  Google Scholar 

    10.
    Ikeda, M. & Tada, R. A 70 million year astronomical time scale for the deep-sea bedded chert sequence (Inuyama, Japan): Implications for Triassic-Jurassic geochronology. Earth Planet. Sci. Lett. 399, 30–43. https://doi.org/10.1016/j.epsl.2014.04.031 (2014).
    ADS  CAS  Article  Google Scholar 

    11.
    Kent, D. V., Olsen, P. E. & Muttoni, G. Astrochronostratigraphic polarity time scale (APTS) for the late triassic and early jurassic from continental sediments and correlation with standard marine stages. Earth-Sci. Rev. 166, 153–180 (2017).
    ADS  CAS  Google Scholar 

    12.
    Olsen, P. E., Kent, D. V. & Whiteside, J. H. Implications of the Newark Supergroup-based astrochronology and geomagnetic polarity time scale (Newark-APTS) for the tempo and mode of the early diversification of the Dinosauria. Earth Environ. Sci. Trans.-R. Soc. Edinb. 101, 201 (2011).
    Google Scholar 

    13.
    Ikeda, M., Tada, R. & Ozaki, K. Astronomical pacing of the global silica cycle recorded in Mesozoic bedded cherts. Nat. Commun. 8, 15532 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    14.
    Schaller, M. F., Wright, J. D. & Kent, D. V. A 30 Myr record of Late Triassic atmospheric pCO2 variation reflects a fundamental control of the carbon cycle by changes in continental weathering. Geol. Soc. Am. Bull. 127, 661–671 (2015).
    ADS  CAS  Google Scholar 

    15.
    Knobbe, T. K. & Schaller, M. F. A tight coupling between atmospheric pCO2 and sea-surface temperature in the Late Triassic. Geology 46, 43–46 (2018).
    ADS  CAS  Google Scholar 

    16.
    Trotter, J. A., Williams, I. S., Nicora, A., Mazza, M. & Rigo, M. Long-term cycles of Triassic climate change: a new δ18O record from conodont apatite. Earth Planet. Sci. Lett. 415, 165–174 (2015).
    ADS  CAS  Google Scholar 

    17.
    Langer, M. C. et al. Untangling the dinosaur family tree. Nature 551, E1 (2017).
    PubMed  Google Scholar 

    18.
    Luo, Z.-X., Gatesy, S. M., Jenkins, F. A., Amaral, W. W. & Shubin, N. H. Mandibular and dental characteristics of Late Triassic mammaliaform Haramiyavia and their ramifications for basal mammal evolution. Proc. Natl. Acad. Sci. 112, E7101–E7109 (2015).
    CAS  PubMed  Google Scholar 

    19.
    Olsen, P. E. et al. Ascent of dinosaurs linked to an iridium anomaly at the Triassic-Jurassic boundary. Science 296, 1305–1307 (2002).
    ADS  CAS  PubMed  Google Scholar 

    20.
    Stocker, M. R., Zhao, L.-J., Nesbitt, S. J., Wu, X.-C. & Li, C. A short-snouted, Middle Triassic phytosaur and its implications for the morphological evolution and biogeography of Phytosauria. Sci. Rep. 7, 46028 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    21.
    Xing, L. et al. An unusual trackway of a possibly bipedal archosaur from the Late Triassic of the Sichuan Basin China. Acta Palaeontol. Polonica 59, 863–871 (2013).
    Google Scholar 

    22.
    Olsen, P. E. & Kent, D. V. Milankovitch climate forcing in the tropics of Pangaea during the Late Triassic. Palaeogeogr. Palaeoclimatol. Palaeoecol. 122, 1–26 (1996).
    Google Scholar 

    23.
    Olsen, P. E. & Kent, D. V. Long-period Milankovitch cycles from the late Triassic and early Jurassic of eastern North America and their implications for the calibration of the Early Mesozoic time-scale and the long-term behaviour of the planets. Philos. Trans. R. Soc. Lond. Ser. A 357, 1761–1786. https://doi.org/10.1098/rsta.1999.0400 (1999).
    ADS  Article  Google Scholar 

    24.
    Kent, D. V. & Olsen, P. E. Magnetic polarity stratigraphy and paleolatitude of the Triassic-Jurassic Blomidon formation in the Fundy Basin (Canada): implications for early Mesozoic tropical climate gradients. Earth Planet. Sci. Lett. 179, 311–324 (2000).
    ADS  CAS  Google Scholar 

    25.
    Schaller, M. F., Wright, J. D. & Kent, D. V. Atmospheric pCO2 perturbations associated with the Central Atlantic magmatic province. Science 331, 1404–1409 (2011).
    ADS  CAS  PubMed  Google Scholar 

    26.
    Schaller, M. F., Wright, J. D., Kent, D. V. & Olsen, P. E. Rapid emplacement of the Central Atlantic Magmatic Province as a net sink for CO 2. Earth Planet. Sci. Lett. 323, 27–39 (2012).
    ADS  Google Scholar 

    27.
    Cornet, B. & Olsen, P. E. A summary of the biostratigraphy of the Newark Supergroup of eastern North America with comments on Early Mesozoic provinciality. In Simposio Sobre Floras del Triasico Tardio, su Fitogeografia y Paleoecologia: Memoria, III Congresso Latinoamericano de Paleontologia, Mexico (ed. Weber, R.) 67–81 (Instituto de Geologia Universidad Nacional Autonoma de Mexico, Mexico City., 1985).
    Google Scholar 

    28.
    Huber, P., Lucas, S. G. & Hunt, A. P. Vertebrate biochronology of the Newark Supergroup Triassic eastern North America. N. M. Mus. Nat. Hist. Sci. Bull. 3, 179–186 (1993).
    Google Scholar 

    29.
    Sugiyama, K. Triassic and Lower Jurassic radiolarian biostratigraphy in the siliceous claystone and bedded chert units of the southeastern Mino Terrane Central Japan. Bull. Mizunami Fossil Mus. 24, 79–193 (1997).
    Google Scholar 

    30.
    Blackburn, T. J. et al. Zircon U–Pb geochronology links the end-Triassic extinction with the Central Atlantic Magmatic Province. Science 340, 941–945 (2013).
    ADS  CAS  PubMed  Google Scholar 

    31.
    Olsen, P. E. et al. Mapping solar system chaos with the Geological Orrery. Proc. Natl. Acad. Sci. 116, 10664–10673 (2019).
    ADS  CAS  PubMed  Google Scholar 

    32.
    Kent, D. V. & Tauxe, L. Corrected Late Triassic latitudes for continents adjacent to the North Atlantic. Science 307, 240–244 (2005).
    ADS  CAS  PubMed  Google Scholar 

    33.
    Ikeda, M. et al. Carbon cycle dynamics linked with Karoo-Ferrar volcanism and astronomical cycles during Pliensbachian-Toarcian (Early Jurassic). Glob. Planet. Change 170, 163–171 (2018).
    ADS  Google Scholar 

    34.
    Kent, D. V. et al. Empirical evidence for stability of the 405-kiloyear Jupiter–Venus eccentricity cycle over hundreds of millions of years. Proc. Natl. Acad. Sci. 115, 6153–6158 (2018).
    ADS  CAS  PubMed  Google Scholar 

    35.
    Hartmann, J., Moosdorf, N., Lauerwald, R., Hinderer, M. & West, A. J. Global chemical weathering and associated P-release—the role of lithology, temperature and soil properties. Chem. Geol. 363, 145–163 (2014).
    ADS  CAS  Google Scholar 

    36.
    Kutzbach, J. & Gallimore, R. Pangaean climates: megamonsoons of the megacontinent. J. Geophys. Res. Atmos. 94, 3341–3357 (1989).
    ADS  Google Scholar 

    37.
    Kutzbach, J. E. Idealized Pangean climates: sensitivity to orbital change. Geol. Soc. Am. Spec. Pap. 288, 41–56 (1994).
    Google Scholar 

    38.
    Nordt, L., Atchley, S. & Dworkin, S. Collapse of the Late Triassic megamonsoon in western equatorial Pangea, present-day American Southwest. Bulletin 127, 1798–1815 (2015).
    CAS  Google Scholar 

    39.
    Donnadieu, Y. et al. A GEOCLIM simulation of climatic and biogeochemical consequences of Pangea breakup. Geochem. Geophys. Geosyst. 7, Q11019 (2006).
    ADS  Google Scholar 

    40.
    Van Dam, J. A. et al. Long-period astronomical forcing of mammal turnover. Nature 443, 687–691 (2006).
    ADS  PubMed  Google Scholar 

    41.
    Whiteside, J. H. et al. Extreme ecosystem instability suppressed tropical dinosaur dominance for 30 million years. Proc. Natl. Acad. Sci. 112, 7909–7913 (2015).
    ADS  CAS  PubMed  Google Scholar 

    42.
    Kent, D. V. & Clemmensen, L. B. Paleomagnetism and cycle stratigraphy of the Triassic Fleming Fjord and Gipsdalen formations of East Greenland. Bull. Geol. Soc. Den. 42, 121–136 (1996).
    Google Scholar 

    43.
    Diedrich, C. Isochirotherium trackways, their possible trackmakers (? Arizonasaurus): intercontinental giant archosaur migrations in the Middle Triassic tsunami-influenced carbonate intertidal mud flats of the European Germanic Basin. Carbonates Evaporites 30, 229–252 (2015).
    CAS  Google Scholar 

    44.
    Teitelbaum, C. S. et al. How far to go? Determinants of migration distance in land mammals. Ecol. Lett. 18, 545–552 (2015).
    PubMed  Google Scholar 

    45.
    Southwood, A. & Avens, L. Physiological, behavioral, and ecological aspects of migration in reptiles. J. Comput. Physiol. B 180, 1–23 (2010).
    Google Scholar 

    46.
    Carter, E. S. & Hori, R. S. Global correlation of the radiolarian faunal change across the Triassic Jurassic boundary. Can. J. Earth Sci. 42, 777–790 (2005).
    ADS  Google Scholar 

    47.
    Ikeda, M., Hori, R. S., Okada, Y. & Nakada, R. Volcanism and deep-ocean acidification across the end-Triassic extinction event. Palaeogeogr. Palaeoclimatol. Palaeoecol. 440, 725–733 (2015).
    Google Scholar 

    48.
    Guex, J. et al. Geochronological constraints on post-extinction recovery of the ammonoids and carbon cycle perturbations during the Early Jurassic. Palaeogeogr. Palaeoclimatol. Palaeoecol. 346, 1–11 (2012).
    Google Scholar 

    49.
    Maron, M. et al. Magnetostratigraphy, biostratigraphy, and chemostratigraphy of the Pignola-Abriola section: new constraints for the Norian-Rhaetian boundary. GSA Bull. 127, 962–974 (2015).
    CAS  Google Scholar 

    50.
    Sell, B. et al. Evaluating the temporal link between the Karoo LIP and climatic–biologic events of the Toarcian Stage with high-precision U–Pb geochronology. Earth Planet. Sci. Lett. 408, 48–56 (2014).
    ADS  CAS  Google Scholar 

    51.
    Burgess, S., Bowring, S., Fleming, T. & Elliot, D. High-precision geochronology links the Ferrar large igneous province with early-Jurassic ocean anoxia and biotic crisis. Earth Planet. Sci. Lett. 415, 90–99 (2015).
    ADS  CAS  Google Scholar 

    52.
    Ritterbush, K. A., Rosas, S., Corsetti, F. A., Bottjer, D. J. & West, A. J. Andean sponges reveal long-term benthic ecosystem shifts following the end-Triassic mass extinction. Palaeogeogr. Palaeoclimatol. Palaeoecol. 420, 193–209 (2015).
    Google Scholar 

    53.
    White, A. F. & Blum, A. E. Effects of climate on chemical weathering in watersheds. Geochim. Cosmochim. Acta 59, 1729–1747 (1995).
    ADS  CAS  Google Scholar 

    54.
    Torrence, C. & Compo, G. P. A practical guide to wavelet analysis. Bull. Am. Meteorol. Soc. 79, 61–78 (1998).
    ADS  Google Scholar 

    55.
    Royer, D. L., Donnadieu, Y., Park, J., Kowalczyk, J. & Goddéris, Y. Error analysis of CO2 and O2 estimates from the long-term geochemical model GEOCARBSULF. Am. J. Sci. 314, 1259–1283 (2014).
    ADS  CAS  Google Scholar 

    56.
    Brenner, G. J. Flowering Plant Origin, Evolution & Phylogeny 91–115 (Springer, Berlin, 1996).
    Google Scholar 

    57.
    Bernardi, M., Gianolla, P., Petti, F. M., Mietto, P. & Benton, M. J. Dinosaur diversification linked with the Carnian Pluvial Episode. Nat. Commun. 9, 1499 (2018).
    ADS  PubMed  PubMed Central  Google Scholar 

    58.
    Li, M. et al. Astronomical tuning and magnetostratigraphy of the Upper Triassic Xujiahe Formation of South China and Newark Supergroup of North America: implications for the Late Triassic time scale. Earth Planet. Sci. Lett. 475, 207–223 (2017).
    ADS  CAS  Google Scholar  More

  • in

    Schistocephalus parasite infection alters sticklebacks’ movement ability and thereby shapes social interactions

    1.
    Poulin, R. Evolutionary Ecology of Parasites (Princeton University Press, Princeton, 2011).
    Google Scholar 
    2.
    Barber, I., Hoare, D. J. & Krause, J. Effects of parasites on fish behaviour: A review and evolutionary perspective. Rev. Fish Biol. Fish. 10, 131–165 (2000).
    Google Scholar 

    3.
    Binning, S. A., Shaw, A. K. & Roche, D. G. Parasites and host performance: Incorporating infection into our understanding of animal movement. Integr. Comp. Biol. 57, 267–280 (2017).
    PubMed  Google Scholar 

    4.
    Lafferty, K. D. & Shaw, J. C. Comparing mechanisms of host manipulation across host and parasite taxa. J. Exp. Biol. 216, 56–66 (2012).
    Google Scholar 

    5.
    McElroy, E. J. & de Buron, I. Host performance as a target of manipulation by parasites: A meta-analysis. J. Parasitol. 100, 399–410 (2014).
    PubMed  Google Scholar 

    6.
    Ward, A. J. W., Hoare, D. J., Couzin, I. D., Broom, M. & Krause, J. The effects of parasitism and body length on positioning within wild fish shoals. J. Anim. Ecol. 71, 10–14 (2002).
    Google Scholar 

    7.
    Krause, J. & Godin, J.-G.J. Influence of parasitism on the shoaling behaviour of banded killifish, Fundulus diaphanus. Can. J. Zool. 72, 1775–1779 (1994).
    Google Scholar 

    8.
    Barber, I., Huntingford, F. A. & Crompton, D. W. T. The effect of Schistocephalus solidus (Cestoda: Pseudophyllidea) on the foraging and shoaling behaviour of three-spined sticklebacks, Gasterosteus aculeatus. Behaviour 132, 1223–1240 (1995).
    Google Scholar 

    9.
    Barber, I. & Scharsack, J. P. The three-spined stickleback-Schistocephalus solidus system: An experimental model for investigating host-parasite interactions in fish. Parasitology 137, 411–424 (2010).
    CAS  PubMed  Google Scholar 

    10.
    Pennycuick, L. Quantitative effects of 3 species of parasites on a population of three-spined sticklebacks, Gasterosteus aculeatus. J. Zool. 165, 143–162 (1971).
    Google Scholar 

    11.
    Barber, I. & Svensson, P. A. Effects of experimental Schistocephalus solidus infections on growth, morphology and sexual development of female three-spined sticklebacks, Gasterosteus aculeatus. Parasitology 126, 359–367 (2003).
    CAS  PubMed  Google Scholar 

    12.
    Walkey, M. & Meakins, R. H. An attempt to balance the energy budget of a host-parasite system. J. Fish Biol. 2, 361–372 (1970).
    Google Scholar 

    13.
    Lester, R. J. G. The influence of Schistocephalus plerocercoids on the respiration of Gasterosteus and a possible resulting effect on the behavior of the fish. Can. J. Zool. 49, 361–366 (1971).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Meakins, R. H. & Walkey, M. The effects of parasitism by the plerocercoid of Schistocephalus solidus Muller 1776 (Pseudophyllidea) on the respiration of the three-spined stickleback Gasterosteus aculeatus L. J. Fish Biol. 7, 817–824 (1975).
    Google Scholar 

    15.
    Arme, C. & Owen, R. W. Infections of the three-spined stickleback, Gasterosteus aculeatus L., with the plerocercoid larvae of Schistocephalus solidus (Müller, 1776), with special reference to pathological effects. Parasitology 57, 301–314 (1967).
    CAS  PubMed  Google Scholar 

    16.
    Barber, I. A non-invasive morphometric technique for estimating cestode plerpcercoid burden in small freshwater fish. J. Fish Biol. 51, 654–658 (1997).
    Google Scholar 

    17.
    Dingemanse, N. J., Oosterhof, C., Van Der Plas, F. & Barber, I. Variation in stickleback head morphology associated with parasite infection. Biol. J. Linn. Soc. 96, 759–768 (2009).
    Google Scholar 

    18.
    Giles, N. Behavioural effects of the parasite Schistocephalus solidus (Cestoda) on an intermediate host, the three-spined stickleback, Gasterosteus aculeatus L. Anim. Behav. 31, 1192–1194 (1983).
    Google Scholar 

    19.
    Milinski, M. Risk of predation of parasitized sticklebacks (Gasterosteus aculeatus L.) under competition for food. Behaviour 93, 203–215 (1985).
    Google Scholar 

    20.
    Godin, J.-G.J. & Sproul, C. D. Risk taking in parasitized sticklebacks under threat of predation: Effects of energetic need and food availability. Can. J. Zool. 66, 2360–2367 (1988).
    Google Scholar 

    21.
    Grécias, L., Valentin, J. & Aubin-Horth, N. Testing the parasite mass burden effect on alteration of host behaviour in the Schistocephalus–stickleback system. J. Exp. Biol. 221, jeb174748 (2018).
    PubMed  Google Scholar 

    22.
    Giles, N. Predation risk and reduced foraging activity in fish: Experiments with parasitized and non-parasitized three-spined sticklebacks, Gasterosteus aculeatus L. J. Fish Biol. 31, 37–44 (1987).
    Google Scholar 

    23.
    Talarico, M. et al. Specific manipulation or systemic impairment? Behavioural changes of three-spined sticklebacks (Gasterosteus aculeatus) infected with the tapeworm Schistocephalus solidus. Behav. Ecol. Sociobiol. 71, 1–10 (2017).
    Google Scholar 

    24.
    Pascoe, D. & Mattey, D. Dietary stress in parasitized and non-parasitized Gasterosteus aculeatus L. Zeitschrift für Parasitenkd. 186, 179–186 (1977).
    Google Scholar 

    25.
    Ness, J. H. & Foster, S. A. Parasite-associated phenotype modifications in threespine stickleback. Oikos 85, 127–134 (1999).
    Google Scholar 

    26.
    Tierney, J. F., Huntingford, F. A. & Crompton, D. W. T. The relationship between infectivity of Schistocephalus solidus (Cestoda) and anti-predator behaviour of its intermediate host, the three-spined stickleback, Gasterosteus aculeatus. Anim. Behav. 46, 603–605 (1993).
    Google Scholar 

    27.
    Barber, I., Walker, P. & Svensson, P. A. Behavioural responses to simulated avian predation in female three-spined sticklebacks: The effect of experimental Schistocephalus solidus infections. Behaviour 141, 1425–1440 (2004).
    Google Scholar 

    28.
    Grécias, L., Hébert, F. O., Berger, C. S., Barber, I. & Aubin-Horth, N. Can the behaviour of threespine stickleback parasitized with Schistocephalus solidus be replicated by manipulating host physiology?. J. Exp. Biol. 220, 237–246 (2017).
    PubMed  Google Scholar 

    29.
    Peacock, S. D. Some Effects of the Cestode Schistocephalus solidus on the Threespine Stickleback Gasterosteus aculeatus (The University of British Columbia, Vancouver, 1972).
    Google Scholar 

    30.
    Blake, R. W., Kwok, P. Y. L. & Chan, K. H. S. Effects of two parasites, Schistocephalus solidus (Cestoda) and Bunodera spp. (Trematoda), on the escape fast-start performance of three-spined sticklebacks. J. Fish Biol. 69, 1345–1355 (2006).
    Google Scholar 

    31.
    Hafer, N. & Milinski, M. An experimental conflict of interest between parasites reveals the mechanism of host manipulation. Behav. Ecol. 27, 617–627 (2016).
    PubMed  Google Scholar 

    32.
    Jolles, J. W., King, A. J. & Killen, S. S. The role of individual heterogeneity in collective animal behaviour. Trends Ecol. Evol. 35, 278–291 (2020).
    PubMed  Google Scholar 

    33.
    Barber, I., Huntingford, F. A. & Crompton, D. W. T. The effect of hunger and cestode parasitism on the shoaling decisions of small freshwater fish. J. Fish Biol. 47, 524–536 (1995).
    Google Scholar 

    34.
    Barber, I. & Huntingford, F. A. Parasite infection alters schooling behaviour: Deviant positioning of helminth-infected minnows in conspecific groups. Proc. R. Soc. B 263, 1095–1102 (1996).
    ADS  Google Scholar 

    35.
    Radabauch, D. C. Changes in minnow, Pimephales promelas Rafinesque, schooling behaviour associated with infections of brain-encysted larvae of the fluke, Ornithodiplostomum ptychocheilus. J. Fish Biol. 16, 621–628 (1980).
    Google Scholar 

    36.
    Hoare, D. J., Ruxton, G. D., Godin, J. G. J. & Krause, J. The social organization of free-ranging fish shoals. Oikos 89, 546–554 (2000).
    Google Scholar 

    37.
    Demandt, N. et al. Parasite-infected sticklebacks increase the risk-taking behaviour of uninfected group members. Proc. R. Soc. B 285, 20180956 (2018).
    PubMed  Google Scholar 

    38.
    Ward, A. J. W., Duff, A. J., Krause, J. & Barber, I. Shoaling behaviour of sticklebacks infected with the microsporidian parasite, Glugea anomala. Environ. Biol. Fishes 72, 155–160 (2005).
    Google Scholar 

    39.
    Krause, J., Godin, J. J. & Brown, D. Phenotypic variability within and between fish shoals. Ecology 77, 1586–1591 (1996).
    Google Scholar 

    40.
    Maximino, C. et al. Measuring anxiety in zebrafish: A critical review. Behav. Brain Res. 214, 157–171 (2010).
    PubMed  Google Scholar 

    41.
    Poulin, R. Parasite Manipulation of Host Behavior: An Update and Frequently Asked Questions. Advances in the Study of Behavior, Vol. 41, (Elsevier Inc., Amsterdam, 2010).

    42.
    Domenici, P. The scaling of locomotor performance in predator-prey encounters: From fish to killer whales. Comp. Biochem. Physiol. Part A 131, 169–182 (2001).
    CAS  Google Scholar 

    43.
    Fish, F. E. Swimming strategies for energy economy. In Fish locomotion: An eco-ethological perspective (eds Domenici, P. & Kapoor, B. G.) 90–122 (Science Publishers, 2010).

    44.
    Binning, S. A., Roche, D. G. & Layton, C. Ectoparasites increase swimming costs in a coral reef fish. Biol. Lett. 9, 20120927–20120927 (2012).
    Google Scholar 

    45.
    Domenici, P. Escape responses in fish: Kinematics, performance and behavior. In Fish locomotion: An eco-ethological perspective (eds Domenici, P. & Kapoor, B. G.) 123–170 (Science Publishers, 2010).

    46.
    Jolles, J. W., Boogert, N. J., Sridhar, V. H., Couzin, I. D. & Manica, A. Consistent individual differences drive collective behavior and group functioning of schooling fish. Curr. Biol. 27, 2862–2868 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    47.
    Tunstrøm, K. et al. Collective states, multistability and transitional behavior in schooling fish. PLoS Comput. Biol. 9, e1002915 (2013).
    MathSciNet  PubMed  PubMed Central  Google Scholar 

    48.
    Jolles, J. W., Laskowski, K. L., Boogert, N. J. & Manica, A. Repeatable group differences in the collective behaviour of stickleback shoals across ecological contexts. Proc. R. Soc. B 285, 20172629 (2018).
    PubMed  Google Scholar 

    49.
    Couzin, I. D., Krause, J., James, R., Ruxton, G. D. & Franks, N. R. Collective memory and spatial sorting in animal groups. J. Theor. Biol. 218, 1–11 (2002).
    MathSciNet  PubMed  Google Scholar 

    50.
    Pettit, B., Ákos, Z., Vicsek, T. & Biro, D. Speed determines leadership and leadership determines learning during pigeon flocking. Curr. Biol. 25, 3132–3137 (2015).
    CAS  PubMed  Google Scholar 

    51.
    Ioannou, C. C., Guttal, V. & Couzin, I. D. Predatory fish select for coordinated collective motion in virtual prey. Science 337, 1212–1215 (2012).
    ADS  CAS  PubMed  Google Scholar 

    52.
    Jones, K. A., Jackson, A. L. & Ruxton, G. D. Prey jitters; Protean behaviour in grouped prey. Behav. Ecol. 22, 831–836 (2011).
    Google Scholar 

    53.
    Tierney, J. F. & Crompton, D. W. Infectivity of plerocercoids of Schistocephalus solidus (Cestoda: Ligulidae) and fecundity of the adults in an experimental definitive host, Gallus gallus. J. Parasitol. 78, 1049–1054 (1992).
    CAS  PubMed  Google Scholar 

    54.
    Berger, C. S. & Aubin-Horth, N. The secretome of a parasite alters its host’s behaviour but does not recapitulate the behavioural response to infection. Proc. R. Soc. B 287, 20200412 (2020).
    PubMed  Google Scholar 

    55.
    Franke, F., Armitage, S. A. O., Kutzer, M. A. M., Kurtz, J. & Scharsack, J. P. Environmental temperature variation influences fitness trade-offs and tolerance in a fish-tapeworm association. Parasites Vectors 10, 1–11 (2017).
    Google Scholar 

    56.
    Hopkins, B. Y. C. A. & Smyth, J. D. Notes on the morphology and life history of Schistocehpaus solidus (Cestoda: Diphyllobothriidae. Parasitology 41, 283–291 (1951).
    CAS  PubMed  Google Scholar 

    57.
    Huntingford, F. A. & Ruiz-Gomez, M. L. Three-spined sticklebacks Gasterosteus aculeatus as a model for exploring behavioural biology. J. Fish Biol. 75, 1943–1976 (2009).
    CAS  PubMed  Google Scholar 

    58.
    Webster, M. M. & Laland, K. N. Evaluation of a non-invasive tagging system for laboratory studies using three-spined sticklebacks Gasterosteus aculeatus. J. Fish Biol. 75, 1868–1873 (2009).
    CAS  PubMed  Google Scholar 

    59.
    Jolles, J. W., Aaron Taylor, B. & Manica, A. Recent social conditions affect boldness repeatability in individual sticklebacks. Anim. Behav. 112, 139–145 (2016).
    PubMed  PubMed Central  Google Scholar 

    60.
    Jolles, J. W. Pirecorder: Controlled and automated image and video recording with the raspberry pi. (2019). https://doi.org/10.5281/zenodo.2529515

    61.
    Krause, J., Reeves, P. & Hoare, D. J. Positioning behaviour in Roach shoals: The role of body length and nutritional state. Behaviour 135, 1031–1039 (1998).
    Google Scholar 

    62.
    Katz, Y., Tunstrøm, K., Ioannou, C. C., Huepe, C. & Couzin, I. D. Inferring the structure and dynamics of interactions in schooling fish. Proc. Natl. Acad. Sci. 108, 18720–18725 (2011).
    ADS  CAS  PubMed  Google Scholar  More

  • in

    Growth dynamics of galls and chemical defence response of Pinus thunbergii Parl. to the pine needle gall midge, Thecodiplosis japonensis Uchida & Inouye (Diptera: Cecidomyiidae)

    1.
    Allison, S. D. & Schultz, J. C. Biochemical responses of chestnut oak to a galling cynipid. J. Chem. Ecol. 31(1), 151–166 (2005).
    CAS  PubMed  Google Scholar 
    2.
    Arimura, G. et al. Effects of feeding Spodoptera littoralis on Lima bean leaves: iv. Diurnal and nocturnal damage differentially initiate plant volatile emission. Plant Physiol. 146(3), 965–973 (2008).
    CAS  PubMed  PubMed Central  Google Scholar 

    3.
    Delan, C. Effects of Apriona Germari Hope damage on the contents of tannin in poplar. Entomol. J. East China. 18(2), 94–100 (2009).
    Google Scholar 

    4.
    Chung, Y., Lee, J. & Lee, B. Effects of pine needle gall midge, Thecodiplosis Japonensis Uchida Et Inouye (Diptera: Cecidomyiidae), infestation on the shoot and needle growth of Pinus densiflora Sieb. Et Zucc. Fri J. For. Sci. (Seoul) 56, 21–29 (1997).
    Google Scholar 

    5.
    Crespi, B. J. & Worobey, M. Comparative analysis of gall morphology in Australian gall thrips: The evolution of extended phenotypes. Evolution 52(6), 1686–1696 (1998).
    PubMed  Google Scholar 

    6.
    Crespi, B. J., Carmean, A. D. A. & Chapman, T. W. Ecology and evolution of galling thrips and their allies. Annu. Rev. Entomol. 42(1), 51–71 (1997).
    CAS  PubMed  Google Scholar 

    7.
    Zhaodi, D. Inclusions Changes in Different Families of Pinus Massoniana Lamb Damaged by Hemiberlesia Pitysophila Takagi. (Fujian Agric. For. Univ, Fuzhou, 2009).
    Google Scholar 

    8.
    Dicke, M. & van Loon, J. J. A. Multitrophic effects of herbivore-induced plant volatiles in an evolutionary context. Entomol. Exp. Appl. 97, 237–249 (2000).
    CAS  Google Scholar 

    9.
    Dsouza, M. R. & Ravishankar, B. Nutritional sink formation in galls of Ficus glomerata Roxb. (Moraceae) by the insect Pauropsylla depressa (Psyllidae, Hemiptera). Trop. Ecol. 55(1), 129–136 (2014).
    Google Scholar 

    10.
    Espírito-Santo, M. M. & Fernandes, G. W. How many species of gall-inducing insects are there on earth, and where are they?. Ann. Entomol. Soc. Am. 100(2), 95–99 (2007).
    Google Scholar 

    11.
    Gagné, R.J., Jaschhof, M. A catalog of the Cecidomyiidae (Diptera) of the world. In 4th Edn. Digital. pp. 762 (2017)

    12.
    Hartley, S. E. Are gall insects large rhizobia?. Oikos 84(2), 333–342 (1999).
    Google Scholar 

    13.
    Zuodong, H. et al. Biological study of Aphidounguis Poniradicicola Zhang et Hu. Acta Agric. Boreali Occidentalis Sin. 8(2), 34–36 (1999).
    Google Scholar 

    14.
    Yunwei, Ju. et al. Physiological and biochemical responses of Distylium racemosum to gall’s formation. J. Fujian Coll. Forestry 32(1), 10–12 (2012).
    Google Scholar 

    15.
    Karley, A., Douglas, A. & Parker, W. Amino acid composition and nutritional quality of potato leaf phloem sap for aphids. J. Exp. Biol. 205(19), 3009–3018 (2002).
    CAS  PubMed  Google Scholar 

    16.
    Koyama, Y., Yao, I. & Akimoto, S. Aphid galls accumulate high concentrations of amino acids: A support for the nutrition hypothesis for gall formation. Entomol. Exp. Appl. 113(1), 35–44 (2004).
    CAS  Google Scholar 

    17.
    Lee, D. K. & Sung, J. H. Variation in photosynthesis and leaf pigments of susceptible Pinus densiflora and resistant Pinus rigida following pine gall midge attack. J. Korean Forestry Soc. 65, 1–11 (1984).
    Google Scholar 

    18.
    Lee, S. H. Biology and Control of Pine Gall Midge (Office of Forestry, Seoul, 1981).
    Google Scholar 

    19.
    Darin, Li., Enguo, W. & Lingwei, L. Study on biology characteristics and population dynamics of invasive Bemisia tabaci (Gennadius). China Plant Protect. Guide 32(1), 17–21 (2012).
    Google Scholar 

    20.
    Nuo, L. et al. Preliminary study on physiology and biochemistry characteristics of eucalyptus galls caused by Leptocy beinvasa LI. Anhui Agric. Sci. 38(8), 4126–4127 (2010).
    Google Scholar 

    21.
    Li, X. C., Schuler, M. A. & Berenhaum, M. R. Jasmonate and salicylate induce expression of herbivore cytochrome P450 genes. Nature 419, 712–715 (2002).
    ADS  CAS  PubMed  Google Scholar 

    22.
    Xiangmei, L. A Comparative Study of the Amino Acids and Defensive Substances of the Gall and Host Plants of Andricus mukaigawae (Hunan Normal University, Changsha, 2019).
    Google Scholar 

    23.
    Xuan, Li. & Zhihong, H. The effect of ulmus gall aphid on the leaf physiology of elm. Jiangsu Agric. Sci. 47(10), 131–134 (2019).
    Google Scholar 

    24.
    Bingjing, L. et al. Determination of total alkaloids in stems, leaves and skins of Polyalthia plagioneura by acid dye colorimetry. Jiangsu Agric. Sci. 42(1), 263–264 (2014).
    Google Scholar 

    25.
    Ling, L., Yunling, M., Zhengliang, Y., Rongpin, Q. & Guanghui, H. Analysis of nutritive compositions of conifer needles from 3 pinaceae trees. J. West China Forestry Sci. 43(2), 117–120 (2014).
    Google Scholar 

    26.
    Youqing, L. & Jianguang, Li. Bionomics and occurrence of Anoplophora glabripennis (Motschulsky). Plant Quarant. 13(1), 5–7 (1999).
    Google Scholar 

    27.
    Ma Shuangmin, Yu., Hong, L. C. & Mingzhi, Y. Plant gall biology. Chin. Bull. Entomol. 45(2), 330–335 (2008).
    Google Scholar 

    28.
    Qi, M. & Wupengmao, X. A preliminary study of biological characteristics of Tetranychid on clove tree. J. Qinghai Univ. (Nat. Sci.) 29(6), 75–79 (2011).
    Google Scholar 

    29.
    Martinez, J. I. Anti-insect effects of the gall wall of Baizongia pistaciae [l], a gall-inducing aphid on Pistacia palaestina boiss. Arthropod Plant Interact. 4(1), 29–34 (2010).
    Google Scholar 

    30.
    Mattson, W. J. Herbivory in relation to plant nitrogen content. Annu. Rev. Ecol. Syst. 11(1), 119–161 (1980).
    Google Scholar 

    31.
    Nam, Y. & Choi, W. I. An empirical predictive model for the spring emergence of Thecodiplosis japonensis (Diptera: Cecidomyiidae): Model construction and validation on the basis of 25 years of field observation data. J. Econ. Entomol. 107(3), 1136–1141 (2014).
    PubMed  Google Scholar 

    32.
    Nisperos-carriedo, M. O., Buslig, B. S. & Shaw, P. E. Simultaneous detection of dehydroascorbic, ascorbic, and some organic acids in fruits and vegetables by HPLC. J. Agric. Food Chem. 40(7), 1127–1130 (1992).
    CAS  Google Scholar 

    33.
    Nyman, T. & Julkunen-tiitto, R. Manipulation of the phenolic chemistry of willows by gall-inducing sawflies. Proc. Natl. Acad. Sci. 97(24), 13184–13187 (2000).
    ADS  CAS  PubMed  Google Scholar 

    34.
    Petersen, M. & Sandström, J. Outcome of indirect competition between two aphid species mediated by responses in their common host plant. Funct. Ecol. 15(4), 525–534 (2001).
    Google Scholar 

    35.
    Pichersky, E. & Gang, D. R. Genetics and biochemistry of secondary metabolites in plants: An evolutionary perspective. Trends Plant Sci. 5, 439–445 (2000).
    CAS  PubMed  Google Scholar 

    36.
    Price, P. W., Waring, G. L. & Fernandes, G. W. Hypotheses on the adaptive nature of galls. Proc. Entomol. Soc. Wash. 88(2), 361–363 (1986).
    Google Scholar 

    37.
    Junde, Q. The relationship between insects and plants. Bull. Biol. 10, 16–18 (1985).
    Google Scholar 

    38.
    Sandstrm, J. & Pettersson, J. Amino acid composition of phloem sap and the relation to intraspecific variation in pea aphid (Acyrthosiphon pisum) performance. J. Insect Physiol. 40, 947–955 (1994).
    Google Scholar 

    39.
    Lakshminarayan, S. (2013). Role of carotenoid cleavage dioxygenases in volatile emissions and insect resistance in Arabidopsis. https://ir.lib.uwo.ca/etd/1846.

    40.
    Shorthouse, J. D. & Rohfritsch, O. Biology of Insect-Induced Galls (Oxford University Press, Oxford, 1992).
    Google Scholar 

    41.
    Singer, A. C., Crowley, D. E. & Thompson, I. P. Secondary plant metabolites in phytoremediation and biotransformation. Trends Biotechnol. 21(3), 123–130 (2003).
    CAS  PubMed  Google Scholar 

    42.
    Skuhravy, V. The development of the needle-shortening gall midge, thecodiplosis brachyntera, on the mountain pine, Pinus mugo Subsp. Mughus (Diptera, Cecidomyiidae). Acta Entomol. Bohemosl. 70(3), 162–167 (1973).
    Google Scholar 

    43.
    Son, D., Eom, T., Seo, J. & Lee, S. Potential resistance factors in pine needles to pine gall midge. J. Korean Forestry Soc. 85(2), 244–250 (1996).
    Google Scholar 

    44.
    Son, D., Eom, T., Seo, J. & Lee, S. A study on resistant substance to pine needle gall midge among phenolic compounds in pine needles. J. Korean Soc. Forest Sci. 85(3), 372–380 (1996).
    Google Scholar 

    45.
    Son, K. Effects of the gregariousness of larvae in galls on the reproductive success of the pine needle gall midge, Thecodiplosis japonensis Uchida Et Inouye (Dipt., Cecidomyiidae). J. Appl. Entomol. 119(1), 267–272 (2009).
    Google Scholar 

    46.
    Sone, K. Ecological studies on the pine needle gall midge, Thecodiplosis japonensis Uchida Et Inouye (Diptera: Cecidomyiidae). III Characteristic features of the infestation and its impacts on the growth of pine trees. Bull. Forestry Forest Products Res. Inst. Japan 349, 71–96 (1987).
    Google Scholar 

    47.
    Soné, K. Population dynamics of the pine needle gall midge, Thecodiplosis japonensis Uchida Et Inouye (Diptera, Cecidomyiidae). J. Appl. Entomol. 103(4), 386–402 (1987).
    Google Scholar 

    48.
    Soné, K. Mortality factors before gall formation by the pine needle gall midge, Thecodiplosis japonensis Uchida Et Inouye (Diptera: Cecidomyiidae). J. Japan. Forestry Soc. 68(1), 32–34 (1986).
    Google Scholar 

    49.
    Stern, D. L. Phylogenetic evidence that aphids, rather than plants, determine gall morphology. Proc. R. Society Lond. Series Biol. Sci. 260(1357), 85–89 (1995).
    ADS  Google Scholar 

    50.
    Stone, G. N. & Cook, J. M. The structure of cynipid oak galls: Patterns in the evolution of an extended phenotype. Proc. R. Society Lond. Series Biol. Sci. 265(1400), 979–988 (1998).
    Google Scholar 

    51.
    Stone, G. N. & Schönrogge, K. The adaptive significance of insect gall morphology. Trends Ecol. Evol. 18(10), 512–522 (2003).
    Google Scholar 

    52.
    Stone, G. N., Schönrogge, K., Atkinson, R. J., Bellido, D. & Pujade-villar, J. The population biology of oak gall wasps (Hymenoptera: Cynipidae). Annu. Rev. Entomol. 47(1), 633–668 (2002).
    CAS  PubMed  Google Scholar 

    53.
    Truernit, E., Schmid, J., Epple, P., Illig, J. & Sauer, N. The sink-specific and stress-regulated arabidopsis stp4 gene: Enhanced expression of a gene encoding a monosaccharide transporter by wounding, elicitors, and pathogen challenge. Plant Cell 8(12), 2169–2182 (1996).
    CAS  PubMed  PubMed Central  Google Scholar 

    54.
    Chenzhu, W. & Junde, Q. Protease inhibitors in plants contributing to resistance to insects: An overview. Acta Entomol. Sin. 40(2), 212–218 (1997).
    Google Scholar 

    55.
    Haoyun, W., Guijie, D., Xie Weibin, Wu. & Feng, Z. G. Growth and photosynthetic pigment content of superiorseedlings of Pinus massoniana. J. Southw. Forestry Univ. 37(5), 42–47 (2017).
    Google Scholar 

    56.
    Mo, W., Yan, J., Zhaojun, M. & Shanchun, Y. Adaptability of larval growth and development in Hyphantria cunea to different host plant secondary metabolites. J. Northe. Forestry Univ. 48(3), 100–104 (2020).
    Google Scholar 

    57.
    Shaolin, W., Minghui, X., Wang Hongqi, Su. & Lianheng, F. K. Bionomics of Acantholyda posticalis and its biological control techniques. Forest Pest Disease 23(2), 23–26 (2004).
    Google Scholar 

    58.
    Wei, W. et al. Effects on secondary metabolite contents in eucalyptus strains damaged by Leptocybe invasa. J. Trop. Subtrop. Bot. 21(6), 521–528 (2013).
    Google Scholar 

    59.
    Weis, A. E., Walton, R. & Crego, C. L. Reactive plant tissue sites and the population biology of gall makers. Annu. Rev. Entomol. 33(1), 467–486 (1988).
    Google Scholar 

    60.
    Wei, S. et al. Enhanced beta-ionone emission in Arabidopsis over-expressing AtCCD1 reduces feeding damage in vivo by the crucifer flea beetle. Environ Entomol. 40, 1622–1630 (2011).
    CAS  PubMed  Google Scholar 

    61.
    Westphal, E., Bronner, R. & Ret, M. L. Changes in leaves of susceptible and resistant solanum dulcamara infested by the gall mite Eriophyes cladophthirus (Acarina, Eriophyoidea). Can. J. Bot. 59(5), 875–882 (1981).
    Google Scholar 

    62.
    Williams, L. E., Lemoine, R. & Sauer, N. Sugar transporters in higher plants—A diversity of roles and complex regulation. Trends Plant Sci. 5(7), 283–290 (2000).
    CAS  PubMed  Google Scholar 

    63.
    Wool, D. & Bar-el, N. Population ecology of the galling aphid forda formicaria von heyden in Israel: Abundance, demography, and gall structure. Israel J. Ecol. Evolut. 41(2), 175–192 (2013).
    Google Scholar 

    64.
    Wool, D. & Ben-zvi, O. Population ecology and clone dynamics of the galling aphid Geoica wertheimae (Sternorrhyncha: Pemphigidae: Fordinae). Eur. J. Entomol. 95(4), 509–518 (1998).
    Google Scholar 

    65.
    Wool, D. & Burstein, M. A galling aphid with extra life-cycle complexity: Population ecology and evolutionary considerations. Popul. Ecol. 33(2), 307–322 (1991).
    Google Scholar 

    66.
    Wool, D. & Manheim, O. Population ecology of the gall-forming aphid, Aploneura lentisci (pass) in Israel. Popul. Ecol. 28(1), 151–162 (1986).
    Google Scholar 

    67.
    Zhijie, Wu., Yinghan, L. & Hong, C. Research progress of resistant active components in Momordica Charantia leaves. Biol. Bull. 1, 16–19 (2009).
    Google Scholar 

    68.
    Hongqiang, Y. & Huajun, G. Journal of plant physiology and molecular biology. Physiol. Funct. Arginine Metab. Plants 33(1), 1–8 (2007).
    Google Scholar 

    69.
    Hongqiang, Y. & Yuling, J. Relationship between polyamines and development of fruit tree. J. Shandong Agric. Univ. 27(4), 134–140 (1996).
    Google Scholar 

    70.
    Mingzhi, Y., Hanbo, Z., Chengchen, Li. & Shuangmin, Ma. Physiological responses of gall tissues on Ivytree leaves induced by thrip. Acta Bot. Yunnanica 32(4), 339–346 (2010).
    Google Scholar 

    71.
    Zhibin, Y., Hai, L. & Lan, G. Determination of total tritepenoidic in leaves and stems of Polyalthia rumphii by Ultraviolet spectrophotometry. Hubei Agric. Sci. 52(5), 1158–1160 (2013).
    Google Scholar 

    72.
    Cao, Y., Dai, H., Cao, W. & Han, Z. Determination of total phenols in Zizyphus jujuba mill by folin-ciocaileu colorimetry. Anhui Agric. Sci. 36(4), 1299–1302 (2008).
    CAS  Google Scholar 

    73.
    Yazaki, K. Abc transporters involved in the transport of plant secondary metabolites. Febs Letters. 580(4), 1183–1191 (2006).
    CAS  PubMed  Google Scholar 

    74.
    Chunling, Z., Zhiwen, Li. & Aiying, G. Studies on biological characteristics of Hyphantria cunea. J. Hebei Forestry Coll. 8(3), 239–243 (1993).
    Google Scholar 

    75.
    Huafeng, Z., Shunli, C., Jianhua, Z. & Silu, Z. Effect of Monochamus alternatus on the composition of chemical materials in needles of Pinus massonina. J. Fujiian Coll. For. 24(1), 28–31 (2004).
    Google Scholar 

    76.
    Ying, Z., Liqing, D., Wenchang, D., Haiping, Li. & Shujun, F. Study on the relationship between Aceria palida Keifer gall and its population. J. Inner Mongol. Agric. Univ. 33(5), 84–86 (2013).
    Google Scholar 

    77.
    Yifan, Z., Shixiong, W. Sustainable research and exploitation on insect resources. In: China Biodiversity Conservation Foundation. The Fifth International Conference on Biodiversity Conservation and Utilization. China Biodiversity Conservation Foundation, pp 238–240. (2005).

    78.
    Xiaohong, Z., Xianshi, L. & Aihua, C. A study on Nai Plum’s flower bud differentiation and its major content of metabolic production. J. Hunan Agric. Univ. 25(1), 31–35 (1990).
    Google Scholar 

    79.
    Lin, Z. Application and research in the bioactivity of Botanical alkaloids against insects. J. Henan Forestry Sci. Technol. 27(1), 32–36 (2007).
    Google Scholar 

    80.
    Yuyong, Z., Li, C. & Wang, J. Impact of free amino acid composition of plant phloem sap and pea aphid itself when it takes food. Xinjiang Agric. Sci. 51(7), 1284–1291 (2014).
    Google Scholar  More

  • in

    Evidence of scavenging behaviour in crested porcupine

    1.
    Kingdon, J. Porcupines (Hystrix). In East African Mammals (ed. Kingdon, J.) 689–695 (Academic Press, London, 1974).
    Google Scholar 
    2.
    Arslan, A. On the indian crested porcupine, Hystrix indica (Kerr, 1972) in Turkey (Mammalia: Rodentia). Pak. J. Biol. Sci. 11(2), 315–317 (2008).
    Article  Google Scholar 

    3.
    Pillay, K. R., Wilson, A. L., Ramesh, T. & Downs, C. T. Digestive parameters and energy assimilation of Cape porcupine on economically important crops. Afr. Zool. 50(4), 321–326. https://doi.org/10.1080/15627020.2015.1116373 (2015).
    Article  Google Scholar 

    4.
    Alkon, P. U. & Saltz, D. Potatoes and the nutritional ecology of crested porcupine in a desert biome. J. Appl. Ecol. 22, 727–737 (1985).
    Article  Google Scholar 

    5.
    Khan, A. A., Mian, A. & Hussain, R. Deterioration impact of Indian crested porcupine, Hystrix indica, on irrigated forest plantations in Punjab Pakistan. Pak. J. Zool. 46(6), 1691–1696 (2014).
    Google Scholar 

    6.
    Laurenzi, A., Bodino, N. & Mori, E. Much ado about nothing: assessing the impact of a problematic rodent on agriculture and native trees. Mammal Res. 61, 65–72 (2016).
    Article  Google Scholar 

    7.
    Felicioli, A., Grazzini, A. & Santini, L. The mounting behaviour of a pair of crested porcupine Hystrix cristata L. Mammalia 61(1), 123–126 (1997).
    Google Scholar 

    8.
    Van Aarde, R. J. Reproduction in the Cape porcupine Hystrix africaeaustralis: an ecological perspective. S Afr. J. Sci. 83, 605–607 (1987).
    Google Scholar 

    9.
    Sever, Z. & Mendelssohn, H. Copulation as a possible mechanism to maintain monogamy in porcupines Hystrix indica. Anim. Behav. 36(5), 1541–1542. https://doi.org/10.1016/S0003-3472(88)80225-2 (1988).
    Article  Google Scholar 

    10.
    Coppola, F., Vecchio, G. & Felicioli, A. Diurnal motor activity and “sunbathing” behaviour in crested porcupine (Hystrix cristata L., 1758). Sci. Rep. 9, 14283. https://doi.org/10.1038/s41598-019-50784-y (2019).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    11.
    Coppola, F., Dari, C., Vecchio, G., Scarselli, D. & Felicioli, A. Co-habitation of settlements between Crested Porcupines (Hystrix cristata), Red Foxes (Vulpes vulpes) and European Badgers (Meles meles). Curr. Sc (in press).

    12.
    Alkon, P. U. & Saltz, D. Foraging time and the northern range limits of Indian crested porcupines (Hystrix indica Kerr). J. Biogeo. 15, 403–408 (1988).
    Article  Google Scholar 

    13.
    Santini, L. The habits and influence on the environment of the old world porcupine Hystrix cristata L. in the northernmost part of its range. In Proceedings of the 9th Vertebrate Pest Conference 34, 149–153 (1980).

    14.
    Sever, Z. & Mendelssohn, H. Spatial movement patterns of porcupines (Hystrix indica). Mammalia 55, 187–206. https://doi.org/10.1515/mamm.1991.55.2.187 (1991).
    Article  Google Scholar 

    15.
    Barthelmess, E. L. Mamm. Species 788, 1–7 (2006).
    Article  Google Scholar 

    16.
    Bruno, E. & Riccardi, C. The diet of the Crested porcupine Hystrix cristata L., 1758 in a Mediterranean rural area. Z. Saugetierkd. 60, 226–236 (1995).
    Google Scholar 

    17.
    Hafeez, S., Khan, G. S., Ashfaq, M. & Khan, Z. H. Food habits of the Indian crested porcupine (Hystrix indica) in Faisalabad Pakistan. Pak. J. Agric. Sci. 48(3), 205–210 (2011).
    Google Scholar 

    18.
    Ori, F., Trappe, J., Leonardi, M., Iotti, M. & Pacioni, G. Crested porcupine (Hystrix cristata): mycophagist spore dispersers of the ectomycorrhizal truffle Tuber aestivum. Mycorrhiza 28, 561–565. https://doi.org/10.1007/s00572-018-0840-1 (2018).
    Article  PubMed  Google Scholar 

    19.
    Chevallier, N. & Ashton, B. A report on the porcupine quill trade in South Africa. Preprint at https://media.withtank.com/23fd88d381/porcupine_quill_trade (2006). Accessed 5 February 2020.

    20.
    Akram, F., Ilyas, O. & Haleem, A. Food and feeding habits of Indian crested porcupine in Pench Tiger Reserve, Madhya Pradesh India. Amb. Sci. https://doi.org/10.21276/ambi.2017.04.1.ra02 (2017).
    Article  Google Scholar 

    21.
    Oussou, C. T. B., Mensah, G. A. & Sinsin, B. Etude de l’écologie du porc-épic (Hystrix cristata) et de son régime alimentaire en milieu naturel. BRAB 53, 45–52 (2006).
    Google Scholar 

    22.
    Duthie, A. G. & Skinner, J. D. Osteophagia in the Cape porcupine Hystrix africaeaustralis. S. Afr. J. Zool. 21(4), 316–318 (1986).
    Google Scholar 

    23.
    Kiibi, J. M. Taphonomic aspects of African porcupines (Hystrix cristata) in the Kenyan Highlands. J. Taphon. 7(1), 21–27 (2009).
    Google Scholar 

    24.
    O’Regan, H. J., Kuman, K. & Clarke, R. J. The likely accumulators of bones: five Cape Porcupine Den assemblages and the role of porcupines in the post-member 6 infill at Sterkfontein South Africa. J. Taphon 9(2), 69–87 (2011).
    Google Scholar 

    25.
    Van Jaarsveld, A. S. & Knight-Eloff, A. K. Digestion in the porcupine Hystrix africaeaustralis. S. Afr. J. Zool. 19(2), 109–112. https://doi.org/10.1080/02541858.1984.11447867 (1984).
    Article  Google Scholar 

    26.
    Gorgas, M. Comparative anatomical studies on the gastrointestinal tract of Sciuromorpha, Hystricomorpha and Caviomorpha (Rodentia). Z. wiss. Zool. 175, 237–404 (1967).
    Google Scholar 

    27.
    Hagen, K. B. et al. Digestive physiology, resting metabolism and methane production of captive Indian crested porcupine (Hystrix indica). J. Anim. Feed Sci. 28(1), 69–77. https://doi.org/10.22358/jafs/102741/2019 (2019).
    Article  Google Scholar 

    28.
    Roth, H. H. Note on the early growth development of Hystrix africaeaustralis. In Zeitschrift fur saugetierkunde (eds Mohr, E. & Rohrs, M.) 313–316 (Verlag Paul Parey, Hamburg und Berlin, 1964).
    Google Scholar 

    29.
    Hutchins, M., Kleiman, D. G., Geist, V. & McDade, M. C. Grzimek’s Animal Life Encyclopedia 2nd edn, Vol. 12–16 (Gale Group, Farmington Hills, MI, 2003).
    Google Scholar 

    30.
    Jori, F., Lopez-Bejar, M. & Houben, P. The biology and use of the African brush-tailed porcupine (Atherurus africanus, Gray, 1842) as a food animal A review. Biodivers. Conserv. 7, 1417–1426 (1998).
    Article  Google Scholar 

    31.
    Marler, P. N., Castro, L. S. G. & Hoevenaars, K. 2015. Mammalian Fauna of the proposed Cleopatra’s Needle Forest Reserve (CNFR): A Camera Trap Study of Palawan’s Mammals. In Proceedings of the 2nd Palawan Research Symposium 2015, 87–93 (2015).

    32.
    White, T. C. R. Mast seeding and mammal breeding: can a bonanza food supply be anticipated?. N.Z. J. Zool. 34, 179–183. https://doi.org/10.1080/03014220709510076 (2007).
    Article  Google Scholar 

    33.
    White, T. C. R. The significance of unripe seeds and animal tissues in the protein nutrition of herbivores. Biol. Rev. 86, 217–224. https://doi.org/10.1111/j.1469-185X.2010.00143 (2011).
    Article  PubMed  Google Scholar 

    34.
    Coppola, F. New knowledge tools for crested porcupine (Hystrix cristata L., 1758) management in the wild. First census model, new behavioural ecology aspects and preliminary investigation on health status. Ph.D. thesis (2020).

    35.
    Coppola, F. et al. First report of Giardia duodenalis infection in the crested porcupine (Hystrix cristata L., 1758). Int. J. Parasitol. Parasites Wildl. 11, 108–113. https://doi.org/10.1016/j.ijppaw.2020.01.006 (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    36.
    Coppola, F. et al. Crested Porcupine (Hystrix cristata L.): a new potential host for pathogenic leptospira among semi-fossorial mammals rested porcupin. Comp. Immunol. Microbiol. Infect. Dis. 70, 101472. https://doi.org/10.1016/j.cimid.2020.101472 (2020).
    Article  PubMed  Google Scholar 

    37.
    Cilia, G., Bertelloni, F., Coppola, F., Turchi, B., Biliotti, C., Poli, A., Parisi, F., Felicioli, A., Cerri, D. & Fratini, F. Isolation of Leptospira serovar Pomona from a crested porcupine (Hystrix cristata, L., 1758). Vet. Med. Sci. https://doi.org/10.1002/vms3.308 (2020).
    Article  PubMed  Google Scholar  More

  • in

    Bile acids drive the newborn’s gut microbiota maturation

    Ethic statement
    All animal experiments were performed in compliance with the German animal protection law (TierSchG) and approved by the local animal welfare committees, the Landesamt für Natur, Umwelt und Verbraucherschutz, North Rhine Westfalia (84–02.04.2016.A207 and 84–02.04.2015.A293). All C57BL6J wildtype mice were bred locally and held under specific pathogen-free conditions at the Institute of Laboratory Animal Science at RWTH Aachen University Hospital. The day of birth was considered day 0, i.e., animals screened at day 1 were approximately 24 h old and verified to have ingested breast milk (abdominal milk spot). Mice were weaned at PND21.
    In vivo study design
    To monitor microbiota and host metabolic development throughout the neonatal period into adulthood, intestinal and hepatic tissues were obtained from C57BL/6 J mice in two different approaches. First, total small intestinal, colon, and liver tissues were obtained from C57BL/6 J mice aged 1, 7, 14, 21, 28, and 56 days (n = 5 per timepoint). In a separate set of experiments, similar tissues were obtained from mice aged 0, 6, 12, 18, and 24 h (n = 10 per timepoint). To rule out potential litter and cage effects, we obtained tissues from a single animal of one litter for a given age group and repeated this by selecting a new animal from the same litter for every other age group (Fig. 1a). Every animal was only examined once at the indicated age (PND). This means that in total 30 animals from 5 litters were used for monitoring the microbiota development between age 1 and 56 days and 46 animals from 10 litters to monitor the microbiota development within the first 24 h. For oral bile acid administration, 7–14 (UDCA and Control, n = 14;, GCA, n = 11; TCA, n = 10; βTMCA, n = 7) PND7 animals received the indicated bile acid (Sigma-Aldrich, Biomol) at a concentration of 70 µg/g body weight or PBS daily for 3 days by oral gavage (average 5 µL) with 100% succession47. Body weight and food/water consumption was monitored daily. To determine the effect of bile acid administration, the intestine was aseptically cut into 10–20 parts and alternately assigned to two collections for microbial profiling and assessment of metabolites. Samples from an additional group of adult 8–12 week old animals (n = 5) were processed in parallel to provide an adult microbiota control. Tissues were transferred into sterile micro-centrifuge tubes and stored at −80 °C before analysis. Liver tissues of 7-, 14-, 21- and 56-day-old germ-free mice were obtained from the Institute for Laboratory Animal Science at Hannover Medical School. Tissues were collected and stored in sterile tubes stored at −80 °C until further analysis.
    DNA isolation and generation of sequence data
    Total metagenomic DNA was isolated from snap frozen small intestinal and colonic tissues by repeated-bead-beating (RBB) combined with chemical lysis plus a column-based purification method48. Approximately 200 mg tissue was added to a 2.0 mL screw-cap tube containing 0.5 g of 0.1 mm zirconia beads (Biospec Products, Bartlesville, OK, USA) and 1 mL ASL lysis buffer from the QIAamp DNA Stool Mini Kit (Qiagen, Hilden, Germany). Samples were incubated at 15 min. at 95 °C and subsequently two successive rounds of bead-beating were employed using a FastPrep®−24 instrument (MP Biomedicals, Inc., USA) at 5.5 ms−1 (3 × 1 min. for each round). To minimize physical damage of DNA in the RBB method, the lysate fraction produced from the first round of bead beating was drawn after centrifugation at full speed (~160,000 × g) for 5 min at 4 °C. A second round of bead beating was performed upon adding 300 μL fresh lysate buffer after which supernatants were pooled. To precipitate nucleic acids, 260 μL 10 M ammonium acetate was added to the lysate tubes, mixed and incubated on ice for 5 min. After centrifugation at 4 °C for 15 min. at full speed, supernatants were transferred to a new 1.5 mL tube to which an equal volume of isopropanol was added. Next, samples were incubated on ice for 30 min., centrifuged at RT for 15. min. at full speed and supernatants were removed by decanting. The nucleic acid pellet was washed with 500 μL 70% EtOH and dried under vacuum for 3 min. The nucleic acid pellet was dissolved in 100 μL TE (Tris-EDTA) buffer. Two microliter DNase-free RNase (1 mg/mL) was added and samples were incubated at 37 °C for 15 min. Next, 15 μL proteinase K and 200 μL AL buffer were added, samples were vortexed for 15 s. and incubated at 70 °C for 10 min. After adding 200 μL ethanol (96–100%) and vortexing, the lysate was transferred to QIAamp spin columns (Qiagen, Hilden, Germany). The DNA was finally purified using the QIAamp DNA Stool Mini Kit according to the manufacturer’s instructions and eluated in 200 μL AE Buffer. For each DNA isolation batch, additional isolation was performed on PCR-grade water as a negative control. Generation of amplicon libraries and sequencing was performed according to a previously published protocol49. Briefly, the V4 region of the 16S rRNA gene was PCR amplified from each DNA sample in duplicate in 50 μL volumes containing 10 pmol of both primers (5′-GTGCCAGCMGCCGCGGTAA*-3′ [515 F] and barcoded 5′-GGACTACHVGGGTWTCTAAT*-3′ [806 R]), 5 μL Accuprime buffer II, 0.2 μL Accuprime Hifi polymerase (Thermo Fisher Scientific, Waltham, WA, USA) and 2 μL DNA. After an initial denaturation step at 94 °C for 3 min., amplification was carried out for 30 s. at 94 °C, 45 s at 50 °C and 1 min. at 72 °C. Amplification was carried out in 35 cycles for PND1-PND56; the samples with a low yield of DNA required 40 cycles (0–24 h) The PCR program ended with a final post-PCR incubation step of 10 min at 72 °C to promote complete synthesis of all PCR products. Pooled amplicons from the duplicate reactions were purified using AMPure XP purification (Agencourt, Massachusetts, USA) according to the manufacturer’s instructions and subsequently quantified by Quant-iT PicoGreen dsDNA reagent kit (Invitrogen, New York, USA). Amplicons were mixed in equimolar concentrations, to ensure equal representation of each sample, and sequenced on an Illumina MiSeq instrument using the V3 reagent kit (2 × 250 cycles). All V4 16S rDNA bacterial sequences generated in this study have been submitted to the Qiita and ENA databases under accession No. 10719 and ERP116798, respectively.
    Sequence analysis
    Data demultiplexing, length and quality filtering, pairing of reads and clustering of reads into Operational Taxonomic Units (OTUs) at 97% sequence identity was performed using the online Integrated Microbial Next Generation Sequencing (IMNGS, www.imngs.org) platform using default settings50. Removal of primers and technical reads resulted in fragments of approximately 250 bases. Sequencing was performed from both the 3′ and 5′ side resulting in sufficient resolution. IMNGS is a UPARSE based analysis pipeline51. Pairing, quality filtering and OTU clustering (97% identity) was done by USEARCH 8.052. The analysis was based on OTUs rather than amplicon sequence variants (ASVs) since we aimed at aggregating taxa at a higher level and wanted to avoid overestimation of prokaryotic diversity due to Intragenomic heterogeneity of 16S rRNA genes53. Chimera filtering was performed by UCHIME (with RDP set 15 as a reference database54. Taxonomic classification was done by RDP classifier version 2.11 training set 15.8 Sequence alignment was performed by MUSCLE and treeing by Fasttree55,56. A total of 21,372,397 V4 reads were generated over two runs. After trimming, quality filtering, removal of potential chimeric reads, de-multiplexing and removal of low abundant operational taxonomic units (OTUs), 15,692,587 sequences belonging to 478 OTUs were retained for downstream analysis. Negative controls were evaluated based on their number of sequences and composition compared to other samples. We used for each batch, sampling blank controls, DNA blank extraction controls and no-template amplification controls, and monitored the lack of contaminant bacterial DNA load herein by a gel-based principle. Moreover, we compared the acquired OTUs composition from the negative controls to our low abundant microbial samples to ensure that our findings were not driven by potential contaminant taxa. Subsequently, samples of the negative controls and with low sequencing depth (less than 6,749 sequences/sample) were excluded from subsequent analysis. For the remaining samples, the number of sequences per sample ranged from 6749 to 239,395 (median 87,227).
    Richness, diversity, taxonomy, and enterotype analyses
    Data normalization, diversity, taxonomical binning and group comparisons were performed using the Rhea package version 1.657. In order to not discard informative data, normalization in Rhea was performed by dividing OTU counts per sample for their total count (sample depth) followed by multiplying all of the obtained relative abundance for the lowest sample depth (6749 reads/sample). Alpha- (observed species and Shannon index) and the generalized Unifrac beta-diversity index were calculated using the Rhea package58. Additional beta-diversity indices (weighted Unifrac, unweighted Unifrac, and Bray–Curtis distance) were calculated using the R package (version 3.6.1.) Phyloseq package version 1.30.059. Ordination of samples according to their microbial composition expressed as Hellinger transformed genus abundance data or beta-diversity indices was visualized using Principal Component Analysis (PCA) and Principal Coordinate Analysis (PCoA), respectively. All ordinations were constructed using the R package Phyloseq and included 95% confidence ellipses. Dirichlet multinomial mixtures models (DMM) were used to calculate genus-level enterotypes60. When including samples from all time-points, Laplace approximation revealed an optimal number of three clusters.
    Downstream microbial analyses and presentation
    Smoothing of the kinetic for dominant taxa (Fig. 1e and f, as well as Supplementary Fig. 1d and e) was generated using the geom_smooth function of the ggplot package 3.2.1. with default settings. The lines reflect the mean values of the relative abundance. The appearance and disappearance of OTUs of the dominant phyla (Actinobacteria, Proteobacteria, Firmicutes, Bacteroidetes) with age was visualized in Sankey-plots (SankeyMATIC.com). For readability of Sankey-plots, only OTUs present in >10% of all samples per timepoint and with a prevalence of >20% in the entire dataset, were included. Ecosystem specific functional metagenome predictions were created by the novel PICRUSt-iMGMC workflow (using PICRUSt version 1.1.3) with the de novo picked OTUs and using mouse metagenome-assembled genomes linked to 16S rRNA genes61. The derived KEGG orthologs were mapped into multiple pathways or modules. Differentially changed KEGG modules were identified using the pathways enrichment analyses from MicrobiomeAnalyst with default settings62. The identification of lactobacillus-related OTUs were performed using EZbiocloud and used to construct a phylogentic tree of lactobacilli by MEGA7 (MUSCLE) for alignment and iTOL v4 for the final annotations (itol.embl.de) with default settings.
    Liver metabolomics and bile acid analyses
    The metabolome analyses were carried out with the AbsoluteIDQ® p180 Kit (Biocrates Life Science AG, Austria. The kit allowed identification and quantification of 188 metabolites from 5 compound classes (acyl carnitines, amino acids, glycerophospho-, and sphingolipids, biogenic amines, and hexoses). The kit used flow injection tandem mass spectrometry (FIA) for the non-polar metabolites and LC-MS/MS for the more polar compounds. The integrated MetIDQ Software (version Boron 2623) streamlined the data analysis by automated calculation of metabolite concentrations. Quantification of analytes utilized stable isotope-labeled or chemically homologous internal standards (IS). Controls were included for 3 different concentration levels. For calibration, the kit contained a calibrator mix at 7 different concentrations. The measurements were carried out with an ABI Sciex API5500 Q-TRAP mass spectrometer via Electrospray ionization (ESI) by Multi Reaction Monitoring (MRM) mode for high specifity and sensitivity. 158 MRM pairs were measured in positive ion mode (13 IS) and 2 MRM pairs were measured in negative mode (1 IS). The following additional chemicals for LC-MS were used: water, Millipore; PITC, Fluka; pyridine, Fluka (p.a.); methanol, Merck; Lichrosolv for LC/MS; acetonitrile, Merck; formic acid, Sigma Aldrich. Metabolites were extracted from liver samples by adding H2O/acetonitrile (1:1,v:v) per mg sample followed by homogenization with a tissue disruptor (10 min, 30 Hz, 4 steel balls). The samples were centrifuged (1400×g, 2 min) and the supernatant analyzed. The targeted analysis was performed by adding 10 μL of extracted liver sample to the AbsoluteIDQ® p180 Kit (Biocrates Life Science AG, Innsbruck, Austria), following the vendor’s instructions63. For bile acid measurements the MS-based Bile Acids Kit (Biocrates Life Sciences AG, Innsbruck, Austria), a 96-well plate format assay, was used following the manufacturer´s instructions with normalization based on mass64. The following settings were used: turbo spray for ion source, 20 for Curtian Gas, medium for CAD Gas, 40 psi for ion source gas 1 and 50 psi for ion source gas 2. For the bile acid measurements we used an ion spray voltage of −4500 V and a temperature of 600 °C; for the p180 kit we used an ion spray voltage of 5500 V and a temperature of 500 °C. All metabolomic data generated in this study have been submitted to the Metabolomics Workbench and have been assigned the Study ID ST001388, ST001389, ST001396, ST001397, the Project ID PR000952 and the Project DOI 10.21228/M8N397.
    The contribution of each metabolite to metabolomic variation was derived from age-constrained redundancy analysis (RDA) based on all metabolites stratified in group levels with additional scaling by normalization based on z-scores (Phyloseq package). Moreover, PCA was used to illustrate changes in the composition of the different metabolic groups during the postnatal period for both PC1 and PC2, and PC2 and PC3. For the bile acids the 2nd and 3rd component were chosen for plotting, since the first component was mainly driven by the strong separation observed between the first and subsequent time points (PND1 vs. PND7–56).
    Multi-omics analyses
    Regularized canonical correlation analyses (rCCA) were performed (Mixomics package 6.10.8)65 to unravel specific correlations between bile acids and OTUs with a minimal presence of 20% in all samples. Samples were excluded from the microbiota-data if they were not measured in the bile acid analyses (i.e. PND1 of litter 1 and 2). Prior to rCCA a hyperbolic sine transformation was used on OTU-counts and a log-transformation for the bile acids. For the estimation of regularization (penalization) parameters λ1 and λ2, the cross-validation procedure (CV) method was used. We used a λ1 = 0.0001, λ2 = 1 with a CV-score = 0.4779644 and 2 components. OTUs with a correlation between −0.3 and 0.3 on the first 2 components were filtered out to optimize the rCCA.
    The Spearman’s rank correlation coefficient was calculated between bile acids (weight corrected) and bacterial genera (relative abundances) with a minimal presence of 20% in all samples. Benjamini and Hochberg FDR correction was performed to correct for multiple testing (p  > 0.05). For the heatmap only significant correlations with adjusted p-value of More