More stories

  • in

    Association of zoonotic protozoan parasites with microplastics in seawater and implications for human and wildlife health

    Jambeck, J. R. et al. Plastic waste inputs from land into the ocean. Science 347, 768–771 (2015).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Andrady, A. L. Microplastics in the marine environment. Mar. Pollut. Bull. 62, 1596–1605 (2011).CAS 
    Article 
    PubMed 

    Google Scholar 
    Avio, C. G., Gorbi, S. & Regoli, F. Plastics and microplastics in the oceans: From emerging pollutants to emerged threat. Mar. Environ. Res. 128, 2–11 (2017).CAS 
    Article 
    PubMed 

    Google Scholar 
    Barboza, L. G. A., Dick Vethaak, A., Lavorante, B. R. B. O., Lundebye, A.-K. & Guilhermino, L. Marine microplastic debris: An emerging issue for food security, food safety and human health. Mar. Pollut. Bull. 133, 336–348 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Van Cauwenberghe, L. & Janssen, C. R. Microplastics in bivalves cultured for human consumption. Environ. Pollut. 193, 65–70 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bucci, K., Tulio, M. & Rochman, C. M. What is known and unknown about the effects of plastic pollution: A meta-analysis and systematic review. Ecol. Appl. 30, e02044 (2020).CAS 
    Article 
    PubMed 

    Google Scholar 
    Worm, B., Lotze, H. K., Jubinville, I., Wilcox, C. & Jambeck, J. Plastic as a Persistent Marine Pollutant. Annu. Rev. Environ. Resour. 42, 1–26 (2017).Article 

    Google Scholar 
    GESAMP. Sources, Fate and Effects of Microplastics in the Marine Environment (Part 2) (2016). http://www.gesamp.org/publications/microplastics-in-the-marine-environment-part-2.Donohue, M. J. et al. Evaluating exposure of northern fur seals, Callorhinus ursinus, to microplastic pollution through fecal analysis. Mar. Pollut. Bull. 138, 213–221 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    Duncan, E. M. et al. Microplastic ingestion ubiquitous in marine turtles. Glob. Change Biol. 25, 744–752 (2019).ADS 
    Article 

    Google Scholar 
    Moore, R. C. et al. Microplastics in beluga whales (Delphinapterus leucas) from the Eastern Beaufort Sea. Mar. Pollut. Bull. 150, 110723 (2020).CAS 
    Article 
    PubMed 

    Google Scholar 
    Bessa, F. et al. Microplastics in gentoo penguins from the Antarctic region. Sci. Rep. 9, 14191 (2019).ADS 
    Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Provencher, J. F., Ammendolia, J., Rochman, C. M. & Mallory, M. L. Assessing plastic debris in aquatic food webs: what we know and don’t know about uptake and trophic transfer. Environ. Rev. 27, 304–317 (2019).Article 

    Google Scholar 
    Bucci, K., Bikker, J., Stevack, K., Watson-Leung, T. & Rochman, C. Impacts to larval fathead minnows vary between preconsumer and environmental microplastics. Environ. Toxicol. Chem. 41, 4 (2021).
    Google Scholar 
    Nelms, S. E., Galloway, T. S., Godley, B. J., Jarvis, D. S. & Lindeque, P. K. Investigating microplastic trophic transfer in marine top predators. Environ. Pollut. 238, 999–1007 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    De-la-Torre, G. E. Microplastics: an emerging threat to food security and human health. J. Food Sci. Technol. 57, 1601–1608 (2020).Article 
    PubMed 

    Google Scholar 
    Teuten, E. L. et al. Transport and release of chemicals from plastics to the environment and to wildlife. Philos. Trans. R. Soc. B 364, 2027–2045 (2009).CAS 
    Article 

    Google Scholar 
    Zettler, E. R., Mincer, T. J. & Amaral-Zettler, L. A. Life in the “plastisphere”: Microbial communities on plastic marine debris. Environ. Sci. Technol. 47, 7137–7146 (2013).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    He, S. et al. Biofilm on microplastics in aqueous environment: Physicochemical properties and environmental implications. J. Hazard. Mater. 1, 127286. https://doi.org/10.1016/j.jhazmat.2021.127286 (2021).CAS 
    Article 

    Google Scholar 
    Kirstein, I. V. et al. Dangerous hitchhikers? Evidence for potentially pathogenic Vibrio spp. on microplastic particles. Mar. Environ. Res. 120, 1–8 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    World Health Organization. Safe Management of Shellfish and Harvest Waters (WHO, 2010).
    Google Scholar 
    Lindsay, D. S. & Dubey, J. P. Long-term survival of Toxoplasma gondii sporulated oocysts in seawater. J. Parasitol. 95, 1019–1020 (2009).Article 
    PubMed 

    Google Scholar 
    Tamburrini, A. & Pozio, E. Long-term survival of Cryptosporidium parvum oocysts in seawater and in experimentally infected mussels (Mytilus galloprovincialis). Int. J. Parasitol. 29, 711–715 (1999).CAS 
    Article 
    PubMed 

    Google Scholar 
    Jones, J. L. et al. Risk factors for Toxoplasma gondii infection in the United States. Clin. Infect. Dis. 49, 878–884 (2009).Article 
    PubMed 

    Google Scholar 
    Robertson, L. J. The potential for marine bivalve shellfish to act as transmission vehicles for outbreaks of protozoan infections in humans: A review. Int. J. Food Microbiol. 120, 201–216 (2007).CAS 
    Article 
    PubMed 

    Google Scholar 
    Shapiro, K. et al. Environmental transmission of Toxoplasma gondii: Oocysts in water, soil and food. Food Waterb. Parasitol. 15, e00049 (2019).Article 

    Google Scholar 
    Miller, M. A., Shapiro, K., Murray, M. J., Haulena, M. J. & Raverty, S. Protozoan parasites of marine mammals. in CRC Handbook of Marine Mammal Medicine (2018).Ward, J. E. & Kach, D. J. Marine aggregates facilitate ingestion of nanoparticles by suspension-feeding bivalves. Mar. Environ. Res. 68, 137–142 (2009).CAS 
    Article 
    PubMed 

    Google Scholar 
    Rose, J. B. Environmental ecology of cryptosporidium and public health implications. Annu. Rev. Public Health 18, 135–161 (1997).CAS 
    Article 
    PubMed 

    Google Scholar 
    Robert-Gangneux, F. & Dardé, M.-L. Epidemiology of and diagnostic strategies for toxoplasmosis. Clin. Microbiol. Rev. 25, 264–296 (2012).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bahia-Oliveira, L., Gomez-Marin, J. & Shapiro, K. Toxoplasma gondii. Global Water Pathogen Project. https://www.waterpathogens.org/book/toxoplasma-gondii (2015).Kreuder, C. et al. Patterns of mortality in southern sea otters (Enhydra lutris nereis) from 1998–2001. J. Wildl. Dis. 39, 495–509 (2003).CAS 
    Article 
    PubMed 

    Google Scholar 
    Shapiro, K. et al. Dual congenital transmission of Toxoplasma gondii and Sarcocystis neurona in a late-term aborted pup from a chronically infected southern sea otter (Enhydra lutris nereis). Parasitology 143, 276–288 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Barbieri, M. M. et al. Protozoal-related mortalities in endangered Hawaiian monk seals Neomonachus schauinslandi. Dis. Aquat. Org. 121, 85–95 (2016).Article 

    Google Scholar 
    Roe, W. D., Howe, L., Baker, E. J., Burrows, L. & Hunter, S. A. An atypical genotype of Toxoplasma gondii as a cause of mortality in Hector’s dolphins (Cephalorhynchus hectori). Vet. Parasitol. 192, 67–74 (2013).CAS 
    Article 
    PubMed 

    Google Scholar 
    Hernandez, E., Nowack, B. & Mitrano, D. M. Polyester textiles as a source of microplastics from households: A mechanistic study to understand microfiber release during washing. Environ. Sci. Technol. 51, 7036–7046 (2017).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Mason, S. A. et al. Microplastic pollution is widely detected in US municipal wastewater treatment plant effluent. Environ. Pollut. 218, 1045–1054 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Sutton, R. et al. Microplastic contamination in the San Francisco Bay, California, USA. Mar. Pollut. Bull. 109, 230–235 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Desforges, J.-P.W., Galbraith, M., Dangerfield, N. & Ross, P. S. Widespread distribution of microplastics in subsurface seawater in the NE Pacific Ocean. Mar. Pollut. Bull. 79, 94–99 (2014).CAS 
    Article 
    PubMed 

    Google Scholar 
    Horn, D., Miller, M., Anderson, S. & Steele, C. Microplastics are ubiquitous on California beaches and enter the coastal food web through consumption by Pacific mole crabs. Mar. Pollut. Bull. 139, 231–237 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    Yu, X. et al. Occurrence and distribution of microplastics at selected coastal sites along the southeastern United States. Sci. Total Environ. 613–614, 298–305 (2018).ADS 
    Article 
    CAS 
    PubMed 

    Google Scholar 
    Collicutt, B., Juanes, F. & Dudas, S. E. Microplastics in juvenile Chinook salmon and their nearshore environments on the east coast of Vancouver Island. Environ. Pollut. 244, 135–142 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    Davidson, K. & Dudas, S. E. Microplastic ingestion by wild and cultured manila clams (Venerupis philippinarum) from Baynes Sound, British Columbia. Arch. Environ. Contam. Toxicol. 71, 147–156 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Waite, H. R., Donnelly, M. J. & Walters, L. J. Quantity and types of microplastics in the organic tissues of the eastern oyster Crassostrea virginica and Atlantic mud crab Panopeus herbstii from a Florida estuary. Mar. Pollut. Bull. 129, 179–185 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Wootton, N., Reis-Santos, P. & Gillanders, B. M. Microplastic in fish: A global synthesis. Rev. Fish. Biol. Fish. 31, 753–771 (2021).Article 

    Google Scholar 
    De-la-Pinta, I. et al. Effect of biomaterials hydrophobicity and roughness on biofilm development. J. Mater. Sci. 30, 77 (2019).
    Google Scholar 
    Rochman, C. M., Hoh, E., Hentschel, B. T. & Kaye, S. Long-term field measurement of sorption of organic contaminants to five types of plastic pellets: implications for plastic marine debris. Environ. Sci. Technol. 47, 1646–1654 (2013).CAS 
    PubMed 

    Google Scholar 
    Lindquist, H. D. A. et al. Autofluorescence of Toxoplasma gondii and related coccidian oocysts. J. Parasitol. 89, 865–867 (2003).Article 
    PubMed 

    Google Scholar 
    Alldredge, A. L., Passow, U. & Logan, B. E. The abundance and significance of a class of large, transparent organic particles in the ocean. Deep Sea Res. Part I 40, 1131–1140 (1993).CAS 
    Article 

    Google Scholar 
    Shapiro, K. et al. Aquatic polymers can drive pathogen transmission in coastal ecosystems. Proc. R. Soc. B 281, 20141287 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bowley, J., Baker-Austin, C., Porter, A., Hartnell, R. & Lewis, C. Oceanic hitchhikers: Assessing pathogen risks from marine microplastic. Trends Microbiol. 29, 107–116 (2021).CAS 
    Article 
    PubMed 

    Google Scholar 
    Nasser, F. & Lynch, I. Secreted protein eco-corona mediates uptake and impacts of polystyrene nanoparticles on Daphnia magna. J. Proteom. 137, 45–51 (2016).CAS 
    Article 

    Google Scholar 
    Savoca, M. S., Wohlfeil, M. E., Ebeler, S. E. & Nevitt, G. A. Marine plastic debris emits a keystone infochemical for olfactory foraging seabirds. Sci. Adv. 2, e1600395 (2016).ADS 
    Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ramsperger, A. F. R. M. et al. Environmental exposure enhances the internalization of microplastic particles into cells. Sci. Adv. 6, 1211 (2020).ADS 
    Article 
    CAS 

    Google Scholar 
    Lusher, A., Hollman, P. C. H. & Mendoza-Hill, J. Microplastics in fisheries and aquaculture: status of knowledge on their occurrence and implications for aquatic organisms and food safety (Food and Agriculture Organization of the United Nations, 2017).
    Google Scholar 
    Tamburri, M. N. & Zimmer-Faust, R. K. Suspension feeding: Basic mechanisms controlling recognition and ingestion of larvae. Limnol. Oceanogr. 41, 1188–1197 (1996).ADS 
    Article 

    Google Scholar 
    Shapiro, K. et al. Simultaneous detection of four protozoan parasites on leafy greens using a novel multiplex PCR assay. Food Microbiol. 84, 103252 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    Choy, C. A. et al. The vertical distribution and biological transport of marine microplastics across the epipelagic and mesopelagic water column. Sci. Rep. 9, 7843 (2019).ADS 
    Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Saley, A. M. et al. Microplastic accumulation and biomagnification in a coastal marine reserve situated in a sparsely populated area. Mar. Pollut. Bull. 146, 54–59 (2019).CAS 
    Article 
    PubMed 

    Google Scholar 
    Shapiro, K. et al. Detection of Toxoplasma gondii oocysts and surrogate microspheres in water using ultrafiltration and capsule filtration. Water Res. 44, 893–903 (2010).CAS 
    Article 
    PubMed 

    Google Scholar  More

  • in

    Role of trade agreements in the global cereal market and implications for virtual water flows

    Link activationContingency tables corresponding to the three cases described in the “Methods” section are shown in Table 1. This Table is quite revealing in several ways. The most interesting aspect is that the highest probability of link establishment occurs when an agreement is activated (Operational Activation in t).Table 1 Contingency tables.Full size tableIn this case, the probability of activation of a new link is 8.8%—namely, the ratio of new activation 7.3% to the total number of links that were not active at year t-1 (82.6%)—which is significantly higher than in the case of links not covered by a commercial agreement (No Trade Agreement), amounting to 1.4%.Therefore, the findings show that operational activation is associated with creating new trade relations between two particular countries. The third set, which considers links where a trade agreement exists in both years (t-1) and t (Trade Agreement in t-1 and t), also shows a consistent activation probability of 6%. This result confirms the assumption that the coverage of a commercial agreement, and not only its implementation, encourages the genesis of new links.Moreover, Table 1 suggests some interesting considerations on trade persistence. To establish these probabilities, we focus on the row totals in which a trade relationship is present at year (t-1), i.e., 28.8% in the case Trade Agreement in t-1 and t. The presence of an agreement influences in a positive way the probability of maintaining a trade relationship. In fact, when a trade agreement is present in both years, (t-1) and t, the probability of preserving the trade relationship is 87.1% ((frac{25.1}{28.8}times {100})), while when a trade agreement is activated at year t, the probability slightly decreases to 81.6%. In cases where trade agreements are missing (No Trade Agreement in t) we observe the probability of retaining a relationship decreases to 77.3%.Another interesting aspect concerns the probability of link deactivation. Once more, the coverage of a trade agreement favors a lower likelihood of deactivation of existing links. The ratio of the percentage of links that were active at year (t-1) and are no more active at year t to the total is 22.7% ((frac{1}{4.4}times {100})) in the case of a lack of agreement. This probability decreases to 18.4% ((frac{3.2}{17.4}times {100})) if we consider only the year of activation of the agreement (Operational Activation), and drops to 12.8% ((frac{3.7}{28.8}times {100})) when looking at agreements present in both years.Together, these results provide insights into the role of trade agreements in the network topology of cereal trade. While the establishment of a trade agreement promotes the potential for new trade links, the presence of the agreement in two consecutive years allows both to maintain an existing relationship and reduce the likelihood of link shutdowns.Flow variationsIn this second part, we study the impact of trade agreements on existing trade flows, analyzing the relationship between the flows at time t and the flows at time (t-1) in each of the three cases described in the “Methods” section—i.e., No trade agreements, Operational Activation in t, and Trade agreement in t-1 and t—measured in US$, Kcal and m(^3) of virtual water.Figure 3Kernel Density scatterplot between trade flows of cereals at time t (on the y-axis) and time (t-1) (on the x-axis) for the three different sets: No trade agreements (column a), Operational Activation in t (b), and Trade agreement in (t-1) and t (c). Panels in the first, second and third row refer to flows in US$, Kcal, and virtual water (m(^3)), respectively. Flow values are shown on a logarithmic scale. The color bar indicates probability densities, and the bisector is highlighted. Notice (i) the higher volumes in the case of flows covered by trade agreement and (ii) a a less relevant increase in volume when the flows are seen in the virtual water lens.Full size imageFigure 3 shows three different scatterplots for each unit of measure (US$ and Kcal and m(^3)). The scatterplots are colored by Kernel Density Estimation (KDE), a non-parametric technique for probability density functions. KDE aims to take a finite sample of data and infer the underlying probability density function. Figure 3 relates the flows at time (t-1) with the flows at time t, both reported on a logarithmic scale since the quantities span several orders of magnitude. Let’s start focusing on flows in terms of dollars and kilocalories. What stands out from the figure is the displacement of the flows toward higher values when they are covered by trade agreements (Trade Agreement in t-1 and t), compared to the case where flows have no trade agreement.We have quantitative evidence of this result by looking at Table 2 where the average flows in both years are shown. The average values of flows in both US$ and Kcal are much higher when there is a trade agreement over time (Trade agreement in t-1 and t). Flows have an average value of (6.13times 10^{7})$, larger than the mean of (3.05times 10^{7})$ achieved by flows not covered by a trade agreement. By comparing the distributions of the two distinct sets with different dimensions by applying the non-parametric Mann-Whitney test, we stand to evaluate this result as extremely significant (p-value approximately 0).Table 2 Average values of trade flows and flow variation index (rho _{ij}) for each of the three sets, in US$ (a), Kcal (b), and Virtual water (VW, m(^3)). The bar indicates the average operator.Full size tableAlso, while operational activation plays a crucial role in creating new links in the global cereal trade, it does not appear to hold central importance in driving flow increases. The average value of flows in both years (t-1) and t are, in fact, smaller than those not covered by trade agreements.The view appears slightly different when we look at the values in terms of virtual water (VW, m(^3)), i.e., the sum of the blue and green components. Flows with a commercial agreement show higher averages values than those not covered by agreements (see panel (c) of Table 2), but the increase is significantly lower than the one recorded in the other two units (US$ and Kcal). The increase recorded in dollars is about 100%, while in terms of virtual water this increase is less than 30%. In the next subsection, we will focus on this peculiar behavior, which reveals a different water content of the goods traded along links covered or not by agreements.Another significant result that emerges from Fig. 3 is the smaller amplitude (around the bisector) of the cloud in the case of link covered by agreements in both years (t-1) and t. This is confirmed by comparing the weighted average of the absolute value of the inter-annual flow variation index (overline{rho _{ij}}_{w}) (weights are the flows traded in the year (t-1)). The index (rho _{ij}) is used to highlight cases where the activation or the presence of the agreement generates a significant flow increase.Larger (rho _{ij}) values correspond to larger average variations from year (t-1) to year t. Accordingly, we observe that in the presence of trade agreement at time (t-1) and t a smaller (rho _{ij}) value of 24.79 percentage points (p.p) is found (see panel (a) of Table 2).Considering all the units (US$, Kcal, and m(^3)), this value is about half of the average inter-annual variation that occurs when there is no trade agreement. Hence, the presence of a commercial agreement over time reduces large fluctuations, stabilizing the year-to-year variations.To shed light on the response of water flows to the occurrence of the agreement, we refer to water productivity (WP)34, both in economic and nutritional terms. Table 3 shows that the Nutritional WP for the total virtual water is, on average, 35% higher in the flows under a trade agreement than in flows that are not under any treaty, while the Economic WP is 62% higher. We also analyze the two virtual water components, blue and green, separately.Interestingly, for blue water in the presence of a trade agreement, the Nutritional WP and the Economic WP for the flows covered by trade agreement are, on average, 68% and 93% higher than for the flows not covered by agreements. In other words, for one cubic meter of water used for grain production, more kilo-calories and dollars are exchanged when an agreement is in place, and this difference is even more significant in terms of blue water.Table 3 Average of nutritional ((mathrm {kcal/m^3})) and economic ((mathrm {US$/m^3})) water productivity (WP) for the total, blue and green virtual water.Full size tableWe also investigate in detail which products contribute most to the imbalance between flows in terms of kcal or water. To this aim, Fig. 4 reports the nutritional WP for each grain item distinguishing whether or not there is a commercial agreement (similar results occur if the economic WP is considered).The figure highlights that the nutritional WP is generally higher in the case where flows are covered by trade agreements (green bars). The most noticeable cases are Maize and Wheat, which are also the most traded products: the value of nutritional WP increases from 1978 (mathrm {kcal/m^3}) (No trade agreement) to 2851 (mathrm {kcal/m^3}) in case of a trade agreement for Wheat, and from 4471 (mathrm {kcal/m^3}) to 5026 for Maize.Figure 4The bar chart shows the nutritional WP for each cereal product in the two sets of Trade agreement in t-1 and t (in green) and No trade agreement (in red). The number over the bars represents the percentage of kcal traded for each product compared to the total kcal of all cereals. Note that green bars are higher than the red ones in 80% of cases.Full size imageA few products have a higher nutritional WP value when the flows are not involved in any treaty, e.g., Rye. This behavior can be traced back to a few flows that dominate the market between countries not linked by trade agreements. For example, trade in Rye in 2014 is attributable to just two major flows in terms of caloric intake relative to water quantity (notably, one between Germany and Japan, the other between Russia and Turkey).Figure 4 clearly shows that grains characterized by greater water efficiency generally move along the links covered by agreements.Performance of trade agreements in increasing flowOur results show that links covered by agreements exhibit larger flows than links not covered by treaties. We also intend to obtain information about the possible flow increase under a specific agreement.As mentioned in the “Methods” section, we selected only those operating links when the agreement came into force to evaluate the variation index ((rho _a)) under a specific treaty. Consequently, since there are trade agreements that came into force before the time interval considered, these are excluded from this analysis. As a result, the total number of agreements selected for this analysis is 99, 61 of which show an increase (positive (rho _{a}) values), while the remaining 38 exhibits a decrease in the flux intensities compared to the overall global trend. We present in Table 5 the results for positive (rho _{a}) variations, while trade agreements with negative (rho _{a}) values are reported in Supplementary Material (5). We provide this analysis in terms of economic flows (US$), but very similar results are obtained if calories (kcal) or virtual water (m(^3)) are chosen as the unit of measure.Table 4 Flow values in millions of dollars in year t and percent changes (rho _{a}) from (t-1) to t for each trade agreement.Full size tableWhat stands out in Table 4 is that most of the positive percentage changes occur in Europe and Central Asia regions. This may be due to long-term commercial activities in Europe, which are supported by the geographical proximity of the countries, as well as the wide variety of political and economic treaties among them. Europe, in fact, is characterized by a fourfold increase in cereal production since the 1960s due to the adoption of the Common Agricultural Policy, which has intensified trade in Europe and towards external markets30.A closer inspection of Table 4 shows that among the agreements with the most significant flows that showed the greatest increases, we find EEA (European Economic Area) in Europe and Central Asia, Japan-ASEAN in East Asia and Pacific, and COMESA in Sub-Saharan Africa.With lower flow values but large increases ((rho _{a})) due to the entry into force of trade agreements, the India-Sri Lanka agreement in South Asia stands out above all others. Also, the treaty signed in 2013 between EU-Colombia and Peru shows significant variations in terms of the percentage of flow increase, but the volume of the corresponding flow is inferior when compared with other trade agreements. On the other hand, the North American Free Trade Agreement (NAFTA), which became effective in 1994, has a lower (rho _{a}) value, but the flows on which the variation is calculated are significantly higher. More

  • in

    Identifying biases in the global placement of river gauges

    Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.This is a summary of: Krabbenhoft, C. A. et al. Assessing placement bias of the global river gauge network. Nat. Sustain. https://doi.org/10.1038/s41893-022-00873-0 (2022). More

  • in

    Short and long-term costs of inbreeding in the lifelong-partnership in a termite

    Shellman-Reeve, J. S. Courting strategies and conflicts in a monogamous, biparental termite. Proc. R. Soc. Lond. Ser. B: Biol. Sci. 266, 137–144 (1999).Article 

    Google Scholar 
    Boomsma, J. J. Beyond promiscuity: mate-choice commitments in social breeding. Philos. Trans. R. Soc. B: Biol. Sci. 368 (2013).Nichols, H. J. The causes and consequences of inbreeding avoidance and tolerance in cooperatively breeding vertebrates. J. Zool. 303, 1–14 (2017).Article 

    Google Scholar 
    Clutton-Brock, T. H. Female transfer and inbreeding avoidance in social mammals. Nature 337, 70–72 (1989).CAS 
    Article 
    PubMed 

    Google Scholar 
    Wolff, J. O. Parents suppress reproduction and stimulate dispersal in opposite-sex juvenile white-footed mice. Nature 359, 409–410 (1992).CAS 
    Article 
    PubMed 

    Google Scholar 
    Abbott, D. In Primate Social Conflict (eds W. A. Mason & S. P. Mendoza) 331–372 (State University of New York Press, 1993).Koenig, W. D., Haydock, J. & Stanback, M. T. Reproductive roles in the cooperatively breeding acorn woodpecker: incest avoidance versus reproductive competition. Am. Nat. 151, 243–255 (1998).CAS 
    Article 
    PubMed 

    Google Scholar 
    Hanby, J. P. & Bygott, J. D. Emigration of subadult lions. Anim. Behav. 35, 161–169 (1987).Article 

    Google Scholar 
    Brooked, M. G., Rowley, I., Adams, M. & Baverstock, P. R. Promiscuity: an inbreeding avoidance mechanism in a socially monogamous species? Behav. Ecol. Sociobiol. 26, 191–199 (1990).Article 

    Google Scholar 
    Amos, B., Schlotterer, C. & Tautz, D. Social structure of pilot whales revealed by analytical DNA proftling. Science 260, 670–672 (1993).CAS 
    Article 
    PubMed 

    Google Scholar 
    Sillero-Zubiri, C., Gottelli, D. & Macdonald, D. W. Male philopatry, extra-pack copulations and inbreeding avoidance in Ethiopian wolves (Canis simensis). Behav. Ecol. Sociobiol. 38, 331–340 (1996).Article 

    Google Scholar 
    Husseneder, C., Simms, D. M. & Ring, D. R. Genetic diversity and genotypic differentiation between the sexes in swarm aggregations decrease inbreeding in the Formosan subterranean termite. Insectes Sociaux 53, 212–219 (2006).Article 

    Google Scholar 
    Blouin, S. F. & Blouin, M. Inbreeding avoidance behaviors. Trends Ecol. Evol. 3, 230–233 (1988).CAS 
    Article 
    PubMed 

    Google Scholar 
    Pusey, A. & Wolf, M. Inbreeding avoidance in animals. Trends Ecol. Evol. 11, 201–206 (1996).CAS 
    Article 
    PubMed 

    Google Scholar 
    Gerlach, G. & Lysiak, N. Kin recognition and inbreeding avoidance in zebrafish, Danio rerio, is based on phenotype matching. Anim. Behav. 71, 1371–1377 (2006).Article 

    Google Scholar 
    Hurst, J. L. et al. Individual recognition in mice mediated by major urinary proteins. Nature 414, 631–634 (2001).CAS 
    Article 
    PubMed 

    Google Scholar 
    Vargo, E. L. & Husseneder, C. In Biology of termites: A modern synthesis (eds D.E. Bignell, Yves Roisin, & Nathan Lo) 133–164 (Springer, 2011).Shellman-Reeve, J. S. Dynamics of biparental care in the dampwood termite, Zootermopsis nevadensis (Hagen): response to nitrogen availability. Behav. Ecol. Sociobiol. 26, 389–397 (1990).Article 

    Google Scholar 
    Cole, E. L., Ilieş, I. & Rosengaus, R. B. Competing physiological demands during incipient colony foundation in a social insect: consequences of pathogenic stress. Front. Ecol. Evol. 6 (2018).Traniello, J. F. A., Rosengaus, R. B. & Savoie, K. The development of immunity in a social insect: evidence for the group facilitation of disease resistance. Proc. Natl Acad. Sci. 99, 6838–6842 (2002).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cremer, S., Armitage, S. A. O. & Schmid-Hempel, P. Social immunity. Curr. Biol. 17, R693–R702 (2007).CAS 
    Article 
    PubMed 

    Google Scholar 
    Rosengaus, R. B., Traniello, J. F. A. & Bulmer, M. In biology of termites: a modern synthesis (eds D. E. Bignell, Yves Roisin & Nathan Lo) 165–191 (Springer, 2011).Cole, E. L., Bayne, H. & Rosengaus, R. B. Young but not defenceless: antifungal activity during embryonic development of a social insect. R. Soc. Open Sci. 7, 191418–191418 (2020).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rosengaus, R. B. & Traniello, J. F. Disease susceptibility and the adaptive nature of colony demography in the dampwood termite Zootermopsis angusticollis. Behav. Ecol. Sociobiol. 50, 546–556 (2001).Article 

    Google Scholar 
    Cole, E. L. & Rosengaus, R. B. Pathogenic dynamics during colony ontogeny reinforce potential drivers of termite eusociality: mate assistance and biparental care. Front. Ecol. Evol. 7 (2019).Chouvenc, T. The relative importance of queen and king initial weights in termite colony foundation success. Insectes Sociaux 66, 177–184 (2019).Article 

    Google Scholar 
    Matsuura, K. & Kobayashi, N. Termite queens adjust egg size according to colony development. Behav. Ecol. 21, 1018–1023 (2010).Article 

    Google Scholar 
    Calleri, D. V., McGrail Reid, E., Rosengaus, R. B., Vargo, E. L. & Traniello, J. F. A. Inbreeding and disease resistance in a social insect: effects of heterozygosity on immunocompetence in the termite Zootermopsis angusticollis. Proc. R. Soc. B: Biol. Sci. 273, 2633–2640 (2006).Article 

    Google Scholar 
    DeHeer, C. J. & Vargo, E. L. An indirect test of inbreeding depression in the termites Reticulitermes flavipes and Reticulitermes virginicus. Behav. Ecol. Sociobiol. 59, 753–761 (2006).Article 

    Google Scholar 
    Aguero, C. M., Eyer, P.-A., Martin, J. S., Bulmer, M. S. & Vargo, E. L. Natural variation in colony inbreeding does not influence susceptibility to a fungal pathogen in a termite. Ecol. Evol. 11, 3072–3083 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Aguero, C., Eyer, P. A. & Vargo, E. L. Increased genetic diversity from colony merging in termites does not improve survival against a fungal pathogen. Sci. Rep. 10, 4212 (2020).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rosengaus, R. B. & Traniello, J. F. Disease risk as a cost of outbreeding in the termite Zootermopsis angusticollis. Proc. Natl Acad. Sci. 90, 6641–6645 (1993).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Eyer, P.-A. et al. Extensive human-mediated jump dispersal within and across the native and introduced ranges of the invasive termite Reticulitermes flavipes. Mol. Ecol. 30, 3948–3964 (2021).Article 
    PubMed 

    Google Scholar 
    Perdereau, E. et al. Global genetic analysis reveals the putative native source of the invasive termite, Reticulitermes flavipes, in France. Mol. Ecol. 22, 1105–1119 (2013).CAS 
    Article 
    PubMed 

    Google Scholar 
    Sinotte, V. M. et al. Female-biased sex allocation and lack of inbreeding avoidance in Cubitermes termites. Ecol. Evolution 11, 5598–5605 (2021).Article 

    Google Scholar 
    Li, G., Gao, Y., Sun, P., Lei, C. & Huang, Q. Factors affecting mate choice in the subterranean termite Reticulitermes chinensis (Isoptera: Rhinotermitidae). J. Ethol. 31, 159–164 (2013).Article 

    Google Scholar 
    Aguilera-Olivares, D., Flores-Prado, L., Véliz, D. & Niemeyer, H. Mechanisms of inbreeding avoidance in the one-piece drywood termite Neotermes chilensis. Insectes Sociaux 62, 237–245 (2015).Article 

    Google Scholar 
    Miyaguni, Y., Agarie, A., Sugio, K., Tsuji, K. & Kobayashi, K. Caste development and sex ratio of the Ryukyu drywood termite Neotermes sugioi and its potential mechanisms. Sci. Rep. 11, 15037 (2021).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nutting, W. L. In Biology of Termites (eds Kumar Krishna & Frances M. Weesner) 233–282 (Academic Press, 1969).Fougeyrollas, R. et al. Dispersal and mating strategies in two neotropical soil-feeding termites, Embiratermes neotenicus and Silvestritermes minutus (Termitidae, Syntermitinae). Insectes Sociaux 65, 251–262 (2018).Article 

    Google Scholar 
    Shellman-Reeve, J. S. Genetic relatedness and partner preference in a monogamous, wood-dwelling termite. Anim. Behav. 61, 869–876 (2001).Article 

    Google Scholar 
    Zhang, Z.-Y. et al. Biochemical, molecular, and morphological variations of flight muscles before and after dispersal flight in a eusocial termite, Reticulitermes chinensis. Insect Sci. 28, 77–92 (2021).CAS 
    Article 
    PubMed 

    Google Scholar 
    Mullins, A. J. et al. Dispersal flights of the Formosan subterranean termite (Isoptera: Rhinotermitidae). J. Econ. Entomol. 108, 707–719 (2015).Article 
    PubMed 

    Google Scholar 
    Goodisman, M. A. D. & Crozier, R. H. Population and colony genetic structure of the primitive termite Mastotermes Darwiniensis. Evolution 56, 70–83 (2002).Article 
    PubMed 

    Google Scholar 
    Schmidt, A. M., Jacklyn, P. & Korb, J. Isolated in an ocean of grass: low levels of gene flow between termite subpopulations. Mol. Ecol. 22, 2096–2105 (2013).Article 
    PubMed 

    Google Scholar 
    Thompson, G. J., Lenz, M., Crozier, R. H. & Crespi, B. J. Molecular-genetic analyses of dispersal and breeding behaviour in the Australian termite Coptotermes lacteus: evidence for non-random mating in a swarm-dispersal mating system. Aust. J. Zool. 55, 219–227 (2007).CAS 
    Article 

    Google Scholar 
    Vargo, E. L. Diversity of termite breeding systems. Insects 10, 52 (2019).Article 
    PubMed Central 

    Google Scholar 
    Tranter, C., LeFevre, L., Evison, S. E. F. & Hughes, W. O. H. Threat detection: contextual recognition and response to parasites by ants. Behav. Ecol. 26, 396–405 (2014).Article 

    Google Scholar 
    Hussain, A., Tian, M.-Y., He, Y.-R., Bland, J. M. & Gu, W.-X. Behavioral and electrophysiological responses of Coptotermes formosanus Shiraki towards entomopathogenic fungal volatiles. Biol. Control 55, 166–173 (2010).Article 

    Google Scholar 
    Yanagawa, A., Imai, T., Akino, T., Toh, Y. & Yoshimura, T. Olfactory cues from pathogenic fungus affect the direction of motion of termites, Coptotermes formosanus. J. Chem. Ecol. 41, 1118–1126 (2015).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rosengaus, R. B., James, L.-T., Hartke, T. R. & Brent, C. S. Mate preference and disease risk in Zootermopsis angusticollis (Isoptera: Termopsidae). Environ. Entomol. 40, 1554–1565 (2011).Article 
    PubMed 

    Google Scholar 
    Beani, L. et al. Cuticular hydrocarbons as cues of sex and health condition in Polistes dominula wasps. Insectes Sociaux 66, 543–553 (2019).Article 

    Google Scholar 
    Waser, P. M., Austad, S. N. & Keane, B. When should animals tolerate inbreeding? Am. Nat. 128, 529–537 (1986).Article 

    Google Scholar 
    Bengtsson, B. O. Avoiding inbreeding: at what cost? J. Theor. Biol. 73, 439–444 (1978).CAS 
    Article 
    PubMed 

    Google Scholar 
    Lehmann, L. & Perrin, N. Inbreeding avoidance through kin recognition: Choosy females boost male dispersal. Am. Nat. 162, 638–652 (2003).Article 
    PubMed 

    Google Scholar 
    Basalingappa, S. Environmental hazards to reproductives of Odontotermes assmuthi Holgrem. Indian Zool. 1, 45–50 (1970).
    Google Scholar 
    Darlington, J., Sands, W. & Pomeroy, D. Distribution and post-settlement survival in the field by reproductive pairs of Hodotermes mossambicus hagen (isoptera, hodotermitida). Insectes Sociaux 24, 353–358 (1977).Article 

    Google Scholar 
    Dial, K. P. & Vaughan, T. A. Opportunistic predation on alate termites in Kenya. Biotropica 19, 185–187 (1987).Article 

    Google Scholar 
    Korb, J. & Salewski, V. Predation on swarming termites by birds. Afr. J. Ecol. 38, 173–174 (2000).Article 

    Google Scholar 
    Schwenke, R. A., Lazzaro, B. P. & Wolfner, M. F. ReproduCtion–immunity Trade-offs In Insects. Annu. Rev. Entomol. 61, 239–256 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Calleri, D. II, Rosengaus, R. & Traniello, J. A. Disease and colony foundation in the dampwood termite Zootermopsis angusticollis: The survival advantage of nestmate pairs. Naturwissenschaften 92, 300–304 (2005).CAS 
    Article 
    PubMed 

    Google Scholar 
    Fei, H. X. & Henderson, G. Comparative study of incipient colony development in the Formosan subterranean termite, Coptotermes formosanus Shiraki (Isoptera,Rhinotermitidae). Insectes Sociaux 50, 226–233 (2003).Article 

    Google Scholar 
    Rosengaus, R. B., Cornelisse, T., Guschanski, K. & Traniello, J. F. A. Inducible immune proteins in the dampwood termite Zootermopsis angusticollis. Naturwissenschaften 94, 25–33 (2007).CAS 
    Article 
    PubMed 

    Google Scholar 
    Rosengaus, R. B., Traniello, J. F. A., Chen, T., Brown, J. J. & Karp, R. D. Immunity in a social insect. Naturwissenschaften 86, 588–591 (1999).CAS 
    Article 

    Google Scholar 
    Sun, Q., Haynes, K. F., Hampton, J. D. & Zhou, X. Sex-specific inhibition and stimulation of worker-reproductive transition in a termite. Sci. Nat. 104, 79 (2017).Article 
    CAS 

    Google Scholar 
    Eyer, P.-A. et al. Inbreeding tolerance as a pre-adapted trait for invasion success in the invasive ant Brachyponera chinensis. Mol. Ecol. 27, 4711–4724 (2018).PubMed 

    Google Scholar 
    Barrett, S. C. H. & Charlesworth, D. Effects of a change in the level of inbreeding on the genetic load. Nature 352, 522 (1991).CAS 
    Article 
    PubMed 

    Google Scholar 
    Crnokrak, P. & Spencer, C. H. B. Perspective: purging the genetic load. A review of the experimental evidence. Evolution 56, 2347–2358 (2002).Article 
    PubMed 

    Google Scholar 
    Day, S. B., Bryant, E. H. & Meffert, L. M. The influence of variable rates of inbreeding on fitness, environmental responsiveness, and evolutionary potential. Evolution 57, 1314–1324 (2003).Article 
    PubMed 

    Google Scholar 
    Syren, R. M. & Luykx, P. Permanent segmental interchange complex in the termite Incisitermes schwarzi. Nature 266, 167–168 (1977).CAS 
    Article 
    PubMed 

    Google Scholar 
    Fontana, F. Multiple reciprocal chromosomal translocations and their role in the evolution of sociality in termites. Ethol. Ecol. Evolution 3, 15–19 (1991).CAS 
    Article 

    Google Scholar 
    Matsuura, K. A test of the haplodiploid analogy hypothesis in the termite Reticulitermes speratus (Isoptera: Rhinotermitidae). Ann. Entomol. Soc. Am. 95, 646–649 (2002).Article 

    Google Scholar 
    Yashiro, T. et al. Enhanced heterozygosity from male meiotic chromosome chains is superseded by hybrid female asexuality in termites. Proc. Natl. Acad. Sci. 118, e2009533118 (2021).Charlesworth, B. & Wall, J. D. Inbreeding, heterozygote advantage and the evolution of neo-X and neo-Y sex chromosomes. Proc. R. Soc. Lond. Ser. B: Biol. Sci. 266, 51–56 (1999).Article 

    Google Scholar 
    Hellemans, S. et al. Widespread occurrence of asexual reproduction in higher termites of the Termes group (Termitidae: Termitinae). BMC Evol. Biol. 19, 131 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vargo, E. L., Labadie, P. E. & Matsuura, K. Asexual queen succession in the subterranean termite Reticulitermes virginicus. Proc. R. Soc. B: Biol. Sci. 279, 813–819 (2012).Article 

    Google Scholar 
    Matsuura, K. et al. Queen succession through asexual reproduction in termites. Science 323, 1687–1687 (2009).CAS 
    Article 
    PubMed 

    Google Scholar 
    Cremer, S., Pull, C. D. & Fürst, M. A. Social immunity: emergence and evolution of colony-level disease protection. Annu. Rev. Entomol. 63, 105–123 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Van Meyel, S., Körner, M. & Meunier, J. Social immunity: why we should study its nature, evolution and functions across all social systems. Curr. Opin. Insect Sci. 28, 1–7 (2018).Article 
    PubMed 

    Google Scholar 
    Cotter, S. C. & Kilner, R. M. Personal immunity versus social immunity. Behav. Ecol. 21, 663–668 (2010).Article 

    Google Scholar 
    Liu, L., Zhao, X.-Y., Tang, Q.-B., Lei, C.-L. & Huang, Q.-Y. The mechanisms of social immunity against fungal infections in eusocial insects. Toxins 11, 244 (2019).CAS 
    Article 
    PubMed Central 

    Google Scholar 
    Chouvenc, T. & Su, N. Y. When subterranean termites challenge the rules of fungal epizootics. Plos One 7, e34484 (2012).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Davis, H. E., Meconcelli, S., Radek, R. & McMahon, D. P. Termites shape their collective behavioural response based on stage of infection. Sci. Rep. 8, 14433–14433 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cassidy, S. T. et al. Disease defences across levels of biological organization: individual and social immunity in acorn ants. Anim. Behav. 179, 73–81 (2021).Article 

    Google Scholar 
    López-Uribe, M. M., Sconiers, W. B., Frank, S. D., Dunn, R. R. & Tarpy, D. R. Reduced cellular immune response in social insect lineages. Biol. Lett. 12, 20150984 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    He, S. et al. Evidence for reduced immune gene diversity and activity during the evolution of termites. Proc. R. Soc. B: Biol. Sci. 288, 20203168 (2021).Article 

    Google Scholar 
    Viljakainen, L. et al. Rapid evolution of immune proteins in social insects. Mol. Biol. Evol. 26, 1791–1801 (2009).CAS 
    Article 
    PubMed 

    Google Scholar 
    Meusemann, K., Korb, J., Schughart, M. & Staubach, F. No evidence for single-copy immune-gene specific signals of selection in termites. Front. Ecol. Evol. 8 (2020).Otani, S., Bos, N. & Yek, S. H. Transitional complexity of social insect immunity. Front. Ecol. Evol. 4 (2016).Barribeau, S. M. et al. A depauperate immune repertoire precedes evolution of sociality in bees. Genome Biol. 16, 83 (2015).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    de Boer, R. A., Vega-Trejo, R., Kotrschal, A. & Fitzpatrick, J. L. Meta-analytic evidence that animals rarely avoid inbreeding. Nat. Ecol. Evol. 5, 949–964 (2021).Article 
    PubMed 

    Google Scholar 
    Szulkin, M., Stopher, K. V., Pemberton, J. M. & Reid, J. M. Inbreeding avoidance, tolerance, or preference in animals? Trends Ecol. Evol. 28, 205–211 (2013).Article 
    PubMed 

    Google Scholar 
    Fox, C. W. & Reed, D. H. Inbreeding depression increases with environmental stress: an experimental study and meta-analysis. Evol. 65, 246–258 (2011).Article 

    Google Scholar 
    Kokko, H., Ots, I. & Tregenza, T. When not to avoid inbreeding. Evolution 60, 467–475 (2006).Article 
    PubMed 

    Google Scholar 
    Zayed, A. & Packer, L. Complementary sex determination substantially increases extinction proneness of haplodiploid populations. Proc. Natl Acad. Sci. USA 102, 10742–10746 (2005).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ross, K. G. & Fletcher, D. J. C. Diploid male production — a significant colony mortality factor in the fire ant Solenopsis invicta (Hymenoptera: Formicidae). Behav. Ecol. Sociobiol. 19, 283–291 (1986).Article 

    Google Scholar 
    Eyer, P.-A., Salin, J., Helms, A. M. & Vargo, E. L. Distinct chemical blends produced by different reproductive castes in the subterranean termite Reticulitermes flavipes. Sci. Rep. 11, 4471 (2021).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kearse, M. et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28, 1647–1649 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Queller, D. C. & Goodnight, K. F. Estimating relatedness using genetic markers. Evolution 43, 258–275 (1989).Article 
    PubMed 

    Google Scholar 
    Wang, J. Coancestry: a program for simulating, estimating and analysing relatedness and inbreeding coefficients. Mol. Ecol. Resour. 11, 141–145 (2011).Article 
    PubMed 

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

    Google Scholar 
    Rosengaus, R. B., Moustakas, J. E., Calleri, D. V. & Traniello, J. F. A. Nesting ecology and cuticular microbial loads in dampwood (Zootermopsis angusticollis) and drywood termites (Incisitermes minor, I. schwarzi, Cryptotermes cavifrons). J. Insect Sci. 3, 31 (2003).Kozich, J. J., Westcott, S. L., Baxter, N. T., Highlander, S. K. & Schloss, P. D. Development of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq Illumina sequencing platform. Appl. Environ. Microbiol. 79, 5112–5120 (2013).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    White, T. J., Burns, T., Lee, S. & Taylor, J. in PCR protocols: A guide to methods and applications (eds. M. A. Innis, D. H. Gelfand, J. J. Snisky, & T. J. White) 315–322 (Academic Press, 1990).Aguero, C. M., Eyer, P.-A., Crippen, T. L. & Vargo, E. L. Reduced environmental microbial diversity on the cuticle and in the galleries of a subterranean termite compared to surrounding soil. Microb. Ecol. 81, 1054–1063 (2021).CAS 
    Article 
    PubMed 

    Google Scholar 
    Bolyen, E. et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat. Biotechnol. 37, 852–857 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hamady, M., Lozupone, C. & Knight, R. Fast UniFrac: facilitating high-throughput phylogenetic analyses of microbial communities including analysis of pyrosequencing and PhyloChip data. ISME J. 4, 17–27 (2010).CAS 
    Article 
    PubMed 

    Google Scholar 
    Therneau, T. & Grambsch, P. Modeling Survival Data: Extending the Cox Model (Springer, 2000).Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar  More

  • in

    Ovaries and testes of Lithobius forficatus (Myriapoda, Chilopoda) react differently to the presence of cadmium in the environment

    Sieńczuk, W. Toksykologia (PZWL Warszawa, 1999) (in Polish).
    Google Scholar 
    Kabata-Pendias, A. & Pendias, H. Biochemia pierwiastków śladowych (PZWL Warszawa, 1999) (in Polish).
    Google Scholar 
    Sharma, H., Rawal, N. & Mathew, B. B. The characteristics, toxicity and effects of cadmium. Int. J. Nanosci. Nanotechnol. 3, 1–9 (2015).
    Google Scholar 
    Duarte, A. et al. (eds) Soil pollution: From Monitoring to Remediation 1st edn. (Academic Press, 2017).
    Google Scholar 
    Zhang, H. & Reynolds, M. Cadmium exposure in living organisms: A short review. Sci. Total Environ. 678, 761–767 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Lane, T. W. et al. A cadmium enzyme from a marine diatom. Nature 435, 42 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Jӓrup, L. Hazards of heavy metal contamination. Br. Med. Bull. 68, 167–182 (2003).Article 

    Google Scholar 
    Massányi, P., Massányi, M., Madeddu, R., Stawarz, R. & Lukáč, N. Effects of cadmium, lead, and mercury on the structure and function of reproductive organs. Toxics 8, 94 (2020).PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Roy, S. Cadmium accumulation in crops and the increasing risk of dietary cadmium exposure: An overview. In Cadmium Tolerance in Plants: Agronomic, Molecular, Signaling, and Omic Approaches (eds Hasanuzzaman, M. et al.) 247–254 (Academic Press, 2019).Chapter 

    Google Scholar 
    Templeton, D. M. & Liu, Y. Multiple roles of cadmium in cell death and survival. Chem. Biol. Interact. 188, 267–275 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Stojsavljević, A. et al. Evaluation of trace metals in thyroid tissues: Comparative analysis with benign and malignant thyroid diseases. Ecotoxicol. Environ. Saf. 183, 109479 (2019).PubMed 
    Article 
    CAS 

    Google Scholar 
    Lewis, J. G. E. The Biology of Centipedes 1st edn. (Cambridge University Press, 1981).Book 

    Google Scholar 
    Hopkin, S. P. Ecophysiology of Metals in Terrestrial Invertebrates 1st edn. (Elsevier Applied Science, 1989).
    Google Scholar 
    Hopkin, S. P. & Read, H. J. The Biology of Millipedes (Oxford University Press, 1992).
    Google Scholar 
    Lipovšek, S., Letofsy-Papst, I., Hofer, F. & Pabst, M. A. Seasonal- and age-dependent changes of the structure and chemical composition of the spherites in the midgut gland of the harvestmen Gyas annulatus (Opiliones). Micron 33, 647–654 (2002).PubMed 
    Article 

    Google Scholar 
    Chajec, Ł, Rost-Roszkowska, M. M., Vilimova, J. & Sosinka, A. Ultrastructure and regeneration of midgut epithelial cells in Lithobius forficatus (Chilopoda, Lithobiidae). Invertebr. Biol. 131, 119–132 (2012).Article 

    Google Scholar 
    Hopkin, S. P., Watson, K., Martin, M. H. & Mould, M. L. The assimilation of heavy metals by Lithobius variegatus and Glomeris marginata (Chilopoda; Diplopoda). Bijdr. Dierkd. 55, 88–94 (1985).
    Google Scholar 
    Adiyodi, K. G. & Adiyodi, R. G. (eds) Reproductive Biology of Invertebrates. Volume I. Oogenesis, Oviposition, and Oosorption (Wiley, 1983).
    Google Scholar 
    Adiyodi, K. G. & Adiyodi, R. G. (eds) Reproductive Biology of Invertebrates. Volume II. Spermatogenesis and Sperm Function (Wiley, 1983).
    Google Scholar 
    Sareen, M. L. & Adiyodi, K. G. Arthropoda – Myriapoda. In Reproductive Biology of Invertebrates. Volume I. Oogenesis, Oviposition, and Oosorption (eds Adiyodi, K. G. & Adiyodi, R. G.) 497–520 (Wiley, 1983).
    Google Scholar 
    Minelli, A. Chilopoda – Reproduction. In Treatise on Zoology – Anatomy, Taxonomy, Biology. The Myriapoda. Vol. 1. Chilopoda (ed. Minelli, A.) 279–294 (Brill, 2011).Chapter 

    Google Scholar 
    Parolini, M. Toxicity of the non-steroidal anti-inflammatory drugs (NSAIDs) acetylsalicylic acid, paracetamol, diclofenac, ibuprofen and naproxen towards freshwater invertebrates: A review. Sci. Total Environ. 740, 140043 (2020).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Nath, V. Oogenesis of Lithobius forficatus. Biol. Rev. 1, 148–157 (1924).Article 

    Google Scholar 
    Nath, V. Spermathogenesis of Lithobius forficatus. Biol. Rev. 1, 270–277 (1925).Article 

    Google Scholar 
    Descamps, M. Etude ultrastructurale des spermatogonies et de la croissance spermatocytaire chez Lithobius forficatus L. (Myriapode Chilopode). Z. Zellforsch. 121, 14–26 (1971).CAS 
    PubMed 
    Article 

    Google Scholar 
    Descamps, M. Le cycle spermatogenétique chez Lithobius forficatus L. (Myriapode, Chilopode). I. Evolution et etude quantitative des populations cellulaires du tes ticle au cours du développement post-embryonnaire. Arch. Zool. Exp. Gen. 112, 199–209 (1971).
    Google Scholar 
    Herbaut, C. Etude cytochimique et ultrastructurale de l’ovogenése chez Lithobius forficatus L. (Myriapode Chilopode). Evolution des constituants cellulaires. Wilhelm Roux’ Arch. 170, 115–134 (1972).CAS 
    Article 

    Google Scholar 
    Descamps, M., Fabre, M. C., Grelle, C. & Gerard, S. Cadmium and lead kinetics during experimental contamination of the centipede Lithobius forficatus L. Arch. Environ. Contam. Toxicol. 31, 350–353 (1996).CAS 
    Article 

    Google Scholar 
    Vandenbulcke, F., Grelle, C., Fabre, M.-C. & Descamps, M. Implication of the midgut of the centipede Lithobius forficatus in the heavy metal detoxification process. Ecotoxicol. Environ. Saf. 41, 258–268 (1998).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rost-Roszkowska, M. et al. Influence of soil contaminated with cadmium on cell death in the digestive epithelium of soil centipede Lithobius forficatus (Myriapoda, Chilopoda). Eur. Zool. J. 87, 242–262 (2020).CAS 
    Article 

    Google Scholar 
    Rost-Roszkowska, M. et al. Effects of short- and long-term exposure to cadmium on salivary glands and fat body of soil centipede Lithobius forficatus (Myriapoda, Chilopoda): Histology and ultrastructure. Micron 137, 102915 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rost-Roszkowska, M. et al. Effects of cadmium on mitochondrial structure and function in different organs: Studies on the soil centipede Lithobius forficatus (Myriapoda, Chilopoda). Eur. Zool. J. 88, 632–664 (2021).CAS 
    Article 

    Google Scholar 
    Włodarczyk, A., Student, S. & Rost-Roszkowska, M. Autophagy and apoptosis in starved and refed Neocaridina davidi (Crustacea, Malacostraca) midgut. Can. J. Zool. 97, 294–303 (2019).Article 

    Google Scholar 
    Bradford, M. M. Rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72, 248–254 (1976).CAS 
    PubMed 
    Article 

    Google Scholar 
    Wieser, W. Conquering terra firma: The copper problem from the isopod’s point of view. Helgolander Wiss. Meeresunters. 15, 282–293 (1967).ADS 
    Article 

    Google Scholar 
    Gräff, S., Berkus, M., Alberti, G. & Köhler, H. R. Metal accumulation strategies in saprophagous and phytophagous soil invertebrates: A quantitative comparison. Biometals 10, 45–53 (1997).Article 

    Google Scholar 
    Siekierska, E. & Urbańska-Jasik, D. The effect of cadmium and selenium ions on the ovary structure in leech Herpobdella octooculata (L.). Folia Morphol. 57, 61 (1998).
    Google Scholar 
    Siekierska, E. & Urbańska-Jasik, D. Cadmium effect on the ovarian structure in earthworm Dendrobaena veneta (Rosa). Environ. Pollut. 120, 289–297 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    Osman, W., El-Samad, L. M., Mokhamer, E.L.-H., El-Touhamy, A. & Shonouda, M. Ecological, morphological, and histological studies on Blaps polycresta (Coleoptera: Tenebrionidae) as biomonitors of cadmium soil pollution. Environ. Sci. Pollut. Res. Int. 22, 14104–14115 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Siekierska, E. & Brzozowa, M. Cadmium effect on the seminal vesicles structure and spermatogenesis in the earthworm Dendrobaena veneta (Rosa). In 8th International Symposium on Earthworm Ecology. Book of abstracts, 231 (2006).Siekierska, E. & Brzozowa, M. Changes in primary and secondary spermatocytes in seminal vesicles in the earthworm Dendrobaena veneta (Rosa) after 10 days of cadmium exposure. Acta Biol. Cracov. Bot. 50, 68 (2008).
    Google Scholar 
    Brzozowa, M. Wpływ kadmu na przebieg spermiogenezy u dżdżownicy Dendrobaena veneta (Rosa). PhD Thesis, University of Silesia in Katowice Poland (2009).Papathanassiou, E. Cadmium accumulation and ultrastructural alterations in oogenesis of the prawn Palaemon serratus (Pennant). Bull. Environ. Contam. Toxicol. 36, 192–198 (1986).CAS 
    PubMed 
    Article 

    Google Scholar 
    Au, D. W. T., Chiang, M. W. L. & Wu, R. Effect of cadmium and phenol on mortality and ultrastructure of sea urchin and mussel spermatozoa. Arcg. Environ. Contam. Toxicol. 38, 455–463 (2000).CAS 
    Article 

    Google Scholar 
    Au, D. W. T., Lee, C. Y., Chan, K. L. & Wu, R. Reproductive impairment of sea urchins upon chronic exposure to cadmium. Part I: Effects on gamete quality. Environ. Pollut. 111, 1–9 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Au, D. W. T., Reunov, A. A. & Wu, R. Reproductive impairment of sea urchins upon chronic exposure to cadmium. Part II: Effects on sperm development. Environ. Pollut. 111, 11–20 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Eckelbarger, K. J. Diversity of metazoan ovaries and vitellogenic mechanisms – implications for life history theory. Proc. Biol. Soc. Wash. 107, 193–218 (1994).
    Google Scholar 
    Suzuki, K. T., Yamamura, M. & Mori, T. Cadmium-binding proteins induced in earthworm. Arch. Environ. Contam. Toxicol. 9, 415–424 (1980).CAS 
    PubMed 
    Article 

    Google Scholar 
    Maroni, G., Wise, J., Young, J. E. & Otto, E. Metallothionein gene duplications and metal tolerance in natural populations of Drosophila melanogaster. Genetics 117, 739–744 (1987).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Luo, M., Finet, C., Cong, H., Wei, H. & Chung, H. The evolution of insect metallothioneins. Proc. R. Soc. B 287, 20202189 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Turbeck, B. O. A study of the concentrically laminated concretions, ‘spherites’, in the regenerative cells of the midgut of Lepidopterous larvae. Tissue Cell. 6, 627–640 (1974).CAS 
    PubMed 
    Article 

    Google Scholar 
    Cruz-Landim, C. Localization of calcium and acid phosphatase in the Malpighian tubules of nurse workers of Melipona quadrifasciata anthidioides Lep. (Hymenoptera, Apidae, Meliponini). Biosci. J. 16, 87–99 (2000).
    Google Scholar 
    Lipovšek, S., Letofsky-Papst, I., Hofer, F., Pabst, M. A. & Devetak, D. Application of analytical electron microscopic methods to investigate the function of spherites in the midgut of the larval antlion Euroleon nostras (Neuroptera: Myrmeleontidae). Microsc. Res. Tech. 75, 397–407 (2012).PubMed 
    Article 
    CAS 

    Google Scholar 
    Pinheiro, D. O., Conte, H. & Gregório, E. A. Spherites in the midgut epithelial cells of the sugarcane borer parasitized by Cotesia flavipes. Biocell 32, 61–67 (2008).Article 

    Google Scholar 
    Rost-Roszkowska, M. M., Kszuk-Jendrysik, M., Marchewka, A. & Poprawa, I. Fine structure of the midgut epithelium in the millipede Telodeinopus aoutii (Myriapoda, Diplopoda) with special emphasis on epithelial regeneration. Protoplasma 255, 43–55 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lipovšek, S. et al. Ultrastructure of spherites in the midgut diverticula and Malpighian tubules of the harvestman Amilenus aurantiacus during the winter diapause. Histochem. Cell Biol. 157, 107–118 (2022).PubMed 
    Article 
    CAS 

    Google Scholar 
    Kramarz, P. Dynamics of accumulation and decontamination of cadmium and zinc in carnivorous invertebrates. 2. The centipede Lithobius mutabilis Koch. Bull. Environ. Contam. Toxicol. 63, 538–545 (1999).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rost-Roszkowska, M. M. et al. Structure of the midgut epithelium in four diplopod species: Histology, histochemistry and ultrastructure. Arthropod Syst. Phylogeny 79, 295–308 (2021).Article 

    Google Scholar 
    Köhler, H.-R. Localization of metals in cells of saprophagous soil arthropods (Isopoda, Diplopoda, Collembola). Microsc. Res. Tech. 56, 393–401 (2002).ADS 
    PubMed 
    Article 

    Google Scholar 
    Cervera, A., Maymó, A. C., Martínez-Pardo, R. & Garcerá, M. D. Vitellogenesis inhibition in Oncopeltus fasciatus females (Heteroptera: Lygaeidae) exposed to cadmium. J. Insect Physiol. 51, 895–911 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Cervera, A., Maymó, A. C., Martínez-Pardo, R. & Garcerá, M. D. Vitellogenin polypeptide levels in one susceptible and one cadmium-resistant strain of Oncopeltus fasciatus (Heteroptera: Lygaeidae), and its role in cadmium resistance. J. Insect Physiol. 52, 158–168 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    Sehgal, A., Osgood, C. & Zimmering, S. Aneuploid in Drosophila. III: Aneuploidogens inhibit in vitro assembly of taxol-purified Drosophila microtubules. Environ. Mol. Mutagen. 16, 217–224 (1990).CAS 
    PubMed 
    Article 

    Google Scholar 
    Li, W., Zhao, Y. & Cou, I. N. Alterations in cytoskeletal protein sulfhydryls and cellular glutathione in cultured cells exposed to cadmium and nickel ions. Toxicology 77, 65–79 (1993).CAS 
    PubMed 
    Article 

    Google Scholar 
    dos Santos, D. C., Gregorio, E. A. & Moreli Silva de Moraes, R. L. Programmed cell death during early oogenesis in the Diatraea saccharalis germarium. Acta Microsc. 16, 311–312 (2007).
    Google Scholar 
    Hoeppner, D. J., Hengartner, M. O. & Schnabel, R. Engulfment genes cooperate with ced-3 to promote cell death in Caenorhabditis elegans. Nature 412, 202–206 (2001).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Hikim, A. P. S. et al. Key apoptotic pathways for heat-induced programmed germ cell death in the testis. Endocrinology 144, 3167–3175 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    Russell, L. D., Chiarini-Garcia, H., Korsmeyer, S. J. & Knudson, C. M. Bax-dependent spermatogonia apoptosis is required for testicular development and spermatogenesis. Biol. Reprod. 66, 950–958 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    Shaha, C., Tripathi, R. & Mishra, D. P. Male germ cell apoptosis: Regulation and biology. Philos. Trans. R. Soc. Lond. B Biol. Sci. 365, 1501–1515 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Devine, P. J., Payne, C. M., McCuskey, M. K. & Hoyer, P. B. Ultrastructural evaluation of oocytes during atresia in rat ovarian follicles. Biol. Reprod. 63, 1245–1252 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    Hussein, M. R. Apoptosis in the ovary: Molecular mechanisms. Hum. Reprod. Update 11, 162–177 (2005).PubMed 
    Article 
    CAS 

    Google Scholar 
    Miller, M. A., Technau, U., Smith, K. M. & Steele, R. E. Oocyte development in Hydra involves selection from competent precursor cells. Dev. Biol. 224, 326–338 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    Matova, N. & Cooley, L. Comparative aspects of animal oogenesis. Dev. Biol. 231, 291–320 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Technau, U., Miller, M. A., Bridge, D. & Steele, R. E. Arrested apoptosis of nurse cells during Hydra oogenesis and embryogenesis. Dev. Biol. 260, 191–206 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    Mpakou, V. E., Nezis, I. P., Stravopodis, D. J., Margaritis, L. H. & Papassideri, I. S. Programmed cell death of the ovarian nurse cells during oogenesis of the silkmoth Bombyx mori. Dev. Growth Differ. 48, 419–428 (2006).PubMed 
    Article 

    Google Scholar 
    Mpakou, V. E. et al. Different modes of programmed cel death during oogenesis of the silkmoth Bombyx mori. Autophagy 4, 97–100 (2008).PubMed 
    Article 

    Google Scholar 
    Mpakou, V. E. et al. Programmed cell death of the ovarian nurse cells during oogenesis of the ladybird beetle Adalia bipunctata (Coleoptera: Coccinellidae). Dev. Growth Differ. 53, 804–815 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    Poprawa, I., Hyra, M., Kszuk-Jendrysik, M. & Rost-Roszkowska, M. M. Ultrastructural changes and programmed cell death of trophocytes in the gonad of Isohypsibius granulifer granulifer Thulin, 1928 (Tardigrada, Eutardigrada, Isohypsibiidae). Micron 70, 26–33 (2015).PubMed 
    Article 

    Google Scholar 
    Janelt, K., Jezierska, M. & Poprawa, I. The female reproductive system and oogenesis in Thulinius ruffoi (Tardigrada, Eutardigrada, Isohypsibiidae). Arthropod. Struct. Dev. 50, 53–63 (2019).PubMed 
    Article 

    Google Scholar 
    Mooyottu, S., Anees, C. & Cherian, S. Ovarian stem cells and neo-oogenesis: A breakthrough in reproductive biology research. Vet. World 4, 89–91 (2011).
    Google Scholar 
    Tiwari, M. et al. Apoptosis in mammalian oocytes: A review. Apoptosis 20, 1019–1025 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Xiu, Y.-R. & Yang, W.-X. Roles of three Es-Caspases during spermatogenesis and cadmium-induced apoptosis in Eriocheir sinensis. Aging 10, 1146–1165 (2018).Article 

    Google Scholar 
    Redza-Dutordoir, M. & Averill-Bates, D. A. Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta 1863, 2977–2992 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Sonakowska, L. et al. Cell death in the epithelia of the intestine and hepatopancreas in Neocaridina heteropoda (Crustacea, Malacostraca). PLoS ONE 11, e0147582 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Włodarczyk, A. et al. The effect of starvation and re-feeding on mitochondrial potential in the midgut of Neocaridina davidi (Crustacea, Malacostraca). PLoS ONE 12, e0173563 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Zorova, L. D. et al. Mitochondrial membrane potential. Anal. Biochem. 552, 50–59 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ossola, J. O. & Tomaro, M. L. Heme oxygenase induction by cadmium chloride: Evidence for oxidative stress involvement. Toxicology 104, 141–147 (1995).CAS 
    PubMed 
    Article 

    Google Scholar 
    Levine, B. & Klionsky, D. J. Development by self-digestion: Molecular mechanisms and biological functions of autophagy. Dev. Cell. 6, 463–477 (2004).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kourtis, N. & Tavernarakis, N. Autophagy and cell death in model organisms. Cell Death Differ. 16, 21–30 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kliosnky, D. et al. Guidelines for the use and interpretation of assays for monitoring autophagy (3rd edition). Autophagy 12, 1–222 (2016).Article 

    Google Scholar 
    Kliosnky, D. et al. Guidelines for the use and interpretation of assays for monitoring autophagy (4th edition). Autophagy 17, 1–382 (2021).Article 

    Google Scholar 
    Velentzas, A. D., Nezis, I. P., Stravopodis, D. J., Papassideri, I. S. & Margaritis, L. H. Apoptosis and autophagy function cooperatively for the efficacious execution of programmed nurse cell death during Drosophila virilis oogenesis. Autophagy 3, 130–132 (2007).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lipovšek, S. et al. Changes in the midgut cells in the European cave spider, Meta menardi, during starvation in spring and autumn. Histochem. Cell Biol. 149, 245–260 (2018).PubMed 
    Article 
    CAS 

    Google Scholar 
    Rost-Roszkowska, M. M. et al. Autophagy and apoptosis in the midgut epithelium of millipedes. Microsc. Microanal. 25, 1004–1016 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Nezis, I. P. et al. Autophagy as a trigger for cell death: Autophagic degradation of inhibitor of apoptosis dBruce controls DNA fragmentation during late oogenesis in Drosophila. Autophagy 6, 1214–1215 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Rost-Roszkowska, M. M., Janelt, K. & Poprawa, I. The role of autophagy in the midgut epithelium of Parachela (Tardigrada). Zoomorphology 137, 501–509 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Leist, M., Single, B., Castoldi, A. F., Kühnle, S. & Nicotera, P. Intracellular adenosine triphosphate (ATP) concentration: A switch in the decision between apoptosis and necrosis. J. Exp. Med. 185, 1481–1486 (1997).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Nikoletopoulou, V., Markaki, M., Palikaras, K. & Tavernarakis, N. Crosstalk between apoptosis, necrosis and autophagy. Biochim. Biophys. Acta. 1833, 3448–3459 (2013).CAS 
    PubMed 
    Article 

    Google Scholar  More

  • in

    Transversal criminality at sea

    Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain
    the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in
    Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles
    and JavaScript. More

  • in

    Global analysis of biosynthetic gene clusters reveals conserved and unique natural products in entomopathogenic nematode-symbiotic bacteria

    General experimental proceduresAll chemicals were purchased from Sigma-Aldrich, Acros Organics or Iris BIOTECH. Isotope-labelled chemicals were purchased from Cambridge Isotope Laboratories. Genomic DNA of selected Xenorhabdus and Photorhabdus strains was isolated using the Qiagen Gentra Puregene Yeast/Bact Kit. DNA polymerases (Taq, Phusion and Q5) and restriction enzymes were purchased from New England Biolabs or Thermo Fisher Scientific. DNA primers were purchased from Eurofins MWG Operon. PCR amplifications were carried out on thermocyclers (SensoQuest). Polymerases were used according to the manufacturers’ instructions. DNA purification was performed from 1% Tris-acetate-EDTA (TAE) agarose gel using an Invisorb Spin DNA Extraction Kit (STRATEC Biomedical AG). Plasmids in E. coli were isolated by alkaline lysis. HPLC-UV-MS analysis was conducted on an UltiMate 3000 system (Thermo Fisher) coupled to an AmaZonX mass spectrometer (Bruker) with an ACQUITY UPLC BEH C18 column (130 Å, 2.1 mm × 100 mm, 1.7-μm particle size, Waters) at a flow rate of 0.6 ml min−1 (5–95% acetonitrile/water with 0.1% formic acid, vol/vol, 16 min, UV detection wavelength 190–800 nm). HPLC-UV-HRMS analysis was conducted on an UltiMate 3000 system (Thermo Fisher) coupled to an Impact II qTof mass spectrometer (Bruker) with an ACQUITY UPLC BEH C18 column (130 Å, 2.1 mm × 100 mm, 1.7-μm particle size, Waters) at a flow rate of 0.4 ml min−1 (5–95% acetonitrile/water with 0.1% formic acid, vol/vol, 16 min, UV detection wavelength 190–800 nm). Flash purification was performed on a Biotage SP1 flash purification system (Biotage) by a C18 main column (Interchim, PF50C18HP-F0080, 120 g) with a self-packed pre-column (Interchim, PF-DLE-F0012, Puriflash dry-load empty F0012 Flash column) coupled with a UV detector. HPLC purification was performed on preparative and semipreparative Agilent 1260 systems coupled to a diode array detector (DAD) and a single quadrupole detector with a C18 ZORBAX Eclipse XDB column (9.4 mm × 250 mm, 5 μm, 3 ml min−1; 21.2 mm × 250 mm, 5 μm, 20 ml min−1; 50 mm × 250 mm, 10 μm, 40 ml min−1). Freeze drying was performed using a BUCHI Lyovapor L-300 Continuous system. NMR experiments were carried out on a Bruker AVANCE 500-, 600- or 700-MHz spectrometer equipped with a 5-mm cryoprobe. 2R,3S-IOC (1) and GameXPeptide A (16) were synthesized by WuXi App Tec following the literature (ref. 73 for 2R,3S-1 and ref. 74 for 16).Genome sequencing, assembly and annotationIsolated DNA was sequenced on the Illumina NextSeq 500 platform. DNA libraries were constructed using the Nextera XT DNA preparation kit (Illumina) and whole-genome sequencing was performed using 2× 150-bp paired-end chemistry. A sequencing depth of >50× was targeted for each sample. Adapters and low-quality ends were trimmed with Trimmomatic 0.39 (ref. 75) and the parameters [2:30:10 LEADING:3 TRAILING:3 SLIDINGWINDOW:4:10 MINLEN:12] using a database of adapter sequences as provided by Illumina. All genomes were assembled using SPAdes v. 3.10.1 (ref. 76) executed with the following parameters: –cov-cutoff auto –careful in paired-end mode plus mate pairs (in cases where accompanying mate-pair libraries were available). Genome annotation was performed using Prokka v. 1.12 (ref. 77) with the following parameters: –usegenus–genus GENUS–addgenes–evalue 0.0001–rfam–kingdom Bacteria–gcode 11–gram –mincontiglen 200. Geneious Prime 2021 was used in genome visualization and analysis.antiSMASH annotations and BiG-FAM preliminary classificationThe antiSMASH 5.0 (ref. 23) web server was employed to mine all the genome sequences for the presence of putative natural-product BGCs. The annotations were conducted using default settings with the extended parameters of ClusterBlast, Cluster Pfam analysis and Pfam-based GO term annotation. The annotated BGCs were summarized for each strain (Supplementary Fig. 1) and visualized in the anvi’o 6.1 (refs. 28,78) layers (Fig. 1 and Supplementary Figs. 3 and 5). We then submitted the antiSMASH job IDs to the biosynthetic gene cluster families database (BiG-FAM 1.0.0)25 for preliminary GCF explorations and classifications of annotated BGCs (Supplementary Table 3), followed by BiG-SCAPE 1.0.0 (ref. 42) refinement with a cutoff of 0.65 (Source Data Fig. 3). The GCFs were double-checked manually via the interactive network (Fig. 3), and corrections were made if necessary. A putative thiopeptide BGC (Xszus_1.region006, Xsze_2.region003, Xsto_4.region001, Xpb_30.3_21.region001, Xmir_10.region001, Xmau_6.region001, Xkoz_3.region001, Xjap_NZ_FOVO01000011.region001, Xish_1.region003, Xhom_ANU1.region005, Xhom_2.region003, Xets_11.region001, XenKK7.region002, XenDL20_c00108_NODE_12.region001, Xekj_19.region001, Xehl_28.region001, Xe30TX1_c0031_NODE_38.region001, Xdo_HBLC131_1.region001, Xdo_FRM16.1.region005, Xbov_NC_013892.1.region004, Ptem_HBLC135_17.region001, Ppb6_4.region001, Plum_TT01_1.region008, Pthr_PT1.1_23.region001, Plau_IT4.1_12.region001, Plum_IL9_35_scf0001.region001, Pbod_HU2.3_20.region001, Plau_HB1.3_105.region001, Plum_EN01_24_scf0009.region001, Pbod_DE6.1_24.region001, Plau_DE2.2_108.region001, Phpb_1.region001, Pbod_LJ_007.region001, Pbod_CN4_25_scf0020.region001, Paeg_BKT4.5_19.region001, P_tem_1.region017 and so on) that exists throughout 45 XP genomes was excluded in the analysis, because it turned out that its annotation by antiSMASH 5.0 is a false positive and early reports suggest that this cluster is responsible for ribosomal methylthiolation79,80. Two BGCs, Xdo_HBLC131_4.region001 encoding the biosynthesis of glidobactins in X. doucetiae HBLC131 and Ptem_HBLC135_2.region002 encoding the biosynthesis of ririwpeptides in P. temperata HBLC135, were artificially integrated into their respective genome by CRAGE39 previously, and thus the two BGCs were also excluded in our analysis.Pangenome analysisBiosynthetic gene cluster boundary definitionThe cluster boundary was defined by antiSMASH with the start nucleotide of the first biosynthetic gene (5′ end) and the stop nucleotide of the last biosynthetic gene (3′ end), and was manually corrected if necessary. Non-structural genes (such as transporters, regulators, transposases and so on) on the outer periphery of an operon were excluded. We compiled a table with contigs of all BGCs encoded by a given genome, BGC start and stop nucleotide positions, BGC classifications by antiSMASH and BiG-SCAPE (see the BiG-SCAPE analysis section), and possible biosynthetic pathways that the BGCs encode (Source Data Fig. 1). These tables would be integrated into the contigs databases of the pangenome for filtering the biosynthetic genes and monitoring distributions of biosynthetic gene homology groups.Interface generationAll genomes were obtained from the National Center for Biotechnology Information (NCBI). Supplementary Table 1 reports their accession numbers. The pangenome analysis herein mainly followed the anvi’o 6.1 pangenomic workflow28,78. After simplifying the header lines of 45 FASTA files for genomes using ‘anvi-script-reformat-fasta’, we converted FASTA files into anvi’o contigs databases by the ‘anvi-gen-contigs-database’ and then decorated the contigs database with hits from HMM models by ‘anvi-run-hmms’. The program ‘anvi-run-ncbi-cogs’ was run to annotate genes in the contigs databases with functions from the NCBI’s Clusters of Orthologous Groups (COGs). Tables of gene caller IDs with start and stop nucleotide positions were exported by ‘anvi-export-table’. By linking the gene caller IDs with BGCs via the start and stop nucleotide positions, genes that fell within a given BGC boundary were considered to be natural product biosynthetic genes (Source Data Fig. 1). Thereafter, the biosynthetic genes were furnished with a classification and a possible compound name, both of which were derived from the BGC that the biosynthetic genes made up. The obtained tables were imported back to contigs databases by ‘anvi-import-functions’. External genome storage was created by ‘anvi-gen-genomes-storage’ to store DNA and amino-acid sequences, as well as functional annotations of each gene. With the genome storage in hand, we used the program ‘anvi-pan-genome’ with the genomes storage database, the flag ‘–use-ncbi-blast’ and the parameter ‘–mcl-inflation 8’. The results were displayed in an interface by ‘anvi-display-pan’. The organization of the pangenome interface as shown in the dendrogram in the centre was represented by ‘presence/absence’ patterns. The core gene bin was characterized by searching the gene homology group (gene homology group represents amino-acid sequences from one or more genomes aligned by muscle81) using filters with ‘Min number of genomes gene homology group occurs, value = 45’. The singleton bin was identified by ‘Max number of genomes gene homology group occurs, value = 1’. The rest of the gene clusters that were neither sorted into the core gene bin nor the singleton bin were appended to the accessory bin. The single-copy-core-gene (scg) bin was found by ‘Min number of genomes gene homology group occurs, value = 45’ and ‘Max number of genes from each genome, value = 1’. The scg bin was refined by ‘Max functional homogeneity index 0.9’ and ‘Min geometric homogeneity index 1’. The resulting protein sequences were exported by ‘anvi-get-sequences-for-gene-clusters’ and aligned using ClustalW 1.2.2, which is incorporated in Geneious Prime 2021. Phylogenetic trees were generated using the Geneious tree builder utilizing the Jukes–Cantor distance model and the unweighted pair group method with arithmetic mean (UPGMA), and subsequently imported back to anvi’o by ‘anvi-import-misc-data’ and visualized by the interface. The statistical data of BGCs obtained from antiSMASH 5.0 (ref. 23) and BiG-SCAPE42 were imported to the layers of the interface by ‘anvi-import-misc-data’ for visualization.Biosynthetic gene and biosynthetic gene cluster filteringThe bin summary (scg, core, accessory and singleton) with BGC classifications was exported by ‘anvi-summarize’ to monitor the distributions of the biosynthetic gene homology group in the pangenomes (Source Data Fig. 1 and Supplementary Data). In the Excel sheets, ‘core’ and ‘scg’ filters were selected from the ‘bin_name’ column, and the ‘(Blank)’ filter from the ‘BGC_classification’ column was unselected. The table was then sorted by ‘genome_name’ and ‘gene_callers_id’ columns in ascending order. This then displayed consecutive core biosynthetic genes that could possibly make up a BGC. The same procedure was used to filter BGCs in the accessory or singleton region.BiG-SCAPE analysisBGCs in all genome sequences obtained from antiSMASH 5.0 (ref. 23) analyses were compared to reference BGCs from MIBiG repository 2.0 (refs. 41,82) using BiG-SCAPE 1.0.0 (ref. 42) with the PFAM database 32.0 (ref. 83). The analysis was conducted using default settings with the mode ‘auto’, mixing all classes and retaining singletons. Networks were computed for raw distance cutoffs of 0.30–0.95 in increments of 0.05. Results were visualized as a network using Cytoscape 3.7.2 (ref. 84) for a cutoff of 0.65 (Fig. 3 and Source Data Fig. 3). Statistical data for the BGCs were analysed and evaluated using Origin 2020b and Excel from Microsoft Office 365.Strain and culture conditionsWild-type strains and the mutants thereof and E. coli (Supplementary Table 14) were cultivated on lysogeny broth (LB) agar plates at 30 °C overnight, and subsequently inoculated into liquid LB culture at 30 °C with shaking at 200 r.p.m. For compound production, the overnight LB culture was transferred into 5 ml of LB, XPP19 or Sf-900 II SFM medium (1:100, vol/vol) with 2% (vol/vol) Amberlite XAD-16 resins, 0.1% l-arabinose as the inducer for mutants with a PBAD promoter, and selective antibiotics such as ampicillin (Am, 100 µg ml−1), kanamycin (Km, 50 µg ml−1) or chloramphenicol (Cm, 34 µg ml−1) at 30 °C, with shaking at 200 r.p.m.Culture extraction and HPLC-UV-MS analysisThe XAD-16 resins were collected after 72 h and extracted with 5 ml of methanol or ethyl acetate. The solvent was dried under rotary evaporators, and the dried extract was resuspended in 500 μl of methanol or acetonitrile/water (1:1 vol/vol for photoxenobactins), of which 5 μl was injected and analysed by HPLC-UV-MS or HPLC-UV-HRMS. Unless otherwise specified, HPLC-UV-MS and HPLC-UV-HRMS chromatograms in the figures are shown on the same scale. Bruker Compass DataAnalysis 4.3 was used for data collection and analysis of chromatography and MS. MetabolicDetec 2.1 was utilized to differentiate MS profiles between induced and non-induced promoter insertion mutants for identifying possible metabolites produced by targeted BGCs.Construction of PBAD promoter insertion mutantsA 500–800-bp section upstream of the target gene (lpcS, pxbF, rdb1A and xvbA) was amplified with a corresponding primer pair as listed in Supplementary Table 15. The resulting fragments were cloned using Hot Fusion85 into a pCEP_kan or pCEP_cm backbone that was amplified by pCEP_Fw and pCEP_Rv. After transformation of a constructed plasmid into E. coli S17-1 λ pir, clones were verified by PCR with primers pCEP-Ve-Fw and pDS132-Ve-Rv. A wild-type strain (X. bovienii SS-2004, X. szentirmaii DSM 16338, X. budapestensis DSM 16342 or X. vietnamensis DSM 22392) or a deletion mutant (X. szentirmaii ∆hfq, X. budapestensis ∆rdb1P or X. budapestensis ∆rdb1P ∆hfq) was used as a recipient strain. The recipient strain was mated with E. coli S17-1 λ pir (donor) carrying a constructed plasmid (Supplementary Table 16). Both strains were grown in the LB medium to an optical density at 600 nm (OD600) of 0.6 to 0.7, and the cells were washed once with fresh LB medium. Subsequently, the donor and recipient strains were mixed on an LB agar plate in ratios of 1:3 and 3:1, and incubated at 37 °C for 3 h followed by incubation at 30 °C for 21 h. After that, the bacterial cell layer was collected with an inoculating loop and resuspended in 2 ml of fresh LB medium. A 200-μl sample of the resuspended culture was spread out on an LB agar plate with Am/Km or Am/Cm and incubated at 30 °C for two days. Individual insertion clones were cultivated and analysed by HPLC-UV-HRMS, and the genotype of all mutants was verified by plasmid- and genome-specific primers.Construction of deletion mutantsA ~1,000-bp upstream and a ~1,000-bp downstream fragment of hfq in X. budapestensis DSM 16342 were amplified using the primer pairs listed in Supplementary Table 15. The amplified fragments were fused using the complementary overhangs introduced by primers and cloned into the pEB17 vector that was linearized with PstI and BglII by Hot Fusion85. Transformation of E. coli S17-1 λ pir with the resulting plasmid (Supplementary Table 16) and conjugation with X. budapestensis DSM 16342, as well as the generation of double crossover mutants via counterselection on LB plates containing 6% sucrose, were carried out as previously described86. The deletion mutant was verified via PCR using the primer pairs listed in Supplementary Table 15, which yielded a ~2,000-bp fragment for mutants genetically equal to the WT strain and a ~1,000-bp fragment for the desired deletion mutant. The same procedure was used to generate Δrdb1P mutants, during which E. coli S17-1 λ pir carrying pEB17 rdb1P was mated with the X. budapestensis DSM 16342 wild-type and X. budapestensis ∆hfq mutant.Labelling experiments for structural elucidation of photoxenobactins C and D by MSThe cultivation of strains for labelling experiments was carried out as described above. For photoxenobactin C (6) labelling experiments, the overnight culture was transferred into LB medium additionally fed with 4-fluorosalicylate-SNAC, l-methionine-(methyl-d3), l-[U-13C,15N]cysteine and l-[U-34S]cysteine at a final concentration of 1 mM. In terms of inverse feeding experiments, cell pellets of the 100-μl overnight culture were washed once with ISOGRO 13C or 13C,15N medium (100 μl) and resuspended in the corresponding isotope labelling medium (100 μl). The feeding culture in the isotope labelling medium (5 ml) was inoculated with a washed overnight culture (50 μl) and additional l-cysteine was added at a final concentration of 1 mM.For photoxenobactin D (7) labelling experiments, the cell pellets of the 100-μl overnight culture were washed once with ISOGRO 13C or 15N medium (100 μl) and then resuspended in the corresponding isotope labelling medium (100 μl). A 5-ml isotope labelling medium was inoculated with a washed overnight culture (50 μl).Isolation and purificationFor photoxenobactin isolation, 10 ml of LB medium was inoculated with a colony of the X. szentirmaii PBAD pxbF ∆hfq mutant from an LB agar plate and cultivated overnight. A 10-ml culture was taken to inoculate 2 × 100 ml of LB medium (OD600 ≈ 0.1). The 2 × 100-ml cultures were incubated overnight and the whole culture volume (200 ml) was used to inoculate a 20-l LB fermenter (Braun) supplemented with 2% XAD-16 and 0.2% arabinose (antifoam was added when required). Fermenter settings were as follows: 30 °C without pH control, three six-blade impellers 150 r.p.m. After 24 h, 10 l of the culture was collected from the fermenter, and the XAD resins were separated from the cells by filtration. (1) The XAD resins were extracted with 2 × 1 l of ethyl acetate with 1% formic acid, and the combined organic phase was dried under reduced pressure. (2) The culture without XAD was centrifuged and the supernatant was extracted with 3 × 5 l of ethyl acetate with 1% formic acid, and the combined organic layers were dried under reduced pressure. (3) The cell pellet was extracted with 2 × 1 l of ethyl acetate with 1% formic acid, and the organic supernatant was dried under reduced pressure. After 48 h, the remaining 10 l of bacterial culture were extracted as described in steps (1) to (3). The combined extracts from 20 l of culture were fractionated by a flash purification system with a C18 column with a gradient elution of acetonitrile/water 20–100% at 20 ml min−1 (every 10% gradient step was performed with five column volumes, except the 60–70% step, which was performed with ten column volumes). Fractions containing photoxenobactins were combined and dried under reduced pressure. Final purification was achieved via preparative and semipreparative HPLCs with a gradient of 30% acetonitrile/water (0–30 min) and 30–100% acetonitrile/water (30–40 min). The fractions were combined in brown flasks and were immediately freeze-dried to afford photoxenobactin A (4, 0.8 mg), photoxenobactin B (5, 0.6 mg), photoxenobactin C (6, 1.2 mg) and photoxenobactin E (8, 2.2 mg).For the isolation and purification of lipocitides A and B, 2% of XAD-16 resins from a 6-l LB culture of the X. bovienii PBAD lpcS mutant induced by l-arabinose were collected after 72 h of incubation at 30 °C with shaking at 120 r.p.m., and were washed with water and extracted with methanol (3 × 1 l) to yield a crude extract (5.3 g after evaporation). The extract was dissolved in methanol and was subjected to preparative HPLC with a C18 column using an acetonitrile/water gradient (0.1% formic acid) for 0–32 min, 55–80%, 40 ml min−1 to afford lipocitides A (17, 4.8 mg) and B (18, 9.0 mg).Two percent of XAD-16 resins from a 12-l LB culture of the X. budapestensis PBAD rdb1A ∆rdb1P ∆hfq mutant induced by l-arabinose were collected after 72 h of incubation at 30 °C with shaking at 120 r.p.m. and washed with water and extracted with methanol (3 × 2 l) to yield a crude extract (15.3 g after evaporation). The extract was subject to a Sephadex LH-20 column eluted with methanol. The fraction (2.8 g) containing pre-rhabdobranins was subjected to preparative HPLC with a C18 column using an acetonitrile/water gradient (0.1% formic acid) for 0–20 min, 15–35%, 40 ml min−1 to afford a fraction (206 mg) mainly containing pre-rhabdobranin D, which was further purified by semipreparative HPLC with a C18 column using an acetonitrile/water gradient (0.1% formic acid) for 0–24 min, 5–53%, 3 ml min−1 to afford pre-rhabdobranin D (27, 59.1 mg).Benzobactin A (28) and its methyl ester (29), which were detected in X. vietnamensis PBAD xvbA, were also produced by Pseudomonas chlororaphis subsp. piscium DSM 21509 (unpublished). Owing to the high production level in Pseudomonas chlororaphis subsp. piscium DSM 21509, 28 and 29 were isolated from the Pseudomonas strain. Four percent of XAD-16 resins from a 12-l XPP culture of Pseudomonas chlororaphis subsp. piscium DSM 21509 PBAD pbzA mutant induced by l-arabinose were collected after 72 h of incubation at 30 °C with shaking at 120 r.p.m., and washed with water and extracted with methanol (3 × 2 l) to yield a crude extract (95.4 g after evaporation). The extract was dissolved in methanol and subjected to preparative HPLC with a C18 column using an acetonitrile/water gradient (0.1% formic acid) for 0–18 min, 5–59%, 20 ml min−1 to afford ten fractions. Fractions 2 (95.6 mg) and 3 (50.7 mg) were further purified by semipreparative HPLC with a C18 column using an acetonitrile/water gradient (0.1% formic acid) for 0–35 min, 5–95%, 3 ml min−1 to afford benzobactin A (28, 3.2 mg) and its methyl ester (29, 0.9 mg), respectively.NMR spectroscopyMeasurements were carried out using 1H and 13C NMR, 1H-13C heteronuclear single quantum coherence (HSQC), 1H-13C heteronuclear multiple bond correlation (HMBC), 1H-1H correlation spectroscopy (COSY), 1H-13C heteronuclear multiple quantum correlation/1H-1H correlation spectroscopy (HMQC-COSY) and 1H-13C heteronuclear single quantum coherence/1H-1H total correlation spectroscopy (HSQC-TOCSY). Chemical shifts (δ) were reported in parts per million (ppm) and referenced to the solvent signals. Data are reported as follows: chemical shift, multiplicity (br = broad, s = singlet, d = doublet, t = triplet, dd = doublet of doublet, m = multiplet and ov = overlapped) and coupling constants (in hertz). Bruker TopSpin 4.0 was used for NMR data collection and spectral interpretation.General synthetic proceduresThe Fmoc protecting group was removed with 2 ml of 40% piperidine/dimethylformamide (DMF; 5 min) followed by 2 ml of 20% piperidine/DMF (10 min). Washings between coupling and deprotection steps were performed with DMF (five syringe volumes) and dichloromethane (DCM) (five syringe volumes). Resin loadings were determined by Fmoc cleavage from a weighted resin sample87. The combined filtrates containing Fmoc cleavage products were quantified spectrophotometrically at 301 nm using a UV–vis spectrophotometer with Hellma absorption cuvettes with a path length of 1 cm. Loadings were calculated (in mmol resin) using Lambert–Beer’s law with ɛ = 7,800 M−1 cm−1: loading (mmol) = ({frac{{{rm{Abs}},{({rm{sample}})}}}{{varepsilon l}}} times V), where ɛ is the molar extinction coefficient, V is the sample volume in liter and l is the optical path length in cm. Final cleavage was achieved by shaking the resin in 2 ml of a mixture of TFA/TIPS/H2O (95:2.5:2.5) for 1 h. The filtrate was then collected and the resin washed three times (2 ml each) with DCM, and the combined filtrates were dried under reduced pressure.Syntheses of lipocitide AFmoc-protected Rink Amide resin (192 mg, 0.52 mmol g−1, 0.1 mmol) was placed in a polypropylene 6-ml syringe vessel fitted with polyethylene porous filter disks and swollen in 3 ml of DMF for 10 min. Subsequently, the Fmoc-protected resin was deprotected and then washed as described in the general synthetic procedures. Fmoc-d-Cit-OH (198.0 mg, 0.5 mmol, 5 equiv.), 1-hydroxy-7-azabenzotriazole (HOAT, 0.83 ml, 0.5 mmol, 5 equiv.), hexafluorophosphate azabenzotriazole tetramethyl uronium (HATU, 190.5 mg, 0.5 mmol, 5 equiv.) and N,N-diisopropylethylamine (DIPEA, 170 μl, 1.0 mmol, 10 equiv.) were dissolved in 1.5 ml of dry DMF. After 5 min, the clear solution was added to the resin and shaken at room temperature overnight. The resin was washed and loading was calculated (79.2%) as described in the general synthetic procedures. Acylation of Fmoc-l-Ala-OH (74.1 mg, 0.24 mmol, 3 equiv.), Fmoc-d-Leu-OH (84.8 mg, 0.24 mmol, 3 equiv.) and myristic acid (54.8 mg, 0.24 mmol, 3 equiv.) were carried out using the abovementioned procedure. Final cleavage was performed as described in the general synthetic procedures, and the crude product (70.8 mg) was purified by HPLC to obtain lipocitide A (17, Supplementary Fig. 100; 24.3 mg, 54.0%) as a white solid.Syntheses of lipocitide B2-CTC resin (63 mg, 1.6 mmol g−1, 0.1 mmol) was placed in a polypropylene 6-ml syringe vessel fitted with polyethylene porous filter disks. The resin was incubated with Fmoc-d-Cit-OH (119.0 mg, 0.3 mmol, 3 equiv.) and DIPEA (153 μl, 0.9 mmol, 9 equiv.) in 1.5 ml of dry DCM at room temperature overnight. The resin was washed and loading was calculated (56.7%) as described in the general synthetic procedures. Acylations of Fmoc-l-Ala-OH (52.9 mg, 0.17 mmol, 3 equiv.), Fmoc-d-Leu-OH (60.1 mg, 0.17 mmol, 3 equiv.) and myristic acid (38.9 mg, 0.24 mmol, 3 equiv.) were performed with additional HOAT (0.47 ml, 0.28 mmol, 5 equiv.), HATU (108 mg, 0.28 mmol, 5 equiv.) and DIPEA (96 μl, 0.56 mmol, 10 equiv.). Final cleavage was carried out as described in the general synthetic procedures, and the crude (54.2 mg) was purified by HPLC to obtain lipocitide B (18, Supplementary Fig. 101; 18.6 mg, 57.6%) as a white solid.Synthesis of S-(2-acetamidoethyl)4-fluoro-2-hydroxybenzothioate (4-fluorosalicylate SNAC)To a solution of 4-fluorosalicylic acid (156 mg, 1.0 mmol, 1.0 equiv.) and hydroxybenzotriazole (HOBt, 162 mg, 1.2 mmol, 1.2 equiv.) in 45 ml of THF, N,N′-dicyclohexylcarbodiimide (DCC, 248 mg, 1.2 mmol, 1.2 equiv.) was added, followed by N-acetylcysteamine (112 µl, 1.0 mmol, 1.0 equiv.). After 1 h at room temperature, K2CO3 (138 mg, 1.0 mmol, 1.2 equiv.) was added and the reaction was stirred for an additional 2 h. The reaction mixture was then filtered and concentrated by rotary evaporation. The solid residue was dissolved in ethyl acetate and washed with sat. NaHCO3 (50 ml) and water (50 ml). The organic layer was dried over MgSO4, concentrated, and purified by flash chromatography (1–10% MeOH in CHCl3) to give 26 mg (10%) S-(2-acetamidoethyl)4-fluoro-2-hydroxybenzothioate (Supplementary Fig. 102).IC50 value determination with the purified yeast 20S proteasome core particleYeast 20S proteasome core particle (yCP) from Saccharomyces cerevisiae was purified according to previously described methods88,89. The concentration of purified yCP was determined spectrophotometrically at 280 nm. yCP (final concentration: 0.05 mg ml−1 in 100 mM Tris-HCl, pH 7.5) was mixed with dimethyl sulfoxide (DMSO) as a control or serial dilutions of IOC (1) in DMSO, thereby not surpassing a final concentration of 10% (vol/vol) DMSO. After an incubation time of 45 min at room temperature, fluorogenic substrates Boc-Leu-Arg-Arg-AMC (AMC, 7-amino-4-methylcoumarin), Z-Leu-Leu-Glu-AMC and Suc-Leu-Leu-Val-Tyr-AMC (final concentration of 200 µM) were added to measure the residual activity of caspase-like (C-L, β1 subunit), trypsin-like (T-L, β2 subunit) and chymotrypsin-like (ChT-L, β5 subunit), respectively. The assay mixture was incubated for another 60 min at room temperature, then diluted 1:10 in 20 mM Tris-HCl, pH 7.5. The AMC molecules released by hydrolysis were measured in triplicate with a Varian Cary Eclipse fluorescence spectrophotometer (Agilent Technologies) at λexc = 360 nm and λem = 460 nm. Relative fluorescence units were normalized to the DMSO-treated control. The calculated residual activities were plotted against the logarithm of the applied inhibitor concentration and fitted with GraphPad Prism 9.0.2. IC50 values were deduced from the fitted data. These depend on enzyme concentration and are comparable within the same experimental settings.Crystallization and structure determination of the yCP in complex with IOC (1)Crystals of the yCP were grown in hanging drops at 20 °C, as previously described88,89. The protein concentration used for crystallization was 40 mg ml−1 in Tris/HCl (20 mM, pH 7.5) and EDTA (1 mM). The drops contained 1 μl of protein and 1 μl of the reservoir solution (30 mM magnesium acetate, 100 mM 2-(N-morpholino)ethanesulfonic acid (pH 6.7) and 10% (wt/vol) 2-methyl-2,4-pentanediol). Crystals appeared after two days and were incubated with 1 at a final concentration of 10 mM for at least 24 h. Droplets were then complemented with a cryoprotecting buffer (30% (wt/vol) 2-methyl-2,4-pentanediol, 15 mM magnesium acetate, 100 mM 2-(N-morpholino)ethanesulfonic acid, pH 6.9) and vitrified in liquid nitrogen. The dataset from the yCP:IOC complex was collected using synchrotron radiation (λ = 1.0 Å) at the X06SA-beamline (Swiss Light Source). X-ray intensities and data reduction were evaluated using the XDS program package version 5 February 2021 (Supplementary Table 17)90. Conventional crystallographic rigid body, positional and temperature factor refinements were carried out with REFMAC5 5.0.32 (ref. 91) and the CCP4 Program Suite 7.1.016 (ref. 92) using coordinates of the yCP structure as the starting model (PDB 5CZ4)50. Model building was performed by the programs SYBYL-X and COOT 0.8.7 (ref. 93). The final coordinates yielded excellent residual factors, as well as geometric bond and angle values. Coordinates were confirmed to fulfil the Ramachandran plot and have been deposited in the RCSB (PDB 7O2L).Haemocyte-spreading assaysSpodoptera exigua larvae were collected from Welsh onion (Allium fistulsum L.) fields in Andong, Korea. Insects were reared in the laboratory under the following conditions: 25 ± 2 °C constant temperature, 16:8 h (light/dark) photoperiod and 60 ± 5% relative humidity. Larvae were reared on an artificial diet94 and 10% sucrose solutions were fed to adult insects. Fifth instar larvae were used in all experiments. For analysing haemocyte behaviours in vivo, fifth instar larvae of S. exigua were co-injected with 1 µl of heat-killed (95 °C for 10 min) E. coli TOP10 (2.4 × 104 cells per larva) with the test compound (0–1,000 ng per larva) by using a Hamilton microsyringe (Reno). At 1 h post-injection, 10 µl of haemolymph from each larva was collected on the glass slide and incubated for 5 min inside a dark wet chamber at room temperature. The medium was replaced with 3.7% of formaldehyde dissolved in phosphate buffered saline (PBS) and incubated for 10 min. After washing three times with PBS, cells were permeabilized with 0.2% Triton X-100 in PBS for 2 min at room temperature. After incubation, the slides were washed with PBS three times. Blocking was performed using 5% skimmed milk (Invitrogen) dissolved in PBS, followed by incubation for 10 min. After washing once with PBS, the cells were incubated with fluorescein isothiocyanate (FITC)-tagged phalloidin in PBS for 1 h at room temperature. After washing three times, the cells were incubated with 4′,6-diamidino-2-phenylindole (DAPI, 1 mg ml−1, Thermo Scientific) in PBS for nucleus staining. Finally, after washing twice in PBS, cells were observed under a fluorescence microscope (DM2500, Leica) at ×400 magnification. Haemocyte spreading was determined by the extension of F-actin out of the original cell boundary. For the in vitro assay, ~100 μl of haemolymph was collected into 400 μl of anticoagulation buffer (ACB; 186 mM NaCl, 17 mM Na2EDTA, 41 mM citric acid, pH 4.5). After adding ACB, the medium was incubated for 30 min on ice. After centrifugation at 300g for 5 min, 400 μl of supernatant was discarded. The rest of the suspension was gently mixed with 200 μl of TC100 insect tissue culture medium (Welgene). From this suspension, 10 µl of haemolymph was collected on the glass slide. The slides were co-injected with 1 µl of E. coli TOP10 (2.4 × 104 cells per larva) with the test compound (0–1,000 ng per larva), followed by the procedure described above. Means were compared by a least squared difference (LSD) test of one-way analysis of variance (ANOVA) using POC GLM of the SAS program (SAS Institute, 1989) and discriminated at type I error = 0.05.Nodulation assaysE. coli TOP10 was heat-killed by incubating at 95 °C for 10 min. Fifth-instar larvae of S. exigua were injected with 1 µl of bacteria (2.4 × 104 cells per larva) using a Hamilton microsyringe along with 1 µl of different concentrations (10, 50, 100, 500 and 1,000 ppm) of inhibitors. Control larvae were injected with bacteria and DMSO. At 8 h after bacterial injection, nodules were counted by dissecting larvae under a stereomicroscope (Stemi SV 11, Zeiss) at ×50 magnification.Phenoloxidase activity assaysThe PO activity from plasma was estimated as previously described95. Briefly, DOPA (l-3,4-dihydroxyphenylalanine) was used as a substrate for determining PO activity from treated larvae plasma. For PO activation, each fifth-instar larva of S. exigua was challenged with 2.4 × 104 cells of heat-killed E. coli TOP10. Different inhibitors were co-injected (1 µg per larva) along with E. coli TOP10. After 8 h of bacterial challenge, haemolymph was collected from treated larvae in a 1.5-ml tube containing a few granules of phenylthiocarbamide (Sigma-Aldrich) to prevent melanization. Haemocytes were separated from plasma by centrifuging at 4 °C for 5 min at 300g. A reaction volume of 200 µl consisted of 180 µl of 10 mM DOPA in PBS (pH 7.4) and 20 µl of plasma. Absorbance was measured using a VICTOR multi-label plate reader (PerkinElmer) at 490 nm. PO activity was expressed as ΔABS per min per µl of plasma. Each treatment was replicated three times with independent samples.Measurement of nitric oxideThe NO was indirectly quantified by measuring its oxidized form, nitrate (NO3−), using the Griess reagent of a Nitrate/Nitrite Colorimetric Assay Kit (Cayman Chemical). Fifth-instar larvae were injected with 1 µl of heat-killed E. coli TOP10 (2.4 × 104 cells per larva) using a Hamilton microsyringe along with 1 µl of the test compound. Haemolymph was collected from each sample 1 h post infection. A 150-μl volume of haemolymph from three L5 larvae was collected and homogenized in 350 μl of 100 mM PBS pH 7.4 with a homogenizer (Ultra-Turrax T8, Ika Laboratory). After centrifugation at 14,000g for 20 min at 4 °C, the supernatant was used to measure the nitrate amounts, and the total protein was measured in each sample by a Bradford assay. The samples were analysed in a 200-μl final reaction volume. Briefly, 80 μl of samples were added to the wells, then 10 μl of enzyme cofactor mixture and 10 μl of nitrate reductase mixture were added. After incubation at room temperature for 1 h, 50 μl of Griess reagent R1 and immediately 50 μl of Griess reagent R2 were added to each well. The plate was left at room temperature for 10 min for colour development. For a standard curve to quantify the nitrate concentrations of the samples, nitrates with final concentrations of 0, 5, 10, 15, 20, 25, 30 and 35 μM in a 200-μl reaction volume were used. The absorbance was recorded at 540 nm on a VICTOR multi-label plate reader. Our measurements used three larvae per sample, and we repeated the treatment with three biological samples.Galleria injection assaysPrecultures of X. szentirmaii DSM wild-type strain and the mutants thereof were grown in LB medium and inoculated into fresh cultures at an OD600 of 0.1. Cells were grown to exponential phase (OD600 ≈ 1) and then diluted to an OD600 of 0.00025. A 5-µl volume of the diluted bacterial culture was injected into the last left pro-leg of the larvae (LB medium as a negative control). G. mellonella larvae were kept at 4 °C for 10 min before injection. After infection, the larvae were incubated at 25 °C. Dead Galleria larvae were frozen at −20 °C, then at −80 °C, and freeze-dried for one day. Freeze-dried larvae were ground. Every injection experiment was aliquoted into two portions, one of which was extracted with 25 ml of acetone/ethyl acetate (vol/vol, 1:1) while the other one was extracted with acetone/methanol. Extracts were dried and resuspended in 3 ml of acetonitrile/water (1:1, vol/vol) with a tenfold dilution for HPLC-MS-UV analysis. To compare the survival percentage of G. mellonella larvae infected with the WT strain and mutants and to determine median lethal time (LT50) values, Kaplan–Meier curves were generated by GraphPad PRISM 8.4.3.Cytotoxicity assaysHepG2 cells (hepatoblastoma cell line; ACC 180, DSMZ) were cultured under conditions recommended by the depositor, and cells were propagated in Dulbecco’s modified Eagle medium supplemented with 10% fetal bovine serum. To determine the cytotoxicity of test compounds, cells were seeded at 6 × 103 cells per well of 96-well plates in 120 μl of complete medium. After 2 h of equilibration, compounds were added in serial dilution in 60 µl of complete medium. Compounds as well as the solvent control and doxorubicin as an in-assay positive control (IC50 of 0.06 ± 0.01 µg ml−1) were tested as duplicates in two independent experiments. After 5 days of incubation, 20 μl of 5 mg ml−1 MTT (thiazolyl blue tetrazolium bromide) in PBS was added per well, and the cells were further incubated for 2 h at 37 °C. The medium was then discarded and cells were washed with 100 μl of PBS before adding 100 μl of 2-propanol/10 N HCl (250:1) to dissolve the formazan granules. The absorbance at 570 nm was measured using a microplate reader (Tecan Infinite M200Pro with Tecan iControl 2.0), and cell viability was expressed as a percentage relative to the respective solvent control. IC50 values were determined by sigmoidal curve fitting using GraphPad PRISM 8.4.3.Statistical analysisIn Fig. 5d,e,g,j,k, means were compared using an LSD test of one-way ANOVA using POC GLM of the SAS program (SAS Institute, 1989) for continuous variables and discriminated at type I error = 0.05. The results were plotted using Sigma Plot 12.0.Reporting SummaryFurther information on research design is available in the Nature Research Reporting Summary linked to this Article. More

  • in

    Assessing placement bias of the global river gauge network

    Vörösmarty, C. J. et al. Global threats to human water security and river biodiversity. Nature 467, 555–561 (2010).Article 
    CAS 

    Google Scholar 
    Ruhi, A., Messager, M. L. & Olden, J. D. Tracking the pulse of the Earth’s fresh waters. Nat. Sustain. 1, 198–203 (2018).Article 

    Google Scholar 
    Pearson, C. Short- and medium-term climate information for water management. World Meteorol. Organ. Bull. 57, 173–177 (2008).
    Google Scholar 
    Tetzlaff, D., Carey, S. K., McNamara, J. P., Laudon, H. & Soulsby, C. The essential value of long-term experimental data for hydrology and water management. Water Resour. Res. 53, 2598–2604 (2017).Article 

    Google Scholar 
    Carlisle, D. M., Wolock, D. M. & Meador, M. R. Alteration of streamflow magnitudes and potential ecological consequences: a multiregional assessment. Front. Ecol. Environ. 9, 264–270 (2011).Article 

    Google Scholar 
    Shrestha, S., Kazama, F. & Newham, L. T. H. A framework for estimating pollutant export coefficients from long-term in-stream water quality monitoring data. Environ. Model. Softw. 23, 182–194 (2008).Article 

    Google Scholar 
    Lepistö, A., Futter, M. N. & Kortelainen, P. Almost 50 years of monitoring shows that climate, not forestry, controls long-term organic carbon fluxes in a large boreal watershed. Glob. Change Biol. 20, 1225–1237 (2014).Article 

    Google Scholar 
    Hester, G., Ford, D., Carsell, K., Vertucci, C. & Stallings, E. A. Flood Management Benefits of USGS Streamgaging Program (National Hydrologic Warning Council, 2006).Xu, H., Xu, C.-Y., Chen, H., Zhang, Z. & Li, L. Assessing the influence of rain gauge density and distribution on hydrological model performance in a humid region of China. J. Hydrol. 505, 1–12 (2013).Article 

    Google Scholar 
    Kiang, J. E., Stewart, D. W., Archfield, S. A., Osborne, E. B. & Eng, K. A National Streamflow Network Gap Analysis (USGS, 2013).Deweber, J. T. et al. Importance of understanding landscape biases in USGS gage locations: implications and solutions for managers. Fisheries 39, 155–163 (2014).Article 

    Google Scholar 
    Tickner, D. et al. Bending the curve of global freshwater biodiversity loss: an emergency recovery plan. BioScience 70, 330–342 (2020).Article 

    Google Scholar 
    Grill, G. et al. Mapping the world’s free-flowing rivers. Nature 569, 215–221 (2019).CAS 
    Article 

    Google Scholar 
    Olden, J. D. et al. Hydrologic classification of Tanzanian rivers to support national water resource policy. Ecohydrology. https://doi.org/10.1002/eco.2282 (2021).Lin, P. et al. Global reconstruction of naturalized river flows at 2.94 million reaches. Water Resour. Res. 55, 6499–6516 (2019).Article 

    Google Scholar 
    Yamazaki, D. et al. MERIT Hydro: a high-resolution global hydrography map based on latest topography dataset. Water Resour. Res. 55, 5053–5073 (2019).Article 

    Google Scholar 
    Beck, H. E. et al. Bias correction of global high-resolution precipitation climatologies using streamflow observations from 9,372 catchments. J. Clim. 33, 1299–1315 (2020).Article 

    Google Scholar 
    Do, H. X., Gudmundsson, L., Leonard, M. & Westra, S. The Global Streamflow Indices and Metadata Archive (GSIM)—part 1: the production of a daily streamflow archive and metadata. Earth Syst. Sci. Data 10, 765–785 (2018).Article 

    Google Scholar 
    Linke, S. et al. Global hydro-environmental sub-basin and river reach characteristics at high spatial resolution. Sci. Data 6, 283 (2019).Article 

    Google Scholar 
    Dobrushin, R. L. Prescribing a system of random variables by conditional distributions. Theory Probab. Appl. 15, 458–486 (1970).Article 

    Google Scholar 
    Schefzik, R., Flesch, J. & Goncalves, A. Fast identification of differential distributions in single-cell RNA-sequencing data with waddR. Bioinformatics 37, 3204–3211 (2021).CAS 
    Article 

    Google Scholar 
    Reid, A. J. et al. Emerging threats and persistent conservation challenges for freshwater biodiversity. Biol. Rev. 94, 849–873 (2019).Article 

    Google Scholar 
    Wittemyer, G., Elsen, P., Bean, W. T., Burton, A. C. O. & Brashares, J. S. Accelerated human population growth at protected area edges. Science 321, 123–126 (2008).CAS 
    Article 

    Google Scholar 
    Colvin, S. A. R. et al. Headwater streams and wetlands are critical for sustaining fish, fisheries, and ecosystem Services. Fisheries 44, 73–91 (2019).Article 

    Google Scholar 
    Chen, K. & Olden, J. D. Threshold responses of riverine fish communities to land use conversion across regions of the world. Glob. Change Biol. 26, 4952–4965 (2020).Article 

    Google Scholar 
    Pardo, I. et al. The European reference condition concept: a scientific and technical approach to identify minimally-impacted river ecosystems. Sci. Total Environ. 420, 33–42 (2012).CAS 
    Article 

    Google Scholar 
    Sauquet, E. et al. Classification and trends in intermittent river flow regimes in Australia, northwestern Europe and USA: a global perspective. J. Hydrol. 597, 126170 (2021).Article 

    Google Scholar 
    Creed, I. F. et al. Enhancing protection for vulnerable waters. Nat. Geosci. 10, 809–815 (2017).CAS 
    Article 

    Google Scholar 
    Abell, R. et al. Freshwater ecoregions of the world: a new map of biogeographic units for freshwater biodiversity conservation. BioScience 58, 403–414 (2008).Article 

    Google Scholar 
    Wilhite, D. A. in Coping with Drought Risk in Agriculture and Water Supply Systems: Drought Management and Policy Development in the Mediterranean, Vol. 26 (eds. Iglesias, A. et al.) 3–19 (Springer Science and Business Media, 2009).Winemiller, K. O. et al. Balancing hydropower and biodiversity in the Amazon, Congo, and Mekong. Science 351, 128–129 (2016).CAS 
    Article 

    Google Scholar 
    Seyfried, M. S. & Wilcox, B. P. Scale and the nature of spatial variability: field examples having implications for hydrologic modeling. Water Resour. Res. 31, 173–184 (1995).Article 

    Google Scholar 
    Hammond, J. C. et al. Spatial patterns and drivers of nonperennial flow regimes in the contiguous United States. Geophys. Res. Lett. 48, e2020GL090794 (2021).Article 

    Google Scholar 
    Messager, M. L. et al. Global prevalence of non-perennial rivers and streams. Nature 594, 391–397 (2021).CAS 
    Article 

    Google Scholar 
    Busch, M. H. et al. What’s in a name? Patterns, trends, and suggestions for defining non-perennial rivers and streams. Water 12, 1980 (2020).Article 

    Google Scholar 
    Zipper, S. C. et al. Pervasive changes in stream intermittency across the United States. Environ. Res. Lett. 16, 084033 (2021).Article 

    Google Scholar 
    Jaeger, K. L., Olden, J. D. & Pelland, N. A. Climate change poised to threaten hydrologic connectivity and endemic fishes in dryland streams. Proc. Natl Acad. Sci. USA 111, 13894–13899 (2014).CAS 
    Article 

    Google Scholar 
    Beaufort, A., Lamouroux, N., Pella, H., Datry, T. & Sauquet, E. Extrapolating regional probability of drying of headwater streams using discrete observations and gauging networks. Hydrol. Earth Syst. Sci. 22, 3033–3051 (2018).Article 

    Google Scholar 
    Argerich, A. et al. Comprehensive multiyear carbon budget of a temperate headwater stream: carbon budget of a headwater stream. J. Geophys. Res. Biogeosci. 121, 1306–1315 (2016).CAS 
    Article 

    Google Scholar 
    Molden, D. J., Shrestha, A. B., Nepal, S. & Immerzeel, W. W. in Water Security, Climate Change and Sustainable Development (eds. Biswas, A. K. & Tortajada, C.) 65–82 (Springer, 2016).Kaletová, T. et al. Relevance of intermittent rivers and streams in agricultural landscape and their impact on provided ecosystem services—a Mediterranean case study. Int. J. Environ. Res. Public Health 16, 2693 (2019).Article 

    Google Scholar 
    Zimmer, M. A. et al. Zero or not? Causes and consequences of zero-flow stream gage readings. WIREs Water 7, e1436 (2020).Article 

    Google Scholar 
    Wine, M. L. Toward strong science to support equitable water sharing in securitized transboundary watersheds. Biologia 9, 907–915 (2020).Article 

    Google Scholar 
    Alsdorf, D. E. GEOPHYSICS: tracking fresh water from space. Science 301, 1491–1494 (2003).CAS 
    Article 

    Google Scholar 
    Benstead, J. P. & Leigh, D. S. An expanded role for river networks. Nat. Geosci. 5, 678–679 (2012).CAS 
    Article 

    Google Scholar 
    Allen, D. C. et al. Citizen scientists document long-term streamflow declines in intermittent rivers of the desert southwest, USA. Freshw. Sci. https://doi.org/10.1086/701483 (2019).Joo, H. et al. Optimal stream gauge network design using entropy theory and importance of stream gauge stations. Entropy 21, 991 (2019).Article 

    Google Scholar 
    Vörösmarty, C. et al. Global water data: a newly endangered species. Eos 82, 54–58 (2001).Article 

    Google Scholar 
    Jordahl, K. et al. Geopandas/geopandas. Zenodo https://doi.org/10.5281/zenodo.3946761 (2020).Lin, P., Pan, M., Wood, E. F., Yamazaki, D. & Allen, G. H. A new vector-based global river network dataset accounting for variable drainage density. Sci. Data 8, 28 (2021).Article 

    Google Scholar 
    Yu, S. et al. Evaluating a landscape-scale daily water balance model to support spatially continuous representation of flow intermittency throughout stream networks. Hydrol. Earth Syst. Sci. 24, 5279–5295 (2020).CAS 
    Article 

    Google Scholar 
    Kennard, M. J. et al. Classification of natural flow regimes in Australia to support environmental flow management. Freshw. Biol. 55, 171–193 (2010).Article 

    Google Scholar 
    Flow/No Flow Observations with Discharge Data from Probabilistic Stream Surveys (US EPA Office of Research and Development, 2021).Rosenbaum, P. R. & Rubin, D. B. The bias due to incomplete matching. Biometrics 41, 103–116 (1985).CAS 
    Article 

    Google Scholar  More