More stories

  • in

    Social transmission in the wild can reduce predation pressure on novel prey signals

    Study siteThe experiment was conducted at Madingley Wood, Cambridgeshire, UK (0◦3.2´E, 52◦12.9´N) during summer 2018. Madingley Wood is an established field site with an ongoing long-term study of the blue tit and great tit populations. During the autumn and winter birds are caught from feeding stations using mist nets and they are fitted with British Trust of Ornithology (BTO) ID rings. Since 2012, blue tits and great tits have been fitted with RFID tags (BTO Special Methods permit to HMR), which enables collecting data remotely about their foraging behavior and social relationships. The study site has 90 nest boxes that are monitored annually during the breeding season. In 2018 chicks (n = 325) fledged successfully from 45 nest boxes (blue tits = 21, great tits = 24) and they were all ringed and fitted with RFID tags when they were approximately 10 days old. Because new juvenile flocks were arriving at our study site throughout the summer, we also conducted several mist-netting and ringing sessions in July and August to maintain a high proportion of blue tits and great tits ringed and RFID tagged for the experiments (on average 89%, see below). The study protocol was approved by the Animal Users Committee at the Department of Zoology, University of Cambridge.Food itemsWe investigated birds’ foraging choices by offering them colored almond flakes at bird feeders that were distributed throughout the wood. Before beginning the experiments, we allowed the birds to become familiar with the food items by providing plain ‘control’ almonds (plain and not colored) in paired feeders (1.5 m apart) at three locations (approximately 170 m from each other). The feeders were surrounded with metal cages to exclude larger birds, and we placed plastic buckets under the feeders to collect any spilled almonds and minimize birds’ opportunities to forage from the ground. We introduced the feeders at the beginning of June when the nestlings had fledged and were beginning to forage independently, and continued to provide these plain almonds in between our learning experiments (Fig. 3c).In the learning experiments, almond flakes were dyed with non-toxic food dye (Classikool Concentrated Droplet Food Colouring). We used three different color pairs: green (Leaf Green) and red (Bright Red), purple (Lavender Purple) and blue (Royal Blue), and orange (Satsuma Orange) and yellow (Dandelion Yellow). Almond flakes were dyed by soaking them for approximately 20 min in a solution of 900 ml of water and 30 ml of food dye and then left to air dry for 48 h. In the avoidance learning experiments, we made half of the almond flakes unpalatable by soaking them for one hour in 67% solution of chloroquine phosphate, following previously established methods from avoidance learning studies with birds in captivity14,23,24,25. The food dye was added to the solution during the last 20 min.Red and green are common colors used by aposematic, or cryptic prey, respectively9. Therefore, we investigated whether blue tits and great tits had initial color biases towards these colors before starting the main experiment. Because we did not want the birds in our study population to have any experience of the colors before the main experiment, this pilot study was conducted in Newbury, which is 130 km from our main study site. Birds were simultaneously presented with two feeders containing red and green almonds (both palatable) for 30 min and the number of almonds of each color taken by blue tits or great tits was recorded using binoculars. The position of the feeders was switched after 15 min to control for any preferences for feeder location, and the test was repeated on 9 different days. We did not find any evidence that birds had initial color preferences (t-test: t = 0, df = 15.69, p = 1). For the other two learning experiments, we chose color pairs that were available as a food dye and as different from red and green in the visible spectrum as possible to avoid generalization across experiments. These color pairs (blue/purple and yellow/orange) had similar contrast ratios as green and red, based on their RGB values (measured from photographs, see Supplementary Information). Although avian and human vision is different, the discriminability of colors is likely to be similar51, and rapid avoidance learning in each experiment shows that all colors were easily distinguishable. This was the main requirement for testing social information use, and subtle differences in color pair discriminability should only introduce noise to our data but not influence our conclusions.Learning experiments with colored almondsWe conducted three avoidance learning experiments with different color pairs throughout the summer: red/green, blue/purple, and yellow/orange (unpalatable/palatable). In addition, we conducted a reversal-learning experiment with the blue/purple color pair by making both colors palatable after birds had acquired avoidance to blue almonds. Each experiment followed a similar protocol, in which birds were presented with colored almonds at the same three feeding stations where they were previously offered plain almonds. Each feeding station had two feeders, where one contained the palatable color and the other contained the unpalatable color (except in the reversal learning test when both colors were palatable). We switched the side of the feeders every day to make sure that birds learned to associate palatability with an almond color and not a feeder position. The feeders were filled at least once a day (or more often if necessary) to make sure that birds always had access to both colors. We continued each avoidance learning experiment until >90% of all recorded visits were to the feeder with palatable almonds, indicating that most birds in the population had learned to discriminate the colors. This took 7 days in the red/green experiment and 8 days in the other two color pairs (blue/purple and yellow/orange). The reversal learning experiment was finished after 9 days when 50% of the visits were to the previously unpalatable color (blue), indicating that most birds had reversed their learned avoidance towards it.Recording visits to feedersWe monitored visits to all feeders using RFID antennas and data loggers (Francis Scientific Instruments, Ltd) that scanned birds’ unique RFID tag codes when they landed on a perch attached to the feeder. During the learning experiments, each day we also recorded videos from all three feeding stations (using Go Pro Hero Action Camera and Canon Legria HF R66 Camcorder). From the videos, we monitored the proportion of blue tits and great tits that did not have RFID tags and were therefore not recorded when visiting the feeders. We calculated the estimated RFID tag coverage for each day of the experiments by watching at least 100 visits to the feeders from the videos (divided equally among the three feeding stations) and recording whether blue tits and great tits had an RFID tag or not. We realized that the number of untagged individuals was very high (approximately 50% of all visiting birds) when we started the experiment with the first color pair (red/green; see Supplementary Fig. 3). We, therefore, stopped the experiment after two days and caught birds from the feeding stations with mist nets to fit RFID tags to new individuals. To maintain a high number of individuals RFID tagged for the other color pairs, we conducted a mist netting session a day before starting each experiment, as well as 4–5 days after it. We always switched the feeders back to containing plain almonds during mist-netting sessions to ensure that this would not interfere with the learning experiments. Apart from the first two days of the red/green experiment, the RFID tag coverage was on average 89% throughout the experiments (varying between 80 and 95%, Supplementary Fig. 3).Birds were recorded every time that they visited the feeders, i.e., landed on the RFID antenna. However, it is possible that birds did not take the almond during every visit. To get an estimate of how often birds landed on the antenna without taking the almond, and whether this differed between palatable and unpalatable colors, we analyzed the visits to the feeders from the video recordings. We watched videos from the five first days of each experiment (i.e., different color pairs) and analyzed 60 visits to each color (divided approximately equally among the three feeding stations). We recorded whether the feeding event happened (birds ate the almond at the feeder or flew away with it) or whether birds left the feeder without sampling the almond. Because the number of visits to the unpalatable feeder was low during the last days of the avoidance learning experiments, we decided not to analyze avoidance learning videos after day five (but recorded visits from all days of the reversal learning experiment). We found that in avoidance learning experiments birds started to ‘reject’ unpalatable almonds after two days, i.e., they sometimes landed on the feeder but flew away without taking the almond (see Supplementary Fig. 4a). This change was not observed at palatable feeders where birds continued to consume almonds at a similar rate as at the beginning of the experiment (Supplementary Fig. 4a). In reversal learning, the proportion of visits that did not include a feeding event did not differ between purple and blue almonds: birds showed similar hesitation towards both colors at the beginning of the experiment, but this wariness decreased when the experiment progressed, with birds taking the almond during most of their visits (Supplementary Fig. 4b).Statistical analyses and model validationForaging choices in learning experimentsWe first analyzed how birds’ foraging choices changed during the learning experiments using generalized linear mixed-effects models with a binomial error distribution. The number of times an individual visited each feeder on each day of the experiment was used as a bounded response variable, and this was explained by species (blue tit/great tit), individuals’ age (juvenile/adult), and day of the experiment (continuous variable), as well as bird identity as a random effect. When analyzing avoidance learning, initial exploration of data suggested that results were similar across all three experiments, so we combined the experiments in the same model. To investigate whether learning curves differed between the species or age groups, the day of the experiment was included as a second-order polynomial term, and we started model selections with models that included a three-way interaction between species, age, and day2. Best-fitting models were selected based on Akaike’s information criterion (see Supplementary Tables 1 and 2).Social networkTo investigate if birds used social information in their foraging choices, we first constructed a social network of the bird population based on their visits to feeders outside of the learning experiments, i.e., when birds were presented with plain almonds (in total 92 days, see Supplementary Information for the robustness of analysis to exclusion of network data before or after the experiment). We used only these data as individuals were likely to vary in their hesitation to visit novel colored almonds. We used a Gaussian mixture model to detect the clusters of visits (‘gathering events’) at the feeders52 and then calculated association strengths between individuals based on how often they were observed in the same group (gambit of the group approach). These associations (network edges) were calculated using the simple ratio index, SRI35.$$frac{x}{x+{y}_{{mathrm{A}}}+{y}_{{mathrm{B}}}+{y}_{{{mathrm{AB}},}}}$$
    (2)
    where x is the number of samples where individuals A and B co-occurred in the same group, yA is the number of samples where only individual A was seen, yB is the number of samples where only individual B was seen, and yAB is the number of samples where both A and B were observed in the same sample but not together. Network associations, therefore, estimated the probability that two individuals were in the same group at a given time, with the values scaled between 0 (never observed in the same group) and 1 (always observed in the same group).Social information use during avoidance learning: model descriptionIf social avoidance learning was occurring, then the more birds observed negative responses of others feeding on the unpalatable feeder, the less likely they would be to choose the unpalatable feeder themselves. Thus, we expected the probability of bird j choosing the unpalatable option at time t to decrease with ({R}_{-,j}left(tright)) (the real number of negative feeding events observed by j prior to time t). Likewise, if appetitive social learning was occurring, then the more birds observed positive responses of others feeding on the palatable feeder, the more likely they would be to choose the palatable feeder themselves (rather than the unpalatable feeder). So, we also expected the probability of j choosing the unpalatable option at time t to decrease as ({R}_{+,j}left(tright),)(the real number of positive events observed by j prior to time t) increased.However, we could not test for an effect of ({R}_{-,j}left(tright)) and ({R}_{+,j}left(tright)) directly, since birds often ate the almond away from the feeder, and therefore the real number of observed feeding events could not be measured. Instead, we aimed to test for a pattern following the social network that is consistent with these social learning processes. We reasoned that the probability that one individual i, observes a specific feeding event by another individual j, was proportional to the network connection between them, aij (probability they are in the same feeding group at a given time). Therefore, in each avoidance learning experiment (i.e., different color pair), we calculated the expected number of negative feeding events observed, prior to each choice (occurring at time t) as$${O}_{-,i}left(tright)={sum }_{j}{N}_{-,j}left(tright){a}_{{ij}},$$
    (3)
    where ({N}_{-,j}left(tright)) was the number of times j had visited unpalatable almonds prior to time t (i ≠ j), and summation is across all birds in the network, and likewise for the expected number of positive feeding events:$${O}_{+,i}left(tright)={sum }_{j}{N}_{+,j}left(tright){a}_{{ij}},$$
    (4)
    where ({N}_{+,j}left(tright)) was the number of times j had visited palatable almonds prior to time t (i ≠ j).We analyzed whether the expected observations of positive and/or negative feeding events of others influenced the foraging choices in the avoidance learning experiments using generalized linear mixed-effects models with a binomial error distribution. We used each choice (i.e., visit a feeder) as a binary response variable (1 = unpalatable chosen, 0 = palatable chosen), with the probability that unpalatable feeder is chosen on feeding event E given by ({p}_{E}={p}_{-,{i}(E)}left({t}_{E}right)), where i(E) is the individual that fed during event E and ({t}_{E}) is the time at which event E occurred. We then modeled the probability of i choosing the unpalatable option at time t as:$${p}_{-,i}left(tright)={rm{logit}}left(alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}left(tright)+{beta }_{{rm{asoc}}-}{N}_{-,i}left(tright)+{beta }_{{rm{soc}}+}{O}_{+,i}left(tright)+{beta }_{s{rm{oc}}-}{O}_{-,i}left(tright)+{{{{rm{B}}}}}_{i}right),$$
    (5)
    where ({N}_{+,i}left(tright)) is the number of times a choosing individual had visited the palatable feeder (positive personal information), ({N}_{-,i}left(tright)) is the number of times a choosing individual had visited the unpalatable feeder (negative personal information), ({O}_{+,i}left(tright)) is the expected number of observed positive (positive social information) and ({O}_{-,i}left(tright)) observed negative feeding events (negative social information). Bird identity was included as a random effect, ({{rm{{B}}}}_{i}) (age and species were later added as variables, see below). Parameters ({beta }_{{rm{asoc}}+}) and ({beta }_{{rm{asoc}}-}) are the effects of asocial learning about the palatable and unpalatable foods, ({beta }_{{rm{soc}}+}) is the effect of social learning about the palatable food, and ({beta }_{{rm{soc}}-})is the effect of social avoidance learning about the unpalatable food. Estimation of these parameters, with associated Wald tests and confidence intervals, allowed us to make inferences about which effects were operating and the size of these effects. To aid model fitting we standardized all predictor variables and then back-transformed the effects to the original scale (see Supplementary Tables 3–5 for the model outputs). To assess the importance of asocial and social effects, we also ran separate models that excluded either asocial or social parameters and compared them to the initial model in Eq. (5) using Akaike’s information criterion (see Supplementary Table 6). However, in most cases, this reduced model fit significantly, and we, therefore, kept all parameters in the final models.Our approach took ({O}_{-,{i}}left(tright)) as a measure of ({R}_{-,j}left(tright)), and ({O}_{+,{i}}left(tright)) as a measure of ({R}_{+,j}left(tright))-, which we termed the ‘expected’ number of observations of each type. Strictly speaking, ({O}_{-,{i}}left(tright)) and ({O}_{+,{i}}left(tright)) were upper limits on the expected number of observations, assuming that birds observed all feeding events in the groups in which they were present, whereas only an unknown proportion of such events (({p}_{o})) was observed. Therefore, the real expected number of negative/positive observations would be (Eleft({R}_{-,j}left(tright)right)={p}_{o}{O}_{-,{i}}left(tright)) and (Eleft({R}_{+,j}left(tright)right)={p}_{o}{O}_{+,{i}}left(tright)) respectively. Thus, the coefficient, ({beta }_{{rm{soc}}-}), for the effect of ({O}_{-,{i}}left(tright)) could be interpreted as ({beta }_{s{rm{oc}}-}={p}_{o}acute{{beta }_{{rm{soc}}-}}) where (acute{{beta }_{s{rm{oc}}-}}) is the effect per observation. Note that since ({p}_{o}le 1), and(,{beta }_{{rm{soc}}-}=acute{{beta }_{s{rm{oc}}-}}{p}_{o}), ({beta }_{s{rm{oc}}-}) is more likely to underestimate than overestimate the effect per observation of a negative feeding event. An analogous argument applies to the coefficient, ({beta }_{{rm{soc}}+}), for the effect of ({O}_{+,{i}}left(tright)).Social information use during avoidance learning: extension to test for species effectsAfter fitting the initial model shown in Eq. (5), we further broke down the model to test whether individuals were more likely to learn socially by observing conspecifics than heterospecifics. This was done by splitting the expected number of observed positive and negative feeding events to observations of conspecifics (({O}_{+{rm{C}},i}left(tright)), ({O}_{-{rm{C}},i}left(tright))) and heterospecifics (({O}_{+{rm{H}},i}left(tright)), ({O}_{-{rm{H}},i}left(tright))), and including these in the model as separate explanatory variables thus:$${p}_{-,i}(t)={rm{logit}}left(begin{array}{c}alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}(t)+{beta }_{{rm{asoc}}-}{N}_{-,i}(t)\ +{beta }_{{rm{soc}},+{rm{H}}}{O}_{+{rm{H}},i}(t)+{beta }_{{rm{soc}},-{rm{H}}}{O}_{-{rm{H}},i}(t)\ +{beta }_{{rm{soc}},+{rm{C}}}{O}_{+{rm{C}},i}(t)+{beta }_{{rm{soc}},-{rm{C}}}{O}_{-{rm{C}},i}(t) \ +{{{{rm{B}}}}}_{i}end{array}right)$$
    (6a)
    with ({beta }_{{rm{soc}},-{rm{H}}}) and ({beta }_{{rm{soc}},-{rm{C}}}) giving the effect of a negative observation of a heterospecific and conspecific, respectively, whereas ({beta }_{{rm{soc}},+{rm{H}}}) and ({beta }_{{rm{soc}},+{rm{C}}}) give the effect of positive observation of a heterospecific and conspecific, respectively. In general –/+ subscripts refer to negative/positive feeding events and C/H subscripts to feeding events by conspecifics/heterospecifics. By re-parameterizing the model thus:$${p}_{-,i}left(tright)={rm{logit}}left(begin{array}{c}alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}left(tright)+{beta }_{{rm{asoc}}-}{N}_{-,i}left(tright) \ +{beta }_{{rm{soc}},{rm{H}}+}{O}_{+,i}left(tright)+{beta }_{{rm{soc}},{rm{H}}-}{O}_{-,i}left(tright)\ +left({beta }_{{rm{soc}},{rm{C}}+}-{beta }_{{rm{soc}},{rm{H}}+}right){O}_{+{rm{C}},i}left(tright)+left({beta }_{{rm{soc}},{rm{C}}-}-{beta }_{{rm{soc}},{rm{H}}-}right){O}_{-{rm{C}},i}left(tright)\ +{{{{rm{B}}}}}_{i}end{array}right)$$
    (6b)
    we were able to test for a difference between observations of negative feeds by conspecifics and heterospecifics (left({beta }_{{rm{soc}},{rm{C}}-}-{beta }_{{rm{soc}},{rm{H}}-}right)) and between observations of positive feeds by conspecifics and heterospecifics (left({beta }_{{rm{soc}},{rm{C}}+}-{beta }_{{rm{soc}},{rm{H}}+}right)).For all experiments there was no evidence for a difference between ({beta }_{{rm{soc}},-{rm{H}}}) and ({beta }_{{rm{soc}},-{rm{C}}}) (yellow/orange: Z = 0.803, p = 0.42; red/green: Z = 0.065, p = 0.95; blue/purple: Z = 1.113, p = 0.27). However, there was some evidence of a difference between ({beta }_{{rm{soc}},+{rm{H}}}) and ({beta }_{{rm{soc}},+{rm{C}}}) in two of the three experiments (yellow/orange: Z = 1.359, p = 0.17; red/green: Z = 1.417, p = 0.16; blue/purple: Z = 0.729, p = 0.47). Consequently, we reduced the model down to:$${p}_{-,i}left(tright)={rm{logit}}left(begin{array}{c}alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}left(tright)+{beta }_{{rm{asoc}}-}{N}_{-,i}left(tright)\ +{beta }_{{rm{soc}},-}{O}_{-,i}left(tright)+{beta }_{{rm{soc}},+{rm{H}}}{O}_{+{rm{H}},i}left(tright)+{beta }_{{rm{soc}},+{rm{C}}}{O}_{+{rm{C}},i}left(tright)\ +{{{{rm{B}}}}}_{i}end{array}right)$$
    (7)
    for further analysis, i.e., with different effects for observations of conspecific/heterospecific positive feeds, but not of negative feeds. We did this for all color combinations (including blue/purple) to allow comparison across experiments (see Table 1). The R code used to run these models can be found in Supplementary data53 in ‘GLMM models Orange Yellow final.r’.Social information use during avoidance learning: simulations to test for a network effectNext, we tested whether the social effects we detected followed the social network. When using a network-based diffusion analysis (NBDA43), researchers can compare a network model with one in which the network has homogeneous connections among all individuals, but we found this to be unreliable for our model. Instead, we used a simulation approach to generate a null distribution for the null hypothesis of homogeneous social effects, taking the size of the social effects from the fitted models. We ran 1000 simulations (using the same procedure described above) for all social effects that were found to be significant in each avoidance learning model (each color pair; see Table 1). The total number of expected observations was kept equal, but we homogenized the observation effect across all birds by replacing the probability of bird i observing a feed by bird j, previously ({a}_{{ij}}), with ({sum }_{i}{a}_{{ij}}/n), where n is the number of birds in the experiment, (i.e., all birds had the same probability of observing each feeding event). The model was fitted to the simulated data each time to extract the Z value (Wald test statistic) of the social effect of interest. The distribution of these values was then used as a null distribution to test whether our observed social effect differed from the effects that did not follow the social network. To this end, we calculated the proportion of simulations that yielded a Z value as extreme or more extreme than that observed (judged by distance in either direction from the mean of the null distribution). The R code used to run these simulations can be found in Supplementary data53 in ‘Simulations to test if network effects follow network Orange Yellow.r’.Social information use during avoidance learning: extension to test for age effectsWe then aimed to test whether each of the three social effects detected differed based on the age class of the observed individual (adult versus juveniles). We, therefore, split the negative expected observations ({O}_{-,i}left(tright)) into the expected observations of adults ({O}_{-{rm{A}},i}left(tright)) and juveniles ({O}_{-{rm{J}},i}left(tright)), each with its associated coefficient in the model ({beta }_{{rm{soc}},-{rm{A}}}) and ({beta }_{{rm{soc}},-{rm{J}}}). Likewise, we split positive observations of conspecifics as ({O}_{+{rm{CA}},i}left(tright)) and ({O}_{+{rm{CJ}},i}left(tright)) and positive observations of heterospecifics as ({O}_{+{rm{HA}},i}left(tright)) and ({O}_{+{rm{HJ}},i}left(tright)) to give the model:$${p}_{-,i}left(tright)={rm{logit}}left(begin{array}{c}alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}left(tright)+{beta }_{{rm{asoc}}-}{N}_{-,i}left(tright)\ +{beta }_{{rm{soc}},-{rm{A}}}{O}_{-{rm{A}},i}left(tright)+{beta }_{{rm{soc}},-{rm{J}}}{O}_{-{rm{J}},i}left(tright)\ +{beta }_{{rm{soc}},+{rm{HA}}}{O}_{+{rm{HA}},i}left(tright)+{beta }_{{rm{soc}},+{rm{HJ}}}{O}_{+{rm{HJ}},i}left(tright)\ +{beta }_{{rm{soc}},+{rm{CA}}}{O}_{+{rm{CA}},i}left(tright)+{beta }_{{rm{soc}},+{rm{CJ}}}{O}_{+{rm{CJ}},i}left(tright)\ +{{{{rm{B}}}}}_{i}end{array}right)$$
    (8a)
    As before, –/+ subscripts refer to negative/positive feeding events, C/H subscripts to feeding events by conspecifics/heterospecifics, and A/J subscripts to feeding events by adults/juveniles. We also fitted a re-parameterized version allowing us to test for a difference between expected observations of adults and observations of juveniles for each of the three social effects:$${p}_{-,i}left(tright)={rm{logit}}left(begin{array}{c}alpha +{beta }_{{rm{asoc}}+}{N}_{+,i}left(tright)+{beta }_{{rm{asoc}}-}{N}_{-,i}left(tright)\ +{beta }_{{rm{soc}},-{rm{J}}}{O}_{-,i}left(tright)+left({{beta }_{{rm{soc}},-{rm{A}}}-beta }_{{rm{soc}},-{rm{J}}}right){O}_{-{rm{A}},i}left(tright)\ +{beta }_{{rm{soc}},+{rm{H}}}{O}_{+{rm{H}},i}left(tright)+left({{beta }_{{rm{soc}},+{rm{HA}}}-beta }_{{rm{soc}},+{rm{HJ}}}right){O}_{+{rm{JA}},i}left(tright)\ +{beta }_{{rm{soc}},+{rm{C}}}{O}_{+{rm{C}},i}left(tright)+left({{beta }_{{rm{soc}},+{rm{CA}}}-beta }_{{rm{soc}},+{rm{CJ}}}right){O}_{+{rm{CA}},i}left(tright)\ +{{{{rm{B}}}}}_{i}end{array}right)$$
    (8b)
    The R code used to run these models can be found in Supplementary data53 in ‘GLMM models Orange Yellow final.r’. The main results of each model are presented in Table 2 and full model outputs in Supplementary Tables 3–5.Social information use during reversal learningTo investigate social information use during reversal learning, we used the order of acquisition diffusion analysis (OADA), a variant of NBDA43, which explores the order in which individuals acquire a behavioral trait44. The rate of social transmission between two individuals is assumed to be linearly proportional to their network connection, and the spread of trait acquisition is therefore predicted to follow the network patterns if individuals are using social information. We used NBDA to investigate whether the order of individuals’ first visit to the previously unpalatable blue almonds (mimics) followed the network. We fitted several different models that included (i) only asocial learning, (ii) social transmission of information following a homogeneous network (equal associations among all individuals), or (iii) social transmission of information following our observed network. Models that included social transmission were further divided into models with equal or different transmission rates from adults and juveniles, and from conspecifics and heterospecifics, by constructing separate networks for each adult/juvenile and conspecific/heterospecific combination. To investigate whether asocial or social learning rates differed between blue tits and great tits, we included species as an individual-level variable. We then compared different social transmission models that assumed that species differed in both asocial and social learning rates, only in asocial or only in social learning rates, or that they did not differ in either (see Table 3). The best-supported model was selected using a model-averaging approach with Akaike’s information criterion corrected for small sample sizes. All analyses were conducted with the software R.3.6.154, using lme455, asnipe56, and NBDA57 packages.Reporting summaryFurther information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Determinants of moult haulout phenology and duration in southern elephant seals

    1.IPCC Climate Change 2014: Synthesis report. In Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (eds Core Writing Team et al.) (IPCC, 2014).
    Google Scholar 
    2.Keogan, K. et al. Global phenological insensitivity to shifting ocean temperatures among seabirds. Nat. Clim. Chang. 8, 313–318 (2018).ADS 
    Article 

    Google Scholar 
    3.Chambers, L. E. et al. Phenological changes in the southern hemisphere. PLoS ONE 8, E75514 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    4.Thackeray, S. J. et al. Phenological sensitivity to climate across taxa and trophic levels. Nature 535, 241–245 (2016).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Forrest, J. & Miller-Rushing, A. J. Toward a synthetic understanding of the role of phenology in ecology and evolution. Philos. Trans. R. Soc. Lond. Series B Biol. Sci. 365, 3101–3112 (2010).Article 

    Google Scholar 
    6.Møller, A. P., Rubolini, D. & Lehikoinen, E. Populations of migratory bird species that did not show a phenological response to climate change are declining. Proc. Natl. Acad. Sci. 105, 16195–16200 (2008).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    7.Hipfner, J. M. Matches and mismatches: Ocean climate, prey phenology and breeding success in a zooplanktivorous seabird. Mar. Ecol. Prog. Ser. 368, 295–304 (2008).ADS 
    Article 

    Google Scholar 
    8.Moyes, K. et al. Advancing breeding phenology in response to environmental change in a wild red deer population. Glob. Change Biol. 17, 2455–2469 (2011).ADS 
    Article 

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

    Google Scholar 
    10.Hauser, D. D. et al. Decadal shifts in autumn migration timing by Pacific Arctic beluga whales are related to delayed annual sea ice formation. Glob. Change Biol. 23, 2206–2217 (2017).ADS 
    Article 

    Google Scholar 
    11.Beltran, R. S., Kirkham, A. L., Breed, G. A., Ward Testa, J. & Burns, J. M. Reproductive success delays moult phenology in a polar mammal. Sci. Rep. 9, 1–12 (2019).CAS 
    Article 

    Google Scholar 
    12.Oosthuizen, W. C. et al. Individual heterogeneity in life-history trade-offs with age at first reproduction in capital breeding elephant seals. Popul. Ecol. 61, 421–435 (2019).Article 

    Google Scholar 
    13.Ling, J. K. The skin and hair of the southern elephant seal, Mirounga leonina (Linn). IV. Annual cycle of pelage follicle activity and moult. Aust. J. Zool. 60, 259–271 (2012).Article 

    Google Scholar 
    14.Johanos, T. C., Becker, B. L. & Ragen, T. J. Annual reproductive cycle of the female Hawaiian monk seal (Monachus schauinslandi). Mar. Mamm. Sci. 10, 13–30 (1994).Article 

    Google Scholar 
    15.Beltran, R. S., Burns, J. M. & Breed, G. A. Convergence of biannual moulting strategies across birds and mammals. Proc. R. Soc. B Biol. Sci. 285, 20180318 (2018).Article 

    Google Scholar 
    16.Boyd, I., Arnbom, T. & Fedak, M. Water flux, body composition, and metabolic rate during moult in female southern elephant seals (Mirounga leonina). Physiol. Zool. 66, 43–60 (1993).Article 

    Google Scholar 
    17.Postma, M., Bester, M. N. & De Bruyn, P. J. N. Spatial variation in female southern elephant seal mass change assessed by an accurate non-invasive photogrammetry method. Antarct. Sci. 25, 731–740 (2013).ADS 
    Article 

    Google Scholar 
    18.Paterson, W. et al. Seals like it hot: Changes in surface temperature of harbour seals (Phoca vitulina) from late pregnancy to moult. J. Therm. Biol. 37, 454–461 (2012).Article 

    Google Scholar 
    19.Slip, D. J. & Burton, H. R. Population status and seasonal haulout patterns of the southern elephant seal (Mirounga leonina) at Heard Island. Antarct. Sci. 11, 38–47 (1999).ADS 
    Article 

    Google Scholar 
    20.Kirkman, S. et al. Variation in the timing of moult in southern elephant seals at Marion Island. S. Afr. J. Wildlife Res. 33, 79–84 (2003).
    Google Scholar 
    21.Ling, J. & Bryden, M. Southern elephant seal Mirounga leonina (Linnaeus), 1758. In Handbook of Marine Mammals Vol. 2 (eds Ridgway, S. H. & Harrison, R. J.) 297–327 (Academic Press, 1981).
    Google Scholar 
    22.Constable, A. J. et al. Climate change and southern ocean ecosystems I: how changes in physical habitats directly affect marine biota. Glob. Change Biol. 20, 3004–3025 (2014).ADS 
    Article 

    Google Scholar 
    23.Rogers, A. et al. Antarctic futures: An assessment of climate-driven changes in ecosystem structure, function, and service provisioning in the Southern Ocean. Ann. Rev. Mar. Sci. 12, 87–120 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    24.Hindell, M. A. & Burton, H. R. Seasonal haul-out patterns of the southern elephant seal (Mirounga leonina L.), at Macquarie Island. J. Mammal. 69, 81–88 (1988).Article 

    Google Scholar 
    25.Postma, M., Bester, M. N. & De Bruyn, P. J. N. Age-related reproductive variation in a wild marine mammal population. Polar Biol. 36, 719–729 (2013).Article 

    Google Scholar 
    26.Le Boeuf, B. J. & Laws, R. M. Elephant seals: An introduction to the genus. In Elephant Seals: Population Ecology, Behavior, and Physiology (eds Le Beouf, B. J. & Laws, R. M.) 1–28 (University Of California Press, 1994).Chapter 

    Google Scholar 
    27.De Bruyn, P. J. N. et al. Sex at sea: Alternative mating system in an extremely polygynous mammal. Anim. Behav. 82, 445–451 (2011).Article 

    Google Scholar 
    28.Oosthuizen, W. C., Pradel, R., Bester, M. N. & De Bruyn, P. J. N. Making use of multiple surveys: Estimating breeding probability using a multievent-robust design capture-recapture model. Ecol. Evolut. 9, 836–848 (2019).Article 

    Google Scholar 
    29.Bester, M. N. et al. The marine mammal programme at the Prince Edward Islands: 38 years of research. Afr. J. Mar. Sci. 33, 511–521 (2011).Article 

    Google Scholar 
    30.Pistorius, P. A., De Bruyn, P. J. N. & Bester, M. N. Population dynamics of southern elephant seals: A synthesis of three decades of demographic research at Marion Island. Afr. J. Mar. Sci. 33, 523–534 (2011).Article 

    Google Scholar 
    31.Pistorius, P. A., Bester, M. N. & Kirkman, S. P. Survivorship of a declining population of southern elephant seals, Mirounga leonina, in relation to age, sex and cohort. Oecologia 121, 201–211 (1999).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    32.Oosthuizen, W. C., Bester, M. N., Altwegg, R., McIntyre, T. & De Bruyn, P. J. N. Decomposing the variance in southern elephant seal weaning mass: Partitioning environmental signals and maternal effects. Ecosphere 6, 139 (2015).Article 

    Google Scholar 
    33.Daniel, R. G., Jemison, L. A., Pendleton, G. W. & Crowley, S. M. Molting phenology of harbor seals on Tugidak Island, Alaska. Mar. Mamm. Sci. 19, 128–140 (2003).Article 

    Google Scholar 
    34.Chaise, L. L. et al. Local weather and body condition influence habitat use and movements on land of molting female southern elephant seals (Mirounga leonina). Ecol. Evol. 8, 6081–6090 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Dingemanse, N. J. & Dochtermann, N. A. Quantifying individual variation in behaviour: Mixed-effect modelling approaches. J. Anim. Ecol. 82, 39–54 (2013).PubMed 
    Article 

    Google Scholar 
    36.Burnham, K. P., Anderson, D. R. & Huyvaert, K. P. AIC model selection and multimodel inference in behavioral ecology: Some background, observations, and comparisons. Behav. Ecol. Sociobiol. 65, 23–35 (2011).Article 

    Google Scholar 
    37.Zuur, A., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer Science & Business Media, 2009).MATH 
    Book 

    Google Scholar 
    38.Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar 
    39.R Core Team. R: A language and environment for statistical computing. (R Foundation for Statistical Computing, Vienna, Austria, 2020). https://www.R-project.org/.40.Nakagawa, S. & Schielzeth, H. Repeatability for Gaussian and non-Gaussian data: A practical guide for biologists. Biol. Rev. 85, 935–956 (2010).PubMed 

    Google Scholar 
    41.Nakagawa, S. & Schielzeth, H. A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol. Evol. 4, 133–142 (2013).Article 

    Google Scholar 
    42.Grosbois, V. et al. Assessing the impact of climate variation on survival in vertebrate populations. Biol. Rev. 83, 357–399 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    43.Forcada, J., Trathan, P. N. & Murphy, E. J. Life history buffering in Antarctic mammals and birds against changing patterns of climate and environmental variation. Glob. Change Biol. 14, 2473–2488 (2008).ADS 
    Article 

    Google Scholar 
    44.Bradshaw, C. J., Hindell, M. A., Sumner, M. D. & Michael, K. J. Loyalty pays: Potential life-history consequences of fidelity to marine foraging regions by southern elephant seals. Anim. Behav. 68, 1349–1360 (2004).Article 

    Google Scholar 
    45.Cotté, C., Park, Y.-H., Guinet, C. & Bost, C.-A. Movements of foraging king penguins through marine mesoscale eddies. Proc. R. Soc. B Biol. Sci. 274, 2385–2391 (2007).Article 

    Google Scholar 
    46.Barbraud, C. & Weimerskirch, H. Antarctic birds breed later in response to climate change. Proc. Natl. Acad. Sci. 103, 6248–6251 (2006).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Hindell, M. A. et al. Long-term breeding phenology shift in royal penguins. Ecol. Evol. 2, 1563–1571 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Zimova, M. et al. Function and underlying mechanisms of seasonal colour moulting in mammals and birds: What keeps them changing in a warming world?. Biol. Rev. 93, 1478–1498 (2018).PubMed 
    Article 

    Google Scholar 
    49.Thompson, P. & Rothery, P. Age and sex differences in the timing of moult in the common seal, Phoca vitulina. J. Zool. 212, 597–603 (1987).Article 

    Google Scholar 
    50.Cronin, M., Gregory, S. & Rogan, E. Moulting phenology of the harbour seal in south-west Ireland. J. Mar. Biol. Assoc. UK 94, 1079–1086 (2014).Article 

    Google Scholar 
    51.Kirkman, S. et al. Variation in the timing of the breeding haulout of female southern elephant seals at Marion Island. Aust. J. Zool. 52, 379–388 (2004).Article 

    Google Scholar 
    52.Hindell, M. A., Slip, D. J. & Burton, H. R. Body mass loss of moulting female southern elephant seals, Mirounga leonina, at Macquarie Island. Polar Biol. 14, 275–278 (1994).Article 

    Google Scholar 
    53.Carlini, A., Marquez, M., Daneri, G. & Poljak, S. Mass changes during their annual cycle in females of southern elephant seals at King George Island. Polar Biol. 21, 234–239 (1999).Article 

    Google Scholar 
    54.Worthy, G. A. J., Morris, P. A., Costa, D. P. & Le Boeuf, B. J. Moult energetics of the northern elephant seal (Mirounga angustirostris). J. Zool. 227, 257–265 (1992).Article 

    Google Scholar 
    55.Arnbom, T., Fedak, M. & Rothery, P. Offspring sex ratio in relation to female size in southern elephant seals, Mirounga leonina. Behav. Ecol. Sociobiol. 35, 373–378 (1994).Article 

    Google Scholar 
    56.Proffitt, K. M., Garrott, R. A., Rotella, J. J. & Wheatley, K. E. Environmental and senescent related variations in Weddell seal body mass: Implications for age-specific reproductive performance. Oikos 116, 1683–1690 (2007).Article 

    Google Scholar 
    57.Bérubé, C. H., Festa-Bianchet, M. & Jorgenson, J. T. Individual differences, longevity, and reproductive senescence in bighorn ewes. Ecology 80, 2555–2565 (1999).Article 

    Google Scholar 
    58.Oosthuizen, W. C., Péron, G., Pradel, R., Bester, M. N. & De Bruyn, P. J. N. Positive early-late life-history trait correlations in elephant seals. Ecology 102, e03288. https://doi.org/10.1002/ecy.3288 (2021).59.Kirby, V. L. & Ortiz, C. L. Hormones and fuel regulation in fasting elephant seals. In Elephant Seals: Population Ecology, Behavior, and Physiology (eds Le Beouf, B. J. & Laws, R. M.) 374–386 (University of California Press, 1994).Chapter 

    Google Scholar 
    60.Yochem, P. K. & Stewart, B. S. Hair and fur. In Encyclopedia of Marine Mammals 3rd edn (eds Würsig, B. et al.) 447–448 (Academic Press, 2018).Chapter 

    Google Scholar 
    61.Ashwell-Erickson, S., Fay, F., Elsner, R. & Wartzok, D. Metabolic correlates of molting and regeneration of pelage in Alaskan harbour and spotted seals (Phoca vitulina and Phoca largha). Can. J. Zool. 64, 1086–1092 (1986).CAS 
    Article 

    Google Scholar 
    62.Laws, R. M. & Sinha, A. A. Reproduction. In Antarctic Seals: Research Methods and Techniques (ed. Laws, R. M.) 228–267 (Cambridge University Press, 1993).Chapter 

    Google Scholar 
    63.Smout, S., King, R. & Pomeroy, P. Environment-sensitive mass changes influence breeding frequency in a capital breeding marine top predator. J. Anim. Ecol. 89, 384–396 (2020).PubMed 
    Article 

    Google Scholar 
    64.Clutton-Brock, T. & Sheldon, B. C. Individuals and populations: the role of long-term, individual-based studies of animals in ecology and evolutionary biology. Trends Ecol. Evol. 25, 562–573 (2010).PubMed 
    Article 

    Google Scholar 
    65.Cordes, L. S. & Thompson, P. M. Variation in breeding phenology provides insights into drivers of long-term population change in harbour seals. Proc. R. Soc. B: Biol. Sci. 280, 20130847 (2013).Article 

    Google Scholar 
    66.Barbraud, C. et al. Contrasted demographic responses facing future climate change in Southern Ocean seabirds. J. Anim. Ecol. 80, 89–100 (2011).PubMed 
    Article 

    Google Scholar 
    67.Atkinson, S., DeMaster, D. P. & Calkins, D. G. Anthropogenic causes of the western Steller sea lion Eumetopias jubatus population decline and their threat to recovery. Mamm. Rev. 38, 1–18 (2008).Article 

    Google Scholar 
    68.Desprez, M., McMahon, C. R., Hindell, M. A., Harcourt, R. & Gimenez, O. Known unknowns in an imperfect world: incorporating uncertainty in recruitment estimates using multi-event capture-recapture models. Ecol. Evol. 3, 4658–4668 (2013).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    69.Griffen, B. D. Reproductive skipping as an optimal life history strategy in the southern elephant seal, Mirounga leonina. Ecol. Evol. 8, 9158–9170 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    70.Rohwer, S., Viggiano, A. & Marzluff, J. M. Reciprocal tradeoffs between moult and breeding in albatrosses. Condor 113, 61–70 (2011).Article 

    Google Scholar 
    71.Grissot, A., Graham, I. M., Quinn, L., Bråthen, V. S. & Thompson, P. M. Breeding status influences timing but not duration of moult in the northern fulmar Fulmarus glacialis. Ibis 162, 446–459 (2019).Article 

    Google Scholar 
    72.Condit, R., Beltran, R.S., Robinson, P.W., Crocker, D.E. & Costa, D.P. Birth timing after the long feeding migration in elephant seals. https://doi.org/10.1101/2020.10.02.324210 (in review). More

  • in

    Genetic diversity, population structure and historical demography of the two-spined yellowtail stargazer (Uranoscopus cognatus)

    1.Cámara, A. & Santero-Sánchez, R. Economic, social, and environmental impact of a sustainable fishereis model in Spain. Sustainability 11, 6311 (2019).Article 

    Google Scholar 
    2.Department of Fisheries Malaysia. Annual Fisheries statistics 2010–2019. https://www.dof.gov.my/index.php/pages/view/82 (2020).3.Department of Fisheries, Thailand. The annual marine fisheries statistics (2010–2019) based on the sample survey. https://elibonline.fisheries.go.th/elib/cgi-bin/opacexe.exe?op=dsp&bid=10498&lang=0&db=Main&pat=&cat=sub&skin=s&lpp=20&catop=edit&scid=zzz (2020).4.Jha, S., Deepti, V., Ravali, V. & Sujatha, K. Studies on some aspects of biology of Uranoscopus cognatus Cantor, 1849 (Pisces: Uranoscopidae) off Visakhapatnam, central eastern coast of India. Indian J. Mar. Sci. 48, 85–92 (2019).
    Google Scholar 
    5.Clark, M. R. et al. The impacts of deep-sea fisheries on benthic communities: A review. ICES J. Mar. Sci. 73(1), 51–69 (2016).Article 

    Google Scholar 
    6.Van Denderen, P. D. et al. Evaluating impacts of bottom trawling and hypoxia on benthic communities at the local, habitat, and regional scale using a modelling approach. ICES J. Mar. Sci. 77(1), 578–589 (2019).
    Google Scholar 
    7.Erdoğan Sağlam, N. & Sağlam, C. Population parameters of stargazer (Uranoscopus scaber Linnaeus, 1758) in the southeastern Black Sea region during the 2011–2012 fishing season. J. Appl. Ichthyol. 29, 1313–1317 (2013).Article 

    Google Scholar 
    8.Matsunuma, M. et al. Fishes of Terengganu: East Coast of Malay Peninsula, Malaysia (National Museum of Nature and Science, 2011).
    Google Scholar 
    9.Vilasri, V. Family Uranoscopidae. In Fishes of Southern Taiwan (eds Koeda, K. & Ho, H. S.) 1097–1105 (National Museum of Marine Biology & Aquarium, 2019).
    Google Scholar 
    10.Starks, E. C. The Osteology and Relationships of the Uranoscopoid Fishes (Stanford University Press, 1923).
    Google Scholar 
    11.Pietsch, T. W. Phylogenetic relationships of trachinoid fishes of the family Uranoscopidae. Copeia 1989, 253–303 (1989).Article 

    Google Scholar 
    12.Kishimoto, H. Uranoscopidae. In FAO Species Identification Guide for Fisheries Purposes (eds Carpenter, K. E. & Niem, V. H.) 3519–3531 (FAO, 2001).
    Google Scholar 
    13.Randall, J. E. & Arnold, R. J. Uranoscopus rosette, a new species of stargazer (Uranoscopidae: Trachinoidei) from the Red Sea. Aqua. Int. J. Ichthyol. 18, 209–219 (2012).
    Google Scholar 
    14.Jung-chen, H. & Hin-Kiu, M. Stargazers (Uranoscopidae) have exceptionally more bile. Kuroshio Sci. 9–1, 17–26 (2015).
    Google Scholar 
    15.Vilasri, V. Comparative anatomy and phylogenetic systematics of the family Uranoscopidae (Actinopterygii: Perciformes). Mem. Fac. Fish. Hokkaido Univ. 55, 1–106 (2013).
    Google Scholar 
    16.Fricke, R., Eschmeyer, W. N. & Van der Laan, R. (eds). Eschmeyer’s catalog of fishes: genera, species, references. http://researcharchive.calacademy.org/research/ichthyology/catalog/fishcatmain.asp (2020).17.Froese, R. & Pauly, D. Uranoscopidae. Fishbase https://www.fishbase.se/Summary/FamilySummary.php?ID=378 (2019).18.Fricke, R. Two new species of stargazers of the genus Uranoscopus (Teleostei: Uranoscopidae) from the western Pacific Ocean. Zootaxa 4476, 157–167 (2018).PubMed 
    Article 

    Google Scholar 
    19.Fricke, R., Jawad, L. A., Al-Kharusi, L. H. & Al-Mamry, J. M. New record and redescription of Uranoscopus crassiceps Alcock, 1890 (Uranoscopidae) From Oman, Arabian Sea, Northwestern Indian Ocean, based on adult specimens. Cybium 37, 143–147 (2013).
    Google Scholar 
    20.Department of Fisheries Malaysia. Valid Local Name of Malaysian Marine Fishes (Department of Fisheries Malaysia, 2009).
    Google Scholar 
    21.Spalding, M. D. et al. Marine ecoregions of the world: A bioregionalization of coastal and shelf areas. Bioscience 57, 573–583 (2007).Article 

    Google Scholar 
    22.Voris, H. K. Maps of Pleistocene sea levels in Southeast Asia: Shorelines, river systems and time durations. J. Biogeogr. 27, 1153–1167 (2000).Article 

    Google Scholar 
    23.Rohfritsch, A. & Borsa, P. Genetic structure of Indian scad mackerel Decapterus russelli: Pleistocene vicariance and secondary contact in the Central Indo-West Pacific Seas. Heredity 95, 315 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    24.Lohman, D. J. et al. Biogeography of the Indo-Australian archipelago. Annu. Rev. Ecol. Evol. Syst. 42, 205–226 (2011).Article 

    Google Scholar 
    25.Crandall, E. D. et al. The molecular biogeography of the Indo-Pacific: Testing hypotheses with multispecies genetic patterns. Glob. Ecol. Biogeogr. 28, 943–960 (2019).Article 

    Google Scholar 
    26.Reece, J. S., Bowen, B. W., Joshi, K., Goz, V. & Larson, A. Phylogeography of two moray eels indicates high dispersal throughout the Indo-Pacific. J. Hered. 101, 391–402 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    27.Akib, N. A. M. et al. High connectivity in Rastrelliger kanagurta: influence of historical signatures and migratory behaviour inferred from mtDNA cytochrome b. PLoS ONE 10, e0119749 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    28.Jamaludin, N. A. et al. Phylogeography of the Japanese scad, Decapterus maruadsi (Teleostei; Carangidae) across the Central Indo-West Pacific: evidence of strong regional structure and cryptic diversity. Mitochondrial DNA A 2, 1–13 (2020).
    Google Scholar 
    29.Gaither, M. R., Toonen, R. J., Robertson, D. R., Planes, S. & Bowen, B. W. Genetic evaluation of marine biogeographical barriers: perspectives from two widespread Indo-Pacific snappers (Lutjanus kasmira and Lutjanus fulvus). J. Biogeogr. 37, 133–147 (2010).Article 

    Google Scholar 
    30.Gaither, M. R. et al. Phylogeography of the reef fish Cephalopholis argus (Epinephelidae) indicates Pleistocene isolation across the Indo-Pacific Barrier with contemporary overlap in the Coral Triangle. BMC Evol. Biol. 11, 189 (2011).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    31.Timm, J. & Kochzius, M. Geological history and oceanography of the Indo-Malay Archipelago shape the genetic population structure in the false clown anemonefish (Amphiprion ocellaris). Mol. Ecol. 17, 3999–4014 (2008).PubMed 
    Article 

    Google Scholar 
    32.Otwoma, L. M. & Kochzius, M. Genetic population structure of the coral reef sea star Linckia laevigata in the Western Indian Ocean and Indo-West Pacific. PLoS ONE 11, 10 (2016).Article 
    CAS 

    Google Scholar 
    33.Williams, S. T., Jara, J., Gomez, E. & Knowlton, N. The marine Indo-West Pacific break: Contrasting the resolving power of mitochondrial and nuclear genes. Integr. Comp. Biol. 42, 941–952 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    34.Supmee, V., Sangthong, P., Songrak, A. & Suppapan, J. Population genetic structure of Asiatic Hard Clam (Meretrix meretrix) in Thailand based on Cytochrome Oxidase subunit I gene sequence. Biodiversitas 21, 2702–2709 (2020).Article 

    Google Scholar 
    35.Hui, M. et al. Comparative genetic population structure of three endangered giant clams (Cardiidae: Tridacna species) throughout the Indo-West Pacific: Implications for divergence, connectivity and conservation. J. Molluscan Stud. 82, 403–414 (2016).Article 

    Google Scholar 
    36.Panithanarak, T., Karuwancharoen, R., Na-Nakorn, U. & Nguyen, T. T. Population genetics of the spotted seahorse (Hippocampus kuda) in Thai waters: Implications for conservation. Zool. Stud. 49, 564–576 (2010).CAS 

    Google Scholar 
    37.Kasim, N. S. et al. Recent population expansion of longtail tuna Thunnus tonggol (Bleeker, 1851) inferred from the mitochondrial DNA markers. PeerJ 8, 9679 (2020).Article 

    Google Scholar 
    38.Canales-Aguirre, C. B., Ferrada-Fuentes, S., Galleguillos, R., Oyarzun, F. X. & Hernández, C. E. Population genetic structure of Patagonian toothfish (Dissostichus eleginoides) in the Southeast Pacific and Southwest Atlantic Ocean. PeerJ 6, e4173 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    39.Sato, M. et al. Genetic structure and demographic connectivity of marbled flounder (Pseudopleuronectes yokohamae) populations of Tokyo Bay. J. Sea Res. 142, 79–90 (2018).ADS 
    Article 

    Google Scholar 
    40.Borsa, P. Genetic structure of round scad mackerel Decapterus macrosoma (Carangidae) in the Indo-Malay archipelago. Mar. Biol. 142, 575–581 (2003).CAS 
    Article 

    Google Scholar 
    41.Eytan, R. I. & Hellberg, M. E. Nuclear and mitochondrial sequence data reveal and conceal different demographic histories and population genetic processes in Caribbean reef fishes. Evolution 64, 3380–3397 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    42.Tan, M. P., Jamsari, A. F. J. & Siti Azizah, M. N. Genotyping of microsatellite markers to study genetic structure of the wild striped snakehead Channa striata in Malaysia. J. Fish. Biol. 88, 1932–1948 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    43.Tan, M. P., Jamsari, A. F. J. & Siti Azizah, M. N. Phylogeographic pattern of the striped snakehead, Channa striata in Sundaland: Ancient river connectivity, geographical and anthropogenic signatures. PLoS ONE 7, 1–11 (2012).
    Google Scholar 
    44.Tan, M. P., Jamsari, A. F. J., Muhlisin, Z. A. & Siti Azizah, M. N. Mitochondrial genetic variation and population structure of the striped snakehead, Channa striata in Malaysia and Sumatra. Indonesia. Biochem. Syst. Ecol. 60, 99–105 (2015).CAS 
    Article 

    Google Scholar 
    45.Haponski, A. E. & Stepien, C. A. Phylogenetic and biogeographical relationships of the Sander pikeperches (Percidae: Perciformes): Patterns across North America and Eurasia. Biol. J. Linn. Soc. Lond. 110, 156–179 (2013).Article 

    Google Scholar 
    46.Milá, B., Van Tassell, J. L., Calderón, J. A., Rüber, L. & Zardoya, R. Cryptic lineage divergence in marine environments: Genetic differentiation at multiple spatial and temporal scales in the widespread intertidal goby Gobiosoma bosc. Ecol. Evol. 7, 5514–5523 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Piganeau, G., Gardner, M. & Eyre-Walker, A. A broad survey of recombination in animal mitochondria. Mol. Biol. Evol. 21, 2319–2325 (2004).CAS 
    PubMed 
    Article 

    Google Scholar 
    48.Avise, J. C. Molecular Markers, Natural History, and Evolution (Sinauer Associates Inc, 2004).
    Google Scholar 
    49.De Mandal, S., Chhakchhuak, L., Gurusubramanian, G. & Kumar, N. S. Mitochondrial markers for identification and phylogenetic studies in insects–A Review. DNA Barcodes 2, 1–9 (2014).CAS 
    Article 

    Google Scholar 
    50.Simon, C. et al. Evolution, weighting, and phylogenetic utility of mitochondrial gene sequences and a compilation of conserved polymerase chain reaction primers. Ann. Entomol. Soc. Am. 87, 651–701 (1994).CAS 
    Article 

    Google Scholar 
    51.Hoofer, S. R., Reeder, S. A., Hansen, E. W. & Van Den Bussche, R. A. Molecular phylogenetics and taxonomic review of noctilionoid and vespertilionoid bats (Chiroptera: Yangochiroptera). J. Mammal. 84, 809–821 (2003).Article 

    Google Scholar 
    52.Hewitt, G. M. Speciation, hybrid zones and phylogeography or seeing genes in space and time. Mol. Ecol. 10, 537–549 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    53.Surya, S. et al. Morphometry and length-weight relationship of Uranoscopus marmoratus Cuvier, 1829 (Family: Uranoscopidae) from Palk Bay, India. Int. J. Biol. Sci. 5, 1–10 (2016).ADS 

    Google Scholar 
    54.Narejo, N. T. Morphometric characters and their relationships in Gudusia chapra (Hamilton) from Keenjhar Lake (Distt: Thatta), Sindh, Pakistan. Pak. J. Zool. 42, 101–104 (2010).
    Google Scholar 
    55.Gan, H. M., Nur Ilham Syahadah, M. Y., Vilasri, V., Tun Nurul Aimi, M. J. & Tan, M. P. Four whole mitogenome sequences of yellowtail stargazers (Uranoscopus cognatus cantor 1849) from East Peninsular Malaysia and West Coast of Thailand. Mitochondrial DNA B 4, 256–258 (2019).Article 

    Google Scholar 
    56.Panjarat, S. Sustainable fisheries in the Andaman Sea coast of Thailand. Division for Ocean Affairs and the Law of the Sea Office of Legal Affairs. (The United Nations, 2008).57.Derrick, B., Noranarttragoon, P., Zeller, D., Teh, L. C. & Pauly, D. Thailand’s missing marine fisheries catch (1950–2014). Front. Mar. Sci. 4, 402 (2017).Article 

    Google Scholar 
    58.Sampantamit, T., Ho, L., Van Echelpoel, W., Lachat, C. & Goethals, P. Links and trade-offs between fisheries and environmental protection in relation to the sustainable development goals in Thailand. Water 12, 399 (2020).Article 

    Google Scholar 
    59.Chong, V., Lee, P. & Lau, C. Diversity, extinction risk and conservation of Malaysian fishes. J. Fish Biol. 76, 2009–2066 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    60.Lim, H. C., Ahmad, A. T., Nuruddin, A. A. & Mohd Nor, S. A. Cytochrome b gene reveals panmixia among Japanese Threadfin Bream, Nemipterus japonicus (Bloch, 1791) populations along the coasts of Peninsular Malaysia and provides evidence of a cryptic species. Mitochondrial DNA A 27, 575–584 (2016).CAS 
    Article 

    Google Scholar 
    61.Nabilsyafiq, M. H. et al. ND5 gene marker reveals recent population expansion of wild pearse’s mudskipper (Periophthalmus novemradiatus Hamilton) inhabits Setiu wetlands in east Peninsular Malaysia. Malays. Appl. Biol. 48, 87–93 (2019).
    Google Scholar 
    62.Zhou, Y. et al. Importance of incomplete lineage sorting and introgression in the origin of shared genetic variation between two closely related pines with overlapping distributions. Heredity 118, 211–220 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    63.Lessios, H. A. The great American schism: divergence of marine organisms after the rise of the Central American Isthmus. Annu. Rev. Ecol. Evol. Syst. 39, 63–91 (2008).Article 

    Google Scholar 
    64.Avise, J. Molecular Markers, Natural History and Evolution (Chapman and Hall, 1994).Book 

    Google Scholar 
    65.Nelson, J. S., Grande, T. C. & Wilson, M. V. Fishes of the World (John Wiley and Sons, 2016).Book 

    Google Scholar 
    66.Young, J. Z. Memoirs: On the autonomic nervous system of the Teleostean Fish Uranoscopus scaber. J. Cell Sci. 2, 491–536 (1931).Article 

    Google Scholar 
    67.Day, J., Clark, J. A., Williamson, J. E., Brown, C. & Gillings, M. Population genetic analyses reveal female reproductive philopatry in the oviparous Port Jackson shark. Mar. Freshw. Res. 70, 986–994 (2019).Article 

    Google Scholar 
    68.Roycroft, E. J., Le Port, A. & Lavery, S. D. Population structure and male-biased dispersal in the short-tail stingray Bathytoshia brevicaudata (Myliobatoidei: Dasyatidae). Conserv. Genet. 20, 717–728 (2019).CAS 
    Article 

    Google Scholar 
    69.King, T. L., Eackles, M. S., Spidle, A. P. & Brockmann, H. J. Regional differentiation and sex-biased dispersal among populations of the horseshoe crab Limulus polyphemus. Trans. Am. Fish. Soc. 134, 441–465 (2005).Article 

    Google Scholar 
    70.Lane, A. & Shine, R. Intraspecific variation in the direction and degree of sex-biased dispersal among sea-snake populations. Mol. Ecol. 20, 1870–1876 (2011).PubMed 
    Article 

    Google Scholar 
    71.Casale, P., Laurent, L., Gerosa, G. & Argano, R. Molecular evidence of male-biased dispersal in loggerhead turtle juveniles. J. Exp. Mar. Biol. Ecol. 267, 139–145 (2002).CAS 
    Article 

    Google Scholar 
    72.Wyrtki, K. Physical Oceanography of the Southeast Asian Waters (University of California, 1961).
    Google Scholar 
    73.Barber, P., Palumbi, S., Erdmann, M. & Moosa, M. Sharp genetic breaks among populations of Haptosquilla pulchella (Stomatopoda) indicate limits to larval transport: patterns, causes, and consequences. Mol. Ecol. 11, 659–674 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    74.Kamarudin, K. R. & Esa, Y. Phylogeny and phylogeography of Barbonymus schwanenfeldii (Cyprinidae) from Malaysia inferred using partial cytochrome b mtDNA gene. J. Trop. Biol. Conserv. 5, 1–13 (2009).
    Google Scholar 
    75.Tan, M. P. et al. Genetic diversity of the Pearse’s Mudskipper Periophthalmus novemradiatus (Perciformes: Gobiidae) and characterization of its complete mitochondrial genome. Thalassas 36, 103–113 (2020).Article 

    Google Scholar 
    76.Roberts, T. R. & Khaironizam, M. Z. Trophic polymorphism in the Malaysian fish Neolissochilus soroides and other old world barbs (Teleostei, Cyprinidae). Nat. Hist. Bull. Siam Soc. 56, 25–53 (2008).
    Google Scholar 
    77.Forsman, A. Rethinking phenotypic plasticity and its consequences for individuals, populations and species. Heredity 115, 276 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    78.Vasileva, E. Morphokaryological variability and divergence of stargazers (Uranoscopus, perciformes) from the Mediterranean Sea basin: I. Divergence and taxonomic state of the Black Sea Stargazer. J. Ichthyol. 52, 476–484 (2012).Article 

    Google Scholar 
    79.Aljanabi, S. M. & Martinez, I. Universal and rapid salt-extraction of high quality genomic DNA for PCR-based techniques. Nucleic Acids Ress 25, 4692–4693 (1997).CAS 
    Article 

    Google Scholar 
    80.Ward, R. D., Zemlak, T. S., Innes, B. H., Last, P. R. & Hebert, P. D. DNA barcoding Australia’s fish species. Philos. Trans. R. Soc. Lond., B Biol. Sci. 360, 1847–1857 (2005).CAS 
    Article 

    Google Scholar 
    81.López, J. A., Chen, W. J. & Ortí, G. Esociform phylogeny. Copeia 2004, 449–464 (2004).Article 

    Google Scholar 
    82.Mat Jaafar, T. N., Taylor, M. I., Mohd Nor, S. A., Bruyn, M. D. & Carvalho, G. R. Comparative genetic stock structure in three species of commercially exploited Indo-Malay Carangidae (Teleosteii, Perciformes). J. Fish Biol. 96, 337–349 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    83.Tamura, K., Stecher, G., Peterson, D., Filipski, A. & Kumar, S. MEGA6: molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 30, 2725–2729 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    84.Farris, J. S., Källersjö, M., Kluge, A. G. & Bult, C. Testing significance of incongruence. Cladistics 10, 315–319 (1994).Article 

    Google Scholar 
    85.Swofford, D. L. PAUP: Phylogenetic Analysis Using Parsimony (and Other Methods), Version 4.0 Beta 10 (Sinauer Associates, 2002).
    Google Scholar 
    86.Librado, P. & Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 25, 1451–1452 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    87.Excoffier, L. & Lischer, H. E. L. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour. 10, 564–567 (2010).PubMed 
    Article 

    Google Scholar 
    88.Hasegawa, M., Kishino, H. & Yano, T. A. Dating of the huma-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160–174 (1985).CAS 
    PubMed 
    Article 

    Google Scholar 
    89.Kimura, M. A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucletide sequences. J. Mol. Evol. 16, 111–120 (1980).ADS 
    CAS 
    Article 

    Google Scholar 
    90.Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 39, 783–791 (1985).PubMed 
    Article 

    Google Scholar 
    91.Bandelt, H. J., Forster, P. & Röhl, A. Median-joining networks for inferring intraspecific phylogenies. Mol. Biol. Evol. 16, 37–48 (1999).CAS 
    PubMed 
    Article 

    Google Scholar 
    92.Dupanloup, I., Schneider, S. & Excoffier, L. A simulated annealing approach to define the genetic structure of populations. Mol. Ecol. 11, 2571–2581 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    93.Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Series B Stat. Methodol. 57, 289–300 (1995).MathSciNet 
    MATH 

    Google Scholar 
    94.Nei, M. Analysis of gene diversity in subdivided populations. Proc. Natl. Acad. Sci. U.S.A. 70, 3321–3323 (1973).ADS 
    CAS 
    PubMed 
    PubMed Central 
    MATH 
    Article 

    Google Scholar 
    95.Lynch, M. & Crease, T. The analysis of population survey data on DNA sequence variation. Mol. Biol. Evol. 7, 377–394 (1990).CAS 
    PubMed 

    Google Scholar 
    96.Hudson, R. R., Slatkin, M. & Maddison, W. P. Estimation of levels of gene flow from DNA sequence data. Genetics 132, 583–589 (1992).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    97.Tajima, F. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123, 585–595 (1989).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    98.Fu, Y.-X. Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. Genetics 147, 915–925 (1997).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    99.Harpending, H. Infertility and forager demography. Am. J. Phys. Anthropol. 93, 385–390 (1994).CAS 
    PubMed 
    Article 

    Google Scholar 
    100.Slatkin, M. & Hudson, R. R. Pairwise comparisons of mitochondrial DNA sequences in stable and exponentially growing populations. Genetics 129, 555–562 (1991).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    101.Rogers, A. R. & Harpending, H. Population growth makes waves in the distribution of pairwise genetic differences. Mol. Biol. Evol. 9, 552–569 (1992).CAS 
    PubMed 

    Google Scholar 
    102.Schneider, S. & Excoffier, L. Estimation of past demographic parameters from the distribution of pairwise differences when the mutation rates vary among sites: Application to human mitochondrial DNA. Genetics 152, 1079–1089 (1999).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    103.Yildirim, Y. Genetic structure of Pleurobranchaea maculata in New Zealand (Massey University, 2016).
    Google Scholar 
    104.Hubbs, C. & Lagler, K. The fishes of the Great Lakes region 213 (The University of Michigan Press, 1958).
    Google Scholar 
    105.Kishimoto, H. A new stargazer, Uranoscopus flavipinnis, from Japan and Taiwan with redescription and neotype designation of U. japonicus. Japan. J. Ichthyol. 34, 1–14 (1987).
    Google Scholar 
    106.Kishimoto, H. Redescription and lectotype designation of the stargazer Uranoscopus kaianus Günther. Copeia 1984, 1009–1011 (1984).Article 

    Google Scholar 
    107.Gomon, M. F. & Johnson, J. A new fringed stargazer (Uranoscopidae: Ichthyscopus) with descriptions of the other Australian species of the genus. Mem. Queensl. Mus. 43, 597–619 (1999).
    Google Scholar 
    108.Rainboth, W. J. Fishes of the Cambodian Mekong (Food and Agriculture Org, 1996).
    Google Scholar 
    109.Imamura, H. & Matsuura, K. Redefinition and phylogenetic relationships of the family Pinguipedidae (Teleostei: Perciformes). Ichthyol. Res. 50, 259–269 (2003).Article 

    Google Scholar 
    110.Hammer, Ø., Harper, D. A. & Ryan, P. D. PAST: Paleontological statistics software package for education and data analysis. Palaeontol. Electron. 4, 9 (2001).
    Google Scholar 
    111.Seah, Y. G., Nabilsyafiq, M. & Mazlan, A. G. Preliminary study on the morphology and biology of coexist Nemipterus furcosus and Nemipterus tambuloides from Terengganu Waters Peninsular Malaysia. J. Fish. Aquat. Sci. 11, 418–424 (2016).Article 

    Google Scholar 
    112.Johnson, R. A. & Wichern, D. W. Multivariate statistical analysis (Prentice Hall Upper Saddle River, 1998).MATH 

    Google Scholar  More

  • in

    Functional differences between TSHR alleles associate with variation in spawning season in Atlantic herring

    AnimalsTissue samples for expression analysis were collected on August 24 and September 9 2016 from a population of spring-spawning herring kept in captivity at University of Bergen; the rearing of captive herring was approved by the Norwegian national animal ethics committee (Forsøksdyrutvalget FOTS ID-5072). The tissue samples used for ATAC-sequencing were collected at Hästskär on June 19 2019 from wild spring-spawning Atlantic herring in the Baltic Sea, which do not require ethical permission.Genome scan and genetic diversity at the TSHR locusWe used a 2 × 2 contingency X2 test to estimate the extent of SNP allele frequency differences between seven spring- and seven autumn-spawning herring populations from the Northeast Atlantic (Supplementary Table 2), and thus, identify the major genomic regions associated with seasonal reproduction. The SNP allele frequencies were generated in a previous study16 and were derived from Pool-seq data. For the X2 test, we formed two groups, spring and autumn spawners, and summed the reads supporting the reference and the alternative alleles for the pools included in each group.To characterize genetic diversity at the TSHR locus, we calculated nucleotide diversity (π) and Tajima’s D for the same seven spring- and seven autumn-spawning Atlantic herring populations used in the genome scan (n = ~41–100 per pool) (Supplementary Table 2). The whole-genome re-sequence data of these pools were previously reported by Han et al.16 (for details of the pools used here see Supplementary Table 2). Unbiased nucleotide diversity π and Tajima’s D were calculated for each pool using the program PoPoolation 1.2.236, which accounts for the truncated allele frequency spectrum of pooled data. In brief, a pileup file of chromosome 15 was generated from each pool BAM file using samtools v.1.1037,38. Indels and 5 bp around indels were removed to exclude spurious SNPs due to misalignments around indels. To minimize biases in the π and Tajima’s D calculations, which are sensitive to sequencing errors and coverage fluctuations39, the coverage of each pileup file was subsampled without replacement to a uniform value based on a per-pool coverage distribution (the target coverage corresponded to the 5th percentile of the coverage frequency distribution, which was used as the minimum coverage allowed for an SNP to be included in the analysis) (Supplementary Table 2). We also calculated the diversity parameters but skipping the coverage subsampling step and obtained very similar results with both approaches (Supplementary Fig. 5), thus, we decided to keep working with the subsampled datasets as coverage subsampling is recommended by the software developers36. To exclude spurious SNPs associated with repetitive sequences and copy number variants, we applied a maximum coverage filter equivalent to the per-pool 99th percentile of the coverage frequency distribution (Supplementary Table 2). A minimum base quality of 20, a minimum mapping quality of 20, and a minor allele count of 2 were required to retain high quality SNPs for further analysis. Both, π and Tajima’s D statistics were calculated using a sliding window approach with a window size of 10 kb and a step size of 2 kb (the selected window-step combination offered a good genomic resolution while reducing the noise from single SNPs, after testing windows of 5, 10, 20, 40, 50, and 100 kb for non-overlapping and overlapping windows with a step size equivalent to 20% of the window size). Only windows with a coverage fraction ≥ 0.5 were included in the computations. In addition, we estimated the effective allele frequency difference, or delta allele frequency (dAF), between spring and autumn spawning groups at the TSHR locus using the formula dAF = abs(mean(spring pools)−mean(autumn pools)). For each of the diversity parameters, we assessed whether the mean differences between sets of SNPs within chr 15: 8.85–8.95 Mb (215 SNPs) and outside (214 635 SNPs) the TSHR region were statistically significant among spring- and autumn-spawning groups using a Wilcoxon test. Data postprocessing, statistical tests, and plotting were performed in the R environment40 (for specific parameters used in PoPoolation, see the associated code to this publication).Identification of the 5.2 kb structural variantSequences spanning the entire TSHR gene plus 10 kb upstream and downstream from two reference assemblies, ASM96633v115 and Ch_v2.0.218, were aligned using BLAST41 and the output was subsequently processed with a custom R script40. Repeats were annotated by CENSOR21 for the region harboring the 5.2 kb structural variant. To validate the structural variant, long-range PCR was performed with genomic DNA from two spring- and autumn-spawning Atlantic herring in a 20 μL reaction containing 0.8 mM dNTPs, 0.3 μM each of the forward and reverse 5.2kb-confirm primers (Supplementary Table 1) and 0.75 U PrimeSTAR GXL DNA Polymerase (TaKaRa) following the program: 95 °C for 3 min, 35 cycles of 98 °C for 10 s, 58 °C for 20 s and 68 °C for 2 min 30 s, and a final extension of 10 min at 68 °C.ATAC-seq analysisBSH and brain without BSH were dissected from two spring-spawning herring caught in the Baltic Sea and transported to the lab on dry ice, then kept at –80 °C before nuclei isolation. ATAC-seq libraries were constructed according to the Omni-ATAC protocol with minor modifications42. Briefly, tissue was homogenized in 2 ml homogenization buffer (5 mM CaCl2, 3 mM Mg(Ac)2, 10 mM Tris-HCl (pH = 7.8), 0.017 mM PMSF, 0.17 mM ß-mercaptoethanol, 320 mM Sucrose, 0.1 mM EDTA and 0.1% NP-40) with a Dounce homogenizer on ice. 400 μl suspension was transferred to a 2 ml tube for the density gradient centrifugation with different concentrations of Iodixanol solution. After centrifugation, a 200 μl fraction containing the nuclei band was collected, stained with Trypan blue and counted with a Countess II Automated Cell Counter (Thermo Fisher Scientific). An aliquot of 100,000–200,000 nuclei was used as input in a 50 μl transposition reaction containing 2 X TD buffer and 100 nM assembled Tn5 transposase for a 30-min incubation at 37 °C. Tagmented DNA was purified with a Zymo clean kit (Zymo Research). Purified DNA was used for an initial pre-amplification for 5 cycles, and the additional amplification cycle was determined by qPCR based on the “R vs Cycle Number” plot43. Amplified libraries were purified with a Zymo clean kit again, and library concentrations and qualities were evaluated using the 2200 TapeStation System (Agilent Technologies).ATAC-seq was performed with a MiniSeq High Output Kit (150 cycles) on a MiniSeq instrument (Illumina) and 7–9 million reads were generated for each ATAC-seq library. Quality control, trimming, mapping, and peak calling of the sequenced reads were conducted following the ENCODE ATAC-seq pipeline (https://www.encodeproject.org/atac-seq/). The trimmed reads were aligned to the Atlantic herring reference genome (Ch_v2.0.2)18 with Bowtie244 and the mapping rate was 85–95%. Duplicate reads, reads with low mapping quality and those aligned to the mitochondria genome were removed. The remaining reads (4–5 million) were subjected to peak calling by MACS245, where 22–32 K peaks were called. Sequenced library qualities were further evaluated by calculating the TSS enrichment score and checking the library complexity with the Non-Redundant Fraction (NRF) and PCR Bottlenecking Coefficients (PBC1 and 2). Finally, conserved peaks between two biological replicates were identified by evaluating the irreproducible discovery rate (IDR).Genotyping of six differentiated variants and haplotype analysisAll six genetic variants, including the 5.2 kb structural variant, two non-coding SNPs, two missense SNPs and the copy number variant of C-terminal 22aa repeat, were genotyped in 45 spring-, 67 autumn-spawning Atlantic herring and 13 Pacific herring. TaqMan Custom SNP assays were performed to genotype the four SNPs in 5 μl reactions with a template of 20 ng genomic DNA (ThermoFisher Scientific). Copy number of the C-terminal 22aa repeat was determined by the PCR product size generated with geno22aa primers. Genotyping of the 5.2 kb structural variant was performed in a PCR reaction containing two forward primers (geno5.2kb-1F and geno5.2kb-2F) and one reverse primer (geno5.2kb-R), which generated PCR products with different sizes between spring and autumn spawners. All the primers used for genotyping are listed in Supplementary Table 1.Tissue expression profiles by quantitative PCRTotal RNA was prepared from gonad, heart, spleen, kidney, gills, intestine, hypothalamus and saccus vasculosus (BSH), and brain without BSH (brain) of six adult spring-spawning Atlantic herring using RNeasy Mini Kit (Qiagen). RNA was then reverse transcribed into cDNA with a High-Capacity cDNA Reverse Transcription Kit (ThermoFisher Scientific). TaqMan Gene Expression assay (ThermoFisher Scientific) containing 0.3 μM primers and 0.25 μM TaqMan probe (Integrated DNA Technologies) was performed to compare the relative expression levels of TSHR among different tissues. qPCR with SYBR Green chemistry was used for TSHB and DIO2 in a 10 μl reaction of SYBR Green PCR Master Mix (ThermoFisher Scientific) and 0.3 μM primers, with a program composed of an initial denaturation for 10 min at 95 °C followed by 40 cycles of 95 °C for 15 s and 60 °C for 1 min. Ct values were first normalized to the housekeeping gene ACTIN, then the average expression for each gene in the gonad was assumed to be 1 for the subsequent calculation of the relative expression in other tissues.Plasmid constructsThe coding sequence for the herring single-chain TSH (scTSH) was designed following a strategy previously used for mammalian gonadotropins46 that contained an in-frame fusion of the cDNA sequences (5′–3′) of herring TSH beta subunit (NCBI: XM_012836756.1) and alpha subunit (NCBI: XM_012822755.1) linked by six histidines and then the C-terminal peptide of the hCG beta subunit. The designed sequence should generate a protein with a size of 30.6 kDa. Both scTSH and spring herring TSHR cDNA sequences were synthesized in vitro and cloned in the expression vector pcDNA3.1 by Genscript (Leiden, Netherlands). pcDNA3.1 plasmid expressing human TSHR was kindly provided by Drs. Gilbert Vassart and Sabine Costagliola (Université libre de Bruxelles, Belgium). Then, the spring herring TSHR and human TSHR plasmids were used as templates for site-directed mutagenesis to generate constructs coding for different mutant herring or human TSHRs. Plasmids for the dual-luciferase assay, including pGL4.29[luc2P/CRE/Hygro] containing cAMP response elements (CREs) to drive the transcription of luciferase gene luc2P and pRL-TK monitoring the transfection efficiency, were purchased from Promega. Five ng of each plasmid was used to transform the XL1-Blue competent cells (Agilent), plasmid DNA was subsequently extracted from 200 ml overnight culture of a single transformant clone using an EndoFree plasmid Maxi Kit (Qiagen).Cell cultureChinese hamster ovary (CHO) (ATCC CCL-61) and human embryonic kidney 293 (HEK293) (ATCC CRL-1573) cells were maintained in DMEM supplemented with 5% (CHO) or 10% FBS (HEK293), 100 U/ml penicillin, 100 μg/ml streptomycin and 292 μg/ml l-Glutamine (ThermoFisher Scientific) at 37 °C with 5% CO2. Epithelioma Papulosum Cyprini (EPC) cells (ATCC CRL-2872) were cultured in EMEM (Sigma) supplemented with 10% FBS, 100 U/ml penicillin, 100 μg/ml streptomycin, 292 μg/ml l-Glutamine and 1 mM Sodium Pyruvate (ThermoFisher Scientific) at 26 °C with 5% CO2.Production of recombinant herring scTSHCHO cells were transfected with the scTSH expression plasmid using Lipofectamine 3000 (Invitrogen), stable clones were subsequently selected with 500 μg/ml G418 (Invitrogen) and screened for producing scTSH by western blot using a polyclonal antisera against the sea bass alpha subunit47. A positive clone was expanded in 225 cm2 cell culture flasks (Corning) in culture medium containing 5% FBS until confluence, then the cells were maintained in serum-free DMEM for hormone production for 7 days at 25 °C48. After 7 days, culture medium containing scTSH or without (negative control) was centrifuged at 15000 x g for 15 min and concentrated by ultrafiltration using Centricon Plus-70 / Ultracel PL-30 (Merck Millipore Ltd.). Then, western blotting was performed to confirm TSH production. Concentrated medium containing herring scTSH was denatured at 94 °C for 5 min in 0.1% SDS and 50 mM 2-mercaptoethanol, and then treated with 2.5 units of peptide-N-glycosidase F (Roche Diagnostics) at 37 °C for 2 h in 20 mM sodium phosphate with 0.5% Nonidet P-40, pH 7.5. All samples were run in 12% SDS-PAGE in the reducing condition and transferred to a PVDF membrane (Immobilon P; Millipore Corp.), then blocked overnight with 5% skimmed milk at 4 °C. After blocking, the membrane was incubated with polyclonal antisera against the sea bass alpha subunit (dilution 1:2000) for 90 min at room temperature, washed, and then further incubated with 1:25000 goat anti-rabbit immunoglobulin G (IgG) horseradish peroxidase conjugate (Bio-Rad Laboratories) for 60 min at room temperature. Immunodetection was performed by chemiluminescence with a Pierce ECL Plus Western Blotting Substrate kit (ThermoFisher Scientific).Cell surface expressionA Rhotag (MNGTEGPNFYVPFSNKTGVVYEE) was inserted at the N-terminus of herring TSHR for flow cytometry analysis of receptor cell surface expression. Anti-Rhotag polyclonal antibody was kindly provided by Drs. Gilbert Vassart and Sabine Costagliola (Université Libre de Bruxelles, Brussels, Belgium). PBS containing 1% BSA and 0.05% sodium azide was prepared as the flow cytometry (FCM) buffer for the washing and antibody incubation steps. 2.2 × 106 EPC cells were seeded in a 100 mm poly-d-Lysine-treated petri dish the day before transfection. Each dish was transfected with 10 μg TSHR or empty pcDNA3.1 expression plasmid using 20 μl jetPRIME transfection reagent in 500 μl jetPRIME transfection buffer (Polyplus transfection). Cells were harvested 24 h after transfection, then washed once in cold PBS and fixed in 2% PFA for 10 min at room temperature. After fixation, cells were washed three times with FCM buffer, then incubated with anti-Rhotag antibody or FCM buffer (negative control) for 1 h at room temperature. Cells were washed again with FCM buffer three times and stained with Alexa Fluor 488-labeled chicken anti-mouse IgG (H + L) antibody (1:200 dilution, ThermoFisher Scientific) or FCM buffer (negative control) for 45 min in the dark. After the fluorescent staining, cells were washed three times and resuspended in FCM buffer before analysis on a CytoFLEX instrument (Beckman Coulter). A minimum of 100,000 events was recorded for each sample, fluorescence intensities of negative control and cells transfected with empty pcDNA3.1 plasmid were used as the background for gating strategy. Cell surface expression was represented by the mean fluorescence intensity of the positively stained cell population.Dual-luciferase reporter assayEPC or HEK293 cells were plated in a 48-well plate at a density of 1 × 105 cells/well the day before transfection. A total of 250 ng plasmid mixture containing pGL4.29[luc2P/CRE/Hygro], TSHR expression plasmid (or empty pcDNA3.1) and pRL-TK with the ratio of 20:5:1 was prepared to transfect each well of cells using jetPRIME transfection reagent (Polyplus). Medium was replaced by fresh medium containing 10% FBS (TSH-induced condition) or serum-free medium (constitutive activity condition) 4 h after transfection. On day three, cells were treated with serum-free medium containing different dilutions of the concentrated scTSH medium for 4 h (TSH-induced condition) or directly subjected to the luminescence measurement without TSH induction (constitutive activity condition). Luminescence was measured using a Dual-Luciferase Reporter assay (Promega) on an Infinite M200 Microplate Reader (Tecan Group Ltd., Switzerland), and luciferase activity was represented as the ratio of firefly (pGL4.29[luc2P/CRE/Hygro]) to Renilla (pRL-TK) luminescence.5′-RACE to identify the herring DIO2 TSSTotal RNA was prepared from brain of a spring-spawning Atlantic herring using the RNeasy Mini Kit (Qiagen). Six μg of the isolated RNA was used for 5′-RACE with a FirstChoiceTM RLM-RACE Kit (ThermoFisher Scientific). One μl cDNA or Outer RACE PCR product was used as PCR template in a 20 μL reaction containing 0.8 mM dNTPs, 0.3 μM of each forward and reverse primer (Supplementary Table 1) and 0.75 U PrimeSTAR GXL DNA Polymerase (TaKaRa). Amplification was carried out with an initial denaturation of 3 min at 95 °C, followed by 35 cycles of 98 °C for 10 s, 58 °C for 20 s and 68 °C for 40 s, and a final extension of 10 min at 68 °C. The final 5′ RACE product was sequenced at Eurofins Genomics (Ebersberg, Germany).Sequence conservation analysisGenomic sequences covering the TSHR locus were extracted from Ensembl Genome Browser for Atlantic herring and 11 other fish species, including Amazon molly (Poecilia formosa), denticle herring (Denticeps clupeoides), goldfish (Carassius auratus), guppy (Poecilia reticulata), Neolamprologus brichardi, Japanese medaka (Oryzias latipes), northern pike (Esox lucius), orange clownfish (Amphiprion percula), spotted gar (Lepisosteus oculatus), three-spined stickleback (Gasterosteus aculeatus) and spotted green pufferfish (Tetraodon nigroviridis). The extracted sequences were firstly aligned using progressiveCactus49,50, and a subsequent alignment was generated using the hal2maf program from halTools51 with Atlantic herring assembly (Ch_v2.0.2)18 as the co-ordinate backbone. This alignment was used for the downstream phastCons score calculation by running phyloFit24 and phastCons25 from the PHAST package with default parameters. Peaks were called by grouping signals with a minimum phastCons score of 0.2 within 500 bp region.Structure modeling of human and herring TSHRsIn order to explore the possible interactions of the variant residues with other receptor interacting proteins and to study intramolecular interactions, we built a structural homology model for the herring TSHR (herrTSHR) complexed with herring TSH and Gs-protein. The TSHR hinge region that harbors the Q370H substitution and the C-terminus containing the 22aa repeat were excluded from the homology model due to the lack of structural templates for these regions. The homology model was constructed by using the following structural templates of evolutionarily related class A GPCRs: (i) the leucine-rich repeat domain (LRRD) complexed with hormone was modeled based on the solved FSHR LRRD – FSH complex structure (Protein Data Bank (PDB) ID: 4AY9)52,53, this part of model included herring TSHR Cys33 – Asn296 and fragments of the hinge region Gln297 – Thr312 and Ser393 – Ile421; (ii) the available structural complex of β2-adrenoreceptor with Gs-protein (PDB ID: 3SN6)54 was used as the template to model the seven-transmembrane helix domain (7TMD) of herring TSHR in the active conformation; (iii) the extracellular loop 2 (ECL2) was built by using the ECL2 of μ-opioid receptor (PDB ID: 6DDE)55. To prepare the template for herring TSHR modeling, the fused T4-lysozyme and bound ligand of β2-adrenoreceptor were deleted, the ECL1 and ECL3 loops were adjusted manually to the loop length of herring TSHR. Due to the lack of third intracellular loop (ICL3) in the β2-adrenoreceptor structure, amino acid residues of herring TSHR ICL3 were manually added to the template. Since herring TSHR does not have the TMH5 proline, which is highly conserved among all class A GPCRs and responsible for the helical kinks and bulges within this region56, we assumed a rather regular (stretched) helix conformation for the herring TSHR TMH5 and therefore replaced the kinked β2-adrenoreceptor TMH5 template with a regular α-helix. Moreover, the ECL2 template was substituted with μOR ECL2 structure because of its higher sequence similarity with herring TSHR in this region. Finally, amino acid residues of this chimeric 7TMD template and FSHR N-terminus were mutated to the corresponding spring herring TSHR residues and sequence of the heterodimeric FSH ligand was substituted by the herring TSH. All homology models were generated by using SYBYL-X 2.0 (Certara, NJ, US). The 7TMD structure was then fused with FSHR N-terminus at position 421. The assembled complex was subsequently optimized by the energy minimization under constrained backbone atoms (the AMBER F99 force field was used), followed by a 2 ns molecular dynamics simulation (MD) of the side chains. The entire TSHR complex was energetically minimized without any constraints until converging at a termination gradient of 0.05 kcal/mol*Å. Next, for autumn herrTSHR modeling, the spring TSHR sequence was substituted with autumn TSHR sequence. For humTSHR, the spring TSHR sequence was substituted with human TSHR, and the herring TSH ligand was replaced by the bovine TSH sequence. Both complex models were energetically minimized until converging at a termination gradient of 0.05 kcal/mol*Å.To investigate the microenvironment around the L471M mutation at TMH2 position 2.51, local short MD’s of 4 ns on Met4712.51 (spring herrTSHR), Leu471 (autumn herrTSHR) or Phe461 (humTSHR) and its surrounding amino acids were performed. During MD simulations, backbone atoms of the entire complexes as well as all side chains, except residues at positions 1.47, 1.51, 1.54, 2.48, 2.52, and 2.55 that form the hydrophobic patch around position 2.51, were constrained.Statistics and reproducibilityResults were presented as the mean + SD (standard deviation) calculated from at least four biological replicates for each experiment, and at least two independent experiments were conducted for each assay. Unpaired two-tailed Student’s t test was performed to calculate the P-values and means were judged as statistically significant when P ≤ 0.05.Reporting summaryFurther information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Multi-year presence of humpback whales in the Atlantic sector of the Southern Ocean but not during El Niño

    1.Clapham, P. J. in Encyclopedia of marine mammals 489–492 (Elsevier, 2018).2.Stevick, P. T. et al. A quarter of a world away: female humpback whale moves 10 000 km between breeding areas. Biol. Lett. 7, 299–302 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    3.Nicol, S. et al. Southern Ocean iron fertilization by baleen whales and Antarctic krill. Fish. Fish. 11, 203–209 (2010).Article 

    Google Scholar 
    4.Smetacek, V. & Nicol, S. Polar ocean ecosystems in a changing world. Nature 437, 362–368 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Dunlop, R. A. Potential motivational information encoded within humpback whale non-song vocal sounds. J. Acoustical Soc. Am. 141, 2204–2213 (2017).Article 

    Google Scholar 
    6.Stimpert, A. K., Au, W. W. L., Parks, S. E., Hurst, T. & Wiley, D. N. Common humpback whale (Megaptera novaeangliae) sound types for passive acoustic monitoring. J. Acoustical Soc. Am. 129, 476–482 (2011).Article 

    Google Scholar 
    7.Van Opzeeland, I., Van Parijs, S., Kindermann, L., Burkhardt, E. & Boebel, O. Calling in the cold: pervasive acoustic presence of humpback whales (Megaptera novaeangliae) in Antarctic coastal waters. PLoS ONE 8, 1–7 (2013).
    Google Scholar 
    8.Siegel, V. Biology and Ecology of Antarctic Krill. (Springer, 2016).9.Atkinson, A. et al. Krill (Euphausia superba) distribution contracts southward during rapid regional warming. Nat. Clim. Change 9, 142–147 (2019).Article 

    Google Scholar 
    10.Loeb, V. J., Hofmann, E. E., Klinck, J. M., Holm-Hansen, O. & White, W. B. ENSO and variability of the Antarctic Peninsula pelagic marine ecosystem. Antarctic Sci. https://doi.org/10.1017/s0954102008001636 (2009).11.Loeb, V. J. & Santora, J. A. Climate variability and spatiotemporal dynamics of five Southern Ocean krill species. Prog. Oceanogr. 134, 93–122 (2015).Article 

    Google Scholar 
    12.Bombosch, A. et al. Predictive habitat modelling of humpback (Megaptera novaeangliae) and Antarctic minke (Balaenoptera bonaerensis) whales in the Southern Ocean as a planning tool for seismic surveys. Deep-Sea Res. Part I: Oceanographic Res. Pap. 91, 101–114 (2014).Article 

    Google Scholar 
    13.Brierley, A. S. et al. Antarctic krill under sea ice: elevated abundance in a narrow band just south of ice edge. Science 295, 1890–1892 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    14.Rettig, S. et al. in 1st International Conference and Exhibition on Underwater Acoustics. (eds Papadakis, J. & Bjorno, L.) 1669–1674 (2013).15.Garland, E. C. et al. Humpback whale song on the Southern Ocean feeding grounds: Implications for cultural transmission. PLoS ONE 8, e79422 (2013).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    16.Stimpert, A. K., Peavey, L. E., Friedlaender, A. S. & Nowacek, D. P. Humpback whale song and foraging behavior on an Antarctic feeding ground. PLoS ONE 7, e51214 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Filun, D. et al. Frozen verses: Antarctic minke whales (Balaenoptera bonaerensis) call predominantly during austral winter. R. Soc. Open Sci. 7, 192112 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    18.Boebel, O. The Expedition PS89 of the Research Vessel POLARSTERN to the Weddell Sea in 2014/2015. (Alfred Wegener Institute, Helmholtz Centre for Polar and Marine Research, Bremerhaven, 2015).19.Burkhardt, E. Whale sightings during Polarstern cruise PS96 (ANT-XXXI/2), https://doi.org/10.1594/PANGAEA.923113 (2020).20.Herr, H., Viquerat, S. & Siebert, U. Aerial cetacean survey Southern Ocean 2014/2015, https://doi.org/10.1594/PANGAEA.894938 (2018).21.Herr, H., Viquerat, S. & Siebert, U. Ship based cetacean survey Southern Ocean 2014/2015, https://doi.org/10.1594/PANGAEA.894873 (2018).22.National Oceanic and Atmospheric Administration & Department of Commerce. Climate Prediction Centre (CPC) Oceanic Nino Index (2019).23.Thomisch, K. et al. Spatio-temporal patterns in acoustic presence and distribution of Antarctic blue whales Balaenoptera musculus intermedia in the Weddell Sea. Endanger. Species Res. 30, 239–253 (2016).Article 

    Google Scholar 
    24.Širović, A. et al. Seasonality of blue and fin whale calls and the influence of sea ice in the Western Antarctic Peninsula. Deep Sea Res. Part II: Topical Stud. Oceanogr. 51, 2327–2344 (2004).Article 

    Google Scholar 
    25.Schall, E. et al. Large-scale spatial variabilities in the humpback whale acoustic presence in the Atlantic sector of the Southern Ocean. R. Soc. Open Sci. 7, 201347 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    26.Loeb, V., Hofmann, E. E., Klinck, J. M. & Holm-Hansen, O. Hydrographic control of the marine ecosystem in the South Shetland-Elephant Island and Bransfield Strait region. Deep Sea Res. Part II: Topical Stud. Oceanogr. 57, 519–542 (2010).Article 

    Google Scholar 
    27.Sallée, J.-B., Speer, K. & Rintoul, S. Zonally asymmetric response of the Southern Ocean mixed-layer depth to the Southern Annular Mode. Nat. Geosci. 3, 273–279 (2010).Article 
    CAS 

    Google Scholar 
    28.Kim, Y. S. & Orsi, A. H. On the variability of Antarctic Circumpolar Current fronts inferred from 1992–2011 altimetry. J. Phys. Oceanogr. 44, 3054–3071 (2014).Article 

    Google Scholar 
    29.Yuan, X. ENSO-related impacts on Antarctic sea ice: a synthesis of phenomenon and mechanisms. Antarct. Sci. 16, 415 (2004).Article 

    Google Scholar 
    30.Meredith, M. P., Murphy, E. J., Hawker, E. J., King, J. C. & Wallace, M. I. On the interannual variability of ocean temperatures around South Georgia, Southern Ocean: Forcing by El Niño/Southern Oscillation and the southern annular mode. Deep Sea Res. Part II: Topical Stud. Oceanogr. 55, 2007–2022 (2008).Article 

    Google Scholar 
    31.Lovenduski, N. S. & Gruber, N. Impact of the Southern Annular Mode on Southern Ocean circulation and biology. Geophys. Res. Lett. 32, 1–4 (2005).Article 

    Google Scholar 
    32.Craig, A. S., Herman, L. M., Gabriele, C. M. & Pack, A. A. Migratory timing of humpback whales (Megaptera novaeangliae) in the central north Pacific varies with age, sex and reproductive status. Behaviour 140, 981–1001 (2003).Article 

    Google Scholar 
    33.Hofmann, E. E., Klinck, J. M., Locarnini, R. A., Fach, B. & Murphy, E. Krill transport in the Scotia Sea and environs. Antarct. Sci. 10, 406–415 (1998).Article 

    Google Scholar 
    34.Barendse, J. et al. Migration redefined? Seasonality, movements and group composition of humpback whales Megaptera novaeangliae off the west coast of South Africa. Afr. J. Mar. Sci. 32, 1–22 (2010).Article 

    Google Scholar 
    35.Witteveen, B. H., Foy, R. J., Wynne, K. M. & Tremblay, Y. Investigation of foraging habits and prey selection by humpback whales (Megaptera novaeangliae) using acoustic tags and concurrent fish surveys. Mar. Mammal. Sci. 24, 516–534 (2008).Article 

    Google Scholar 
    36.Brown, M. R., Corkeron, P. J., Hale, P. T., Schultz, K. W. & Bryden, M. M. Evidence for a sex-segregated migration in the humpback whale (Megaptera novaeangliae). Proc. R. Soc. Lond. B 259, 229–234 (1995).CAS 
    Article 

    Google Scholar 
    37.International Whaling Commission. Report on the workshop on the comprehensive assessment of Southern Hemisphere humpback whales. J. Cetacea. Res. Manag. Spec. Issue 3, 1–50 (2011).
    Google Scholar 
    38.Findlay, K. P. et al. Humpback whale “super-groups” – A novel low-latitude feeding behaviour of Southern Hemisphere humpback whales (Megaptera novaeangliae) in the Benguela Upwelling System. PLOS ONE 12, e0172002 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    39.Gridley, T., Silva, M., Wilkinson, C., Seakamela, S. & Elwen, S. H. Song recorded near a super-group of humpback whales on a mid-latitude feeding ground off South Africa. J. Acoustical Soc. Am. 143, EL298–EL304 (2018).CAS 
    Article 

    Google Scholar 
    40.Ross-Marsh, E., Elwen, S., Prinsloo, A., James, B. & Gridley, T. Singing in South Africa: monitoring the occurrence of humpback whale (Megaptera novaeangliae) song near the Western Cape. Bioacoustics, 1–17, https://doi.org/10.1080/09524622.2019.1710254 (2020).41.Cai, W. et al. Increasing frequency of extreme El Niño events due to greenhouse warming. Nat. Clim. Change 4, 111–116 (2014).Article 

    Google Scholar 
    42.Bengtson Nash, S. M. et al. Signals from the south; humpback whales carry messages of Antarctic sea‐ice ecosystem variability. Glob. Change Biol. 24, 1500–1510 (2018).Article 

    Google Scholar 
    43.Baumgartner, M. F. & Mussoline, S. E. A generalized baleen whale call detection and classification system. J. Acoustical Soc. Am. 129, 2889–2902 (2011).Article 

    Google Scholar 
    44.Klinck, H. et al. Long-range underwater vocalizations of the crabeater seal (Lobodon carcinophaga). J. Acoustical Soc. Am. 128, 474–479 (2010).Article 

    Google Scholar 
    45.Risch, D. et al. Mysterious bio-duck sound attributed to the Antarctic minke whale (Balaenoptera bonaerensis). Biol. Lett. 10, 20140175 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Schall, E. & Van Opzeeland, I. Calls produced by Ecotype C killer whales (Orcinus orca) off the Eckstrom iceshelf, Antarctica. Aquat. Mamm. 43, 117–126 (2017).Article 

    Google Scholar 
    47.Van Opzeeland, I. et al. Acoustic ecology of Antarctic pinnipeds. Mar. Ecol. Prog. Ser. 414, 267–291 (2010).Article 

    Google Scholar 
    48.Dunlop, R. A., Cato, D. H. & Noad, M. J. Non-song acoustic communication in migrating humpback whales (Megaptera novaeangliae). Mar. Mammal. Sci. 24, 613–629 (2008).Article 

    Google Scholar 
    49.Schall, E. & El-Gabbas, A. Humpback-whale-acoustic-detection-and-environmental-modelling, https://github.com/elenaschall/Humpback-whale-acoustic-detection-and-environmental-modelling (GitHub, GitHub, 2021).50.Bioacoustics, Research & Program. Raven Pro: Interactive Sound Analysis Software (Version 1.5) http://ravensoundsoftware.com/ (The Cornell Lab of Ornithology, Ithaca, NY, 2014).51.Cavalieri, D., Parkinson, C., Gloersen, P. & Zwally, H. Sea ice concentrations from Nimbus-7 SMMR and DMSP SSM/I-SSMIS passive microwave data, version 1. Boulder, Colorado USA, NASA National Snow and Ice Data Center Distributed Active Archive Center 10, https://doi.org/10.5067/8GQ8LZQVL0VL (1996).52.Greene, C. A. Daily Antarctic sea ice concentration (2020).53.Marshall, G. J. Trends in the southern annular mode from observations and reanalyses. J. Clim. 16, 4134–4143 (2003).Article 

    Google Scholar 
    54.Marshall, G. & National Center for Atmospheric Research Staff (Eds). The climate data guide: Marshall Southern Annular Mode (SAM) index (Station-based), (2019).55.R Core Team. R: A language and environment for statistical computing, https://www.R-project.org/ (R Foundation for Statistical Computing, Vienna, Austria, 2018)56.National Oceanic and Atmospheric Administration (NOAA) & Climate Prediction Centre (CPC). Oceanic Nino Index (2019).57.Wood, S. N. Generalized additive models: an introduction with R (CRC press, 2017).58.Pinheiro, J., Bates, D., DebRoy, S. & Sarkar, D. _nlme: Linear and nonlinear mixed effects models. R package version 3.1-145 (R CoreTeam, 2020).59.Hyndman, R. et al. forecast: Forecasting functions for time series and linear models_. R. package version 8.11, http://pkg.robjhyndman.com/forecast (2020)..60.Schall, E. et al. Humpback whale acoustic presence in the Atlantic sector of the Southern Ocean, https://doi.org/10.5061/dryad.ncjsxkss0 (2021).61.Wessel, P. & Smith, W. H. A global, self‐consistent, hierarchical, high‐resolution shoreline database. J. Geophys. Res.: Solid Earth 101, 8741–8743 (1996).Article 

    Google Scholar 
    62.Amante, C. & Eakins, B. W. ETOPO1 arc-minute global relief model: procedures, data sources and analysis, https://doi.org/10.7289/V5C8276M (2009).63.Spreen, G., Kaleschke, L. & Heygster, G. Sea ice remote sensing using AMSR-E 89-GHz channels. J. Geophys Res. Oceans 113, C02S03 (2008).Article 

    Google Scholar  More

  • in

    The impact of large and small dams on malaria transmission in four basins in Africa

    Study areaFour major river basins, located across different sub-regions of SSA, were selected for this study: Limpopo, Omo-Turkana, Volta, and Zambezi (Fig. 1). These basins were selected to (i) foster inclusion of enable different African regions and (ii) ensure focus on basins with sufficient data availability.Figure 1source malaria data23 on ArcGIS software (version 10.5. 1, Environmental Systems Research Institute Inc, Redlands, CA, USA, 2016)].Distribution of large and small dams in Limpopo, Volta, Zambezi and Omo-Turkana basins by malaria stability zone. [The figure was made using open-Full size imageThe Limpopo River basin is located in southern Africa. Draining an area of approximately 408,000 km2, the Limpopo River basin is distributed among South Africa (45%), Botswana (20%), Zimbabwe (15%) and Mozambique (20%). About 14 million people live in this basin. The climate of the Limpopo River basin varies along the path of the river from a temperate climate in the west to a subtropical climate at the river mouth in Mozambique. The hydrology of the Limpopo River basin is influenced by the highly seasonal distribution of rainfall over the catchment. About 95% of rain falls between October and April with a peak normally in February. Temperature varies from 30 to 34 °C in summer and 22–26 °C in winter15.The Volta River basin is located in West Africa with a population of over 23 million. Draining an area of 409,000 km2 the basin is spread across six countries: Benin (4%), Burkina Faso (42%), Cote d’Ivoire (3%), Ghana (41%), Mali (4%) and Togo (6%). Average annual rainfall varies across the basin from approximately 1600 mm in the southeast, to about 360 mm in the north. Annual mean temperatures in the basin vary from 27 to 30 °C16. The main rainy season is between March and October.The Zambezi River basin is located in southern Africa. Draining an area of 1.34 million km2, the basin is spread across eight countries: Angola (19%), Botswana (1%), Namibia (1%) Benin (4%), Zimbabwe (16%), Zambia (42%), Tanzania (2%), Malawi (8%) and Mozambique (12%). The population of the Zambezi basin is estimated to be about 32 million. Annual rainfall in the basin ranges from 550 mm in the south to 1800 mm in the north. The annual mean temperatures ranges from 18 °C at higher elevations in the south of the basin to 26 °C for low elevations in the delta in Mozambique17.The Omo-Turkana Basin covers approximately 131,000 km2, stretching from southern Ethiopia to northern Kenya. Hydrologically, the basin is dominated by Lake Turkana, with the Omo River, which drains the Ethiopian portion of the basin, supplying 90% of the inflow to the lake. The basin is home to approximately 15 million people, the majority of whom live in the Ethiopian highlands, in the north. The annual mean temperature ranges from 24 °C in the north to 29 °C in the south. The mean annual rainfall ranges from 250 mm in the south to 500 mm in the north18.Data sourcesDam dataSmall damsData on location and size of small dams are not readily available in either global or regional data sets. The European Commission’s Joint Research Center (JRC) Yearly Water Classification History v1.0 data set was used to identify water bodies in each of the four basins19. Water bodies less than 100 ha and greater than 2 ha were identified. All were checked with Google Earth images to distinguish between reservoirs and natural water bodies (Supplementary Fig. S1). Ultimately, a total of 4907 small dams located in the four basins were identified and included in the analyses.Large damsFor large dams, the FAO African Dams Database20, International Commission for Large dams (ICOLD)21 and the International Rivers Database22, which together contain 1286 georeferenced African large dams, were utilized. The accuracy of dam locations was first verified with Google Earth. When the location of a dam did not precisely match the coordinates stipulated in either of the two databases, manual corrections were made by adjusting the coordinates of a dam to its location as shown in Google Earth (see Supplementary Information). Dams for which precise locations could not be determined, as well as dams without reservoirs (i.e., run-of-river schemes), were removed. Ultimately, across the four basins, a total of 258 large dams with confirmed georeferenced locations were identified and included in the analyses.Perimeters of large and small dam reservoirsReservoir perimeters of both large and small dams were extracted from the European Commission’s Joint Research Center (JRC) global surface water datasets19, published through the Google Earth Engine. This dataset includes maps of the location and temporal variability in maximum perimeter records of the global surface water coverage from 1984 to 2015. In this study, the maximum perimeter records were used in each year of 2000, 2005, 2010 and 2015. The data were exported to ArcGIS.Data on anopheles mosquito distributionData for vector distribution were obtained from the Malaria Atlas Project (MAP) database23. The MAP database contains a georeferenced illustration of the major malaria vector species in different malaria-endemic areas in Africa.Malaria dataAnnual malaria incidence data were obtained from the MAP database. We acquired data for the years 2000, 2005, 2010 and 2015. These years were selected to align with updates to Worldpop population data24, which are recomputed every five years. MAP produced a 1 km resolution continuous map of annual malaria incidence for Africa based on 33,761 studies across the region. We imported these data to ArcGIS for analyses. Annual malaria incidence was determined as the number of cases per 1000 population. To ascertain the impact of dams on malaria incidence rates as a function of distance from the reservoir perimeter, we created two distance zones: 0–5 km (at risk) and 5–10 km (control). When distance zones were overlapping for two or more nearby dams, areas were assigned to the closest distance cohort. Populations residing more than 5 km from a reservoir perimeter (large or small) were considered to be free of risk from dam induced malaria transmission because the maximum mosquitoes’ flight range is considered to be  0.1 malaria cases per 1000 population), unstable (≤ 0.1 malaria cases per 1000 population) and no malaria (zero malaria incidence) based on the level of malaria incidence in each of the four years: 2000, 2005, 2010, and 2015. The number of dams in each of the three stability categories for each of the four years was determined, as well as the population at-risk of dam-related malaria (i.e.,  More

  • in

    Seasonal influence on the bathymetric distribution of an endangered fish within a marine protected area

    1.Lejeusne, C., Chevaldonné, P., Pergent-Martini, C., Boudouresque, C. F. & Pérez, T. Climate change effects on a miniature ocean: The highly diverse, highly impacted Mediterranean Sea. Trends Ecol. Evol. 25, 250–260 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Fulton, E. A. et al. Modelling marine protected areas: Insights and hurdles. Philos. Trans. R. Soc. B Biol. Sci. 370, 20140278 (2015).Article 

    Google Scholar 
    3.Roff, J. & Zacharias, M. Marine Conservation Ecology (Earthscan, 2011).
    Google Scholar 
    4.Mitcheson, Y. S. D. et al. A global baseline for spawning aggregations of reef fishes. Conserv. Biol. 22, 1233–1244 (2008).Article 

    Google Scholar 
    5.Salinas-de-León, P., Rastoin, E. & Acuña-Marrero, D. First record of a spawning aggregation for the tropical eastern Pacific endemic grouper Mycteroperca olfax in the Galapagos Marine Reserve. J. Fish Biol. 87, 179–186 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Bueno, L. S. et al. Evidence for spawning aggregations of the endangered Atlantic goliath grouper Epinephelus itajara in southern Brazil. J. Fish Biol. 89, 876–889 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Kumar, V. et al. Biological clocks and regulation of seasonal reproduction and migration in birds. Physiol. Biochem. Zool. 83, 827–835 (2010).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    8.van Haren, H. & Compton, T. J. Diel vertical migration in deep sea plankton is finely tuned to latitudinal and seasonal day length. PLoS ONE 8, e64435 (2013).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    9.Horký, P. & Slavík, O. Diel and seasonal rhythms of asp Leuciscus aspius (L.) in a riverine environment. Ethol. Ecol. Evol. 29, 449–459 (2017).Article 

    Google Scholar 
    10.Falcón, J., Besseau, L., Sauzet, S. & Boeuf, G. Melatonin effects on the hypothalamo–pituitary axis in fish. Trends Endocrinol. Metab. 18, 81–88 (2007).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    11.Oliveira, C. et al. Monthly day/night changes and seasonal daily rhythms of sexual steroids in Senegal sole (Solea senegalensis) under natural fluctuating or controlled environmental conditions. Comp. Biochem. Physiol. A: Mol. Integr. Physiol. 152, 168–175 (2009).ADS 
    Article 
    CAS 

    Google Scholar 
    12.Wuitchik, D. M. et al. Seasonal temperature, the lunar cycle and diurnal rhythms interact in a combinatorial manner to modulate genomic responses to the environment in a reef-building coral. Mol. Ecol. 28, 3629–3641 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.Sanchez-Cardenas, C. et al. Pituitary growth hormone network responses are sexually dimorphic and regulated by gonadal steroids in adulthood. PNAS 107, 21878–21883 (2010).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    14.Stock, C. A. et al. Seasonal sea surface temperature anomaly prediction for coastal ecosystems. Prog. Oceanogr. 137, 219–236 (2015).ADS 
    Article 

    Google Scholar 
    15.Bisagni, J. J. Salinity variability along the eastern continental shelf of Canada and the United States, 1973–2013. Cont. Shelf Res. 126, 89–109 (2016).ADS 
    Article 

    Google Scholar 
    16.Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).Article 

    Google Scholar 
    17.Dorts, J. et al. Evidence that elevated water temperature affects the reproductive physiology of the European bullhead Cottus gobio. Fish Physiol. Biochem. 38, 389–399 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    18.Arfuso, F. et al. Water temperature influences growth and gonad differentiation in European sea bass (Dicentrarchus labrax, L. 1758). Theriogenology 88, 145–151 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    19.Aspillaga, E. et al. Thermal stratification drives movement of a coastal apex predator. Sci. Rep. 7, 526 (2017).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    20.Martin, T. L. & Huey, R. B. Why, “suboptimal” is optimal: Jensen’s inequality and ectotherm thermal preferences. Am. Nat. 171, E102–E118 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Freitas, C., Olsen, E. M., Moland, E., Ciannelli, L. & Knutsen, H. Behavioral responses of Atlantic cod to sea temperature changes. Ecol. Evol. 5, 2070–2083 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    22.Harmelin, J.-G. & Marinopoulos, J. Recensementde la Population de corbs (Sciaena umbra, Linneaus1758: Pisces) du Parc National de Port-Cros (Méditerrannée, France) par Inventaires Visuels 265–275 (1993).23.Coll, J., Linde, M., García-Rubies, A., Riera, F. & Grau, A. M. Spear fishing in the Balearic Islands (west central Mediterranean): Species affected and catch evolution during the period 1975–2001. Fish. Res. 70, 97–111 (2004).Article 

    Google Scholar 
    24.Lloret, J. et al. Spearfishing pressure on fish communities in rocky coastal habitats in a Mediterranean marine protected area. Fish. Res. 94, 84–91 (2008).Article 

    Google Scholar 
    25.Harmelin, J.-G. Statut du corb (Sciaena umbra) en méditerranée. In Les Espèces Marines à Protéger en Méditerranée (eds Boudouresque, C. F. et al.) 219–227 (GiS Posidonie Publ., 1991).
    Google Scholar 
    26.Mayol, J., Grau, A. M., Riera, F. & Oliver, J. Llista Ver-mella dels Peixos de les Balears (2000).
    Google Scholar 
    27.Chao, L. The IUCN Red List of Threatened Species 2020: e.T198707A130230194Sciaena umbra (2020). https://doi.org/10.2305/IUCN.UK.2020-2.RLTS.T198707A130230194.en.28.Forcada, A. et al. Effects of habitat on spillover from marine protected areas to artisanal fisheries. Mar. Ecol. Prog. Ser. 379, 197–211 (2009).ADS 
    Article 

    Google Scholar 
    29.Franco, A. D., Bussotti, S., Navone, A., Panzalis, P. & Guidetti, P. Evaluating effects of total and partial restrictions to fishing on Mediterranean rocky-reef fish assemblages. Mar. Ecol. Prog. Ser. 387, 275–285 (2009).ADS 
    Article 

    Google Scholar 
    30.Le Préfet de la region Provence-Alpes-Côte d’Azur, Préfet de la zone de défense et de sécurité Sud, Préfet des Bouches-du-Rhône. http://www.dirm.mediterranee.developpement-durable.gouv.fr/IMG/pdf/ap_corb_med_continentale_20_dec_2018-2.pdf (Accessed 20 December 2018).31.Harmelin-Vivien, M. et al. Effects of reserve protection level on the vulnerable fish species Sciaena umbra and implications for fishing management and policy. Glob. Ecol. Conserv. 3, 279–287 (2015).Article 

    Google Scholar 
    32.Chakroun-Marzouk, N. & Ktari, M.-H. Le Corb des côtes Tunisiennes, Sciaena umbra (Sciaenidae): Cycle Sexuel, Age et Croissance 15 (2003).33.Derbal, F. & Kara, M. H. Régime Alimentaire du corb Sciaena umbra (Sciaenidae) des côtes de l’est Algérien 9 (2007).34.Engin, S. & Seyhan, K. Age, growth, sexual maturity and food composition of Sciaena umbra in the south-eastern Black Sea, Turkey. J. Appl. Ichthyol. 25, 96–99 (2009).Article 

    Google Scholar 
    35.Botsford, L. W. et al. Connectivity, sustainability, and yield: Bridging the gap between conventional fisheries management and marine protected areas. Rev. Fish Biol. Fish. 19, 69–95 (2009).Article 

    Google Scholar 
    36.Alós, J. & Cabanellas-Reboredo, M. Experimental acoustic telemetry experiment reveals strong site fidelity during the sexual resting period of wild brown meagre, Sciaena umbra. J. Appl. Ichthyol. 28, 606–611 (2012).Article 

    Google Scholar 
    37.Jadot, C., Donnay, A., Acolas, M. L., Cornet, Y. & Bégout Anras, M. L. Activity patterns, home-range size, and habitat utilization of Sarpa salpa (Teleostei: Sparidae) in the Mediterranean Sea. ICES J. Mar. Sci. 63, 128–139 (2006).Article 

    Google Scholar 
    38.Jorgensen, S. J. et al. Limited movement in blue rockfish Sebastes mystinus: Internal structure of home range. Mar. Ecol. Prog. Ser. 327, 157–170 (2006).ADS 
    Article 

    Google Scholar 
    39.Kerwath, S. E., Götz, A., Attwood, C. G., Sauer, W. H. H. & Wilke, C. G. Area utilisation and activity patterns of roman Chrysoblephus laticeps (Sparidae) in a small marine protected area. Afr. J. Mar. Sci. 29, 259–270 (2007).Article 

    Google Scholar 
    40.Collins, A. B., Heupel, M. R. & Motta, P. J. Residence and movement patterns of cownose rays Rhinoptera bonasus within a south-west Florida estuary. J. Fish Biol. 71, 1159–1178 (2007).Article 

    Google Scholar 
    41.Abecasis, D. & Erzini, K. Site fidelity and movements of gilthead sea bream (Sparus aurata) in a coastal lagoon (Ria Formosa, Portugal). Estuar. Coast. Shelf Sci. 79, 758–763 (2008).ADS 
    Article 

    Google Scholar 
    42.Afonso, P. et al. A multi-scale study of red porgy movements and habitat use, and its application to the design of marine reserve networks. In Tagging and Tracking of Marine Animals with Electronic Devices (eds Nielsen, J. L. et al.) 423–443 (Springer, 2009).Chapter 

    Google Scholar 
    43.March, D., Palmer, M., Alós, J., Grau, A. & Cardona, F. Short-term residence, home range size and diel patterns of the painted comber Serranus scriba in a temperate marine reserve. Mar. Ecol. Prog. Ser. 400, 195–206 (2010).ADS 
    Article 

    Google Scholar 
    44.Zeller, D. C. Ultrasonic telemetry: Its application to coral reef fisheries research. Fish. Bull. 97, 1058–1065 (1999).
    Google Scholar 
    45.Heupel, M. R., Semmens, J. M. & Hobday, A. J. Automated acoustic tracking of aquatic animals: Scales, design and deployment of listening station arrays. Mar. Freshw. Res. 57, 1–13 (2006).Article 

    Google Scholar 
    46.Lowe, C. G., Topping, D. T., Cartamil, D. P. & Papastamatiou, Y. P. Movement patterns, home range, and habitat utilization of adult kelp bass Paralabrax clathratus in a temperate no-take marine reserve. Mar. Ecol. Prog. Ser. 256, 205–216 (2003).ADS 
    Article 

    Google Scholar 
    47.Kaunda-Arara, B. & Rose, G. A. Homing and site fidelity in the greasy grouper Epinephelus tauvina (Serranidae) within a marine protected area in coastal Kenya. Mar. Ecol. Prog. Ser. 277, 245–251 (2004).ADS 
    Article 

    Google Scholar 
    48.Parsons, D. & Egli, D. Fish movement in a temperate marine reserve: New insights through application of acoustic tracking. Mar. Technol. Soc. J. 39, 56–63 (2005).Article 

    Google Scholar 
    49.Topping, D. T., Lowe, C. G. & Caselle, J. E. Home range and habitat utilization of adult California sheephead, Semicossyphus pulcher (Labridae), in a temperate no-take marine reserve. Mar. Biol. 147, 301–311 (2005).Article 

    Google Scholar 
    50.Pastor, J. et al. Acoustic telemetry survey of the dusky grouper (Epinephelus marginatus) in the Marine Reserve of Cerbère-Banyuls: Informations on the territoriality of this emblematic species. C.R. Biol. 332, 732–740 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.D’Anna, G., Giacalone, V. M., Pipitone, C. & Badalamenti, F. Movement pattern of white seabream, Diplodus sargus (L., 1758) (Osteichthyes, Sparidae) acoustically tracked in an artificial reef area. Ital. J. Zool. 78, 255–263 (2011).Article 

    Google Scholar 
    52.La Mesa, G., Consalvo, I., Annunziatellis, A. & Canese, S. Movement patterns of the parrotfish Sparisoma cretense in a Mediterranean marine protected area. Mar. Environ. Res. 82, 59–68 (2012).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    53.Picciulin, M. et al. Passive acoustic monitoring of Sciaena umbra on rocky habitats in the Venetian littoral zone. Fish. Res. 145, 76–81 (2013).Article 

    Google Scholar 
    54.Lenfant, P., Louisy, P. & Licari, M.-L. Recensement des Mérous bruns (Epinephelus marginatus) de la Réserve Naturelle de Cerbère-Banyuls (France, Méditerranée) Effectué en Septembre 2001, aprés 17 Années de Protection 10 (2003).55.Koeck, B., Gudefin, A., Romans, P., Loubet, J. & Lenfant, P. Effects of intracoelomic tagging procedure on white seabream (Diplodus sargus) behavior and survival. J. Exp. Mar. Biol. Ecol. 440, 1–7 (2013).Article 

    Google Scholar 
    56.Garcia, J., Mourier, J. & Lenfant, P. Spatial behavior of two coral reef fishes within a Caribbean marine protected area. Mar. Environ. Res. 109, 41–51 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Percie du Sert, N. et al. The ARRIVE guidelines 2.0: Updated guidelines for reporting animal research. PLoS Biol. 18(7), e3000410. https://doi.org/10.1371/journal.pbio.3000410 (2020).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    58.Grau, A., Linde, M. & Grau, A. M. Reproductive biology of the vulnerable species Sciaena umbra Linnaeus, 1758 (Pisces: Sciaenidae). Sci. Mar. 73, 67–81 (2009).Article 

    Google Scholar 
    59.McKinzie, M. K., Jarvis, E. T. & Lowe, C. G. Fine-scale horizontal and vertical movement of barred sand bass, Paralabrax nebulifer, during spawning and non-spawning seasons. Fish. Res. 150, 66–75 (2014).Article 

    Google Scholar 
    60.Kassambara, A. & Mundt, F. factoextra: Extract and Visualize the Results of Multivariate Data Analyses (2017).61.Zuur, A., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer, 2009).MATH 
    Book 

    Google Scholar 
    62.Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D. & R Core Team. nlme: Linear and Nonlinear Mixed Effects Models (2018).63.Wood, S. N. Generalized Additive Models: An Introduction with R 2nd edn. (Chapman and Hall/CRC, 2017).MATH 
    Book 

    Google Scholar 
    64.R Core Team. R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, 2018).65.Wearmouth, V. J. & Sims, D. W. Chapter 2 sexual segregation in marine fish, reptiles, birds and mammals: Behaviour patterns, mechanisms and conservation implications. Adv. Mar. Biol. 54, 107–170 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    66.Haraldstad, Ø. & Jonsson, B. Age and sex segregation in habitat utilization by brown trout in a Norwegian Lake. Trans. Am. Fish. Soc. 112, 27–37 (1983).Article 

    Google Scholar 
    67.L’Abée-Lund, J. H., Langeland, A., Jonsson, B. & Ugedal, O. Spatial segregation by age and size in Arctic Charr: A trade-off between feeding possibility and risk of predation. J. Anim. Ecol. 62, 160–168 (1993).Article 

    Google Scholar 
    68.Oxenford, H. A. & Hunte, W. Feeding habits of the dolphinfish (Coryphaena hippurus) in the eastern Caribbean. Sci. Mar. 63, 303–315 (1999).Article 

    Google Scholar 
    69.Sarà, G. et al. Effect of boat noise on the behaviour of bluefin tuna Thunnus thynnus in the Mediterranean Sea. Mar. Ecol. Prog. Ser. 331, 243–253 (2007).ADS 
    Article 

    Google Scholar 
    70.Codarin, A., Wysocki, L. E., Ladich, F. & Picciulin, M. Effects of ambient and boat noise on hearing and communication in three fish species living in a marine protected area (Miramare, Italy). Mar. Pollut. Bull. 58, 1880–1887 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    71.Picciulin, M., Sebastianutto, L., Codarin, A., Calcagno, G. & Ferrero, E. A. Brown meagre vocalization rate increases during repetitive boat noise exposures: A possible case of vocal compensation. J. Acoust. Soc. Am. 132, 3118–3124 (2012).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    72.McCormick, M. I., Allan, B. J. M., Harding, H. & Simpson, S. D. Boat noise impacts risk assessment in a coral reef fish but effects depend on engine type. Sci. Rep. 8, 3847 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    73.Robichaud, D. & Rose, G. A. Sex differences in cod residency on a spawning ground. Fish. Res. 60, 33–43 (2003).Article 

    Google Scholar 
    74.Fiorentino, F. et al. On a spawning aggregation of the brown meagre Sciaena umbra L. 1758 (Sciaenidae, Osteichthyes) in the Maltese waters (Sicilian Channel—Central Mediterranean). Rapp. Commun. Int. Mer Médit. 36, 266 (2001).
    Google Scholar 
    75.Furukawa, S. et al. Vertical movements of Pacific bluefin tuna (Thunnus orientalis) and dolphinfish (Coryphaena hippurus) relative to the thermocline in the northern East China Sea. Fish. Res. 149, 86–91 (2014).Article 

    Google Scholar 
    76.Claireaux, G., Webber, D., Kerr, S. & Boutilier, R. Physiology and behaviour of free-swimming Atlantic cod (Gadus morhua) facing fluctuating temperature conditions. J. Exp. Biol. 198, 49–60 (1995).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    77.Armstrong, J. B. et al. Diel horizontal migration in streams: Juvenile fish exploit spatial heterogeneity in thermal and trophic resources. Ecology 94, 2066–2075 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    78.Bobe, J. & Labbé, C. Egg and sperm quality in fish. Gen. Comp. Endocrinol. 165, 535–548 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    79.Pepin, P. Effect of temperature and size on development, mortality, and survival rates of the pelagic early life history stages of marine fish. Can. J. Fish. Aquat. Sci. 48, 503–518 (1991).Article 

    Google Scholar 
    80.Guevara-Fletcher, C., Alvarez, P., Sanchez, J. & Iglesias, J. Effect of temperature on the development and mortality of European hake (Merluccius merluccius L.) eggs from southern stock under laboratory conditions. J. Exp. Mar. Biol. Ecol. 476, 50–57 (2016).Article 

    Google Scholar 
    81.Dubrovský, M. et al. Multi-GCM projections of future drought and climate variability indicators for the Mediterranean region. Reg. Environ Change 14, 1907–1919 (2014).Article 

    Google Scholar 
    82.Pankhurst, N. W. & Munday, P. L. Effects of climate change on fish reproduction and early life history stages. Mar. Freshw. Res. 62, 1015–1026 (2011).CAS 
    Article 

    Google Scholar 
    83.McKenzie, D. J. et al. Conservation physiology of marine fishes: State of the art and prospects for policy. Conserv. Physiol. 4, 046 (2016).Article 

    Google Scholar 
    84.Fabi, G., Panfili, M. & Spagnolo, A. Note on feeding of Sciaena umbra L. (Asteichthyes:Sciaenidae) in the central Adriatic Sea. Rapp. Comm. Int. Mer Médit. 35, 426 (1998).
    Google Scholar 
    85.Fabi, G., Manoukian, S. & Spagnolo, A. Feeding behavior of three common fishes at an artificial reef in the northern Adriatic Sea. Bull. Mar. Sci. 78, 39–56 (2006).
    Google Scholar 
    86.Ramcharitar, J., Gannon, D. P. & Popper, A. N. Bioacoustics of fishes of the family Sciaenidae (croakers and drums). Trans. Am. Fish. Soc. 135, 1409–1431 (2006).Article 

    Google Scholar 
    87.Mesa, M. L., Colella, S., Giannetti, G. & Arneri, E. Age and growth of brown meagre Sciaena umbra (Sciaenidae) in the Adriatic Sea. Aquat. Living Resour. 21, 153–161 (2008).Article 

    Google Scholar 
    88.Picciulin, M. et al. Diagnostics of nocturnal calls of Sciaena umbra (L., fam. Sciaenidae) in a nearshore Mediterranean marine reserve. Bioacoustics 22, 109–120 (2013).Article 

    Google Scholar 
    89.Schmidt, M. B. & Gassner, H. Influence of scuba divers on the avoidance reaction of a dense vendace (Coregonus albula L.) population monitored by hydroacoustics. Fish. Res. 82, 131–139 (2006).Article 

    Google Scholar 
    90.Moffitt, E. A., Botsford, L. W., Kaplan, D. M. & O’Farrell, M. R. Marine reserve networks for species that move within a home range. Ecol. Appl. 19, 1835–1847 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar  More

  • in

    Antibiotic treatment increases yellowness of carotenoid feather coloration in male greenfinches (Chloris chloris)

    1.Hill, G. E. Plumage coloration is a sexually selected indicator of male quality. Nature 350, 337 (1991).ADS 
    Article 

    Google Scholar 
    2.Cantarero, A., Pérez-Rodríguez, L., Romero-Haro, A. Á., Chastel, O. & Alonso-Alvarez, C. Carotenoid-based coloration predicts both longevity and lifetime fecundity in male birds, but testosterone disrupts signal reliability. PLoS ONE 14, e0221436. https://doi.org/10.1371/journal.pone.0221436 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    3.Zahavi, A. Mate selection—A selection for a handicap. J. Theor. Biol. 53, 205–214 (1975).CAS 
    Article 

    Google Scholar 
    4.Alonso-Alvarez, C. & Galván, I. Free radical exposure creates paler carotenoid-based ornaments: A possible interaction in the expression of black and red traits. PLoS ONE 6 (2011).5.Schantz, T. V., Bensch, S., Grahn, M., Hasselquist, D. & Wittzell, H. Good genes, oxidative stress and condition–dependent sexual signals. Proc. R. Soc. Lond. Ser. B: Biol. Sci. 266, 1–12 (1999).Article 

    Google Scholar 
    6.Tomášek, O. et al. Opposing effects of oxidative challenge and carotenoids on antioxidant status and condition-dependent sexual signalling. Sci. Rep. 6, 23546. https://doi.org/10.1038/srep23546 (2016).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    7.Sild, E., Sepp, T., Männiste, M. & Hõrak, P. Carotenoid intake does not affect immune-stimulated oxidative burst in greenfinches. J. Exp. Biol. 214, 3467–3473 (2011).CAS 
    Article 

    Google Scholar 
    8.Mohr, A. E., Girard, M., Rowe, M., McGraw, K. J. & Sweazea, K. L. Varied effects of dietary carotenoid supplementation on oxidative damage in tissues of two waterfowl species. Comp. Biochem. Physiol. B: Biochem. Mol. Biol. 231, 67–74. https://doi.org/10.1016/j.cbpb.2019.02.003 (2019).CAS 
    Article 

    Google Scholar 
    9.Costantini, D. & Møller, A. Carotenoids are minor antioxidants for birds. Funct. Ecol. 22, 367–370 (2008).Article 

    Google Scholar 
    10.Simons, M. J. P., Cohen, A. A. & Verhulst, S. What does carotenoid-dependent coloration tell? Plasma carotenoid level signals immunocompetence and oxidative stress state in birds—A meta-analysis. PLoS ONE 7, e43088. https://doi.org/10.1371/journal.pone.0043088 (2012).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    11.Hill, G. E. et al. Plumage redness signals mitochondrial function in the house finch. Proc. R. Soc. B 286, 20191354 (2019).CAS 
    Article 

    Google Scholar 
    12.Hill, G. E. Condition-dependent traits as signals of the functionality of vital cellular processes. Ecol. Lett. 14, 625–634 (2011).Article 

    Google Scholar 
    13.del Cerro, S. et al. Carotenoid-based plumage colouration is associated with blood parasite richness and stress protein levels in blue tits (Cyanistes caeruleus). Oecologia 162, 825–835. https://doi.org/10.1007/s00442-009-1510-y (2010).ADS 
    Article 
    PubMed 

    Google Scholar 
    14.Hõrak, P. et al. How coccidian parasites affect health and appearance of greenfinches. J. Anim. Ecol. 73, 935–947 (2004).Article 

    Google Scholar 
    15.Weaver, R. J., Santos, E. S., Tucker, A. M., Wilson, A. E. & Hill, G. E. Carotenoid metabolism strengthens the link between feather coloration and individual quality. Nat. Commun. 9, 73 (2018).ADS 
    Article 

    Google Scholar 
    16.Tyczkowski, J. K., Hamilton, P. B. & Ruff, M. D. Altered metabolism of carotenoids during pale-bird syndrome in chickens infected with Eimeria acervulina. Poult. Sci. 70, 2074–2081. https://doi.org/10.3382/ps.0702074 (1991).CAS 
    Article 
    PubMed 

    Google Scholar 
    17.Joyner, L. et al. Amino-acid malabsorption and intestinal leakage of plasma-proteins in young chicks infected with Eimeria acervulina. Avian Pathol. 4, 17–33 (1975).CAS 
    PubMed 

    Google Scholar 
    18.Sharma, V. & Fernando, M. Effect of Eimeria acervulina infection on nutrient retention with special reference to fat malabsorption in chickens. Can. J. Comp. Med. 39, 146 (1975).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Pout, D. D. Villous atrophy and coccidiosis. Nature 213, 306–307 (1967).ADS 
    CAS 
    Article 

    Google Scholar 
    20.Sanches, A. W. D. et al. Basal and infectious enteritis in broilers under the I See inside methodology: A chronological evaluation. Front. Vet. Sci. 6, 512. https://doi.org/10.3389/fvets.2019.00512 (2020).Article 
    PubMed 

    Google Scholar 
    21.Russell, J. Jr. & Ruff, M. Eimeria spp.: Influence of coccidia on digestion (amylolytic activity) in broiler chickens. Exp. Parasitol. 45, 234–240 (1978).Article 

    Google Scholar 
    22.Kouwenhoven, B. & van der Horst, C. J. Disturbed intestinal absorption of vitamin A and carotenes and the effect of a low pH during Eimeria acervulina infection in the domestic fowl (Gallus domesticus). Z. Parasitenkd. 38, 152–161 (1972).CAS 
    Article 

    Google Scholar 
    23.Ruff, M. D. & Fuller, H. L. Some mechanisms of reduction of carotenoid levels in chickens infected with Eimeria acervulina or E. tenella. J. Nutr. 105, 1447–1456 (1975).CAS 
    Article 

    Google Scholar 
    24.Swayne, D. E., Getzy, D., Slemons, R. D., Bocetti, C. & Kramer, L. Coccidiosis as a cause of transmural lymphocytic enteritis and mortality in captive Nashville warblers (Vermivora ruficapilla). J. Wildl. Dis. 27, 615–620 (1991).CAS 
    Article 

    Google Scholar 
    25.Gosbell, M. C., Olaogun, O. M., Luk, K. & Noormohammadi, A. H. Investigation of systemic isosporosis outbreaks in an aviary of greenfinch (Carduelis chloris) and goldfinch (Carduelis carduelis) and a possible link with local wild sparrows (Passer domesticus). Aust. Vet. J. 98, 338–344 (2020).CAS 
    Article 

    Google Scholar 
    26.Baeta, R., Faivre, B., Motreuil, S., Gaillard, M. & Moreau, J. Carotenoid trade-off between parasitic resistance and sexual display: An experimental study in the blackbird (Turdus merula). Proc. R. Soc. B Biol. Sci. 275, 427–434 (2008).CAS 
    Article 

    Google Scholar 
    27.Amin, A., Bilic, I., Liebhart, D. & Hess, M. Trichomonads in birds—A review. Parasitology 141, 733–747 (2014).Article 

    Google Scholar 
    28.Robinson, R. A. et al. Emerging infectious disease leads to rapid population declines of common British birds. PLoS ONE 5 (2010).29.Chavatte, J.-M. et al. An outbreak of trichomonosis in European greenfinches Chloris chloris and European goldfinches Carduelis carduelis wintering in Northern France. Parasite 26, 21–21. https://doi.org/10.1051/parasite/2019022 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    30.Huyghebaert, G., Ducatelle, R. & Immerseel, F. V. An update on alternatives to antimicrobial growth promoters for broilers. Vet. J. 187, 182–188. https://doi.org/10.1016/j.tvjl.2010.03.003 (2011).CAS 
    Article 
    PubMed 

    Google Scholar 
    31.Singer, R. S. & Hofacre, C. L. Potential impacts of antibiotic use in poultry production. Avian Dis. 50, 161–172, 112 (2006).Article 

    Google Scholar 
    32.Miles, R. D., Butcher, G. D., Henry, P. R. & Littell, R. C. Effect of antibiotic growth promoters on broiler performance, intestinal growth parameters, and quantitative morphology1. Poult. Sci. 85, 476–485. https://doi.org/10.1093/ps/85.3.476 (2006).CAS 
    Article 
    PubMed 

    Google Scholar 
    33.Oh, S., Lillehoj, H. S., Lee, Y., Bravo, D. & Lillehoj, E. P. Dietary antibiotic growth promoters down-regulate intestinal inflammatory cytokine expression in chickens challenged with LPS or co-infected with Eimeria maxima and Clostridium perfringens. Front. Vet. Sci. https://doi.org/10.3389/fvets.2019.00420 (2019).Article 
    PubMed 

    Google Scholar 
    34.Meitern, R., Lind, M. A., Karu, U. & Hõrak, P. Simple and noninvasive method for assessment of digestive efficiency: Validation of fecal steatocrit in greenfinch coccidiosis model. Ecol. Evol. 6, 8756–8763 (2016).Article 

    Google Scholar 
    35.Surai, P., Speake, B. & Sparks, N. Carotenoids in avian nutrition and embryonic development. 1. Absorption, availability and levels in plasma and egg yolk. J. Poultry Sci. 38, 1–27 (2001).CAS 
    Article 

    Google Scholar 
    36.Madonia, C., Hutton, P., Giraudeau, M. & Sepp, T. Carotenoid coloration is related to fat digestion efficiency in a wild bird. Sci. Nat. 104, 96. https://doi.org/10.1007/s00114-017-1516-y (2017).CAS 
    Article 

    Google Scholar 
    37.Hõrak, P. & Männiste, M. Viability selection affects black but not yellow plumage colour in greenfinches. Oecologia 180, 23–32 (2016).ADS 
    Article 

    Google Scholar 
    38.Saks, L., McGraw, K. & Hõrak, P. How feather colour reflects its carotenoid content. Funct. Ecol. 17, 555–561 (2003).Article 

    Google Scholar 
    39.Sepp, T. et al. Coccidian infection causes oxidative damage in greenfinches. PLoS ONE 7 (2012).40.Männiste, M. & Hõrak, P. Emerging infectious disease selects for darker plumage coloration in greenfinches. Front. Ecol. Evol. 2, 4 (2014).Article 

    Google Scholar 
    41.Hackstein, J. H. et al. Parasitic apicomplexans harbor a chlorophyll a-D1 complex, the potential target for therapeutic triazines. Parasitol. Res. 81, 207–216 (1995).CAS 
    PubMed 

    Google Scholar 
    42.Krautwald-Junghanns, M.-E., Zebisch, R. & Schmidt, V. Relevance and treatment of coccidiosis in domestic pigeons (Columba livia forma domestica) with particular emphasis on toltrazuril. Journal of Avian Medicine and Surgery, 1–5 (2009).43.Löfmark, S., Edlund, C. & Nord, C. E. Metronidazole is still the drug of choice for treatment of anaerobic infections. Clin. Infect. Dis. 50, S16–S23. https://doi.org/10.1086/647939 (2010).CAS 
    Article 
    PubMed 

    Google Scholar 
    44.Cramp, S. & Perrins, C. Handbook of the Birds of the Western Palearctic. Volume IV. Terns to Woodpeckers (ed. Cramp, S.), 353–363 (1994).45.Stradi, R., Celentano, G., Rossi, E., Rovati, G. & Pastore, M. Carotenoids in bird plumage—I. The carotenoid pattern in a series of Palearctic Carduelinae. Comp. Biochem. Physiol. Part B: Biochem. Mol. Biol. 110, 131–143 (1995).Article 

    Google Scholar 
    46.Stradi, R. The colour of flight: carotenoids in bird plumages. (Solei Gruppo Editoriale Informatico, 1998).47.McGraw, K., Hill, G., Stradi, R. & Parker, R. The effect of dietary carotenoid access on sexual dichromatism and plumage pigment composition in the American goldfinch. Comp. Biochem. Physiol. B: Biochem. Mol. Biol. 131, 261–269 (2002).CAS 
    Article 

    Google Scholar 
    48.Sepp, T., Karu, U., Sild, E., Männiste, M. & Hõrak, P. Effects of carotenoids, immune activation and immune suppression on the intensity of chronic coccidiosis in greenfinches. Exp. Parasitol. 127, 651–657. https://doi.org/10.1016/j.exppara.2010.12.004 (2011).CAS 
    Article 
    PubMed 

    Google Scholar 
    49.Hõrak, P. et al. Dexamethasone inhibits corticosterone deposition in feathers of greenfinches. Gen. Comp. Endocrinol. 191, 210–214 (2013).Article 

    Google Scholar 
    50.Endler, J. A. On the measurement and classification of colour in studies of animal colour patterns. Biol. J. Lin. Soc. 41, 315–352 (1990).Article 

    Google Scholar 
    51.Lessells, C. & Boag, P. T. Unrepeatable repeatabilities: A common mistake. Auk 104, 116–121 (1987).Article 

    Google Scholar 
    52.Hõrak, P., Saks, L., Karu, U. & Ots, I. Host resistance and parasite virulence in greenfinch coccidiosis. J. Evol. Biol. 19, 277–288 (2006).Article 

    Google Scholar 
    53.Jenni-Eiermann, S. & Jenni, L. Plasma metabolite levels predict individual body-mass changes in a small long-distance migrant, the Garden Warbler. Auk 111, 888–899 (1994).Article 

    Google Scholar 
    54.Saint-Georges-Chaumet, Y. & Edeas, M. Microbiota–mitochondria inter-talk: Consequence for microbiota–host interaction. Pathogens Dis. https://doi.org/10.1093/femspd/ftv096 (2015).Article 

    Google Scholar 
    55.Franco-Obregón, A. & Gilbert, J. A. The microbiome-mitochondrion connection: Common ancestries, common mechanisms, common goals. mSystems https://doi.org/10.1128/mSystems.00018-17 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    56.Paterson, S. The immunology and ecology of co-infection. Mol. Ecol. 22, 2603–2604 (2013).CAS 
    Article 

    Google Scholar 
    57.Quillfeldt, P. et al. Prevalence and genotyping of Trichomonas infections in wild birds in central Germany. PLoS ONE 13, e0200798–e0200798. https://doi.org/10.1371/journal.pone.0200798 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    58.Kinnula, H., Mappes, J. & Sundberg, L.-R. Coinfection outcome in an opportunistic pathogen depends on the inter-strain interactions. BMC Evol. Biol. 17, 77. https://doi.org/10.1186/s12862-017-0922-2 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    59.Gill, H. & Paperna, I. Proliferative visceral Isospora (atoxoplasmosis) with morbid impact on the Israeli sparrow Passer domesticus biblicus Hartert, 1904. Parasitol. Res. 103, 493. https://doi.org/10.1007/s00436-008-0986-4 (2008).Article 
    PubMed 

    Google Scholar 
    60.Shojadoost, B., Vince, A. R. & Prescott, J. F. The successful experimental induction of necrotic enteritis in chickens by Clostridium perfringens: A critical review. Vet. Res. 43, 74. https://doi.org/10.1186/1297-9716-43-74 (2012).CAS 
    Article 
    PubMed 

    Google Scholar 
    61.Williams, R. Intercurrent coccidiosis and necrotic enteritis of chickens: rational, integrated disease management by maintenance of gut integrity. Avian Pathol. 34, 159–180 (2005).CAS 
    Article 

    Google Scholar 
    62.Freeman, C. D., Klutman, N. E. & Lamp, K. C. Metronidazole. Drugs 54, 679–708. https://doi.org/10.2165/00003495-199754050-00003 (1997).CAS 
    Article 
    PubMed 

    Google Scholar 
    63.Hill, G. E. Energetic constraints on expression of carotenoid-based plumage coloration. J. Avian Biol. 31, 559–566 (2000).Article 

    Google Scholar 
    64.Hill, G. E. Cellular respiration: The nexus of stress, condition, and ornamentation. Integr. Comp. Biol. 54, 645–657 (2014).Article 

    Google Scholar 
    65.Ianiro, G., Tilg, H. & Gasbarrini, A. Antibiotics as deep modulators of gut microbiota: Between good and evil. Gut 65, 1906. https://doi.org/10.1136/gutjnl-2016-312297 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    66.Heiss, C. N. & Olofsson, L. E. Gut microbiota-dependent modulation of energy metabolism. J. Innate Immun. 10, 163–171. https://doi.org/10.1159/000481519 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    67.Lind, M.-A., Hõrak, P., Sepp, T. & Meitern, R. Corticosterone levels correlate in wild-grown and lab-grown feathers in greenfinches (Carduelis chloris) and predict behaviour and survival in captivity. Horm. Behav. 118, 104642 (2020).CAS 
    Article 

    Google Scholar 
    68.Sepp, T., Sild, E. & Horak, P. Hematological condition indexes in greenfinches: Effects of captivity and diurnal variation. Physiol. Biochem. Zool. 83, 276–282 (2010).CAS 
    Article 

    Google Scholar  More