More stories

  • in

    The power of community science to quantify ecological interactions in cities

    1.
    United Nations, Department of Economic and Social Affairs, Population Division. World Urban. Prospect. 2018 Highlights (2019).
    2.
    Sebastián-González, E., Dalsgaard, B., Sandel, B. & Guimarães, P. R. Macroecological trends in nestedness and modularity of seed-dispersal networks: Human impact matters. Glob. Ecol. Biogeogr. 24, 293–303 (2015).
    Article  Google Scholar 

    3.
    Eötvös, C. B., Magura, T. & Lövei, G. L. A meta-analysis indicates reduced predation pressure with increasing urbanization. Landsc. Urban Plan. 180, 54–59 (2018).
    Article  Google Scholar 

    4.
    Anders, J. L. et al. Comparison of the intestinal helminth community of the large Japanese field mouse (Apodemus speciosus) between urban, rural, and natural sites in Hokkaido, Japan. Parasitol. Int. 70, 51–57 (2019).
    Article  Google Scholar 

    5.
    Abrams, P. A. & Rowe, L. The effects of predation on the age and size of maturity of prey. Evolution 50, 1052–1061 (1996).
    Article  Google Scholar 

    6.
    Sheriff, M. J., Krebs, C. J. & Boonstra, R. The ghosts of predators past: Population cycles and the role of maternal programming under fluctuating predation risk. Ecology 91, 2983–2994 (2010).
    Article  Google Scholar 

    7.
    Vollset, K. W., Barlaup, B. T., Skoglund, H., Normann, E. S. & Skilbrei, O. Salmon lice increase the age of returning Atlantic salmon. Biol. Lett. 10, 20140085 (2014).
    Article  Google Scholar 

    8.
    Medina, F. M. et al. A global review of the impacts of invasive cats on island endangered vertebrates. Glob. Change Biol. 17, 3503–3510 (2011).
    ADS  Article  Google Scholar 

    9.
    Devictor, V., Whittaker, R. J. & Beltrame, C. Beyond scarcity: Citizen science programmes as useful tools for conservation biogeography. Divers. Distrib. 16, 354–362 (2010).
    Article  Google Scholar 

    10.
    Hurlbert, A. H. & Liang, Z. Spatiotemporal variation in avian migration phenology: Citizen science reveals effects of climate change. PLoS ONE 7, e31662 (2012).
    ADS  CAS  Article  Google Scholar 

    11.
    Li, E. et al. An urban biodiversity assessment framework that combines an urban habitat classification scheme and citizen science data. Front. Ecol. Evol. 7, 277 (2019).
    ADS  CAS  Article  Google Scholar 

    12.
    Berger, J. Fear, human shields and the redistribution of prey and predators in protected areas. Biol. Lett. 3, 620–623 (2007).
    Article  Google Scholar 

    13.
    Geffroy, B., Samia, D. S. M., Bessa, E. & Blumstein, D. T. How nature-based tourism might increase prey vulnerability to predators. Trends Ecol. Evol. 30, 755–765 (2015).
    Article  Google Scholar 

    14.
    Loss, S. R., Will, T. & Marra, P. P. The impact of free-ranging domestic cats on wildlife of the United States. Nat. Commun. 4, 1396 (2013).
    ADS  Article  Google Scholar 

    15.
    Tablado, Z. & Jenni, L. Determinants of uncertainty in wildlife responses to human disturbance. Biol. Rev. https://doi.org/10.1111/brv.12224 (2015).
    Article  PubMed  Google Scholar 

    16.
    Becker, M. & Buchholz, S. The sand lizard moves downtown—habitat analogues for an endangered species in a metropolitan area. Urban Ecosyst. 19, 361–372 (2016).
    Article  Google Scholar 

    17.
    Chavez-Zichinelli, C. A. et al. How stressed are birds in an urbanizing landscape? Relationships between the physiology of birds and three levels of habitat alteration. Condor 115, 84–92 (2013).
    Article  Google Scholar 

    18.
    Jiménez-Peñuela, J., Ferraguti, M., Martínez-de la Puente, J., Soriguer, R. & Figuerola, J. Urbanization and blood parasite infections affect the body condition of wild birds. Sci. Total Environ. 651, 3015–3022 (2019).
    ADS  Article  Google Scholar 

    19.
    Birnie-Gauvin, K., Peiman, K. S., Gallagher, A. J., de Bruijn, R. & Cooke, S. J. Sublethal consequences of urban life for wild vertebrates. Environ. Rev. 24, 416–425 (2016).
    Article  Google Scholar 

    20.
    Møller, A. P. Successful city dwellers: A comparative study of the ecological characteristics of urban birds in the western Palearctic. Oecologia 159, 849–858 (2009).
    ADS  Article  Google Scholar 

    21.
    Bradley, C. A. & Altizer, S. Urbanization and the ecology of wildlife diseases. Trends Ecol. Evol. 22, 95–102 (2007).
    Article  Google Scholar 

    22.
    Lazic, M. M., Carretero, M. A., Zivkovic, U. & Crnobrnja-Isailovic, J. City life has fitness costs: Lower body condition and increased parasite intensity in urban common wall lizards Podarcis muralis. Salamandra 53, 10–17 (2016).
    Google Scholar 

    23.
    Giraudeau, M., Mousel, M., Earl, S. & Mcgraw, K. Parasites in the city: Degree of urbanization predicts poxvirus and coccidian infections in house finches (Haemorhous mexicanus). PLoS ONE 9, e86747 (2014).
    ADS  Article  Google Scholar 

    24.
    Kellner, K. F., Page, L. K., Downey, M. & Mccord, S. E. Effects of urbanization on prevalence of Baylisascaris procyonis in intermediate host populations. J. Wildl. Dis. 48, 1083–1087 (2012).
    Article  Google Scholar 

    25.
    Geue, D. & Partecke, J. Reduced parasite infestation in urban Eurasian blackbirds (Turdus merula): A factor favoring urbanization?. Can. J. Zool. 86, 1419–1425 (2008).
    Article  Google Scholar 

    26.
    Arnold, E. N. Caudal autotomy as a defense. In Biology of the Reptilia (eds. Gans, C. & Huey, R. B.) 235–273 (Liss, 1988).

    27.
    Higham, T. E., Russell, A. P. & Zani, P. A. Integrative biology of tail autotomy in lizards. Physiol. Biochem. Zool. 86, 603–610 (2013).
    Article  Google Scholar 

    28.
    Itescu, Y., Schwarz, R., Meiri, S. & Pafilis, P. Intraspecific competition, not predation, drives lizard tail loss on islands. J. Anim. Ecol. 86, 66–74 (2017).
    Article  Google Scholar 

    29.
    Bowker, R. W. A Comparative Behavioral Study and Taxonomic Analyses of Gerrhonotine Lizards (Arizona State University, Tempe, 1988).
    Google Scholar 

    30.
    Bateman, P. W. & Fleming, P. A. To cut a long tail short: A review of lizard caudal autotomy studies carried out over the last 20 years. J. Zool. 277, 1–14 (2009).
    Article  Google Scholar 

    31.
    Tyler, R. K., Winchell, K. M. & Revell, L. J. Tails of the city: Caudal autotomy in the tropical lizard, Anolis cristatellus, in urban and natural areas of Puerto Rico. J. Herpetol. 50, 435–441 (2016).
    Article  Google Scholar 

    32.
    Arnold, E. N. Evolutionary aspects of tail shedding in lizards and their relatives. J. Nat. Hist. 18, 127–169 (1984).
    Article  Google Scholar 

    33.
    Bellairs, A. D. A. & Bryant, S. V. Autotomy and regeneration in reptiles. In Biology of the Reptilia (eds. Gans, C. & Billet, F.) 301–410 (Wiley, Hoboken, 1985).

    34.
    Chapple, D. G. & Swain, R. Inter-populational variation in the cost of autotomy in the metallic skink (Niveoscincus metallicus). J. Zool. 264, 411–418 (2004).
    Article  Google Scholar 

    35.
    Wright, S. A., Lane, R. S. & Clover, J. R. Infestation of the Southern Alligator Lizard (Squamata: Anguidae) by Ixodes pacificus (Acari: Ixodidae) and its susceptibility to Borrelia burgdorferi. J. Med. Entomol. 35, 1044–1049 (1998).
    CAS  Article  Google Scholar 

    36.
    Tanner, D. & Perry, J. Road effects on abundance and fitness of Galápagos lava lizards (Microlophus albemarlensis). J. Environ. Manag. 85, 270–278 (2007).
    Article  Google Scholar 

    37.
    Main, A. R. & Bull, M. The impact of tick parasites on the behaviour of the lizard Tiliqua rugosa. Oecologia 122, 574–581 (2000).
    ADS  CAS  Article  Google Scholar 

    38.
    Stanley, T. R. et al. Changes in capture rates and body size among vertebrate species occupying an insular urban habitat reserve. Conserv. Sci. Pract. 2, e245 (2020).
    Google Scholar 

    39.
    Pellitteri-Rosa, D. et al. Urbanization affects refuge use and habituation to predators in a polymorphic lizard. Anim. Behav. 123, 359–367 (2017).
    Article  Google Scholar 

    40.
    Rebolo-Ifrán, N. et al. Links between fear of humans, stress and survival support a non-random distribution of birds among urban and rural habitats. Sci. Rep. 5, 13723 (2015).
    ADS  Article  Google Scholar  More

  • in

    Assessing spatio-temporal patterns and driving force of ecosystem service value in the main urban area of Guangzhou

    This paper uses Landsat TM/OLI images from 1987, 1993, 1999, 2005, 2011 and 2017 by the weight vector AdaBoost (WV AdaBoost) multi-classification algorithm extracting LULC data sets, and the spatiotemporal patterns of LULC over these periods were analysed. The spatiotemporal change patterns and driving force of ESV was estimated. The effect of LULC dynamics on the ESV was evaluated. The flow chart is shown in Fig. 1.
    Figure 1

    The framework of this paper.

    Full size image

    Study area
    Guangzhou is located at 112° 57′ ~ 114° 3′ E, 22° 26′ ~ 23° 56′ N in the southeast part of Guangdong Province in the northern margin of the Pearl River Delta. It has a total land area of 7434.40 km2. The topography is high in the northeast and low in the southwest, with low mountains and hills in the north and plains in the south and a rich geomorphology. The level of urbanization is high, the land use structure is complex and the land use pattern is changing rapidly. Regional public infrastructure, commercial service land and external transportation land is increasing. Export-oriented industrial agglomeration areas are developing continuously, and the characteristics of export-oriented land use are obvious. Guangzhou has a subtropical oceanic monsoon climate, with an annual average temperature of 20–22 °C and an annual rainfall of about 1720 mm. Guangzhou includes eleven districts (Fig. 2a). This paper focuses on the main urban area of Guangzhou as the research object, a decision based on the city government’s overall urban development strategy for Guangzhou (2010–2020), the main urban area of Guangzhou includes five districts—Liwan, Yuexiu, Haizhu, Baiyun and Tianhe (Fig. 2a). In 2017, the per capita GDP will exceed 136,100 yuan for the first time (Fig. 2b). The population density is large, and the intensity of development is extraordinary. The population has increased from 2.73 million in 1987 to 8.05 million in 2017 (Fig. 2c). The industrial structure has been constantly optimized and improved. The proportion of secondary industries decreased from 31.46% in 1987 to 11.28% in 2017, while the proportion of tertiary industries increased from 61.9% in 1987 to 89.61% in 2017.
    Figure 2

    Location of study area and GDP and population. (a) location, (b) GDP, (c) population. (Software: Arc Map 10.5.0, http://www.esri.com. OriginPro 2017C SR2, https://www.originlab.com/).

    Full size image

    Data and data processing
    Data
    The data used include Landsat series images, an administrative zoning map of Guangzhou, a local historical land use map and social and economic statistics from 1987 to 2017. To explore the socio-economic and natural factors driving the ESV change for different land use types, we examined ten factors—the normalized vegetation index (NDVI), year-end population, elevation, slope, distance from roads, distance from railways, land cover types, GDP, secondary industry GDP and investment in fixed assets.
    The Landsat images were taken from the United States Geological Survey website (http://glovis.usgs.gov/), and we downloaded cloud-free Landsat 5 TM and Landsat 8 OLI images (path 122, row 44). Images taken in the dry season (October, November and December) were selected because there was less cloud cover, the change in surface reflectance was much smaller than in other seasons and the image quality was higher. The spatial resolution of imaging is 30 m, and the data of Landsat 5 TM (8 December 1987, 24 December 1993, 25 December 1999, 23 November 2005, 19 November 2011) and Landsat 8 OLI (23 October 2017) were collected.
    The elevation data are provided by the Japan Aerospace Exploration Agency (http://www.eorc.jaxa.jp/ALOS/en/aw3d30/data/index.htm); the horizontal resolution is 30 m, and the elevation accuracy is 5 m. The road and railway data are provided by Openstreetmap (download.geofabrik.de/), and the distance between roads and railways were calculated. The land use types are provided by Tsinghua University (http://data.ess.tsinghua.edu.cn/) with a resolution of 10 m. GDP, secondary industry GDP and investment in fixed assets were provided by Guangzhou Statistics Bureau28.
    Considering the size of the study area and the resolution of the data, we used Fishnet in ArcGIS10.5 to establish a 500 × 500 m grid covering the study area. The values of ten factors were extracted to the grid centre points, and 3915 points were obtained.
    Remote sensing extraction of LULC types
    To obtain high-precision LULC data, the surface features based on the original remote sensing image were enhanced using the index model (the Index-based Built-up Index (IBI)29, the modified normalized difference water index (MNDWI)30 and the soil-adjusted vegetation index (SAVI)). The humidity, brightness and green chroma indices were transformed using the tasselled-cap method23. Using the six indices, six images were calculated and superposed into a new multi-band image in the order of IBI, SAVI, MNDWI, brightness, green chroma and humidity (referred to as the six-index image). This was superposed with the original image (band 1, 2, 3, 4, 5, 7) to create an enhanced 12-band image as the data source for LULC classification. The image stretching function was then used to stretch the composite image to better distinguish the characteristics of land use types. After feature enhancement, the LULC types in the main urban area of Guangzhou could be divided into seven types—forest, water body, grassland, cultivated land, high reflectivity building, low reflectivity building and bare land. Of these, high reflectance buildings and low reflectance buildings are sub-categories of the built-up area category. The spectrum of high reflectivity buildings is similar to that of bare land; therefore, they are classified separately to avoid erroneous classification31. Based on Google Earth and TM/OLI images, the interpretation marks of LULC types were established (Table 1). To obtain pure and representative samples, the K-means method of unsupervised classification was utilized to cluster the pixels of the ISM image into seven unlabelled classes and to eliminate abnormal clustering. We then selected uniform points and drew polygons on them. Each polygon contains more than 100 pixels, and each class contains between five and seven polygons. All polygons were saved in a shape file as a mask file, which is utilized to extract pure pixels as training samples.
    Table 1 Interpretation mark of land cover types.
    Full size table

    Dou and Chen31 suggest using the weight vector AdaBoost (WV AdaBoost) multi-classification algorithm for LULC classification. Compared with AdaBoost, it provides higher classification accuracy and stability. WV AdaBoost includes a C4.5 decision tree, a Naïve Bayes neural network and an artificial neural network. In this study, we use Naïve Bayes-based WV AdaBoost to classify LULC based on Landsat-enhanced images. To obtain a highly accurate LULC classification, it is necessary to post-process the image after classification, overlay the early land use map during the research period, check the incorrectly classified area, and filter and correct some segments. After image classification, the random selection of 200 pixels in each class was checked, and the correctness of the classification and the evaluated accuracy were confirmed.
    The average classification accuracy of WV AdaBoost based on Naïve Bayes on the original image is 85.02%, and the kappa coefficient is 0.839. The WV AdaBoost algorithm is based on Naïve Bayes processes for the 12-band images of the new combination (enhanced 12-band image). The average classification accuracy is 88.86%, and the kappa coefficient is 0.871. The classification results still need to be strengthened by post-processing, which achieves good classification accuracy. The final average classification accuracy is 91.97%. The kappa coefficient is 0.907. These classification results agree with the ensuing analysis of LULC.
    Methods
    Based on Landsat TM/OLI data from 1987 to 2017, a transfer matrix, a land use change index, an ESV evaluation index, a sensitivity model (CS) and a geo-detector (P) were used to analyze the response of ESVs to LULC evolution.
    Transfer matrix
    The transfer matrix reflects the dynamic process information about mutual transformation LULC types at the beginning (T) and the end (T + 1) of a specified period of time in a certain region (Fig. 3). The general form is:

    $${S}_{ij}=left[begin{array}{c}{s}_{11} {s}_{12}dots {s}_{1n}\ {s}_{21} {s}_{22}dots {s}_{2n}\ dots dots dots dots \ { s}_{n1} {s}_{n2}dots {s}_{nn}end{array}right]$$
    (1)

    where S stands for area, and i,j (i,j = 1,2,…, n) represents LULC types before and after the transfer.
    Figure 3

    Land use transfer process. T the beginning of land use types, T + 1 the end of land use types; A, B, C, D, E, F: different land use types.

    Full size image

    The LULC analyzing indices
    (1)
    Land use dynamics (RS)
    Land use dynamics describe the rate and magnitude of LUCC. The general form is:

    $$RS=frac{{U}_{b}-{U}_{a}}{{U}_{a}}times frac{1}{T}times 100mathrm{%},$$
    (2)

    where ({U}_{a}) is the area of a certain land class at the beginning (km2), ({U}_{b}) is the area of a certain land class at the end (km2), and T is the study period.

    (2)
    Spatial dynamics of land use (RSS)
    Spatial dynamics of land use describe the degree of spatial change in a certain land use type. The general form is:

    $$RSS=frac{{U}_{in}-{U}_{out}}{{U}_{a}}times frac{1}{T}times 100mathrm{%},$$
    (3)

    where ({U}_{in}) is the sum area of other types converted to this type in study period T, and ({U}_{out}) is the sum area of a certain type converted to other types in study period T. ({U}_{a}) is the area of a certain type at the beginning.

    (3)
    Land use change state index (PS)

    The land use change state index represents the trend and state of LUCC. The general form is:

    $$PS=frac{{U}_{in}-{U}_{out}}{{U}_{in}+{U}_{out}},$$
    (4)

    where ({U}_{in}) is the sum area of other types converted to this type in study period T, and ({U}_{out}) is the sum area of a certain type converted to other types in study period T.
    Calculation of ecosystem service value
    In this paper, the improved ecosystem services valuation method based on the unit area value equivalent factor proposed by Xie et al.2 is employed to evaluate ESV. Therefore, the LULC types of the main urban area of Guangzhou were associated to the corresponding representative biomes (Table 2). The most representative biomes used as a proxy for each LULC type are: (1) cropland for cultivated land, (2) tropical forest for forest, (3) grassland for grassland, and (4) water system and wetland for water body. (5) bare land for bare land.
    Table 2 Parameters of ESV for different land use types in the main urban area of Guangzhou.
    Full size table

    The LULC types are not exactly identical with the representative biomes. For example, cultivated land in this study accounts areas used for paddy fields and dry land. Therefore, the ESV index of cultivated land is calculated by weighting the area ratio of paddy field and dry land in the statistical yearbook. The average value of the ESV index of the water system and wetland is adopted for water body. The ESV index of build-up area is 0. the ESV index of land use types is shown in Table 2. The value of the universal equivalent factor ESV (D value) of the improved ESV method is 340.65 thousand yuan/km2.
    In this study, the ESV of each land use unit area is based on the research methods of Costanza3 and Xie et al.2. The calculation formula is:

    $$ESV=sum {A}_{i}times {VC}_{i}$$
    (5)

    $${ESV}_{f}=sum {A}_{i}times {VC}_{fi},$$
    (6)

    where ESV is the total value (yuan) of ecosystem services, ({A}_{i}) (km2) is the area of class i, and VCi is the ESV coefficient (yuan/km· year) corresponding to class i. ESVf is the single ESV, and VCfi is the value coefficient of the single service function.
    Sensitivity analysis model
    The sensitivity model is used to calculate the response of ESV to the change of value coefficient (VC)14 by adjusting the 50% of the ESV coefficient of each land use type up and down and determining the change in ESV over time and the degree of dependence on the value coefficient. The calculation formula is as follows:

    $$CS=frac{{(ESV}_{j}-{ESV}_{i})/{ESV}_{i}}{{(VC}_{jk}-{VC}_{ik})/{VC}_{ik}},$$
    (7)

    where ESV is the estimated ESV, VC is the value coefficient, i and j are the initial value and adjusted value (50% up and down adjustment) and K is the land use type. When CS ≥ 1, ESV is elastic relative to VC; when CS ≤ 1, ESV is inelastic. The larger the CS value is, the more critical the accuracy of the ESV index.
    Grid analysis method
    Grid analysis method is used to divided the study area into regular grid matrixes with the same size and no overlap, and takes grid as the research object to express and statistical unit in geospatial space9. It uses regular square grid as ESV’s spatial statistical unit instead of irregular land-use map spots to ensure the capacity invariance within the unit, which not only highlights the spatial distribution characteristics of ESV, but also facilitates the spatial quantitative statistics of ESV.
    The premise of calculating ESV spatial differentiation is to determine the size of grid cell. In this study, based on ArcGIS10.5 software, the area of each land use type in four grid units with side length of 0.5 km, 1 km, 2 km and 3 km was extracted respectively, and then compare the degree of area change, namely the coefficient of variation. The grid of this scale is the optimal grid unit size for the spatial differentiation of ESV in the study area. Finally, the grid analysis method is introduced to construct a (0.5 × 0.5) km square grid as a geospatial statistical unit. By using the equal spacing system sampling method, the study area is divided into 2024 square grids that do not overlap each other (0.5 × 0.5) km, and the grid matrix is composed of these grids. Through the intersection operation of grid matrix and land use data of each research year, the area of various land use types in each grid is counted, and the spatial heterogeneity of land use types and ESV is analyzed.
    Geo-detector
    Combined with GIS spatial superposition technology and set theory, a statistical method proposed by Wang et al.32 can detect spatial heterogeneity and reveal the driving force to identify the interaction between multiple factors. This model is widely used to analyze the influence mechanism of social economic factors and natural environmental factors. The geo-detector consists of four detectors—a differentiation and factor detector, an interaction detector, a risk area detector and an ecological detector. In this paper, factor detection and interaction detection are utilized to detect and analyze the driving force of ESVs in the main urban area of Guangzhou.
    (1) Factor detector: Factor detection can identify the explanatory power of each spatial driving factor in landscape type change, and its model is as follows:

    $$P=1-frac{1}{{ndelta }^{2}}sum_{i=1}^{m}{n}_{i}{{delta }^{2}}_{i},$$
    (8)

    where P is the explanatory power index of ESV influencing factors; ni is the number of samples in the secondary area; n is the total number of samples; m is the number of samples in the secondary area; ({delta }^{2}) is the variance in land use type change in the whole area; and ({{delta }^{2}}_{i}) is the variance in land use type in the secondary area. Thus, the model is established, assuming ({{delta }^{2}}_{i}) ≠ 0.
    The range of values for P is [0,1]. When P = 0, it shows that the spatial distribution of ESV changes is random. The larger the P value is, the greater the impact of longitudinal driving factors on ESV changes.
    (2) Interactive detector: Interaction detection can be used to identify the interaction between different spatial drivers. When detecting the interaction of X1 and X2, the explanatory power of the dependent variable Y will increase or decrease. The evaluation method is used to calculate the q value of two factors, X1 and X2, for Y, respectively, q (X1) and q (X2); to calculate the q value of their interaction, q (X1 ∩ X2); and to compare q (X1), q (X2) and q (X1 ∩ X2). The five results of the interactive detector are given in Table 3.
    Table 3 Types of interactions between two covariates.
    Full size table More

  • in

    Fecundity determines the outcome of founding queen associations in ants

    In this study, we used the black garden ant Lasius niger to investigate the benefits and factors of pleometrosis, the transitory association between founding queens. The monitoring of colonies founded by one or two queens showed that pleometrosis increased and accelerated offspring production. Then, the experimental pairing of L. niger founding queens revealed that in pairs of queens of different fecundity but similar size, the most fecund queen was more likely to survive. Our experiment could not detect a similar effect of size when controlling for fecundity. Finally, we found that queens associated preferentially with less fecund queens.
    Our findings of pleometrosis benefiting offspring production are in line with the literature for this, and other ant species3,7,9,10,12,22,23. Interestingly, we only detected these benefits at the colony level, as pleometrosis had either no effect or a negative influence on the per capita offspring production9,12,22. However, colony-level measurements are more relevant in the case of pleometrosis, as the queen that survives the association inherits all the offspring produced during colony foundation. In the field, colonies with a faster, more efficient worker production would have a competitive advantage over neighbouring founding colonies3,4. This is especially true for L. niger, which shows high density of founding colonies that compete for limiting resources and raid the brood of other colonies10. Thus, the competitive advantage provided by pleometrosis likely enhances colony growth and survival.
    The increased and faster production of workers in colonies with two queens may stem from a nutritional boost for the larvae. L. niger founding queens do not forage, and produce the first cohort of workers from their own metabolic reserves. Larvae have been observed to cannibalize both viable and non-viable (trophic) eggs24. We found that colonies with two queens produced more eggs, but that this did not translate in them having more larvae. However, more of these larvae became pupae—and ultimately workers. In addition, while the time to produce the first egg and larva did not differ between colonies with one and two queens, the first pupa and worker were produced faster when two queens were present, consistent with a shorter larval stage. We propose that larvae in pleometrotic colonies developed faster and were more likely to reach pupation because they had more eggs that provided nutrients, boosting the development rate of the first workers.
    These benefits of pleometrosis are only inherited by the queens that survive, it is thus important to understand the factors that determine queen survival in pleometrotic associations. Although this question has been relatively well studied3,16,17,18,19,20,21, it has remained challenging to disentangle the effects of correlated factors. For example, we found that size, which has been reported to predict queen survival16,19, correlated with fecundity, which would itself be confounded with the parentage of workers in the first cohort produced. To address this issue, we disentangled size and fecundity experimentally, and used foreign workers that developed from pupae collected in field colonies to prevent any potential nepotistic behaviour.
    We found that fecundity, but not size, determined queen survival. The finding that, despite being of similar size, more fecund queens are more likely to survive indicates that the outcome of pleometrosis is not the mere consequence of physical dominance. The higher fecundity could reflect a better health condition, which may give the advantage to the more fecund queen in direct fights3,15, or if workers initiate the fights. Natural selection may have favoured workers that skew aggression toward the less fecund queen, both because this queen would be less efficient at building a colony, and because the workers would be more likely to be the offspring of the more fecund queen. The latter would not necessarily involve direct nepotistic behaviours (the workers would not behave according to parentage, but to fecundity), which have remained elusive in social insects in general25,26,27, and in pleometrotic associations in particular16,17,20. Despite regular behavioural observations, we did not observe who initiated aggression in our experiments, and it remains unclear whether the queens and/or the workers are responsible for the onset of fights. Consistently with previous studies16,23, we found that a certain proportion of queen death occurred before worker emergence, suggesting that worker presence is not required for queen execution. Finally, we cannot rule out that the least fecund queens were more likely to die because of a weaker health status, possibly combined with the stress of being associated with another, healthier queen.
    Although it has not been directly reported before, our finding that fecundity determines queen survival is consistent with previous reports of weight being associated with queen survival17, more fecund queens being more aggressive28, cuticular hydrocarbon profiles differing between surviving and culled queens21, and between more and less fecund queens28. We could not directly support previous reports of size correlating with survival16,19. This could be because in those studies, size could have been confounded with fecundity, and/or because we lacked the statistical power to detect such effect in our experiment.
    Pleometrosis provides clear benefits, but these benefits are only inherited by the surviving queens, and the losing queens pay the great cost of dying without contributing to the next generation. Natural selection should thus favour queens that decide whether or not to join a pleometrotic association based on the relative benefits compared to individual foundation—these may differ across ecological contexts29—and the likelihood of surviving the association. As fecundity appears to determine queen survival in L. niger, queens may have evolved the ability to choose among potential partners according to their fecundity. Our results are consistent with this hypothesis, as queens preferentially associated with partners that would later produce fewer eggs, possibly because they were less fecund, and therefore less healthy and easier to eliminate. This suggests that founding queens may assess the fecundity of potential partners, possibly via their cuticular hydrocarbon profile28. This result further supports our finding that fecundity plays an important role in pleometrotic associations. It is important to note that this difference in egg production could have alternative explanations. First, it could stem from more fecund queens having no interest in forming an association because they are able to start a competitive colony alone. Second, it could be a consequence, rather than a cause, of the outcome of the choice experiment. We cannot rule out that entering an association with another queen and/or leaving this association prematurely at the end of the choice experiment may have been stressful for the chosen queens, and affected their later production of eggs. We could not detect any difference between chosen and not chosen queens in the number of larvae and pupae produced, which are likely influenced by factors other than fecundity (e.g., brood care behaviour). Interestingly, we did not find that queens chose according to size, consistent with our finding that size may not affect which queen survives the pleometrotic association.
    Our study informs on the benefits and factors of pleometrosis, and highlights the role of fecundity in the decision to associate with another queen, and in determining which queen survives the association. As such, it contributes to a better understanding of the onset and outcome of pleometrosis, a classic case of cooperation between unrelated animals. More

  • in

    Composition and acquisition of the microbiome in solitary, ground-nesting alkali bees

    1.
    Dharampal, P. S., Hetherington, M. C. & Steffan, S. A. Microbes make the meal: Oligolectic bees require microbes within their host pollen to thrive. Ecol. Entomol. https://doi.org/10.1111/een.12926 (2020).
    Article  Google Scholar 
    2.
    Sonnenburg, J. L. & Bäckhed, F. Diet-microbiota interactions as moderators of human metabolism. Nature 535, 56–64 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    3.
    Suzuki, T. A. Links between natural variation in the microbiome and host fitness in wild mammals. Integr. Comp. Biol. 57, 756–769 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Zheng, D., Liwinski, T. & Elinav, E. Interaction between microbiota and immunity in health and disease. Cell Res. 30, 492–506 (2020).
    PubMed  PubMed Central  Article  Google Scholar 

    5.
    Kwong, W. K., Mancenido, A. L. & Moran, N. A. Immune system stimulation by the native gut microbiota of honey bees. R. Soc. Open Sci. 4, 170003 (2017).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    6.
    Bo, T.-B. et al. Coprophagy prevention alters microbiome, metabolism, neurochemistry, and cognitive behavior in a small mammal. ISME J. https://doi.org/10.1038/s41396-020-0711-6 (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    7.
    Sarkar, A. et al. The role of the microbiome in the neurobiology of social behaviour. Biol. Rev. 95, 12603 (2020).
    Article  Google Scholar 

    8.
    Vernier, C. L. et al. The gut microbiome defines social group membership in honey bee colonies. Sci. Adv. 6, 3431 (2020).
    ADS  Article  CAS  Google Scholar 

    9.
    Lemoine, M. M., Engl, T. & Kaltenpoth, M. Microbial symbionts expanding or constraining abiotic niche space in insects. Curr. Opin. Insect Sci. 39, 14–20 (2020).
    PubMed  Article  Google Scholar 

    10.
    Engel, P. et al. The bee microbiome: Impact on bee health and model for evolution and ecology of host-microbe interactions. MBio. https://doi.org/10.1128/mBio.02164-15 (2006).
    Article  Google Scholar 

    11.
    Daisley, B. A., Chmiel, J. A., Pitek, A. P., Thompson, G. J. & Reid, G. Missing microbes in bees: How systematic depletion of key symbionts erodes immunity. Trends Microbiol. https://doi.org/10.1016/j.tim.2020.06.006 (2020).
    Article  PubMed  Google Scholar 

    12.
    Bonilla-Rosso, G. & Engel, P. Functional roles and metabolic niches in the honey bee gut microbiota. Curr. Opin. Microbiol. 43, 69–76 (2018).
    CAS  PubMed  Article  Google Scholar 

    13.
    Zheng, H., Powell, J. E., Steele, M. I., Dietrich, C. & Moran, N. A. Honeybee gut microbiota promotes host weight gain via bacterial metabolism and hormonal signaling. Proc. Natl. Acad. Sci. 114, 4775–4780 (2017).
    CAS  PubMed  Article  Google Scholar 

    14.
    Zheng, H. et al. Metabolism of toxic sugars by strains of the bee gut symbiont Gilliamella apicola. MBio. https://doi.org/10.1128/mBio.01326-16 (2016).
    Article  PubMed  PubMed Central  Google Scholar 

    15.
    Engel, P. & Moran, N. A. Functional and evolutionary insights into the simple yet specific gut microbiota of the honey bee from metagenomic analysis. Gut Microbes 4, 60–65 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    16.
    Lee, F. J., Rusch, D. B., Stewart, F. J., Mattila, H. R. & Newton, I. L. G. Saccharide breakdown and fermentation by the honey bee gut microbiome. Environ. Microbiol. 17, 796–815 (2015).
    CAS  PubMed  Article  Google Scholar 

    17.
    Anderson, K. E. et al. Hive-stored pollen of honey bees: Many lines of evidence are consistent with pollen preservation, not nutrient conversion. Mol. Ecol. 23, 5904–5917 (2014).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    18.
    Dharampal, P. S., Carlson, C., Currie, C. R. & Steffan, S. A. Pollen-borne microbes shape bee fitness. Proc. R. Soc. B Biol. Sci. 286, 20182894 (2019).
    CAS  Article  Google Scholar 

    19.
    Rothman, J. A., Leger, L., Graystock, P., Russell, K. & McFrederick, Q. S. The bumble bee microbiome increases survival of bees exposed to selenate toxicity. Environ. Microbiol. 21, 1462–2920. https://doi.org/10.1111/1462-2920.14641 (2019).
    CAS  Article  Google Scholar 

    20.
    Wu, Y. et al. Honey bee ( Apis mellifera ) gut microbiota promotes host endogenous detoxification capability via regulation of P450 gene expression in the digestive tract. Microb. Biotechnol. 13, 1201–1212 (2020).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    21.
    Praet, J. et al. Large-scale cultivation of the bumblebee gut microbiota reveals an underestimated bacterial species diversity capable of pathogen inhibition. Environ. Microbiol. 20, 214–227 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Forsgren, E., Olofsson, T. C., Vásquez, A. & Fries, I. Novel lactic acid bacteria inhibiting Paenibacillus larvae in honey bee larvae. Apidologie 41, 99–108 (2010).
    Article  Google Scholar 

    23.
    Cariveau, D. P., Elijah Powell, J., Koch, H., Winfree, R. & Moran, N. A. Variation in gut microbial communities and its association with pathogen infection in wild bumble bees (Bombus). ISME J. 8, 2369–2379 (2014).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    24.
    Raymann, K., Shaffer, Z. & Moran, N. A. Antibiotic exposure perturbs the gut microbiota and elevates mortality in honeybees. PLoS Biol. 15, e2001861 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    25.
    Schwarz, R. S., Moran, N. A. & Evans, J. D. Early gut colonizers shape parasite susceptibility and microbiota composition in honey bee workers. Proc. Natl. Acad. Sci. 113, 9345–9350 (2016).
    CAS  PubMed  Article  Google Scholar 

    26.
    Maes, P. W., Rodrigues, A. P., Oliver, R., Mott, B. M. & Anderson, K. E. Diet related gut bacterial dysbiosis correlates with impaired development, increased mortality and Nosema disease in the honey bee (Apis mellifera). Mol. Ecol. 25, 5439–5450 (2016).
    CAS  PubMed  Article  Google Scholar 

    27.
    Koch, H. & Schmid-Hempel, P. Socially transmitted gut microbiota protect bumble bees against an intestinal parasite. Proc. Natl. Acad. Sci. 108, 19288–19292 (2011).
    ADS  CAS  PubMed  Article  Google Scholar 

    28.
    Evans, J. D. & Lopez, D. L. Bacterial probiotics induce an immune response in the honey bee (Hymenoptera: Apidae). J. Econ. Entomol. 97, 752–756 (2004).
    CAS  PubMed  Article  Google Scholar 

    29.
    Emery, O., Schmidt, K. & Engel, P. Immune system stimulation by the gut symbiont Frischella perrara in the honey bee (Apis mellifera). Mol. Ecol. 26, 2576–2590 (2017).
    CAS  PubMed  Article  Google Scholar 

    30.
    Engel, P., Martinson, V. G. & Moran, N. A. Functional diversity within the simple gut microbiota of the honey bee. Proc. Natl. Acad. Sci. 109, 11002–11007 (2012).
    ADS  CAS  PubMed  Article  Google Scholar 

    31.
    Kwong, W. K. & Moran, N. A. Gut microbial communities of social bees. Nat. Rev. Micro 14, 374–384 (2016).
    CAS  Article  Google Scholar 

    32.
    McFrederick, Q. S. & Rehan, S. M. Characterization of pollen and bacterial community composition in brood provisions of a small carpenter bee. Mol. Ecol. 25, 2302–2311 (2016).
    CAS  PubMed  Article  Google Scholar 

    33.
    McFrederick, Q. S. et al. Flowers and wild megachilid bees share microbes. Microb. Ecol. 73, 188–200 (2017).
    PubMed  Article  Google Scholar 

    34.
    McFrederick, Q. S. et al. Environment or kin: whence do bees obtain acidophilic bacteria?. Mol. Ecol. 21, 1754–1768 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    35.
    McFrederick, Q. S., Wcislo, W. T., Hout, M. C. & Mueller, U. G. Host species and developmental stage, but not host social structure, affects bacterial community structure in socially polymorphic bees. FEMS Microbiol. Ecol. 88, 398–406 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    36.
    Graystock, P., Rehan, S. M. & McFrederick, Q. S. Hunting for healthy microbiomes: Determining the core microbiomes of Ceratina, Megalopta, and Apis bees and how they associate with microbes in bee collected pollen. Conserv. Genet. 18, 701–711 (2017).
    Article  Google Scholar 

    37.
    McFrederick, Q. S. et al. Specificity between lactobacilli and hymenopteran hosts is the exception rather than the rule. Appl. Environ. Microbiol. 79, 1803–1812 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    38.
    Sanders, J. G. et al. Stability and phylogenetic correlation in gut microbiota: Lessons from ants and apes. Mol. Ecol. 23, 1268–1283 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    39.
    Kwong, W. K. et al. Dynamic microbiome evolution in social bees. Sci. Adv. 3, 1–17 (2017).
    Article  Google Scholar 

    40.
    Rothman, J. A., Andrikopoulos, C., Cox-Foster, D. & McFrederick, Q. S. Floral and foliar source affect the bee nest microbial community. Microb. Ecol. 78, 506–516 (2019).
    PubMed  Article  Google Scholar 

    41.
    Cohen, H., McFrederick, Q. S. & Philpott, S. M. Environment shapes the microbiome of the blue orchard bee, Osmia lignaria. Microb. Ecol. 80, 897–907 (2020).
    PubMed  Article  PubMed Central  Google Scholar 

    42.
    Muñoz-Colmenero, M. et al. Differences in honey bee bacterial diversity and composition in agricultural and pristine environments—A field study. Apidologie. https://doi.org/10.1007/s13592-020-00779-w (2020).
    Article  Google Scholar 

    43.
    Kapheim, K. M. et al. Caste-specific differences in hindgut microbial communities of honey bees (Apis mellifera). PLoS ONE 10, 1–14 (2015).
    Article  CAS  Google Scholar 

    44.
    Elijah Powell, J., Eiri, D., Moran, N. A. & Rangel, J. Modulation of the honey bee queen microbiota: Effects of early social contact. PLoS ONE 13, 1–14 (2018).
    Google Scholar 

    45.
    Tarpy, D. R., Mattila, H. R. & Newton, I. L. G. Development of the honey bee gut microbiome throughout the queen-rearing process. Appl. Environ. Microbiol. 81, 3182–3191 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    46.
    Dong, Z. X. et al. Colonization of the gut microbiota of honey bee (Apis mellifera) workers at different developmental stages. Microbiol. Res. 231, 126370 (2020).
    CAS  PubMed  Article  Google Scholar 

    47.
    D’Alvise, P. et al. The impact of winter feed type on intestinal microbiota and parasites in honey bees. Apidologie 49, 252–264 (2018).
    Article  CAS  Google Scholar 

    48.
    Huang, S. K. et al. Influence of feeding type and Nosema ceranae infection on the gut microbiota of Apis cerana workers. mSystems 3, 177–195 (2018).
    Article  Google Scholar 

    49.
    Rothman, J. A., Carroll, M. J., Meikle, W. G., Anderson, K. E. & McFrederick, Q. S. Longitudinal effects of supplemental forage on the Honey Bee (Apis mellifera) microbiota and inter- and intra-colony variability. Microb. Ecol. 76, 814–824 (2018).
    CAS  PubMed  Article  Google Scholar 

    50.
    Zhang, Y. et al. Nosema ceranae infection enhances Bifidobacterium spp. abundances in the honey bee hindgut. Apidologie 50, 353–362 (2019).
    Article  Google Scholar 

    51.
    Danforth, B. N., Minckley, R. L. & Neff, J. L. The Solitary Bees (Princeton University Press, Princeton, 2019).
    Google Scholar 

    52.
    Santos, P. K. F., Arias, M. C. & Kapheim, K. M. Loss of developmental diapause as prerequisite for social evolution in bees. Biol. Lett. 15, 20190398 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    53.
    Harmon-Threatt, A. Influence of nesting characteristics on health of wild bee communities. Annu. Rev. Entomol. 65, 39–56 (2020).
    CAS  PubMed  Article  Google Scholar 

    54.
    Johansen, C., Mayer, D., Stanford, A. & Kious, C. Alkali Bees: Their Biology and Management for Alfalfa Seed Production in the Pacific Northwest (Publication, Pacific Northwest Cooperative Extension Service, Genesee, 1982).
    Google Scholar 

    55.
    Cane, J. H. A native ground-nesting bee (Nomia melanderi) sustainably managed to pollinate alfalfa across an intensively agricultural landscape. Apidologie 39, 315–323 (2008).
    Article  Google Scholar 

    56.
    Cane, J. H. Pollinating bees (Hymenoptera: Apiformes) of U.S. alfalfa compared for rates of pod and seed set. J. Econ. Entomol. 95, 22–27 (2002).
    PubMed  Article  Google Scholar 

    57.
    Batra, S. W. & Bohart, G. E. Alkali bees: Response of adults to pathogenic fungi in brood cells. Science 165, 607 (1969).
    ADS  CAS  PubMed  Article  Google Scholar 

    58.
    Galbraith, D. A. et al. Investigating the viral ecology of global bee communities with high-throughput metagenomics. Sci. Rep. 8, 1–11 (2018).
    CAS  Article  Google Scholar 

    59.
    Bohart, G. E., Stephen, W. P. & Eppley, E. K. The biology of Heterostylum robustum (Diptera: Bombyliidae), a parasite of the alkali bee. Ann. Entomol. Soc. Am. 53, 425–435 (1960).
    Article  Google Scholar 

    60.
    Johansen, C. A., Mayer, D. F. & Eves, J. D. Biology and management of the alkali bee, Nomia melanderi Cockrell (Hymenoptera: Halictidae).Melanderii Cockrell (Hymenoptera: Halictidae).Melanderi (Washington State Entomology, Pullman, 1978).
    Google Scholar 

    61.
    Johansen, C. A. & Mayer, D. F. Pollinator Protection: A Bee and Pesticide Handbook (Wicwas Press, Kalamazoo, 1990).
    Google Scholar 

    62.
    Stephen, W. P. Solitary bees in North American agriculture: A perspective. In For Nonnative Crops, Whence Pollinators of the Future? (eds Strickler, K. & Cane, J. H.) 41–66 (Entomological Society of America, Annapolis, 2003).
    Google Scholar 

    63.
    Kapheim, K. M. et al. Draft genome assembly and population genetics of an agricultural pollinator, the solitary alkali bee (Halictidae: Nomia melanderi). G3 9, 625–634 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    64.
    Batra, S. W. T. Aggression, territoriality, mating and nest aggregation of some solitary bees (Hymenoptera: Halictidae, Megachilidae, Colletidae, Anthophoridae). J. Kansas Entomol. Soc. 51, 547–559 (1978).
    Google Scholar 

    65.
    Mayer, D. F. & Miliczky, E. R. Emergence, male behavior, and mating in the alkali bee, Nomia melanderi Cockerell (Hymenoptera: Halictidae). J. Kansas Entomol. Soc. 71, 61–68 (1998).
    Google Scholar 

    66.
    Kapheim, K. M. & Johnson, M. M. Juvenile hormone, but not nutrition or social cues, affects reproductive maturation in solitary alkali bees (Nomia melanderi). J. Exp. Biol. https://doi.org/10.1242/jeb.162255 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    67.
    Koch, H. & Schmid-Hempel, P. Bacterial communities in central European bumble bees: Low diversity and high specificity. Microb. Ecol. 62, 121–133 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    68.
    Martinson, V. G., Moy, J. & Moran, N. A. Establishment of characteristic gut bacteria during development of the honeybee worker. Appl. Environ. Microbiol. 78, 2830–2840 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    69.
    Powell, J. E., Martinson, V. G., Urban-Mead, K. & Moran, N. A. Routes of acquisition of the gut microbiota of the honey bee Apis mellifera. Appl. Environ. Microbiol. 80, 7378–7387 (2014).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    70.
    Kapheim, K. M. & Johnson, M. M. Support for the reproductive ground plan hypothesis in a solitary bee: Links between sucrose response and reproductive status. Proc. R. Soc. B Biol. Sci. 284, 20162406 (2017).
    Article  CAS  Google Scholar 

    71.
    R Core Team. R: A Language and Environment for Statistical Computing (2019).

    72.
    McMurdie, P. J. & Holmes, S. phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    73.
    Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. BioRxiv. https://doi.org/10.1101/221499 (2017).
    Article  Google Scholar 

    74.
    Jari Oksanen, F. et al. vegan: Community Ecology Package (2019).

    75.
    McMurdie, P. J. & Holmes, S. Waste not, want not: Why rarefying microbiome data is inadmissible. PLoS Comput. Biol. 10, e1003531 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    76.
    Anderson, M. J., Ellingsen, K. E. & McArdle, B. H. Multivariate dispersion as a measure of beta diversity. Ecol. Lett. 9, 683–693 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    77.
    Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
    Article  Google Scholar 

    78.
    Lenth, R. emmeans: Estimated Marginal Means, Aka Least-Squares Means (2020).

    79.
    Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    80.
    Lahti, L. & Shetty, S. Microbiome R Package (2012).

    81.
    Zheng, J. et al. A taxonomic note on the genus Lactobacillus: Description of 23 novel genera, emended description of the genus Lactobacillus beijerinck 1901, and union of Lactobacillaceae and Leuconostocaceae. Int. J. Syst. Evol. Microbiol. 70, 2782–2858 (2020).
    CAS  PubMed  Article  Google Scholar 

    82.
    Rummel, P. S. et al. Maize root and shoot litter quality controls short-term emissions and bacterial community structure of arable soil. Biogeosciences 17, 1181–1198 (2020).
    ADS  CAS  Article  Google Scholar 

    83.
    McFrederick, Q. S., Vuong, H. Q. & Rothman, J. A. Lactobacillus micheneri sp. nov., Lactobacillus timberlakei sp. nov. and Lactobacillus quenuiae sp. nov., lactic acid bacteria isolated from wild bees and flowers. Int. J. Syst. Evol. Microbiol. 68, 1879–1884 (2018).
    CAS  PubMed  Article  Google Scholar 

    84.
    Wittouck, S., Wuyts, S., Meehan, C. J., van Noort, V. & Lebeer, S. A genome-based species taxonomy of the Lactobacillus genus complex. mSystems. https://doi.org/10.1128/mSystems.00264-19 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    85.
    Engel, P. & Moran, N. A. The gut microbiota of insects—Diversity in structure and function. FEMS Microbiol. Rev. 37, 699–735 (2013).
    CAS  PubMed  Article  Google Scholar 

    86.
    Cane, J. H., Dobson, H. E. M. & Boyer, B. Timing and size of daily pollen meals eaten by adult females of a solitary bee (Nomia melanderi) (Apiformes: Halictidae). Apidologie 48, 17–30 (2016).
    Article  CAS  Google Scholar 

    87.
    Engel, P., Bartlett, K. D. & Moran, N. A. The bacterium Frischella perrara causes scab formation in the gut of its honeybee host. MBio 6, 1–8 (2015).
    Article  CAS  Google Scholar 

    88.
    Martinson, V. G. et al. A simple and distinctive microbiota associated with honey bees and bumble bees. Mol. Ecol. 20, 619–628 (2011).
    PubMed  Article  Google Scholar 

    89.
    Vásquez, A. & Olofsson, T. C. The lactic acid bacteria involved in the production of bee pollen and bee bread. J. Apic. Res. 48, 189–195 (2009).
    Article  Google Scholar 

    90.
    Vuong, H. Q. & McFrederick, Q. S. Comparative genomics of wild bee and flower isolated Lactobacillus reveals potential adaptation to the bee host. Genome Biol. Evol. 11, 2151–2161 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar  More

  • in

    Sludge amendment accelerating reclamation process of reconstructed mining substrates

    1.
    Kuang, X. Y., Cao, Y. G., Luo, G. B. & Huang, Y. H. Responses of melilotus officinalis growth to the composition of different topsoil substitute materials in the reclamation of open-pit mining grassland area in Inner Mongolia. Materials 12, 1–21. https://doi.org/10.3390/ma12233888 (2019).
    CAS  Article  Google Scholar 
    2.
    Cheng, W., Bian, Z. F., Dong, J. H. & Lei, S. G. Soil properties in reclaimed farmland by filling subsidence basin due to underground coal mining with mineral wastes in China. Trans. Nonferrous Metals Soc. China. 24, 2627–2635. https://doi.org/10.1016/S1003-6326(14)63392-6 (2014).
    CAS  Article  Google Scholar 

    3.
    Du, T., Wang, D. M., Bai, Y. J. & Zhang, Z. Z. Optimizing the formulation of coal gangue planting substrate using wastes: The sustainability of coal mine ecological restoration. Ecol. Eng. 143, 1–10. https://doi.org/10.1016/j.ecoleng.2019.105669 (2020).
    Article  Google Scholar 

    4.
    Yin, N. N., Zhang, Z., Wang, L. P. & Qian, K. M. Variations in organic carbon, aggregation, and enzyme activities of gangue-fly ash-reconstructed soils with sludge and arbuscular mycorrhizal fungi during 6-year reclamation. Environ. Sci. Pollut. Res. 23, 17840–17849. https://doi.org/10.1007/s11356-016-6941-5 (2016).
    CAS  Article  Google Scholar 

    5.
    Clemente, R. et al. Combination of soil organic and inorganic amendments helps plants overcome trace element induced oxidative stress and allows phytostabilisation. Chemosphere 223, 223–231. https://doi.org/10.1016/j.chemosphere.2019.02.056 (2019).
    ADS  CAS  Article  PubMed  Google Scholar 

    6.
    Wu, D. et al. Integrated application of sewage sludge, earthworms and Jatropha curcas on abandoned rare-earth mine land soil. Chemosphere 214, 47–54. https://doi.org/10.1016/j.chemosphere.2018.09.087 (2019).
    ADS  CAS  Article  PubMed  Google Scholar 

    7.
    Blankinship, J. C., Fonte, S. J., Six, J. & Schimel, J. P. Plant versus microbial controls on soil aggregate stability in a seasonally dry ecosystem. Geoderma 272, 39–50. https://doi.org/10.1016/j.geoderma.2016.03.008 (2016).
    ADS  Article  Google Scholar 

    8.
    Wang, S., Li, T., Zheng, Z. & Chen, H. Y. H. Soil aggregate-associated bacterial metabolic activity and community structure in different aged tea plantations. Sci. Total Environ. 654, 1023–1032. https://doi.org/10.1016/j.scitotenv.2018.11.032 (2019).
    ADS  CAS  Article  PubMed  Google Scholar 

    9.
    Parvin, S., Van Geel, M., Yeasmin, T., Lievens, B. & Honnay, O. Variation in arbuscular mycorrhizal fungal communities associated with lowland rice (Oryza sativa) along a gradient of soil salinity and arsenic contamination in Bangladesh. Sci. Total Environ. 686, 546–554. https://doi.org/10.1016/j.scitotenv.2019.05.450 (2019).
    ADS  CAS  Article  PubMed  Google Scholar 

    10.
    Lanfranco, L., Fiorilli, V., Venice, F. & Bonfante, P. Strigolactones cross the kingdoms: plants, fungi, and bacteria in the arbuscular mycorrhizal symbiosis. J. Exp. Bot. 69, 2175–2188. https://doi.org/10.1093/jxb/erx432 (2018).
    CAS  Article  PubMed  Google Scholar 

    11.
    Singh, A. K., Rai, A., Pandey, V. & Singh, N. Contribution of glomalin to dissolve organic carbon under different land uses and seasonality in dry tropics. J. Environ. Manag. 192, 142–149. https://doi.org/10.1016/j.jenvman.2017.01.041 (2017).
    CAS  Article  Google Scholar 

    12.
    Zhang, J., Ekblad, A., Sigurdsson, B. D. & Wallander, H. The influence of soil warming on organic carbon sequestration of arbuscular mycorrhizal fungi in a sub-arctic grassland. Soil Biol. Biochem. 147, 1–9. https://doi.org/10.1016/j.soilbio.2020.107826 (2020).
    CAS  Article  Google Scholar 

    13.
    Singh, A. K., Rai, A. & Singh, N. Effect of long term land use systems on fractions of glomalin and soil organic carbon in the Indo-Gangetic plain. Geoderma 277, 41–50. https://doi.org/10.1016/j.geoderma.2016.05.004 (2016).
    ADS  CAS  Article  Google Scholar 

    14.
    Xiao, L. et al. Effects of freeze-thaw cycles on aggregate-associated organic carbon and glomalin-related soil protein in natural-succession grassland and Chinese pine forest on the Loess Plateau. Geoderma 334, 1–8. https://doi.org/10.1016/j.geoderma.2018.07.043 (2019).
    ADS  CAS  Article  Google Scholar 

    15.
    Qian, K. M., Wang, P. & Yin, N. Effects of AMF on soil enzyme activity and carbon sequestration capacity in reclaimed mine soil. Int. J. Min. Sci. Technol. 22, 553–557. https://doi.org/10.1016/j.ijmst.2012.01.019 (2012).
    CAS  Article  Google Scholar 

    16.
    Ahirwal, J. & Maiti, S. K. Assessment of carbon sequestration potential of revegetated coal mine overburden dumps: A chronosequence study from dry tropical climate. J. Environ. Manag. 201, 369–377. https://doi.org/10.1016/j.jenvman.2017.07.003 (2017).
    CAS  Article  Google Scholar 

    17.
    Yuan, Y., Zhao, Z. Q., Li, X. Z., Wang, Y. Y. & Bai, Z. K. Characteristics of labile organic carbon fractions in reclaimed mine soils: Evidence from three reclaimed forests in the Pingshuo opencast coal mine, China. Sci. Total Environ. 613, 1196–1206. https://doi.org/10.1016/j.scitotenv.2017.09.170 (2018).
    ADS  CAS  Article  PubMed  Google Scholar 

    18.
    Hassan, W. et al. Labile organic carbon fractions, regulator of CO2 emission: Effect of plant residues and water regimes. Clean: Soil, Air, Water 44, 1358–1367. https://doi.org/10.1002/clen.201400405 (2016).
    CAS  Article  Google Scholar 

    19.
    Cheng, X. R., Yu, M. K. & Wang, G. G. Effects of thinning on soil organic carbon fractions and soil properties in Cunninghamia lanceolata stands in Eastern China. Forests. 8, 1–21. https://doi.org/10.3390/f8060198 (2017).
    Article  Google Scholar 

    20.
    Zhong, Y. Q. W., Yan, W. M. & Shangguan, Z. P. Soil carbon and nitrogen fractions in the soil profile and their response to long-term nitrogen fertilization in a wheat field. CATENA 135, 38–46. https://doi.org/10.1016/j.catena.2015.06.018 (2015).
    CAS  Article  Google Scholar 

    21.
    Wang, Y., Ling, C., Chen, Y., Jiang, X. R. & Chen, G. Q. Microbial engineering for easy downstream processing. Biotechnol. Adv. 37, 1–9. https://doi.org/10.1016/j.biotechadv.2019.03.004 (2019).
    CAS  Article  PubMed  Google Scholar 

    22.
    Ye, G. P. et al. Manure over crop residues increases soil organic matter but decreases microbial necromass relative contribution in upland Ultisols: Results of a 27-year field experiment. Soil Biol. Biochem. 134, 15–24. https://doi.org/10.1016/j.soilbio.2019.03.018 (2019).
    CAS  Article  Google Scholar 

    23.
    Du, R. et al. Advanced nitrogen removal with simultaneous Anammox and denitrification in sequencing batch reactor. Bioresour. Technol. 162, 316–322. https://doi.org/10.1016/j.biortech.2014.03.041 (2014).
    CAS  Article  PubMed  Google Scholar 

    24.
    Luna, L. et al. Restoration techniques affect soil organic carbon, glomalin and aggregate stability in degraded soils of a semiarid Mediterranean region. CATENA 143, 256–264. https://doi.org/10.1016/j.catena.2016.04.013 (2016).
    CAS  Article  Google Scholar 

    25.
    Hao, X. H. et al. Effect of long-term application of inorganic fertilizer and organic amendments on soil organic matter and microbial biomass in three subtropical paddy soils. Nutr. Cycl. Agroecosyst. 81, 17–24. https://doi.org/10.1007/s10705-007-9145-z (2008).
    Article  Google Scholar 

    26.
    Fokom, R. et al. Glomalin related soil protein, carbon, nitrogen and soil aggregate stability as affected by land use variation in the humid forest zone of south Cameroon. Soil Tillage Res. 120, 69–75. https://doi.org/10.1016/j.still.2011.11.004 (2012).
    Article  Google Scholar 

    27.
    Wang, W. et al. Glomalin changes in urban-rural gradients and their possible associations with forest characteristics and soil properties in Harbin City, Northeastern China. J. Environ. Manag. 224, 225–234. https://doi.org/10.1016/j.jenvman.2018.07.047 (2018).
    CAS  Article  Google Scholar 

    28.
    Anirwal, J., Kumar, A., Pietrzykowski, M. & Maiti, S. K. Reclamation of coal mine spoil and its effect on Technosol quality and carbon sequestration: A case study from India. Environ. Sci. Pollut. Res. 25, 27992–28003. https://doi.org/10.1007/s11356-018-2789-1 (2018).
    CAS  Article  Google Scholar 

    29.
    Sun, S., Li, S., Avera, B. N., Strahm, B. D. & Badgley, B. D. Soil bacterial and fungal communities show distinct recovery patterns during forest ecosystem restoration. Appl. Environ. Microbiol. 83, 1–14. https://doi.org/10.1128/AEM.00966-17 (2017).
    CAS  Article  Google Scholar 

    30.
    Amir, H. et al. Arbuscular mycorrhizal fungi and sewage sludge enhance growth and adaptation of Metrosideros laurifolia on ultramafic soil in New Caledonia: A field experiment. Sci. Total Environ. 651, 334–343. https://doi.org/10.1016/j.scitotenv.2018.09.153 (2019).
    ADS  CAS  Article  PubMed  Google Scholar 

    31.
    Klabi, R. et al. Interaction between legume and arbuscular mycorrhizal fungi identity alters the competitive ability of warm-season grass species in a grassland community. Soil Biol. Biochem. 70, 176–182. https://doi.org/10.1016/j.soilbio.2013.12.019 (2014).
    CAS  Article  Google Scholar 

    32.
    Prasad, R., Tuteja, N. & Varma, A. Mycorrhiza—Nutrient Uptake, Biocontrol Ecorestoration 161–163 (Springer International Publishing, Cham, 2017).
    Google Scholar 

    33.
    Becher, P. G. et al. Developmentally regulated volatiles geosmin and 2-methylisoborneol attract a soil arthropod to Streptomyces bacteria promoting spore dispersal. Nat. Microbiol. 5, 821–829. https://doi.org/10.1038/s41564-020-0697-x (2020).
    CAS  Article  PubMed  Google Scholar 

    34.
    Houston, T. F. & Bougher, N. L. Records of hypogeous mycorrhizal fungi in the diet of some Western Australian bolboceratine beetles (Coleoptera: Geotrupidae, Bolboceratinae). Aust. J. Entomol. 49, 49–55. https://doi.org/10.1111/j.1440-6055.2009.00720.x (2010).
    Article  Google Scholar 

    35.
    Han, X. G. et al. Dynamics of arbuscular mycorrhizal fungi in relation to root colonization, spore density, and soil properties among different spreading stages of the exotic plant threeflower beggarweed (Desmodium triflorum) in a Zoysia tenuifolia lawn. Weed Sci. 67, 689–701. https://doi.org/10.1017/wsc.2019.50 (2019).
    Article  Google Scholar 

    36.
    Sandeep, S., Manjaiah, K. M., Mayadevi, M. R. & Singh, A. K. Monitoring temperature sensitivity of soil organic carbon decomposition under maize-wheat cropping systems in semi-arid India. Environ. Monit. Assess. 188, 1–15. https://doi.org/10.1007/s10661-016-5455-4 (2016).
    CAS  Article  Google Scholar 

    37.
    Ahirwal, J. & Maiti, S. K. Development of Technosol properties and recovery of carbon stock after 16 years of revegetation on coal mine degraded lands, India. CATENA 166, 114–123. https://doi.org/10.1016/j.catena.2018.03.026 (2018).
    CAS  Article  Google Scholar 

    38.
    Stumpf, L., Pauletto, E. A. & Pinto, L. F. S. Soil aggregation and root growth of perennial grasses in a constructed clay minesoil. Soil Tillage Res. 161, 71–78. https://doi.org/10.1016/j.still.2016.03.005 (2016).
    Article  Google Scholar 

    39.
    Helliwell, J. R. et al. The emergent rhizosphere: imaging the development of the porous architecture at the root-soil interface. Sci. Rep. 7, 1–10. https://doi.org/10.1038/s41598-017-14904-w (2017).
    CAS  Article  Google Scholar 

    40.
    Fu, W. J. et al. The carbon storage in moso bamboo plantation and its spatial variation in Anji County of southeastern China. J. Soils Sedim. 14, 320–329. https://doi.org/10.1007/s11368-013-0665-7 (2014).
    CAS  Article  Google Scholar 

    41.
    Zhang, Z. H., Wang, Q., Wang, H., Nie, S. M. & Liang, Z. W. Effects of soil salinity on the content, composition, and ion binding capacity of glomalin-related soil protein (GRSP). Sci. Total Environ. 581, 657–665. https://doi.org/10.1016/jscitotenv.2016.12.176 (2017).
    ADS  Article  PubMed  Google Scholar 

    42.
    Janos, D. P., Garamszegi, S. & Beltran, B. Glomalin extraction and measurement. Soil Biol. Biochem. 40, 728–739. https://doi.org/10.1016/j.soilbio.2007.10.007 (2008).
    CAS  Article  Google Scholar 

    43.
    Dave, B. P. & Soni, A. Diversity of halophilic Archaea at salt pans around Bhavnagar Coast, Gujarat. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 83, 225–232. https://doi.org/10.1007/s40011-012-0124-z (2013).
    Article  Google Scholar 

    44.
    Magurran, A. E. Ecological Diversity and Its Measurement 61–80 (Princeton University Press, Princeton, 1988).
    Google Scholar  More

  • in

    Characterization of a green Stentor with symbiotic algae growing in an extremely oligotrophic environment and storing large amounts of starch granules in its cytoplasm

    Distribution of Stentor pyriformis in Japan and its optimal culture conditions
    S. pyriformis was described by Johnson in 18936. This algae-bearing Stentor has separated spherical macronuclei without pigmentation, which certainly differentiates it from other Stentor species (see Table S1, Fig. 5B). While the most common algae-bearing Stentor, S. polymorphus assumes a slender trumpet shape (often shortened), S. pyriformis never resembles such a slender trumpet, but assumes a pear or short conical shape, even when it is swimming6. Presence or absence of colored pigmentation is also a prominent characteristic for separating Stentor species. Among algae-bearing Stentor spp., S. polymorphus and S. pyriformis only are considered colorless species, whereas colored species are S. amethystinus, S. fuliginosus, S. araucanus, and S. tartari8 (Table S1). Therefore, S. pyriformis is a clearly discernible species; however, it remains underexplored. Indeed, we could only find one paper on the new habitats of S. pyriformis7, with the exception of the paper of species consolidation of this genus8. We confirmed the presence of S. pyriformis at 23 locations (Fig. 1A). This indicates that S. pyriformis is by no means a rare organism. We assume one of the reasons why S. pyriformis has been poorly studied is the difficulty of cultivation. In fact, Johnson6 noted that he could not keep them more than a month and never observed any cells in fission. In addition, after five years of failure, it was finally possible to culture S. pyriformis for more than several months. Because of objectively unfounded data that we could not include in the distribution data (Fig. 1A), we noticed the wetlands where we found S. pyriformis were limited to small ponds or bogs locating near the mountain peak or along the ridge (Fig. 1B). That is, the ponds depending on rainfall without inflowing rivers. Because there is no nutrient flowing in, waters in these ponds showed noticeable oligotrophic tendency, i.e., extremely low electric conductivity (Fig. 1A), which gave us some clues on culture.
    The most important point of culture for S. pyriformis was keeping the medium lower electric conductivity. We use the KCM medium diluted by 2% with Milli-Q water, and changed medium once a week. A non-photosynthetic cryptophyte, Chilomonas paramecium was selected for food. We selected the food so that it would not itself grow in the culture medium. Growing organisms, like photosynthetic algae, seemed to cause damage to S. pyriformis. Using this culture method, S. pyriformis can be maintained for more than four years (see Table 1). For the organisms not easy to grow in culture, Professor Michael Melkonian mentioned no protist is ‘uncultivable’, there is just human failure30. Here, it just became possible to culture S. pyriformis 120 years after its discovery; however, this method does not always work. S. pyriformis appears to be extremely fragile and disintegrates when any variables are unintentionally altered, that is, the culture is still unstable. When its condition deteriorates, the cells divide unevenly in such a way that a part of the cell is broken. When this happens, the cells become spherical, and the drug drops to the bottom of the dish. It retains this shape for more than a month, but eventually disappears. The doubling time of S. pyriformis remains 3 to 4 weeks, even under favorable conditions (data not shown). We occasionally encountered the blooming of S. pyriformis all over the bottom of the ponds (Fig. 1C). S. pyriformis, therefore, does not seem to be a particularly slow growing species, but our culture method appears to be far from the optimal culture conditions for them. Three S. pyriformis strains used in this study are available from the authors upon request.
    Ultrastructure
    In this study, we compared the conventional chemical fixation method with the rapid-freezing fixation method for electron microscopic observation. As a result, large vacuoles were observed in the cytoplasm when chemical fixation was used, but not by rapid freezing. Instead, many multi-vesicular bodies were observed in the cytoplasm. The quick-freezing and freeze-substitution method is considered superior in that it can prevent deformation of the intracellular structure compared to chemical fixation31. Therefore, it is possible that the originally existing multi-vesicular bodies were artificially disintegrated by chemical fixation, and the constituent biological membranes fused together, eventually forming large vacuoles. To the authors’ knowledge, no intracellular structure similar to the multi-vesicular body in S. pyriformis has been reported in protists. As multi-vesicular bodies of S. pyriformis could only be observed using the freeze-substitution method, similar granules may also be found in other protists if the same technique is used for electron microscopy. In animals, on the other hand, aggregates of secretory vesicles resembling the multi-vesicular bodies of S. pyriformis are present in cardiac telocytes32. The extracellular vesicles form multi-vesicular structures of about 1 μm in diameter and contain materials for intercellular communication that are involved in cardiac physiology and regeneration. Because S. pyriformis cells often form aggregates at the bottom of the pond, some chemicals may be released from the multi-vesicular body, attracting nearby cells and forming aggregates.
    Observation by the freeze-substitution method revealed that the symbiosome membrane was in close contact with the symbiotic chlorella. Furthermore, fluffy projections were observed on the cell wall of the symbiotic chlorella. These characteristics were consistent with those of C. variabilis, which is symbiotic in the cells of P. bursaria9. The only difference was that in S. pyriformis, the symbiotic chlorella cells were scattered in the cytoplasm, whereas the symbiotic Chlorella in P. bursaria were anchored directly below the cell surface.
    Storage granules
    The iodine in Lugol’s solution selectively binds to α-1, 4-linked glucose found in polysaccharides, such as starch33 and glycogen34. The color stained with Lugol’s solution reflects the type of glucose polymer. Starches with high amylose content stain blue-violet (cf. Fig. 4B), high amylopectin stains red–purple, and glycogen stains reddish brown (Table 2). The granules in the cytoplasm of S. pyriformis stained reddish brown with Lugol’s solution (Fig. 4A), suggesting that these granules are composed of α-1,4-linked glucans with high number of α-1,6-linked branch points, either amylopectin-rich starch or glycogen. The pyrenoid of Chlorella spp. is surrounded by a starch sheath of two large plates35. As shown in Fig. 4F,G, the image contrast formed by electron staining of the starch granule in the chloroplast (arrow) was lost by treatment with Lugol’s solution. Although the detailed mechanism is unknown, this observation suggests that electron-stained heavy metals (osmium, lead, and lanthanoid ions) bound to the granules may have been eliminated by iodine in Lugol’s solution. The cytoplasmic granules of S. pyriformis showed the same staining properties as the starch granules in the chloroplasts of symbiotic chlorella, suggesting that both types of granules share chemical characteristics as polysaccharides.
    Alveolates make up one of the most diverse and largest groups of protists. They include three major taxa: dinoflagellates, ciliates, and apicomplexan protozoa. All three alveolate lineages store glucose in an α-1,4-linked glucose chain with α-1, 6 branches. Ciliates are known to synthesize glycogen granules. For example, Tetrahymena has glycogen granules between 35 and 40 nm in diameter, each granule being a collection of small γ-granules of 2–3 nm in size36. Dinoflagellates and apicomplexans typically produce more complex and larger spherical starch particles, usually greater than 1 μm in size37,38. Amylopectin-rich starch and glycogen are very similar polysaccharides, but they differ in granule size and birefringence (Table 2). Starch granules are large, birefringent, and have a high refractive index, but glycogen does not exhibit birefringence, and its granules generally have a size of 300 nm or less. When observed with a polarizing microscope, the starch granules show a Maltese cross pattern. This pattern is derived from the radial arrangement of amylose and amylopectin molecules in granules and is one of the criteria for starch identification. Since the cytoplasmic granules of S. pyriformis are large in size (1–3 μm) and show a typical Maltese cross pattern as shown in Fig. 4E, these granules are likely to be starch granules rich in amylopectin.
    Phylogeny of S. pyriformis and its morphology
    Relationships of Stentor species were not clearly resolved. BI and ML analyses indicated basal diverging of the S. pyriformis + S. amethystinus clade from others, but NJ analysis did not indicate so (Fig. 5). Recent phylogenetic analyses inclusive of Stentor species also indicated basal diverging of S. amethystinus from the others; however, the monophyly of the others is not highly supported21,22. Therefore, the one thing that can be said is that S. pyriformis is closely related to S. amethystinus.
    For the identification of Stentor species, the shape of macronucleus, presence or absence of cortical pigmentation, and symbiotic algae are very important and iconic characteristics8,19. S. pyriformis and S. amethystinus share beaded macronuclei and the presence of symbiotic chlorella (Table S1, Fig. 5B). Pigmentation is present in S. amethystinus, but not in S. pyriformis. Pigmentation is a noticeable characteristic, which tinctures the whole body of Stentor cells. The pigment is thought to function as a defense against predators39. However, the kind of pigment compound depends on the species40, and the relationship between pigment possession and phylogeny is poor (Fig. 5). Of note, colorless vesicles exist in S. pyriformis (Fig. 2D). The short and fat shape is also a common characteristic for S. pyriformis and S. amethystinus, in this genus with many elongated trumpet shape species6,8.
    Symbiotic algae in S. pyriformis
    Algae-targeted PCR products from whole cells of S. pyriformis were sequenced directly, and clear peaks were obtained for each. This shows that all or nearly all of the algal symbionts in each Stentor cell are unified, regardless of samples under long-term culture or nature. In addition, all symbionts were closely related to C. variabilis (Fig. S3), which has been known as a representative symbiont of P. bursaria (Oligohymenophorea), the model organism of multi-algae retaining protists (MARP41) style symbioses. For the chlorellacean species, the diversity of ITS2 sequence comparisons has often been adopted. For two organisms to compare, ITS2 sequence differences (gaps are counted as a fifth character) usually fall either less than 2% for single species or more than 10% for different species42,43. This characteristic simply encourages a species concept. The ITS2 sequences of S. pyriformis algae differ only by one nucleotide site out of 248 sites from those of P. bursaria algae (Fig. 6A), which strongly suggests the symbiotic chlorella of S. pyriformis are also C. variabilis. Several Stentor species retain coccoid green algae8 (Table S1), but only three algal sequences have been published. Two algal sequences from S. polymorphus belonged to different clades from Chlorellaceae44,45. As for the other algal sequence of S. amethystinus, the symbiont may belong to Chlorellaceae46. This sequence (EF589816) is short (991 bp) and only covers a part of SSU rDNA; therefore, it was not included in our phylogenetic analyses (Fig. S3). The sequence differs from C. variabilis with 10 base changes and 3 indels, indicating that it is not C. variabilis.
    Figure 6

    Sequence differences of SSU, ITS1, 5.8S and ITS2 rRNA gene (without group I introns) among Chlorella variabilis. “PbS-gt” indicates Paramecium bursaria symbiont genotypes. Genotype 1 includes SAG 211-6, ATCC 50258 (NC64A), NIES-2541, and some other US and Japanese strains. Genotype 2 is the alga of Chinese P. bursaria strain Cs2, and genotype 3 is the alga of Australian P. bursaria strain MRBG1. For further information, see Hoshina et al.53. “SpS” indicates the algal sequence of Stentor pyriformis strains collected from Japan. (A). Different positions. Numerals represent the nucleotide number in aligned sequences (2462 aligned sites). (B). Distance tree of above four types of sequences. (C). E23_2 helix of SSU rRNA structure that includes hemi-CBC at the alignment position 656. (D). Deformation of ITS1 Helix 1 associated with the mutations including several nucleotide insertions.

    Full size image

    In the case of P. bursaria-C. variabilis symbiosis, C. variabilis has been shown to be vastly different from other free-living species. C. variabilis demands organic nitrogen compounds47 and leaks nearly half of the photosynthate to outside algal cells48,49. Furthermore, they are sensitive to the C. variabilis virus (CvV; so-called ‘NC64A virus’), which is abundant in natural freshwater50,51,52. Therefore, C. variabilis should be considered an already evolved species that is unable to survive without the protection of the host cell53.
    Four C. variabilis rDNA sequences obtained from S. pyriformis were identical, with the exception of a nucleotide position in the S1512 intron. Here, the regions without group I introns, i.e., SSU, ITS1, 5.8S, and ITS2 rDNA, are compared among C. variabilis sequences of S. pyriformis and of P. bursaria. Several published sequences cover the above SSU-ITS region, of which varieties are shown as P. bursaria symbiont genotype (PbS-gt) 1 to 3 (Fig. 6A). Due to the small number of sequences, it is still unknown whether these genotypes depend on (or are related to) their living regions. Genotype 1 was from USA and Japan, genotype 2 was from China, and genotype 3 was from Australia. All available sequences for S. pyriformis symbionts were obtained in this study, and they were all from Japan. As a result, all sequences of S. pyriformis symbionts were aggregated into a single genotype SpS, which was distantly related to all P. bursaria symbionts, including those from Japan (Fig. 6B). Five variable sites are found in SSU rDNA among C. variabilis genotypes, of which four are concentrated to that of the symbionts of S. pyriformis (SpS) (Fig. 6A). C/T substitution at alignment position 656 will be a hemi-compensatory base change (hemi-CBC) at the E23_2 helix of SSU rRNA structure (Fig. 6C), whereas the other four sites are at single strand regions (data not shown). Mutations (1821–1828) including comparatively large indels were seen in ITS1 region (Fig. 6A). It was found that all these mutations are assembled in helix 1 (for chlorellacean ITS1 structure, see Bock et al.54,55). Thermodynamic analysis via Mfold56,57 predicted that PbS sequences form linear helix 1 similar to the other chlorellacean species, but SpS sequences including the additional nucleotides may form a dichotomous branching of helix 1 (Fig. 6D).
    The group I introns inserted in SSU rDNA of S. pyriformis symbionts are identical to those of P. bursaria symbionts28,58 in terms of numbers (three introns) and insertion sites (S943, S1367 and S1512). The sequences of S943 and S1512 introns are matched more than 99%. However, with respect to the S1367 intron, a large length gap was found (168 nucleotides) at the tip of P8 (Fig. S4). This section has been indicated as a homing endonuclease gene remnant28, and those of S. pyriformis symbionts are presumed to be a more degenerated form than those of P. bursaria symbionts.
    At any rate, the symbiotic algae of S. pyriformis were found to be C. variabilis. Because S. pyriformis never lost the symbiotic algae in four years of culture, and all four algae had nearly identical genetic characteristics, the symbiotic relationships between S. pyriformis and C. variabilis can be regarded as stable, or permanent. Although S. pyriformis and P. bursaria share C. variabilis as their endosymbionts, considering the genetic differences depending on their host species, the sharing event has not happened recently. Symbiont sharing among various host species has also been known for some ciliates41,59 (Carolibrandtia ciliaticola in Fig. S3), and a script to spread a particular algal symbiont has been suggested41. Given the physiological characters of C. variabilis (mentioned above), this algal species might be an ideal algal symbiont, and it will be no surprise if the other protists also retained C. variabilis as their algal partners. Research on the symbiotic algae that other Stentor spp. have and on host and regional dependencies are awaited.
    Adaptation of S. pyriformis to oligotrophic environment in highland marsh
    In Japan, S. pyriformis lives only in alpine ponds (Fig. 1), where the winter is cold, and the surface of the pond is always covered with ice. The water in these ponds has low electrical conductivity (~ 10 μS/cm), and there are few living organisms except S. pyriformis, meaning that only little food is available in wintertime. The reason this ciliate is rich in stored carbohydrate granules may be due to its need for nutrients to survive such harsh winter environments.
    Preliminary studies suggest that many protists, especially ciliates, may make starch. Large amounts of cytoplasmic granules that show a Maltese cross were observed in chlorella-bearing ciliates such as P. bursaria, while only a small amount of such granules was observed in Euplotes aediculatus, Paramecium caudatum, Blepharisma japonicum, and Tetrahymena pyriformis. Protists with symbiotic algae seem to produce particularly large amounts of stored carbohydrate granules in the cytoplasm, but the mechanism of starch synthesis may be widely shared by ciliates.
    P. bursaria has been shown to be more resistant to starvation conditions than the aposymbiotic strain of the same species13. Under food-deprived conditions, P. bursaria was interpreted to have survived by digesting symbiotic algae. Resting cyst formation and cannibalism are known as other strategies for protozoans to survive starvation conditions60. This study suggests that the use of carbohydrate granules stored in cells may be another possible strategy for ciliates to survive harsh environments such as highland oligotrophic bogs. More

  • in

    Intraspecific differences in the invasion success of the Argentine ant Linepithema humile Mayr are associated with diet breadth

    1.
    Abril, S. & Gómez, C. Aggressive behaviour of the two European Argentine ant supercolonies (Hymenoptera: Formicidae) towards displaced native ant species of the northeastern Iberian Peninsula. Myrmecol. News 14, 99–106 (2010).
    Google Scholar 
    2.
    Blight, O. et al. Differences in behavioural traits among native and introduced colonies of an invasive ant. Biol. Invasions 19, 1389–1398 (2017).
    Article  Google Scholar 

    3.
    Jun, G., Wei, D., Qiong, W. & Hong-liang, L. Thermal tolerance for two cohorts of a native and an invasive freshwater turtle species. Acta Herpetol. 13, 83–88 (2018).
    Google Scholar 

    4.
    Jackson, M. C. & Britton, J. R. Divergence in the trophic niche of sympatric freshwater invaders. Biol. Invasions 16, 1095–1103 (2014).
    Article  Google Scholar 

    5.
    Pettitt-Wade, H., Wellband, K. W., Heath, D. D. & Fisk, A. T. Niche plasticity in invasive fishes in the Great Lakes. Biol. Invasions 17, 2565–2580 (2015).
    Article  Google Scholar 

    6.
    Pyšek, P. & Richardson, D. M. Invasive species, environmental change and management, and health. Annu. Rev. Environ. Resour. 35, 25–55 (2010).
    Article  Google Scholar 

    7.
    Cadotte, M. W., Murray, B. R. & Lovett-Doust, J. Ecological patterns and biological invasions: Using regional species inventories in macroecology. Biol. Invasions 8, 809–821 (2006).
    Article  Google Scholar 

    8.
    Pyšek, P. & Richardson, D. M. Traits associated with invasiveness in alien plants: Where do we stand?. Biol. Invasions 193, 97–125 (2007).
    Article  Google Scholar 

    9.
    Falconer, D. S. & Mackay, T. F. C. Introduction to Quantitative Genetics (Longman Group, London, 1996).
    Google Scholar 

    10.
    Lowe, S., Browne, S., Boudjelas, M. S. & De Poorter, M. 100 of the World’s Worst Invasive Alien Species: A Selection From The Global Invasive Species Database. Encyclopedia of Biological Invasions vol. 12 (Invasive Species Specialist Group, 2000).

    11.
    Suarez, A. V., Holway, D. A. & Case, T. J. Patterns of spread in biological invasions dominated by long-distance jump dispersal: Insights from Argentine ants. Proc. Natl. Acad. Sci. U.S.A. 98, 1095–1100 (2001).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Roura-Pascual, N. et al. Relative roles of climatic suitability and anthropogenic influence in determining the pattern of spread in a global invader. Proc. Natl. Acad. Sci. U.S.A. 108, 220–225 (2011).
    ADS  CAS  PubMed  Article  Google Scholar 

    13.
    Hölldobler, B. & Wilson, E. O. The Ants (Harvard University Press, Cambridge, 1990).
    Google Scholar 

    14.
    Holway, D. A., Lach, L., Suarez, A. V., Tsutsui, N. D. & Case, T. J. The causes and consequences of ant invasions. Annu. Rev. Ecol. Syst. 33, 181–233 (2002).
    Article  Google Scholar 

    15.
    Suarez, A. V., Tsutsui, N. D., Holway, D. A. & Case, T. J. Behavioral and genetic differentiation between native and introduced populations of the Argentine ant. Biol. Invasions 1, 43–53 (1999).
    Article  Google Scholar 

    16.
    Giraud, T., Pedersen, J. S. & Keller, L. Evolution of supercolonies: The Argentine ants of southern Europe. Proc. Natl. Acad. Sci. U.S.A. 99, 6075–6079 (2002).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    17.
    Pedersen, J. S., Krieger, M. J. B., Vogel, V., Giraud, T. & Keller, L. Native supercolonies of unrelated individuals in the invasive Argentine ant. Evolution 60, 782–791 (2006).
    PubMed  Article  Google Scholar 

    18.
    Inoue, M. N. et al. Recent range expansion of the Argentine ant in Japan. Divers. Distrib. 19, 29–37 (2013).
    Article  Google Scholar 

    19.
    Sunamura, E. et al. Four mutually incompatible Argentine ant supercolonies in Japan: inferring invasion history of introduced Argentine ants from their social structure. Biol. Invasions 11, 2329–2339 (2009).
    Article  Google Scholar 

    20.
    Sunamura, E. et al. Intercontinental union of Argentine ants: behavioral relationships among introduced populations in Europe, North America, and Asia. Insectes Soc. 56, 143–147 (2009).
    Article  Google Scholar 

    21.
    Thomas, M. L., Payne-Makrisâ, C. M., Suarez, A. V., Tsutsui, N. D. & Holway, D. A. When supercolonies collide: Territorial aggression in an invasive and unicolonial social insect. Mol. Ecol. 15, 4303–4315 (2006).
    PubMed  Article  Google Scholar 

    22.
    Tsutsui, N. D., Suarez, A. V., Holway, D. A. & Case, T. J. Reduced genetic variation and the success of an invasive species. Proc. Natl. Acad. Sci. 97, 5948–5953 (2000).
    ADS  CAS  PubMed  Article  Google Scholar 

    23.
    Corin, S. E., Abbott, K. L., Ritchie, P. A. & Lester, P. J. Large scale unicoloniality: The population and colony structure of the invasive Argentine ant (Linepithema humile) in New Zealand. Insectes Soc. 54, 275–282 (2007).
    Article  Google Scholar 

    24.
    Hayasaka, D. et al. Different acute toxicity of fipronil baits on invasive Linepithema humile supercolonies and some non-target ground arthropods. Ecotoxicology 24, 1221–1228 (2015).
    CAS  PubMed  Article  Google Scholar 

    25.
    R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.r-project.org/ (2019).

    26.
    Rudnick, D. & Resh, V. Stable isotopes, mesocosms and gut content analysis demonstrate trophic differences in two invasive decapod crustacea. Freshw. Biol. 50, 1323–1336 (2005).
    Article  Google Scholar 

    27.
    Jackson, M. C. et al. Population-level metrics of trophic structure based on stable isotopes and their application to invasion ecology. PLoS ONE 7, 1–12 (2012).
    Google Scholar 

    28.
    Sunamura, E., Nishisue, K., Terayama, M. & Tatsuki, S. Invasion of four Argentine ant supercolonies into Kobe Port, Japan: Their distributions and effects on indigenous ants (Hymenoptera: Formicidae). Sociobiology 50, 659–674 (2007).
    Google Scholar 

    29.
    Nakahama, N. et al. Identification of the mitochondrial DNA haplotype of an invasive Linepithema humile (Mayr, 1868) (Hymenoptera: Formicidae) population of a new location in Japan for its effective eradication. Entomol. News 128, 217–225 (2019).
    Article  Google Scholar 

    30.
    Sato, K., Sakamoto, H., Hirata, M., Ozaki, M. & Higashi, S. Household and Structural Insects Relationship Among Establishment Durations , Kin Relatedness , Aggressiveness , and Distance Between Populations of Eight Invasive Argentine Ant (Hymenoptera : Formicidae) Supercolonies in Japan. 110, 1676–1684 (2017).

    31.
    Layman, C. A., Arrington, D. A., Montana, C. G. & Post, D. M. Can stable isotope ratios provide for community-wide measures of trophic structure?. Ecol. Soc. Am. 89, 2358–2359 (2007).
    Google Scholar 

    32.
    Jackson, A. L., Inger, R., Parnell, A. C. & Bearhop, S. Comparing isotopic niche widths among and within communities: SIBER—Stable Isotope Bayesian Ellipses in R. J. Anim. Ecol. 80, 595–602 (2011).
    PubMed  Article  Google Scholar 

    33.
    Post, D. M. Using stable isotopes to estimate trophic position: Models, methods, and assumptions. Ecology 83, 703–718 (2002).
    Article  Google Scholar 

    34.
    VanderZanden, M. J. & Rasmussen, J. B. Primary consumer δ13C and δ15N and the trophic position of aquatic consumers. Ecology 80, 1395–1404 (1999).
    Article  Google Scholar 

    35.
    Cerling, T. E., Harris, J. M. & Leakey, M. G. Browsing and grazing in elephants: The isotope record of modern and fossil proboscideans. Oecologia 120, 364–374 (1999).
    ADS  PubMed  Article  Google Scholar 

    36.
    Tipple, B. J. & Pagani, M. The early origins of terrestrial C4 photosynthesis. Annual Review of Earth and Planetary Sciences (2007).

    37.
    Takeda, T., Ueno, O., Samejima, M. & Ohtani, T. An investigation for the occurrence of C4 photosynthesis in the Cyperaceae from Australia. Bot. Mag. Tokyo 98, 393–411 (1985).
    Article  Google Scholar 

    38.
    Hyodo, F., Kohzu, A. & Tayasu, I. Linking aboveground and belowground food webs through carbon and nitrogen stable isotope analyses. Ecol. Res. 25, 745–756 (2010).
    CAS  Article  Google Scholar 

    39.
    Hishi, T., Hyodo, F., Saitoh, S. & Takeda, H. The feeding habits of collembola along decomposition gradients using stable carbon and nitrogen isotope analyses. Soil Biol. Biochem. 39, 1820–1823 (2007).
    CAS  Article  Google Scholar 

    40.
    Suehiro, W. et al. Radiocarbon analysis reveals expanded diet breadth associates with the invasion of a predatory ant. Sci. Rep. 7, 1–10 (2017).
    CAS  Article  Google Scholar 

    41.
    Tillberg, C. V., Holway, D. A., LeBrun, E. G. & Suarez, A. V. Trophic ecology of invasive Argentine ants in their native and introduced ranges. Proc. Natl. Acad. Sci. U. S. Am. 104, 20856–20861 (2007).
    ADS  CAS  Article  Google Scholar 

    42.
    Roeder, K. A. & Kaspari, M. From cryptic herbivore to predator: Stable isotopes reveal consistent variability in trophic levels in an ant population. Ecology 98, 297–303 (2017).
    PubMed  Article  Google Scholar 

    43.
    Post, D. M. et al. Getting to the fat of the matter: models, methods and assumptions for dealing with lipids in stable isotope analyses. Oecologia 152, 179–189 (2007).
    ADS  PubMed  Article  Google Scholar 

    44.
    Tayasu, I., Hirasawa, R., Ogawa, N. O., Ohkouchi, N. & Yamada, K. New organic reference materials for carbon- and nitrogen-stable isotope ratio measurements provided by Center for Ecological Research, Kyoto University, and Institute of Biogeosciences, Japan Agency for Marine-Earth Science and Technology. Limnology 12, 261–266 (2011).
    CAS  Article  Google Scholar 

    45.
    Pettitt-Wade, H., Wellband, K. W. & Fisk, A. T. Inconsistency for the niche breadth invasion success hypothesis in aquatic invertebrates. Limnol. Oceanogr. 63, 144–159 (2018).
    ADS  Article  Google Scholar  More

  • in

    Genetic patterns in Mugil cephalus and implications for fisheries and aquaculture management

    1.
    Garibaldi, L. The FAO global capture production database: A six-decade effort to catch the trend. Mar. Pol. 36, 760–768 (2012).
    Article  Google Scholar 
    2.
    FAO (Food and Agriculture Organization). The State of World Fisheries and Aquaculture 2018. in Meeting The Sustainable Development Goals. (FAO, Rome, 2018).

    3.
    Grant, W. S., Jasper, J., Bekkevold, D. & Adkison, M. Responsible genetic approach to stock restoration, sea ranching and stock enhancement of marine fishes and invertebrates. Rev. Fish Biol. Fish. 27, 615–649 (2017).
    Article  Google Scholar 

    4.
    Christie, M. R., Marine, M. L., French, R. A., Waples, R. S. & Blouin, M. S. Effective size of a wild salmonid population is greatly reduced by hatchery supplementation. Heredity 109, 254–260 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    5.
    Ryman, N. & Laikre, L. Effects of supportive breeding on the genetically effective population size. Conserv. Biol. 5, 325–329 (1991).
    Article  Google Scholar 

    6.
    Waples, R. S., Hindar, K., Karlsson, S. & Hard, J. J. Evaluating the Ryman–Laikre effect for marine stock enhancement and aquaculture. Curr. Zool. 62, 617–627 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    7.
    Sun, X. & Hedgecock, D. Temporal genetic change in North American Pacific oyster populations suggests caution in seascape genetics analyses of high gene-flow species. Mar. Ecol. Prog. Ser. 565, 79–93 (2017).
    ADS  Article  Google Scholar 

    8.
    Bacheler, N. M., Wong, R. A. & Buckel, J. A. Movements and mortality rates of striped mullet in North Carolina. N. Am. J. Fish. Manage. 25, 361–373 (2005).
    Article  Google Scholar 

    9.
    Whitfield, A. K., Panfili, J. & Durand, J.-D. A global review of the cosmopolitan flathead mullet Mugil cephalus Linnaeus 1758 (Teleostei: Mugilidae), with emphasis on the biology, genetics, ecology and fisheries aspects of this apparent species complex. Rev. Fish Biol. Fish. 22, 641–681 (2012).
    Article  Google Scholar 

    10.
    Hsu, C.-C., Chang, C.-W., Iizuka, Y. & Tzeng, W.-N. A growth check deposited at estuarine arrival in otoliths of juvenile flathead mullet (Mugil cephalus L.). Zool. Stud. 48(3), 315–323 (2009).
    Google Scholar 

    11.
    Antuofermo, E. et al. First evidence of intersex condition in extensively reared mullets from Sardinian lagoons (central-western Mediterranean, Italy). Ital. J. Anim. Sci. 16, 283–291 (2017).
    Article  Google Scholar 

    12.
    Heras, S., Roldán, M. I. & Castro, M. G. Molecular phylogeny of Mugilidae fishes revised. Rev. Fish Biol. Fish. 19, 217–231 (2009).
    Article  Google Scholar 

    13.
    Heras, S., Maltagliati, F., Fernández, M. V. & Roldán, M. I. Shaken not stirred: A molecular contribution to the systematics of genus Mugil (Teleostei, Mugilidae). Integr. Zool. 11, 263–281 (2016).
    PubMed  Article  Google Scholar 

    14.
    Shen, K.-N., Jamandre, B. W., Hsu, C.-C., Tzeng, W.-N. & Durand, J.-D. Plio-Pleistocene sea level and temperature fluctuations in the northwestern Pacific promoted speciation in the globally-distributed flathead mullet Mugil cephalus. BMC Evol. Biol. 11, 83. https://doi.org/10.1186/1471-2148-11-83 (2011).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    15.
    Durand, J.-D. et al. Systematics of the grey mullets (Teleostei: Mugiliformes: Mugilidae): Molecular phylogenetic evidence challenges two centuries of morphology-based taxonomy. Mol. Phylogenet. Evol. 64, 73–92 (2012).
    PubMed  Article  Google Scholar 

    16.
    Rossi, A. R., Capula, M., Crosetti, D., Campton, D. E. & Sola, L. Genetic divergence and phylogenetic inferences in five species of Mugilidae (Pisces: Perciformes). Mar. Biol. 131, 213–218 (1998).
    CAS  Article  Google Scholar 

    17.
    Blel, H. et al. Selection footprint at the first intron of the Prl gene in natural populations of the flathead mullet (Mugil cephalus, L. 1758). J. Exp. Mar. Biol. Ecol. 387, 60–67 (2010).
    CAS  Article  Google Scholar 

    18.
    Durand, J., Blel, H., Shen, K., Koutrakis, E. & Guinand, B. Population genetic structure of Mugil cephalus in the Mediterranean and Black Seas: A single mitochondrial clade and many nuclear barriers. Mar. Ecol.-Prog. Ser. 474, 243–261 (2013).
    ADS  Article  Google Scholar 

    19.
    Šegvić-Bubić, T. et al. Range expansion of the non-native oyster Crassostrea gigas in the Adriatic Sea. Acta Adriat. 57(2), 321–330 (2016).
    Google Scholar 

    20.
    Piras, P. et al. A case study on the labeling of bottarga produced in Sardinia from ovaries of grey mullets (Mugil cephalus and Mugil capurrii) caught in Eastern Central Atlantic coasts. Ital. J. Food. Saf. 7(1), 6893. https://doi.org/10.4081/ijfs.2018.6893 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    21.
    Miggiano, E. et al. Isolation and characterization of microsatellite loci in the striped mullet, Mugil cephalus. Mol. Ecol. Notes 5, 323–326 (2005).
    CAS  Article  Google Scholar 

    22.
    Jinliang, W. A parsimony estimator of the number of populations from a STRUCTURE‐like analysis. Mol. Ecol. Resour. 19, 970–981 (2019).
    Article  CAS  Google Scholar 

    23.
    FAO (Food and Agriculture Organization). Code of Conduct for Responsible Fisheries (FAO, Rome, 1995).
    Google Scholar 

    24.
    Mai, A. C. G. et al. Microsatellite variation and genetic structuring in Mugil liza (Teleostei: Mugilidae) populations from Argentina and Brazil. Estuar. Coast. Shelf Sci. 149, 80–86 (2014).
    ADS  Article  Google Scholar 

    25.
    Pacheco-Almanzar, E., Simons, J., Espinosa-Perez, H., Chiappa-Carrara, X. & Ibanez, A. L. Can the name Mugil cephalus (Pisces: Mugilidae) be used for the species occurring in the north western Atlantic?. Zootaxa 4109, 381–390 (2016).
    PubMed  Article  Google Scholar 

    26.
    Waples, R. S. & Gaggiotti, O. What is a population? An empirical evaluation of some genetic methods for identifying the number of gene pools and their degree of connectivity. Mol. Ecol. 15, 1419–1439 (2006).
    CAS  PubMed  Article  Google Scholar 

    27.
    Hauser, L. & Carvalho, G. R. Paradigm shifts in marine fisheries genetics: Ugly hypotheses slain by beautiful facts. Fish Fish. 9, 333–362 (2008).
    Article  Google Scholar 

    28.
    Waples, R. S. & England, P. R. Estimating contemporary effective population size on the basis of linkage disequilibrium in the face of migration. Genetics 189, 633–644 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    29.
    Murenu, M., Olita, A., Sabatini, A., Follesa, M. C. & Cau, A. Dystrophy effects on the Liza ramada (Risso, 1826) (Pisces, Mugilidae) population in the Cabras lagoon (Central-Western Sardinia). Chem. Ecol. 20, 425–433 (2004).
    Article  Google Scholar 

    30.
    Ryman, N., Laikre, L. & Hössjer, O. Do estimates of contemporary effective population size tell us what we want to know?. Mol. Ecol. 28, 1904–1918 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    31.
    Guinand, B. et al. Candidate gene variation in gilthead sea bream reveals complex spatiotemporal selection patterns between marine and lagoon habitats. Mar. Ecol. Prog. Ser. 558, 115–127 (2016).
    ADS  CAS  Article  Google Scholar 

    32.
    Chaoui, L. et al. Microsatellite length variation in candidate genes correlates with habitat in the gilthead sea bream Sparus aurata. Mol. Ecol. 21, 5497–5511 (2012).
    CAS  PubMed  Article  Google Scholar 

    33.
    González-Wangüemert, M. & Pérez-Ruzafa, Á. In two waters: contemporary evolution of lagoonal and marine white seabream (Diplodus sargus) populations. Mar. Ecol. 33, 337–349 (2012).
    ADS  Article  Google Scholar 

    34.
    Cardona, L. Effects of salinity on the habitat selection and growth performance of Mediterranean flathead grey mullet Mugil cephalus (Osteichthyes, Mugilidae). Estuar. Coast. Shelf Sci. 50, 727–737 (2000).
    ADS  Article  Google Scholar 

    35.
    Fortunato, R. C., Galán, A. R., Alonso, I. G., Volpedo, A. & Durà, V. B. Environmental migratory patterns and stock identification of Mugil cephalus in the Spanish Mediterranean Sea, by means of otolith microchemistry. Estuar. Coast. Shelf. Sci. 188, 174–180 (2017).
    ADS  Article  CAS  Google Scholar 

    36.
    Jones, A. G., Small, C. M., Paczolt, K. A. & Ratterman, N. L. A practical guide to methods of parentage analysis. Mol. Ecol. Resour. 10, 6–30 (2010).
    PubMed  Article  Google Scholar 

    37.
    Taylor, H. R. The use and abuse of genetic marker-based estimates of relatedness and inbreeding. Ecol. Evol. 5, 3140–3150 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    38.
    Coppinger, C. R. et al. Assessing the genetic diversity of catface grouper Epinephelus andersoni in the subtropical Western Indian Ocean. Fish. Res. 218, 186–197 (2019).
    Article  Google Scholar 

    39.
    Cushman, E. L. et al. Development of a standardized molecular tool and estimation of genetic measures for responsible aquaculture-based fisheries enhancement of American Shad in North and South Carolina. Trans. Am. Fish. Soc. 148, 148–162 (2019).
    Article  Google Scholar 

    40.
    Waples, R. S., Punt, A. E. & Cope, J. M. Integrating genetic data into management of marine resources: How can we do it better?. Fish. Fish. 9, 423–449 (2008).
    Article  Google Scholar 

    41.
    Iacchei, M. et al. Combined analyses of kinship and FST suggest potential drivers of chaotic genetic patchiness in high gene-flow populations. Mol. Ecol. 22, 3476–3494 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    42.
    Bernardi, G., Beldade, R., Holbrook, S. J. & Schmitt, R. J. Full-sibs in cohorts of newly settled coral reef fishes. PLoS ONE 7(e44953), 2012. https://doi.org/10.1371/journal.pone.0044953 (2012).
    CAS  Article  Google Scholar 

    43.
    Como, S., van der Velde, G. & Magni, P. Temporal variation in the trophic levels of secondary consumers in a Mediterranean coastal lagoon (Cabras lagoon, Italy). Estuaries Coasts 41, 218–232 (2018).
    CAS  Article  Google Scholar 

    44.
    Floris, R., Manca, S. & Fois, N. Microbial ecology of intestinal tract of gilthead sea bream (Sparus aurata Linnaeus, 1758) from two coastal lagoons of Sardinia (Italy). Transit. Waters Bullet. 7(2), 4–12. https://doi.org/10.1285/i1825229Xv7n2p4 (2013).
    Article  Google Scholar 

    45.
    Merella, P. & Garippa, G. Metazoan parasites of grey mullets (Teleostea: Mugilidae) from the Mistras lagoon (Sardinia-Western Mediterranean). Sci. Mar. 65, 201–206 (2001).
    Article  Google Scholar 

    46.
    Pitacco, V. et al. Spatial patterns of macrobenthic alpha and beta diversity at different scales in Italian transitional waters (Central Mediterranean). Estuar. Coast. Shelf Sci. 222, 126–138 (2019).
    ADS  Article  Google Scholar 

    47.
    Cioffi, F. & Gallerano, F. From rooted to floating vegetal species in lagoons as a consequence of the increases of external nutrient load: An analysis by model of the species selection mechanism. Appl. Math. Model. 30, 10–37 (2006).
    MATH  Article  Google Scholar 

    48.
    Wasko, A. P., Martins, C., Oliveira, C. & Foresti, F. Non-destructive genetic sampling in fish. An improved method for DNA extraction from fish fins and scales. Hereditas 138, 161–165 (2003).
    PubMed  Article  Google Scholar 

    49.
    Waples, R. S. Testing for Hardy–Weinberg proportions: Have we lost the plot?. J. Hered. 106, 1–19 (2015).
    PubMed  Article  Google Scholar 

    50.
    Rousset, F. genepop’007: A complete re-implementation of the genepop software for Windows and Linux. Mol. Ecol. Resour. 8, 103–106 (2008).
    PubMed  Article  Google Scholar 

    51.
    Kinnison, M. T., Bentzen, P., Unwin, M. J. & Quinn, T. P. Reconstructing recent divergence: Evaluating nonequilibrium population structure in New Zealand chinook salmon. Mol. Ecol. 11, 739–754 (2002).
    CAS  PubMed  Article  Google Scholar 

    52.
    Benjamini, Y. & Yekutieli, D. The control of the false discovery rate in multiple testing under dependency. Ann. Stat. 29(4), 1165–1188 (2001).
    MathSciNet  MATH  Article  Google Scholar 

    53.
    R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, 2018).
    Google Scholar 

    54.
    Cossu, P. et al. Influence of genetic drift on patterns of genetic variation: The footprint of aquaculture practices in Sparus aurata (Teleostei: Sparidae). Mol. Ecol. 28, 3012–3024 (2019).
    PubMed  Article  Google Scholar 

    55.
    Van Oosterhout, C., Hutchinson, W. F., Wills, D. P. M. & Shipley, P. Micro-checker: software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes 4, 535–538 (2004).
    Article  CAS  Google Scholar 

    56.
    Chapuis, M.-P. & Estoup, A. Microsatellite null alleles and estimation of population differentiation. Mol. Biol. Evol. 24, 621–631 (2007).
    CAS  PubMed  Article  Google Scholar 

    57.
    Dąbrowski, M. J. et al. Reliability assessment of null allele detection: Inconsistencies between and within different methods. Mol. Ecol. Resour. 14, 361–373 (2014).
    PubMed  Article  CAS  Google Scholar 

    58.
    Foll, M. & Gaggiotti, O. A genome-scan method to identify selected loci appropriate for both dominant and codominant Markers: A Bayesian perspective. Genetics 180, 977–993 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    59.
    Kauer, M. O., Dieringer, D. & Schlötterer, C. A microsatellite variability screen for positive selection associated with the “Out of Africa” habitat expansion of drosophila melanogaster. Genetics 165, 1137–1148 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    60.
    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 

    61.
    Paris, M. et al. Genome scan in the mosquito Aedes rusticus: population structure and detection of positive selection after insecticide treatment. Mol. Ecol. 19, 325–337 (2010).
    PubMed  Article  Google Scholar 

    62.
    Keenan, K., McGinnity, P., Cross, T. F., Crozier, W. W. & Prodöhl, P. A. diveRsity: An R package for the estimation and exploration of population genetics parameters and their associated errors. Methods Ecol. Evol. 4, 782–788 (2013).
    Article  Google Scholar 

    63.
    Piry, S., Luikart, G. & Cornuet, J.-M. Computer note. BOTTLENECK: A computer program for detecting recent reductions in the effective size using allele frequency data. J. Hered. 90, 502–503 (1999).
    Article  Google Scholar 

    64.
    Do, C. et al. NeEstimator v2: Re-implementation of software for the estimation of contemporary effective population size (Ne) from genetic data. Mol. Ecol. Resour. 14, 209–214 (2014).
    CAS  PubMed  Article  Google Scholar 

    65.
    Waples, R. S. & Do, C. Linkage disequilibrium estimates of contemporary Ne using highly variable genetic markers: A largely untapped resource for applied conservation and evolution. Evol. Appl. 3, 244–262 (2010).
    PubMed  Article  Google Scholar 

    66.
    Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analysis of population structure. Evolution 38, 1358–1370 (1984).
    CAS  PubMed  Google Scholar 

    67.
    Jost, L. GST and its relatives do not measure differentiation. Mol. Ecol. 17, 4015–4026 (2008).
    PubMed  Article  Google Scholar 

    68.
    Ryman, N. & Palm, S. POWSIM: A computer program for assessing statistical power when testing for genetic differentiation. Mol. Ecol. Notes 6, 600–602 (2006).
    Article  Google Scholar 

    69.
    Blouin, M. S., Parsons, M., Lacaille, V. & Lotz, S. Use of microsatellite loci to classify individuals by relatedness. Mol. Ecol. 5, 393–401 (1996).
    CAS  PubMed  Article  Google Scholar 

    70.
    Kraemer, P. & Gerlach, G. Demerelate: Calculating interindividual relatedness for kinship analysis based on codominant diploid genetic markers using R. Mol. Ecol. Resour. 17, 1371–1377 (2017).
    CAS  PubMed  Article  Google Scholar 

    71.
    Kalinowski, S. T., Wagner, A. P. & Taper, M. L. ml-relate: A computer program for maximum likelihood estimation of relatedness and relationship. Mol. Ecol. Notes 6, 576–579 (2006).
    CAS  Article  Google Scholar 

    72.
    Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959 (2000).
    CAS  PubMed  PubMed Central  Google Scholar 

    73.
    Falush, D., Stephens, M. & Pritchard, J. K. Inference of population structure using multilocus genotype data: Linked loci and correlated allele frequencies. Genetics 164, 1567–1587 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    74.
    Francis, R. M. pophelper: An R package and web app to analyse and visualize population structure. Mol. Ecol. Resour. 17, 27–32 (2017).
    CAS  PubMed  Article  Google Scholar 

    75.
    Jombart, T., Devillard, S. & Balloux, F. Discriminant analysis of principal components: A new method for the analysis of genetically structured populations. BMC Genet. 11, 94. https://doi.org/10.1186/1471-2156-11-94 (2010).
    Article  PubMed  PubMed Central  Google Scholar 

    76.
    Jombart, T. adegenet: A R package for the multivariate analysis of genetic markers. Bioinformatics 24, 1403–1405 (2008).
    CAS  PubMed  Article  Google Scholar  More