More stories

  • in

    Community confounding in joint species distribution models

    Historically, species distributions have been modeled independently from each other due to unavailability of multispecies datasets and computational restraints. However, ecological datasets that provide insights about collections of organisms have become prevalent over the last decade thanks to efforts like Long Term Ecological Research Network (LTER), National Ecological Observatory Network (NEON), and citizen science surveys1. In addition, technology has improved our ability to fit modern statistical models to these datasets that account for both species environmental preferences and interspecies dependence. These advancements have allowed for the development of joint species distribution models (JSDM)2,3,4 that can model dependence among species simultaneously with environmental drivers of occurrence and/or abundance.Species distributions are shaped by both interspecies dynamics and environmental preferences5,6,7,8. JSDMs integrate both sources of variability and adjust uncertainty to reflect that multiple confounded factors can contribute to similar patterns in species distributions. Some have proposed that JSDMs not only account for biotic interactions but also correct estimates of association between species distributions and environmental drivers3,9, while others claim JSDMs cannot disentangle the roles of interspecies dependence and environmental drivers5. We address why JSDMs can provide inference distinct from their concomitant independent SDMs, how certain parameterizations of a JSDM induce confounding between the environmental and random species effects, and when deconfounding these effects may be appealing for computation and interpretation.Because of the prevalence of occupancy data for biomonitoring in ecology, we focus our discussion of community confounding in JSDMs on occupancy models, although we also consider a JSDM for species density data in the simulation study. The individual species occupancy model was first formulated by MacKenzie et al.10 and has several joint species extensions4,11,12,13,14,15,16. We chose to investigate the impacts of community confounding on the probit model since it has been widely used in the analysis of occupancy data4,13,17. We also developed a joint species extension to the Royle-Nichols model18 and consider community confounding in that model.We use the probit and Royle-Nichols occupancy models to improve our understanding of montaine mammal communities in what follows. We show that including unstructured random species effects in either occupancy model induces confounding between the fixed environmental and random species effects. We demonstrate how to orthogonalize these effects in the model and compare the resulting inference compared to models where species are treated independently.Unlike previous approaches that have applied restricted regression techniques similar to ours, we use it in the context of well-known ecological models for species occupancy and intensity. While such approaches have been discussed in spatial statistics and environmental science, they have not been adopted in settings involving the multivariate analysis of community data. We draw parallels between restricted spatial regression and restricted JSDMs but also highlight where the methods differ in goals and outcomes. We find that the computational benefits conferred by performing restricted spatial regression also hold for some joint species distribution models.Royle-Nichols joint species distribution modelWe present a JSDM extension to the Royle-Nichols model18. The Royle-Nichols model accounts for heterogeneity in detection induced by the species’ latent intensity, a surrogate related to true species abundance. Abundance, density, and occupancy estimation often requires an explicit spatial region that is closed to emmigration and immigration. In our model, the unobservable intensity variable helps us explain heterogeneity in the frequencies we observe a species at different sites without making assumptions about population closure. In the “Model” section, we further discuss the distinctions between abundance and intensity in the Royle-Nichols model.The Royle-Nichols model utilizes occupancy survey data but provides inference distinct from the basic occupancy model10. In the Royle-Nichols model, we estimate individual detection probability for homogeneous members of the population, whereas in an occupancy model, we estimate probability of observing at least one member of the population given that the site is occupied. Furthermore, the Royle-Nichols model allows us to relate environmental covariates to the latent intensity associated with a species at a site, while in an occupancy model, environmental covariates are associated with the species latent probability of occupancy at a site. Species intensity and occupancy may be governed by different mechanisms, and inference from an intensity model can be distinct from that provided by an occupancy model19,20,21. Cingolani et al.20 proposed that, in plant communities, certain environmental filters preclude species from occupying a site and an additional set of filters may regulate if a species can flourish. Hence, certain covariates that were unimportant in an occupancy model may improve predictive power in an intensity model.Community confoundingSpecies distributions are shaped by environment as well as competition and mutualism within the community8,22,23. Community confounding occurs when species distributions are explained by a convolution of environmental and interspecies effects and can lead to inferential differences between a joint and single species distribution model as well as create difficulties for fitting JSDMs. Former studies have incorporated interspecies dependence into an occupancy model4,11,12,13,14,15,16, and others have addressed spatial confounding1,17,24,25, but none of these explicitly addressed community confounding. However, all Bayesian joint occupancy models naturally attenuate the effects of community confounding due to the prior on the regression coefficients. The prior, assuming it is proper, induces regularization on the regression coefficients26 that can lessen the inferential and computational impacts of confounding27. Furthermore, latent factor models like that described by Tobler et al.4 restrict the dimensionality of the random species effect which should also reduce confounding with the environmental effects.We address community confounding by formulating a version of our model that orthogonalizes the environmental effects and random species effects. Orthogonalizing the fixed and random effects is common practice in spatial statistics and often referred to as restricted spatial regression27,28,29,30,31. Restricted regression has been applied to spatial generalized linear mixed models (SGLMM) for observations (varvec{y},) which can be expressed as$$begin{aligned} varvec{y}&sim [varvec{y}|varvec{mu }, varvec{psi }], end{aligned}$$
    (1)
    $$begin{aligned} g(varvec{mu })&= varvec{X}varvec{beta } + varvec{eta }, end{aligned}$$
    (2)
    $$begin{aligned} varvec{eta }&sim mathcal {N}(varvec{0}, varvec{Sigma }), end{aligned}$$
    (3)
    where (g(cdot )) is a link function, (varvec{psi }) are additional parameters for the data model, and (varvec{Sigma }) is the covariance matrix of the spatial random effect. In the SGLMM, prior information facilitates the estimation of (varvec{eta },) which would not be estimable otherwise due to its shared column space with (varvec{beta })30. This is analogous to applying a ridge penalty to (varvec{eta },) which stabilizes the likelihood. Another method for fitting the confounded SGLMM is to specify a restricted version:$$begin{aligned} varvec{y}&sim [varvec{y}|varvec{mu }, varvec{psi }], end{aligned}$$
    (4)
    $$begin{aligned} g(varvec{mu })&= varvec{X}varvec{delta } + (varvec{I}-varvec{P}_{varvec{X}})varvec{eta }, end{aligned}$$
    (5)
    $$begin{aligned} varvec{eta }&sim mathcal {N}(varvec{0}, varvec{Sigma }), end{aligned}$$
    (6)
    where (varvec{P}_{varvec{X}}=varvec{X}(varvec{X}varvec{X})^{-1}varvec{X}’) is the projection matrix onto the column space of (varvec{X}.) In the unrestricted SGLMM, the regression coefficients (varvec{beta }) and random effect (varvec{eta }) in (1) compete to explain variability in the latent mean (varvec{mu }) in the direction of (varvec{X})27. In the restricted model, however, all variability in the direction of (varvec{X}) is explained solely by the regression coefficients (varvec{delta }) in (4)31, and (varvec{eta }) explains residual variation that is orthogonal to (varvec{X}). We refer to (varvec{beta }) as the conditional effects because they depend on (varvec{eta }), and (varvec{delta }) as the unconditional effects.Restricted regression, as specified in (4), was proposed by Reich et al.28. Reich et al.28 described a disease-mapping example in which the inclusion of a spatial random effect rendered one covariate effect unimportant that was important in the non-spatial model. Spatial maps indicated an association between the covariate and response, making inference from the spatial model appear untenable. Reich et al.28 proposed restricted spatial regression as a method for recovering the posterior expectations of the non-spatial model and shrinking the posterior variances which tend to be inflated for the unrestricted SGLMM.Several modifications of restricted spatial regression have been proposed30,32,33,34,35. All restricted spatial regression methods seek to provide posterior means (text {E}left( delta _j|varvec{y}right)) and marginal posterior variances (text {Var}left( delta _j|varvec{y}right)), (j=1,…,p) that satisfy the following two conditions36:

    1.

    (text {E}left( varvec{delta }|varvec{y}right) = text {E}left( varvec{beta }_{text {NS}}|varvec{y}right)) and,

    2.

    (text {Var}left( beta _{text {NS,}j}|varvec{y}right) le text {Var}left( delta _{j}|varvec{y}right) le text {Var}left( beta _{text {Spatial,}j}|varvec{y}right)) for (j=1,…,p),

    where (varvec{beta }_{NS}) and (varvec{beta }_{Spatial}) are the regression coefficients corresponding to the non-spatial and unrestricted spatial models, respectively.The inferential impacts of spatial confounding on the regression coefficients has been debated. Hodges and Reich29 outlined five viewpoints on spatial confounding and restricted regression in the literature and refuted the two following views:

    1.

    Adding the random effect (varvec{eta }) corrects for bias in (varvec{beta }) resulting from missing covariates.

    2.

    Estimates of (varvec{beta }) in a SGLMM are shrunk by the random effect and hence conservative.

    The random effect (varvec{eta }) can increase or decrease the magnitude of (varvec{beta }), and the change may be galvanized by mechanisms not related to missing covariates. Therefore, we cannot assume the regression coefficients in the SGLMM will exceed those of the restricted model, nor should we regard the estimates in either model as biased due to misspecification. Confounding in the SGLMM causes (text {Var}left( beta _j|varvec{y}right) ge text {Var}left( delta _j|varvec{y}right)), (j=1,…,p), because of the shared column space of the fixed and random effects. Thus, we refer to the conditional coefficients as conservative with regard to their credible intervals, not their posterior expectations.Reich et al.28 argued that restricted spatial regression should always be applied because the spatial random effect is generally added to improve predictions and/or correct the fixed effect variance estimate. While it may be inappropriate to orthogonalize a set of fixed effects in an ordinary linear model, orthogonalizing the fixed and random effect in a spatial model is permissible because the random effect is generally not of inferential interest. Paciorek37 provided the alternative perspective that, if confounding exists, it is inappropriate to attribute all contested variability in (varvec{y}) to the fixed effects. Hanks et al.31 discussed factors for deciding between the unrestricted and restricted SGLMM on a continuous spatial support. The restricted SGLMM leads to improved computational stability, but the unconditional effects are less conservative under model misspecification and more prone to type-S errors: The Bayesian analogue of Type I error. Fitting the unrestricted SGLMM when the fixed and random effects are truly orthogonal does not introduce bias, but it will increase the fixed effect variance. Given these considerations, Hanks et al.31 suggested a hybrid approach where the conditional effects, (varvec{beta }), are extracted from the restricted SGLMM. This is possible because the restricted SGLMM is a reparameterization of the unrestricted SGLMM. This hybrid approach leads to improved computational stability but yields the more conservative parameter estimates. We describe how to implement this hybrid approach for joint species distribution models in the “Community confounding” section.Restricted regression has also been applied in time series applications. Dominici et al.38 debiased estimates of fixed effects confounded by time using restricted smoothing splines. Without the temporal random effect, Dominici et al.38 asserted all temporal variation in the response would be wrongly attributed to temporally correlated fixed effects. Houseman et al.39 used restricted regression to ensure identifiability of a nonparametric temporal effect and highlighted certain covariate effects that were more evident in the restricted model (i.e., the unconditional effects’ magnitude was greater). Furthermore, restricted regression is implicit in restricted maximum likelihood estimation (REML). REML is often employed for debiasing the estimate of the variance of (varvec{y}) in linear regression and fitting linear mixed models that are not estimable in their unrestricted format40. Because REML is generally applied in the context of variance and covariance estimation, considerations regarding the effects of REML on inference for the fixed effects are lacking in the literature.In ecological science, JSDMs often include an unstructured random effect like (varvec{eta }) in (1) to account for interspecies dependence, and hence can also experience community confounding between (varvec{X}) and (varvec{eta }) analogous to spatial confounding. Unlike a spatial or temporal random effect, we consider random species effects to be inferentially important, rather than a tool solely for improving predictions or catch-all for missing covariates. An orthogonalization approach in a JSDM attributes contested variation between the fixed effects (environmental information) and random effect (community information) to the fixed effect.We describe how to orthogonalize the fixed and random species effects in a suite of JSDMs and present a method for detecting community confounding. In the simulation study, we test the efficacy of our method for detecting confounding, show that community confounding can lead to computational difficulties similar to those caused by spatial confounding31, and highlight that, for some models, restricted regression can improve model fitting. We also investigate the inferential implications of community confouding and restricted regression in JSDMs by comparing outputs from the SDM, unrestricted JSDM, and restricted JSDM of the Royle-Nichols and probit occupancy models fit to mammalian camera trap data. Lastly, we discuss other inferential and computational methods for confounded models and consider their appropriateness for joint species distribution modeling. More

  • in

    Air exposure moderates ocean acidification effects during embryonic development of intertidally spawning fish

    Abiotic parametersThe temperature regime experienced by the embryos was purposefully natural and therefore varied between the three air exposure treatments. The subtidal treatment, where embryos were continuously submerged in water, remained around 9.5 °C for the duration of the experiment, while the intertidal treatments experienced dips in temperature during outside air exposure down to 2.5 °C and 0.8 °C, for the low and high intertidal respectively (Fig. 3A). Accumulated thermal units (ATU; days × temperature post collection until hatch) for each air exposure treatment were 79.6, 75.4 and 65.3 for subtidal, low intertidal and high intertidal, respectively. Despite differences in thermal regime, peak hatch was on the same day (March 14, 2021) for all air exposure and CO2 treatments, estimated at 11 dpf.Figure 3Temperature and pH experienced by the herring embryos and larvae throughout the experiment. Hourly measurements of (A) air/water temperature experienced by herring embryos in each of the tidal treatments (subtidal: continuous immersion in 9.5 °C water; low intertidal: 2 × 2 h air exposure; high intertidal: 5 + 9 h air exposure) and (B) pH levels in the tanks for each of the CO2 treatments (greens = control, 400 µatm CO2, yellows = medium, 1600 µatm CO2, reds = high, 3000 µatm CO2); dots are pH levels measured in the individual jars during larval incubation.Full size imageThe pH levels in the tanks were measured hourly and were stable over the course of the embryonic incubation period, with no overlap between treatments, although there was some overlap between individual jars. Control treatment was consistently around a pH of 8, the medium treatment had a pH of 7.4 and the high CO2 treatment had a mean pH of 7.1 (Table 1). After hatch, when the larvae were transferred to the jars, circulation and gas exchange between jars and tank were not as high and CO2 accumulated in the jars over time, leading to pH levels deviating from tank pH levels (Fig. 3B). Although oxygen levels remained high (7–9 mg/L), the pH dropped from a mean 8–7.6 in the control on two occasions, and was brought back up with a partial water exchange from the incubation tank water. The pH in the medium and high CO2 treatments were not as affected (Fig. 3B), however, final water chemistry measurements after completion of the experiment (2 days post water exchange) revealed much higher CO2 levels in all treatments (Table 1: day 15).Table 1 Mean water parameters for each treatment (mean of 3 tanks ± S.D.) at the beginning (day 1, 2021-03-06) and end (day 6, 2021-03-12) of embryonic incubation and mean parameters in the jars (N = 9) at the end of larval incubation (day 15, 2021-03-19); Temperature (T), salinity, pCO2, total CO2 (TCO2) measured at distinct sampling intervals with the BoL; total alkalinity (TA) and pH (on the total scale) calculated with CO2SYS.Full size tableEffect of air exposure and CO2 treatment during embryonic developmentNeither embryonic survival nor growth were significantly affected by treatment in our experiment. Percent daily embryonic mortality was low and not significantly affected by CO2 treatment or air exposure (CO2: p = 0.088, F2 = 2.45; Tide: p = 0.11, F2 = 2.19; CO2*Tide: p = 0.18, F2 = 1.59) . Egg diameter at 6 dpf was also not significantly affected by treatment (CO2: p = 0.38, X2 (2, N = 30) = 1.92; Tide: p = 0.83, X2 (2, N = 30) = 0.33; CO2*Tide: p = 0.08, X2 (2, N = 30) = 8.25). Metabolic rate, as indicated by embryonic heart rate, was significantly affected by air exposure at 6 dpf (p  *; 0.1  >).Full size image More

  • in

    α-cyanobacteria possessing form IA RuBisCO globally dominate aquatic habitats

    Martin WF, Bryant DA, Beatty JT. A physiological perspective on the origin and evolution of photosynthesis. FEMS Microbiol Rev. 2018;42:205–31.CAS 
    PubMed 
    Article 

    Google Scholar 
    Partensky F, Blanchot J, Vaulot D. Differential distribution and ecology of Prochlorococcus and Synechococcus in oceanic waters: a review. Bull Oceanogr Monaco, no Spec. 1999;19:457–76.
    Google Scholar 
    Zwirglmaier K, Jardillier L, Ostrowski M, Mazard S, Garczarek L, Vaulot D, et al. Global phylogeography of marine Synechococcus and Prochlorococcus reveals a distinct partitioning of lineages among oceanic biomes. Environ Microbiol. 2008;10:147–61.PubMed 

    Google Scholar 
    Callieri C. Picophytoplankton in freshwater ecosystems: the importance of small-sized phototrophs. Freshw Rev. 2008;1:1–28.Article 

    Google Scholar 
    Stal LJ. Physiological ecology of cyanobacteria in microbial mats and other communities. New Phytol. 1995;131:1–32.CAS 
    PubMed 
    Article 

    Google Scholar 
    Rikkinen J. Cyanobacteria in terrestrial symbiotic systems. In: Hallenbeck PC editor. Modern topics in the phototrophic prokaryotes. Switzerland: Springer; 2017. p. 243–94.Badger MR, Price GD, Long BM, Woodger FJ. The environmental plasticity and ecological genomics of the cyanobacterial CO2 concentrating mechanism. J Exp Bot. 2006;57:249–65.CAS 
    PubMed 
    Article 

    Google Scholar 
    Rae BD, Long BM, Badger MR, Price GD. Functions, compositions, and evolution of the two types of carboxysomes: polyhedral microcompartments that facilitate CO2 fixation in cyanobacteria and some proteobacteria. Microbiol Mol Biol Rev. 2013;77:357–79.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bar-On YM, Phillips R, Milo R. The biomass distribution on Earth. Proc Natl Acad Sci USA. 2018;115:6506–11.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Buitenhuis ET, Li WKW, Vaulot D, Lomas MW, Landry MR, Partensky F, et al. Picophytoplankton biomass distribution in the global ocean. Earth Syst Sci Data. 2012;4:37–46.Article 

    Google Scholar 
    Flombaum P, Gallegos JL, Gordillo RA, Rincón J, Zabala LL, Jiao N, et al. Present and future global distributions of the marine Cyanobacteria Prochlorococcus and Synechococcus. Proc Natl Acad Sci USA. 2013;110:9824–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Field CB, Behrenfeld MJ, Randerson JT, Falkowski P. Primary production of the biosphere: integrating terrestrial and oceanic components. Science. 1998;281:237–40.CAS 
    PubMed 
    Article 

    Google Scholar 
    Garcia-Pichel F, Belnap J, Neuer S, Schanz F. Estimates of global cyanobacterial biomass and its distribution. Arch Hydrobiol Suppl Algol Stud. 2003;109:213.
    Google Scholar 
    Scanlan DJ, Ostrowski M, Mazard S, Dufresne A, Garczarek L, Hess WR, et al. Ecological genomics of marine picocyanobacteria. Microbiol Mol Biol Rev. 2009;73:249–99.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Doré H, Farrant GK, Guyet U, Haguait J, Humily F, Ratin M, et al. Evolutionary mechanisms of long-term genome diversification associated with niche partitioning in marine picocyanobacteria. Front Microbiol. 2020;11:2129.Article 

    Google Scholar 
    Dufresne A, Ostrowski M, Scanlan DJ, Garczarek L, Mazard S, Palenik BP, et al. Unraveling the genomic mosaic of a ubiquitous genus of marine cyanobacteria. Genome Biol. 2008;9:R90.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Badger MR, Hanson D, Price GD. Evolution and diversity of CO2 concentrating mechanisms in cyanobacteria. Funct Plant Biol. 2002;29:161–73.CAS 
    PubMed 
    Article 

    Google Scholar 
    Whitehead L, Long BM, Price GD, Badger MR. Comparing the in vivo function of α-carboxysomes and β-carboxysomes in two model cyanobacteria. Plant Physiol. 2014;165:398–411.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Castenholz RW, Wilmotte A, Herdman M, Rippka R, Waterbury JB, Iteman I, et al. Phylum BX. cyanobacteria. In: Boone DR, Castenholz RW, Garrity GM editors. Bergey’s manual of systematic bacteriology. New York, NY: Springer; 2001. p. 473–599.Cabello‐Yeves PJ, Picazo A, Camacho A, Callieri C, Rosselli R, Roda-Garcia JJ, et al. Ecological and genomic features of two widespread freshwater picocyanobacteria. Environ Microbiol. 2018;20:3757–71.PubMed 
    Article 
    CAS 

    Google Scholar 
    Di Cesare A, Cabello-Yeves PJ, Chrismas NAM, Sánchez-Baracaldo P, Salcher MM, Callieri C, et al. Genome analysis of the freshwater planktonic Vulcanococcus limneticus sp. nov. reveals horizontal transfer of nitrogenase operon and alternative pathways of nitrogen utilization. BMC Genomics. 2018;19:259.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Sánchez-Baracaldo P, Bianchini G, Di Cesare A, Callieri C, Chrismas NAM. Insights into the evolution of picocyanobacteria and phycoerythrin genes (mpeBA and cpeBA). Front Microbiol. 2019;10:a45.Article 

    Google Scholar 
    Callieri C, Mandolini E, Bertoni R, Lauceri R, Picazo A, Camacho A, et al. Atlas of picocyanobacteria monoclonal strains from the collection of CNR-IRSA, Italy. J Limnol. 2021;80:2002.Article 

    Google Scholar 
    Herdman M, Castenholz RW, Iteman I, Waterbury JB, Rippka R. Subsection I (Formerly Chroococcales Wettstein 1924, emend. Rippka, Deruelles, Waterbury, Herdman and Stanier 1979). In: Boone DR, Castenholz RW, Garrity GM, editors. Bergey’s manual of systematic bacteriology, Vol 1, 2nd ed., The archaea and the deeply branching and phototrophic bacteria. New York: Springer; 2001. p. 493–514.Cabello-Yeves PJ, Haro-Moreno JM, Martin-Cuadrado A, Ghai R, Picazo A, Camacho A, et al. Novel Synechococcus genomes reconstructed from freshwater reservoirs. Front Microbiol. 2017;8:1151.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Badger MR, Price GD. CO2 concentrating mechanisms in cyanobacteria: molecular components, their diversity and evolution. J Exp Bot. 2003;54:609–22.CAS 
    PubMed 
    Article 

    Google Scholar 
    Wheatley NM, Sundberg CD, Gidaniyan SD, Cascio D, Yeates TO. Structure and identification of a pterin dehydratase-like protein as a ribulose-bisphosphate carboxylase/oxygenase (RuBisCO) assembly factor in the α-carboxysome. J Biol Chem. 2014;289:7973–81.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Huang F, Kong WW, Sun Y, Chen T, Dykes GF, Jiang Y, et al. Rubisco accumulation factor 1 (Raf1) plays essential roles in mediating Rubisco assembly and carboxysome biogenesis. Proc Natl Acad Sci USA. 2020;117:17418–28.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kerfeld CA, Melnicki MR. Assembly, function and evolution of cyanobacterial carboxysomes. Curr Opin Plant Biol. 2016;31:66–75.CAS 
    PubMed 
    Article 

    Google Scholar 
    Kupriyanova E, Pronina N, Los D. Carbonic anhydrase—a universal enzyme of the carbon-based life. Photosynthetica. 2017;55:3–19.CAS 
    Article 

    Google Scholar 
    DiMario RJ, Machingura MC, Waldrop GL, Moroney JV. The many types of carbonic anhydrases in photosynthetic organisms. Plant Sci. 2018;268:11–17.Heinhorst S, Cannon GC. A novel evolutionary lineage of carbonic anhydrase (epsilon class) is a component of the carboxysome shell. J Bacteriol. 2004;186:623–30.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Sawaya MR, Cannon GC, Heinhorst S, Tanaka S, Williams EB, Yeates TO, et al. The structure of beta-carbonic anhydrase from the carboxysomal shell reveals a distinct subclass with one active site for the price of two. J Biol Chem. 2006;281:7546–55.CAS 
    PubMed 
    Article 

    Google Scholar 
    Heinhorst S, Williams EB, Cai F, Murin CD, Shively JM, Cannon GC. Characterization of the carboxysomal carbonic anhydrase CsoSCA from Halothiobacillus neapolitanus. J Bacteriol. 2006;188:8087–94.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Omata T, Takahashi Y, Yamaguchi O, Nishimura T. Structure, function and regulation of the cyanobacterial high-affinity bicarbonate transporter, BCT1. Funct Plant Biol. 2002;29:151–9.CAS 
    PubMed 
    Article 

    Google Scholar 
    Omata T, Price GD, Badger MR, Okamura M, Gohta S, Ogawa T. Identification of an ATP-binding cassette transporter involved in bicarbonate uptake in the cyanobacterium Synechococcus sp. strain PCC 7942. Proc Natl Acad Sci USA. 1999;96:13571–6.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Shelden MC, Howitt SM, Price GD. Membrane topology of the cyanobacterial bicarbonate transporter, BicA, a member of the SulP (SLC26A) family. Mol Membr Biol. 2010;27:12–22.PubMed 
    Article 
    CAS 

    Google Scholar 
    Price GD, Howitt SM. The cyanobacterial bicarbonate transporter BicA: its physiological role and the implications of structural similarities with human SLC26 transporters. Biochem Cell Biol. 2011;89:178–88.CAS 
    PubMed 
    Article 

    Google Scholar 
    Price GD, Woodger FJ, Badger MR, Howitt SM, Tucker L. Identification of a SulP-type bicarbonate transporter in marine cyanobacteria. Proc Natl Acad Sci USA. 2004;101:18228–33.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bonfil DJ, Ronen-Tarazia M, Sültemeyer D, Lieman-Hurwitza J, Schatz D, Kaplan A. A putative HCO−3 transporter in the cyanobacterium Synechococcus sp. strain PCC 7942. FEBS Lett. 1998;430:236–40.CAS 
    PubMed 
    Article 

    Google Scholar 
    Price GD, Shelden MC, Howitt SM. Membrane topology of the cyanobacterial bicarbonate transporter, SbtA, and identification of potential regulatory loops. Mol Membr Biol. 2011;28:265–75.CAS 
    PubMed 
    Article 

    Google Scholar 
    Shibata M, Katoh H, Sonoda M, Ohkawa H, Shimoyama M, Fukuzawa H. Genes essential to sodium-dependent bicarbonate transport in cyanobacteria: function and phylogenetic analysis. J Biol Chem. 2002;277:18658–64.CAS 
    PubMed 
    Article 

    Google Scholar 
    Zhang P, Battchikova N, Jansen T, Appel J, Ogawa T, Aro EM. Expression and functional roles of the two distinct NDH-1 complexes and the carbon acquisition complex NdhD3/NdhF3/CupA/Sll1735 in Synechocystis sp PCC 6803. Plant Cell. 2004;16:3326–40.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Price GD, Badger MR, Woodger FJ, Long BM. Advances in understanding the cyanobacterial CO2-concentrating-mechanism (CCM): functional components, Ci transporters, diversity, genetic regulation and prospects for engineering into plants. J Exp Bot. 2008;59:1441–61.CAS 
    PubMed 
    Article 

    Google Scholar 
    Battchikova N, Eisenhut M, Aro E-M. Cyanobacterial NDH-1 complexes: novel insights and remaining puzzles. Biochim Biophys Acta Bioenerg. 2011;1807:935–44.CAS 
    Article 

    Google Scholar 
    Koester RP, Pignon CP, Kesler DC, Willison RS, Kang M, Shen Y. Transgenic insertion of the cyanobacterial membrane protein ictB increases grain yield in Zea mays through increased photosynthesis and carbohydrate production. PLoS ONE. 2021;16:e0246359.Johnson ZI, Zinser ER, Coe A, McNulty NP, Woodward EMS, Chisholm SW. Niche partitioning among Prochlorococcus ecotypes along ocean-scale environmental gradients. Science. 2006;311:1737–40.CAS 
    PubMed 
    Article 

    Google Scholar 
    Callieri C, Coci M, Corno G, Macek M, Modenutti B, Balseiro E. Phylogenetic diversity of nonmarine picocyanobacteria. FEMS Microbiol Ecol. 2013;85:293–301.CAS 
    PubMed 
    Article 

    Google Scholar 
    Schallenberg LA, Pearman JK, Burns CW, Wood SA. Spatial abundance and distribution of picocyanobacterial communities in two contrasting lakes revealed using environmental DNA metabarcoding. FEMS Microbiol Ecol. 2021;97:fiab075.CAS 
    PubMed 
    Article 

    Google Scholar 
    Mózes A, Présing M, Vörös L. Seasonal dynamics of picocyanobacteria and picoeukaryotes in a large shallow lake (Lake Balaton, Hungary). Int Rev Hydrobiol. 2006;91:38–50.Article 
    CAS 

    Google Scholar 
    Vörös L, Callieri C, Balogh KV, Bertoni R. Freshwater picocyanobacteria along a trophic gradient and light quality range. Hydrobiologia. 1998;369/370:117–25.Watanabe MF, Harada K, Carmichael WW, Fujiki H. Toxic microcystis. Boca Raton, FL: CRC Press; 1995.Stockner J, Callieri C, Cronberg G. Picoplankton and other non-bloom-forming cyanobacteria in lakes. In: Whitton BA, Potts M editors. The ecology of cyanobacteria. The Netherlands: Springer; 2000. p. 195–231.Flamholz AI, Prywes N, Moran U, Davidi D, Bar-On YM, Oltrogge LM. Revisiting trade-offs between Rubisco kinetic parameters. Biochemistry. 2019;58:3365–76.CAS 
    PubMed 
    Article 

    Google Scholar 
    Filazzola A, Mahdiyan O, Shuvo A, Ewins C, Moslenko L, Sadid T. A database of chlorophyll and water chemistry in freshwater lakes. Sci Data. 2020;7:1–10.Article 

    Google Scholar 
    Cabello‐Yeves PJ, Zemskaya TI, Zakharenko AS, Sakirko MV, Ivanov VG, Ghai R, et al. Microbiome of the deep Lake Baikal, a unique oxic bathypelagic habitat. Limnol Oceanogr. 2019.Rodrigo MA, Miracle MR, Vicente E. The meromictic Lake La Cruz (Central Spain). Patterns of stratification. Aquat Sci. 2001;63:406–16.Article 

    Google Scholar 
    Camacho A, Picazo A, Miracle MR, Vicente E. Spatial distribution and temporal dynamics of picocyanobacteria in a meromictic karstic lake. Arch Hydrobiol Suppl Algol Stud. 2003;109:171–84.
    Google Scholar 
    Vicente E, Camacho A, Rodrigo MA. Morphometry and physico-chemistry of the crenogenic meromictic Lake El Tobar (Spain). Int Ver für Theor und Angew Limnol Verhandlungen. 1993;25:698–704.CAS 

    Google Scholar 
    Camacho A, Miracle MR, Vicente E. Which factors determine the abundance and distribution of picocyanobacteria in inland waters? A comparison among different types of lakes and ponds. Arch für Hydrobiol. 2003;157:321–38.Article 

    Google Scholar 
    Kaźmierczak J, Kempe S, Kremer B, López-García P, Moreira D, Tavera R. Hydrochemistry and microbialites of the alkaline crater lake Alchichica, Mexico. Facies. 2011;57:543–70.Article 

    Google Scholar 
    Ghai R, Mizuno CM, Picazo A, Camacho A, Rodriguez‐Valera F. Key roles for freshwater Actinobacteria revealed by deep metagenomic sequencing. Mol Ecol. 2014;23:6073–90.CAS 
    PubMed 
    Article 

    Google Scholar 
    Cabello-Yeves PJ, Ghai R, Mehrshad M, Picazo A, Camacho A, Rodriguez-Valera F. Reconstruction of diverse Verrucomicrobial genomes from metagenome datasets of freshwater reservoirs. Front Microbiol. 2017;8:2131.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    de Hoyos C, Negro AI, Aldasoro JJ. Cyanobacteria distribution and abundance in the Spanish water reservoirs during thermal stratification. Limnetica. 2004;23:119–32.Article 

    Google Scholar 
    Raven J, Caldeira K, Eldefield H, Hoegh-Guldberg O, Liss P, Riebesell U, et al. Ocean acidification due to increasing atmospheric carbon dioxide. London: The Royal Society; 2005.Mangan NM, Flamholz A, Hood RD, Milo R, Savage DF. pH determines the energetic efficiency of the cyanobacterial CO2 concentrating mechanism. Proc Natl Acad Sci USA. 2016;113:E5354–62.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Tadesse I, Green FB, Puhakka JA. Seasonal and diurnal variations of temperature, pH and dissolved oxygen in advanced integrated wastewater pond system® treating tannery effluent. Water Res. 2004;38:645–54.CAS 
    PubMed 
    Article 

    Google Scholar 
    Gao Y, Zhang Z, Liu X, Yi N, Zhang L, Song W, et al. Seasonal and diurnal dynamics of physicochemical parameters and gas production in vertical water column of a eutrophic pond. Ecol Eng. 2016;87:313–23.Article 

    Google Scholar 
    Schindler DW. Recent advances in the understanding and management of eutrophication. Limnol Oceanogr. 2006;51:356–63.Article 

    Google Scholar 
    Bosak T, Bush JWM, Flynn MR, Liang B, Ono S, Petroff AP, et al. Formation and stability of oxygen‐rich bubbles that shape photosynthetic mats. Geobiology. 2010;8:45–55.CAS 
    PubMed 
    Article 

    Google Scholar 
    Zeebe RE, Wolf-Gladrow D. CO2 in seawater: equilibrium, kinetics, isotopes. Amsterdam: Elsevier Science B.V.; 2001.Rae BD, Förster B, Badger MR, Price GD. The CO2-concentrating mechanism of Synechococcus WH5701 is composed of native and horizontally-acquired components. Photosynth Res. 2011;109:59–72.CAS 
    PubMed 
    Article 

    Google Scholar 
    Rippka R, Deruelles J, Waterbury JB, Herdman M, Stanier RY. Generic assignments, strain histories and properties of pure cultures of cyanobacteria. Microbiology. 1979;111:1–61.Article 

    Google Scholar 
    Martín-Cuadrado A-B, López-García P, Alba J-C, Moreira D, Monticelli L, Strittmatter A, et al. Metagenomics of the deep Mediterranean, a warm bathypelagic habitat. PLoS ONE. 2007;2:e914.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–20.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19:455–77.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hyatt D, Chen G-L, LoCascio PF, Land ML, Larimer FW, Hauser LJ, et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinforma. 2010;11:1.Article 
    CAS 

    Google Scholar 
    Buchfink B, Xie C, Huson DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods. 2015;12:59–60.CAS 
    PubMed 
    Article 

    Google Scholar 
    Kanehisa M, Goto S. KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 2000;28:27–30.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Overbeek R, Olson R, Pusch GD, Olsen GJ, Davis JJ, Disz T, et al. The SEED and the Rapid Annotation of microbial genomes using Subsystems Technology (RAST). Nucleic Acids Res. 2013;42:D206–14.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Tatusov RL, Natale DA, Garkavtsev IV, Tatusova TA, Shankavaram UT, Rao BS, et al. The COG database: new developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res. 2001;29:22–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Haft DH, Loftus BJ, Richardson DL, Yang F, Eisen JA, Paulsen IT, et al. TIGRFAMs: a protein family resource for the functional identification of proteins. Nucleic Acids Res. 2001;29:41–3.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kang D, Li F, Kirton E, Thomas A, Egan R, An H, et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ Prepr. 2019;7:e27522v1.
    Google Scholar 
    Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 2015;25:1043–55.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Parks DH, Chuvochina M, Waite DW, Rinke C, Skarshewski A, Chaumeil P-A, et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat Biotechnol. 2018;36:996–1004.CAS 
    PubMed 
    Article 

    Google Scholar 
    Asnicar F, Thomas AM, Beghini F, Mengoni C, Manara S, Manghi P, et al. Precise phylogenetic analysis of microbial isolates and genomes from metagenomes using PhyloPhlAn 3.0. Nat Commun. 2020;11:1–10.Article 
    CAS 

    Google Scholar 
    Garczarek L, Guyet U, Doré H, Farrant GK, Hoebeke M, Brillet-Guéguen L, et al. Cyanorak v2.1: a scalable information system dedicated to the visualization and expert curation of marine and brackish picocyanobacteria genomes. Nucleic Acids Res. 2021;49:D667–76.CAS 
    PubMed 
    Article 

    Google Scholar 
    Erwin PM, Thacker RW. Cryptic diversity of the symbiotic cyanobacterium Synechococcus spongiarum among sponge hosts. Mol Ecol. 2008;17:2937–47.CAS 
    PubMed 
    Article 

    Google Scholar 
    Usher KM, Toze S, Fromont J, Ku J, Sutton DC. A new species of cyanobacterial symbiont from the marine sponge Chondrilla nucula. Symbiosis. 2004.Holtman CK, Chen Y, Sandoval P, Gonzales A, Nalty MS, Thomas TL, et al. High-throughput functional analysis of the Synechococcus elongatus PCC 7942 genome. DNA Res. 2005;12:103–15.CAS 
    PubMed 
    Article 

    Google Scholar 
    Chen M-Y, Teng W-K, Zhao L, Hu C-X, Zhou Y-K, Han B-P, et al. Comparative genomics reveals insights into cyanobacterial evolution and habitat adaptation. ISME J. 2021;15:211–27.PubMed 
    Article 

    Google Scholar 
    Katoh K, Misawa K, Kuma K, Miyata T. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002;30:3059–66.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Price MN, Dehal PS, Arkin AP. FastTree 2–approximately maximum-likelihood trees for large alignments. PLoS ONE. 2010;5:e9490.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Rozen S, Skaletsky H. Primer3 on the WWW for general users and for biologist programmers. Methods Mol. Biol. 2000;132:365–86.CAS 
    PubMed 

    Google Scholar  More

  • in

    Effectiveness of management zones for recovering parrotfish species within the largest coastal marine protected area in Brazil

    Hughes, T. P. et al. Climate change, human impacts, and the resilience of coral reefs. Science 301(5635), 929–933 (2003).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    Hoegh-Guldberg, O. E. et al. Coral reefs under rapid climate change and ocean acidification. Science 318(5857), 1737–1742 (2007).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    Soares, M. et al. The flourishing and vulnerabilities of zoantharians on Southwestern Atlantic reefs. Mar. Environ. Res. 173(3), 105535 (2021).Ban, N. C. et al. Designing, implementing and managing marine protected areas: Emerging trends and opportunities for coral reef nations. J. Exp. Mar. Biol. Ecol. 408(1–2), 21–31 (2011).Article 

    Google Scholar 
    Magris, R. A., Pressey, R. L., Mills, M., Vila-Nova, D. A. & Floeter, S. Integrated conservation planning for coral reefs: Designing conservation zones for multiple conservation objectives in spatial prioritisation. Glob. Ecol. Conserv. 11, 53–68 (2017).Article 

    Google Scholar 
    Vercammen, A. et al. Evaluating the impact of accounting for coral cover in large-scale marine conservation prioritizations. Divers. Distrib. 25(10), 1564–1574 (2019).Article 

    Google Scholar 
    Giakoumi, S., Grantham, H. S., Kokkoris, G. D. & Possingham, H. P. Designing a network of marine reserves in the Mediterranean Sea with limited socio-economic data. Biol. Conserv. 144(2), 753–763 (2011).Article 

    Google Scholar 
    Gill, D. A. et al. Capacity shortfalls hinder the performance of marine protected areas globally. Nature 543(7647), 665–669 (2017).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    Magris, R. A. et al. A blueprint for securing Brazil’s marine biodiversity and supporting the achievement of global conservation goals. Divers. Distrib. 27(2), 198–215 (2021).Article 

    Google Scholar 
    Day, J. C. Zoning—lessons from the Great Barrier Reef marine park. Ocean Coast. Manag. 45(2–3), 139–156 (2002).Article 

    Google Scholar 
    Agardy, T. Ocean Zoning: Making Marine Management More Effective (Earthscan, 2010).Makino, A., Klein, C. J., Beger, M., Jupiter, S. D. & Possingham, H. P. Incorporating conservation zone effectiveness for protecting biodiversity in marine planning. PLoS ONE 8(11), e78986 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    Villa, F., Tunesi, L. & Agardy, T. Zoning marine protected areas through spatial multiple-criteria analysis: The case of the Asinara Island National Marine Reserve of Italy. Conserv. Biol. 16(2), 515–526 (2002).Article 

    Google Scholar 
    Muhl, E. K., Esteves Dias, A. C. & Armitage, D. Experiences with governance in three marine conservation zoning initiatives: Parameters for assessment and pathways forward. Front. Mar. Sci. 7, 629 (2020).Article 

    Google Scholar 
    Beger, M. et al. Integrating regional conservation priorities for multiple objectives into national policy. Nat. Commun. 6(1), 1–8 (2015).Article 
    CAS 

    Google Scholar 
    Ban, N. C. et al. A social–ecological approach to conservation planning: Embedding social considerations. Front. Ecol. Environ. 11(4), 194–202 (2013).Article 

    Google Scholar 
    Teh, L. C., Teh, L. S. & Jumin, R. Combining human preference and biodiversity priorities for marine protected area site selection in Sabah, Malaysia. Biol. Conserv. 167, 396–404 (2013).Article 

    Google Scholar 
    Sarker, S., Rahman, M. M., Yadav, A. K. & Islam, M. M. Zoning of marine protected areas for biodiversity conservation in Bangladesh through socio-spatial data. Ocean Coast. Manag. 173, 114–122 (2019).Article 

    Google Scholar 
    Day, J. C., Kenchington, R. A., Tanzer, J. M. & Cameron, D. S. Marine zoning revisited: How decades of zoning the Great Barrier Reef has evolved as an effective spatial planning approach for marine ecosystem-based management. Aquat. Conserv. Mar. Freshw. Ecosyst. 29, 9–32 (2019).Article 

    Google Scholar 
    Claudet, J. et al. Assessing the effects of marine protected area (MPA) on a reef fish assemblage in a northwestern Mediterranean marine reserve: Identifying community-based indicators. Biol. Conserv. 130(3), 349–369 (2006).Article 

    Google Scholar 
    Emslie, M. J. et al. Expectations and outcomes of reserve network performance following re-zoning of the Great Barrier Reef Marine Park. Curr. Biol. 25(8), 983–992 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    McClure, E. C. et al. Higher fish biomass inside than outside marine protected areas despite typhoon impacts in a complex reefscape. Biol. Cons. 241, 108354 (2020).Article 

    Google Scholar 
    Bender, M. G. et al. Local ecological knowledge and scientific data reveal overexploitation by multigear artisanal fisheries in the Southwestern Atlantic. PLoS ONE 9(10), e110332 (2014).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    Hamilton, R. J. et al. Hyperstability masks declines in bumphead parrotfish (Bolbometopon muricatum) populations. Coral Reefs 35(3), 751–763 (2016).Article 
    ADS 

    Google Scholar 
    Pereira, P. H. C., Ternes, M. L. F., Nunes, J. A. C. & Giglio, V. J. Overexploitation and behavioral changes of the largest South Atlantic parrotfish (Scarus trispinosus): Evidence from fishers’ knowledge. Biol. Conserv. 254, 108940 (2021).Article 

    Google Scholar 
    Mumby, P. J. et al. Fishing, trophic cascades, and the process of grazing on coral reefs. Science 311(5757), 98–101 (2006).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    Mumby, P. J. & Harborne, A. R. Marine reserves enhance the recovery of corals on Caribbean reefs. PLoS ONE 5(1), e8657 (2010).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    Topor, Z. M., Rasher, D. B., Duffy, J. E. & Brandl, S. J. Marine protected areas enhance coral reef functioning by promoting fish biodiversity. Conserv. Lett. 12(4), e12638 (2019).Article 

    Google Scholar 
    Liu, C., White, M. & Newell, G. Measuring and comparing the accuracy of species distribution models with presence–absence data. Ecography 34(2), 232–243 (2011).CAS 
    Article 

    Google Scholar 
    Miranda, R. J. et al. Integrating long term ecological research (LTER) and marine protected area management: Challenges and solutions. Oecol. Aust. 24(2), 279–300 (2020).Article 

    Google Scholar 
    ICMBIO. Plano de Manejo da Área de Proteção Ambiental Costa dos Corais. ICMBio/MMA (2021).Jones, K. R. et al. Area requirements to safeguard Earth’s marine species. One Earth 2(2), 188–196 (2020).Article 
    ADS 

    Google Scholar 
    Figueiredo, M. S. & Grelle, C. E. V. Predicting global abundance of a threatened species from its occurrence: Implications for conservation planning. Divers. Distrib. 15(1), 117–121 (2009).Article 

    Google Scholar 
    Pearce, J. & Ferrier, S. The practical value of modelling relative abundance of species for regional conservation planning: A case study. Biol. Conserv. 98(1), 33–43 (2001).Article 

    Google Scholar 
    Ferreira, H. M., Magris, R. A., Floeter, S. R. & Ferreira, C. E. Drivers of ecological effectiveness of marine protected areas: A meta-analytic approach from the Southwestern Atlantic Ocean (Brazil). J. Environ. Manag. 301, 113889 (2021).Article 

    Google Scholar 
    Mills, M. et al. Real-world progress in overcoming the challenges of adaptive spatial planning in marine protected areas. Biol. Conserv. 181, 54–63 (2015).Article 

    Google Scholar 
    Bennett, N. J. et al. Local support for conservation is associated with perceptions of good governance, social impacts, and ecological effectiveness. Conserv. Lett. 12(4), e12640 (2019).Article 

    Google Scholar 
    Oldekop, J. A., Holmes, G., Harris, W. E. & Evans, K. L. A global assessment of the social and conservation outcomes of protected areas. Conserv. Biol. 30(1), 133–141 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Emslie, M. J. et al. Decades of monitoring have informed the stewardship and ecological understanding of Australia’s Great Barrier Reef. Biol. Conserv. 252, 108854 (2020).Article 

    Google Scholar 
    Gerhardinger, L. C., Godoy, E. A., Jones, P. J., Sales, G. & Ferreira, B. P. Marine protected dramas: The flaws of the Brazilian national system of marine protected areas. Environ. Manag. 47(4), 630–643 (2011).Article 
    ADS 

    Google Scholar 
    Oliveira, E. A., Martelli, H., Silva, A. C. S. E., Martelli, D. R. B. & Oliveira, M. C. L. Science funding crisis in Brazil and COVID-19: Deleterious impact on scientific output. Anais Acad. Bras. Ciênc. 92, 1–2 (2020).
    Floeter, S. R., Halpern, B. S. & Ferreira, C. E. L. Effects of fishing and protection on Brazilian reef fishes. Biol. Conserv. 128(3), 391–402 (2006).Article 

    Google Scholar 
    Bender, M. G., Floeter, S. R. & Hanazaki, N. Do traditional fishers recognise reef fish species declines? Shifting environmental baselines in E astern B razil. Fish. Manag. Ecol. 20(1), 58–67 (2013).Article 

    Google Scholar 
    Hoey, A. S. & Bonaldo, R. M. (eds) Biology of Parrotfishes (CRC Press, Boca Raton, 2018).
    Google Scholar 
    Frédou, T. & Ferreira, B. P. Bathymetric trends of Northeastern Brazilian snappers (Pisces, Lutjanidae): Implications for the reef fishery dynamic. Braz. Arch. Biol. Technol. 48(5), 787–800 (2005).Article 

    Google Scholar 
    Guerra, A. S. Wolves of the Sea: Managing human-wildlife conflict in an increasingly tense ocean. Mar. Policy 99, 369–373 (2019).Article 

    Google Scholar 
    Hawkins, J. P. & Roberts, C. M. Effects of fishing on sex-changing Caribbean parrotfishes. Biol. Cons. 115(2), 213–226 (2004).Article 

    Google Scholar 
    Tuya, F. et al. Effect of fishing pressure on the spatio-temporal variability of the parrotfish, Sparisoma cretense (Pisces: Scaridae), across the Canarian Archipelago (eastern Atlantic). Fish. Res. 7(1), 24–33 (2006).Article 

    Google Scholar 
    Steneck, R. S., Arnold, S. N. & Mumby, P. J. Experiment mimics fishing on parrotfish: Insights on coral reef recovery and alternative attractors. Mar. Ecol. Prog. Ser. 506, 115–127 (2014).Article 
    ADS 

    Google Scholar 
    Taylor, B. M., Trip, E. D., & Choat, J. H. Dynamic demography: Investigations of life-history variation in the parrotfishes. In Biology of Parrotfishes 69–98 (CRC Press, 2018).Moura, R. L. & Francini-Filho, R. B. Reef and Shore Fishes of the Abrolhos Region, Brazil Vol. 38, 40–55 (RAP Bulletin of Biological Assessment, Washington, 2005).
    Google Scholar 
    Francini-Filho, R. B., Moura, R. L., Ferreira, C. M. & Coni, E. O. Live coral predation by parrotfishes (Perciformes: Scaridae) in the Abrolhos Bank, eastern Brazil, with comments on the classification of species into functional groups. Neotrop. Ichthyol. 6, 191–200 (2008).Article 

    Google Scholar 
    Freitas, M. O. et al. Age, growth, reproduction and management of Southwestern Atlantic’s largest and endangered herbivorous reef fish, Scarus trispinosus Valenciennes, 1840. PeerJ 7, e7459 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pinheiro, H. T. et al. An inverted management strategy for the fishery of endangered marine species. Front. Mar. Sci. 8, 172 (2021).Article 

    Google Scholar 
    Correia, M. D. Scleractinian corals (Cnidaria: Anthozoa) from reef ecosystems on the Alagoas coast, Brazil. J. Mar. Biol. Assoc. U. K. 91, 659–668 (2011).CAS 
    Article 

    Google Scholar 
    Santos, D. K. F., Rufino, R. D., Luna, J. M., Santos, V. A. & Sarubbo, L. A. Biosurfactants: Multifunctional biomolecules of the 21st century. Int. J. Mol. Sci. 17(3), 401 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    de Oliveira, S. et al. Oil spill in South Atlantic (Brazil): Environmental and governmental disaster. Mar. Policy 115, 103879 (2020).Article 

    Google Scholar 
    Teixeira, L. M. P. & Creed, J. C. A decade on: An updated assessment of the status of marine non-indigenous species in Brazil. Aquat. Invasions 15(1), 30–43 (2020).Article 

    Google Scholar 
    Braga, M. D. A. et al. Retirement risks: Invasive coral on old oil platform on the Brazilian equatorial continental shelf. Mar. Pollut. Bull. 165, 112156 (2021).CAS 
    PubMed 
    Article 

    Google Scholar 
    Luiz, O. J. et al. Multiple lionfish (Pterois spp.) new occurrences along the Brazilian coast confirm the invasion pathway into the Southwestern Atlantic. Biol. Invasions 23, 3013–3019 (2021).Article 

    Google Scholar 
    Maida, M., & Ferreira, B. P. Coral reefs of Brazil: An overview. In Proceedings of the 8th International Coral Reef Symposium, Vol. 1, 263–274 (Smithsonian Tropical Research Institute Panamá, 1997).Pereira, P. H. C., Macedo, C. H., Nunes, J. D. A. C., Marangoni, L. F. D. B. & Bianchini, A. Effects of depth on reef fish communities: Insights of a “deep refuge hypothesis” from Southwestern Atlantic reefs. PLoS ONE 13(9), e0203072 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    ICMBIO. Plano de Manejo da Área de Proteção Ambiental Costa dos Corais (ICMBio/MMA, 2013).Hill, J. & Wilkinson, C. E. Methods for Ecological Monitoring of Coral Reefs Vol. 117 (Australian Institute of Marine Science, Townsville, 2004).
    Google Scholar 
    Dalapicolla, J. Tutorial de modelos de distribuição de espécies: guia prático usando o MaxEnt e o ArcGIS 10. Laboratório de Mastozoologia e Biogeografia. Universidade Federal do Espírito Santo, Vitória. Retrieved, 6 (2016).Phillips, S. J., Dudík, M., & Schapire, R. E. A maximum entropy approach to species distribution modeling. In Proceedings of the Twenty-First International Conference on Machine learning, Vol. 83 (2004).Phillips, S. J., Anderson, R. P. & Schapire, R. E. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190(3–4), 231–259 (2006).Article 

    Google Scholar 
    Anderson, R. P. & Martınez-Meyer, E. Modeling species’ geographic distributions for preliminary conservation assessments: An implementation with the spiny pocket mice (Heteromys) of Ecuador. Biol. Conserv. 116(2), 167–179 (2004).Article 

    Google Scholar 
    Phillips, S. J., Anderson, R. P., Dudík, M., Schapire, R. E. & Blair, M. E. Opening the black box: An open-source release of Maxent. Ecography 40(7), 887–893 (2017).Article 

    Google Scholar 
    Rodrigues, E. D. C., Rodrigues, F. A., Rocha, R. L. A. & Corrêa, P. L. P. An adaptive maximum entropy approach for modeling of species distribution. Mem. WTA 108–117 (2010).Rodrigues, E. S. D. C., Rodrigues, F. A., Ricardo, L. D. A., Corrêa, P. L. & Giannini, T. C. Evaluation of different aspects of maximum entropy for niche-based modeling. Procedia Environ. Sci. 2, 990–1001 (2010).Article 

    Google Scholar 
    Hattab, T. et al. The use of a predictive habitat model and a fuzzy logic approach for marine management and planning. PLoS ONE 8(10), e76430 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    Galante, P. J. et al. The challenge of modeling niches and distributions for data-poor species: A comprehensive approach to model complexity. Ecography 41(5), 726–736 (2018).Article 

    Google Scholar 
    Silber, G. K. et al. Projecting marine mammal distribution in a changing climate. Front. Mar. Sci. 4, 413 (2017).Article 

    Google Scholar 
    Perkins-Taylor, I. E. & Frey, J. K. Predicting the distribution of a rare chipmunk (Neotamias quadrivittatus oscuraensis): Comparing MaxEnt and occupancy models. J. Mammal. 101(4), 1035–1048 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lee, C. M., Lee, D. S., Kwon, T. S., Athar, M. & Park, Y. S. Predicting the global distribution of Solenopsis geminata (Hymenoptera: Formicidae) under climate change using the MaxEnt model. Insects 12(3), 229 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Possingham, H., Ball, I. & Andelman, S. Mathematical methods for identifying representative reserve networks. In Quantitative methods for conservation biology 291–306 (Springer, New York, 2000).Terrell, G. R. & Scott, D. W. Variable kernel density estimation.  Ann. Stat. 20(3), 1236–1265 (1992).
    O’Brien, S. H., Webb, A., Brewer, M. J. & Reid, J. B. Use of kernel density estimation and maximum curvature to set Marine Protected Area boundaries: Identifying a Special Protection Area for wintering red-throated divers in the UK. Biol. Conserv. 156, 15–21 (2012).Article 

    Google Scholar 
    Fielding, A. H. & Bell, J. F. A review of methods for the assessment of prediction errors in conservation presence/absence models. Environ. Conserv. 24, 38–49 (1997).Article 

    Google Scholar  More

  • in

    Effects of solar irradiance noise on a complex marine trophic web

    This section is devoted to show results and to highlight eventual effects of the interplay between the nonlinearity characterizing the system dynamics and the presence of noisy fluctuations for the irradiance variable.Analysis of experimental dataThe need of taking into account noisy fluctuations of such an environmental variable is well demonstrated in Fig. 1. In the first panel (a) the experimental time behaviour of the irradiance is shown. This noisy curve is based on the experimental data (purple points) of the Boussole buoy located in the Gulf of Lion, collected over a period of nine years, precisely from 2004 to 2013. The time series of the experimental data presents quite a few gaps in time due to the malfunction of the buoy. This aspect has been remedied by merging the experimental data with those of the OASIM model validated for the Boussole site61 (yellow points). The latter is a multispectral atmospheric radiative transfer model that is in turn forced by experimental-model data based on ECMWF ERAINTERIM reanalyses which provide, for example, cloud cover data. The radiative model is partly stochastic since it considers the effects stemming from the presence of clouds, averaged along a single day (this explains why the yellow points are slightly less scattered). We see that the OASIM model accurately reproduces the profile which emerges from the experimental data. Further, we stress that the experimental data are only used in this initial analysis. In the biogeochemical simulations the irradiance signal is fully reconstructed starting from a realistic seasonal cycle combined with a range of different random fluctuations, and the information from OASIM is not used. In the second panel (b) the daily (black points) as well as the three-month (red points) running mean of the experimental series are plotted. Figure 1c shows the irradiance noisy fluctuations (INF) which have been obtained by subtracting the three-month running mean curve (3MRM, red curve in Fig. 1b) from the daily running mean one (DRM, black curve in Fig. 1b) and normalizing with respect to the mean of the 3MRM ((overline{3MRM})), namely (INF = (DRM – 3MRM) / overline{3MRM}). We see that a seasonal overall trend with higher oscillations during the winter time can be seen, implying that the characteristics of the noise may change over the year. Moreover, a slight imbalance between positive and negative values of the noisy fluctuations (that is, different values of the maximum fluctuation intensity) is present. The physical reason for the occurrence of such an aspect can be ascribed to the fact that the maximum value of solar irradiance corresponds to that measured during a sunny day. Conversely, the minimum level tends to zero corresponding to a dense darkness. While the former is close to the mean value of the solar irradiance (most of all in summer), the latter is much further away and then a natural asymmetry arises in the random fluctuations. However, it should be noted that, apart from the intense spikes, the asymmetry is not so pronounced, as proved by the mean value (red line in Fig. 1c) which is practically zero, namely (0.4%) of the (overline{3MRM}). Therefore, basing on this last observation, to model the noise affecting the irradiance dynamics, as a first approximation we consider a symmetric Gaussian autocorrelated noise as described in the next subsection.On the basis of such experimental results, we postulate the hypothesis that random fluctuations of light cannot be neglected, most of all in the study of ecological systems where light profoundly determines the system dynamics, governing fundamental processes at the basis of of the food web.Figure 1(a) Experimental data (purple points) of the stochastic solar irradiance collected by the Boussole buoy in a time-window of 9 years (2004-2013); the yellow points are the data generated by the OASIM model used to fill the gaps present in the experimental time-series due to malfunctioning of the buoy. (b) Daily (black points) and three-month (red points) running mean of the light curve in panel (a). (c) Irradiance noisy fluctuations (INF), obtained by subtracting the three-month running mean curve (3MRM) from the daily running mean one (DRM) and normalizing with respect to the mean value of 3MRM ((overline{3MRM})), namely (INF = (DRM – 3MRM) / overline{3MRM}); the red line represents the mean value of such fluctuations. Data already presented and validated in61.Full size imageSolar irradianceThe solar irradiance forcing is derived considering a deterministic seasonal oscillation combined with an Ornstein-Uhlenbeck process. The coefficient of variation (CV) of simulated light forcing, Fig. 2, (CV=sigma / mu) ((mu) and (sigma) being mean value and standard deviation calculated over both time and numerical realizations), is shown for 231 (D-tau) pairs. D and (tau) represent the intensity of a Gaussian noise source and the auto-correlation time of the fluctuations, respectively (see Eqs. (2) and (3)).Each pixel represents the mean value on time of CV calculated with respect to 1000 different stochastic realizations. Figure 2Coefficient of variation ((CV=sigma / mu)) of irradiance resulting from numerical integration of model equations for 231 (D-tau) different scenarios.Full size imageIt is easy to see the agreement between the results obtained from the numerical integration and the theoretical ones derivable from Eq. (5) by putting (text {var}{F_L(0)}=0) and (t gg 1), getting (sigma ^2_L=D / 2tau). In Fig. 2, indeed, the maximum values of (sigma) lie in the upper left part of the plot corresponding to small (high) values of (tau) (D). As it is clear the values of D have been chosen in order to obtain a relative standard deviation ranging from (5%mu) to (60%mu). We underline that, in this case, it is possible to interchangeably consider (sigma) and CV since the dependence of CV on D and (tau) does not differ from that of (sigma) (meaning that the dependence of (sigma) is not altered by dividing by (mu)) (results not shown).Effects on population dynamicsIn this section the noise-induced effects on the population dynamics are examined. The nine planktonic populations present a different qualitative behaviour of the CV, compared to that of the irradiance. In this case, the CV is characterized by a strong non-monotonic dependence on the parameter (tau). This aspect can be appreciated in Fig. 3 where different curves of CV versus the time correlation parameter are shown for different fixed values of D.Figure 3Coefficient of variation ((CV=sigma / mu)) of the nine planktonic populations resulting from numerical integration of model equations plotted versus the considered values of (tau); the different curves are related to different values of the noise intensity D.Full size imageThe existence of a maximum value for CV can be appreciated for each species. Although the qualitative behaviour is the same for all strains, particular attention has to be payed on diatoms and nanoflagellates. All the other species, indeed, present a percent variation of standard deviation between (2%) and (15%). In the case of nanoflagellates, instead, the D-dependent range is (20-90%), while diatoms reach values over the (100%) for the highest values of D. Therefore, these two species, in particular, and the whole system, in general, are extremely sensitive to the auto-correlation time which characterizes the noise.We note that the different curves related to the different selected values of D approach the horizontal axis, tending asymptotically to vanish as (tau) increases. Such a behaviour can be explained by the fact that high values of (tau) give rise to a more correlated dynamics, so that (tau rightarrow infty) implies fully correlated time-behaviours corresponding to the deterministic case. In this instance, then, all the different realizations give the same results, making the standard deviation vanish. The same happens, independently of the value of (tau), for low values of noise intensity for which the corresponding curves approach the same almost vanishing value (see orange, gray and yellow lines). Differently from the previous case, when (tau rightarrow 0) the noise tends to a delta-correlated noise, that is a white noise; for (tau ne 0), instead, the noise spectrum is not flat, being characterized by a Cauchy-Lorentz distribution. The strong nonmonotonicity of CV with respect to (tau), emerging when there are relatively high values of CV, implies a greater variability of the system biomass. Lower values of CV indicate that the system dynamics is less influenced by the presence of noise where very little or no differences with respect to the deterministic case are present. Conversely, high values of CV clearly demonstrate the remarkable signature of the presence of an impacting noise source. It is interesting to note that the noise influence on the ecosystem strongly depends on both (tau) and D, that is, just an intense noise is not enough to generate a greater response of the ecosystem. In particular, experimental data are characterized by a CV approximately equal to 0.361, which corresponds to values of D and (tau) lying on the diagonal strip in Fig. 2 ranging from ((tau ,D)=(0.5,10^4)) to ((tau ,D)=(365,10^7)). Finally we note the presence of a noise suppression effect. High values of D, indeed, can generate slight effects when the correlation time (tau) does not take on suitable values.The results shown here are an extension of the previous work by Benincà et al.56. There, the authors analyse a simpler, less realistic model of two interacting populations, whose dynamics is affected by a randomly fluctuating temperature. In that case, moreover, the deterministic oscillations of the temperature are suppressed, and the system exhibits intrinsic Lotka-Volterra oscillations whose frequency match with the characteristic one(s) of the noise. On the contrary, here, the observed maximum response (see Fig. 3) cannot be interpreted as a synchronization effect, since our model does not present intrinsic Lotka-Volterra-like oscillations and the periodic population variability is only due to the deterministic forcing(s).The nonmonotonic behaviour of the CV can be then interpreted as the signature of the intimate interplay between the ecological system and the noise. This interplay, indeed, has a pivotal role in both determining the dynamics of the populations and defining the characteristics of the ecosystem.In Fig. 3 it can be observed that the value of (tau) for which CV is maximum strongly depends on the noise intensity D. In particular, it is possible to note that the peaks in Fig. 3 move towards higher values of (tau) as the noise intensity increases. Thus, Fig. 3 demonstrates that the maximum-response effect to the random fluctuations is sensitive to the noise intensity D.However, it is important to underline that the response of the system to the noisy signal does not depend on the yearly oscillations induced by the deterministic forcings. Indeed, by considering constant the deterministic part of all external forcings (temperature, irradiance, wind and salinity), the non monotonic behaviour of CV with respect to both (tau) and D is still present, provided that the populations are not extinct (plot not shown). In this scenario indeed, besides dinoflagellates, diatoms and nanoflagellates are practically extinct as well, exhibiting thus a constant vanishing variance. All the other strains, instead, present qualitatively the same nonmonotonicity with only slight differences (shift of the peaks and different mean values of the CV curves), probably due to the extinction of diatoms and nanoflagellates which causes relevant differences in the system dynamics. More specifically, the system’s response seems to depend on both the noise intensity and the correlation time (see Fig. 3).In this scenario (absence of seasonal driving) we have studied the dependence on both parameters D and (tau) of the probability density functions (PDFs) of the non-vanishing populations. In Fig. 4, the PDFs of bacteria (B1), picophytoplankton (P3), microzooplankton (Z5) and etherotrophic nanoflagellates (Z6) are plotted for (tau =0.5) and eight different values of the parameter D.Figure 4Dependence of the probability density functions of non-vanishing populations on the parameter D for (tau =0.5). The curves are normalized within the interval taken into account. For this reason the relative peaks of the curves in the bottom panels have different values compared to those of the top panels. However, the figure aims at showing the existence of the value of the noise intensity for which the system is more sensitive as well as the generation of a stationary out-of-equilibrium state induced by the noise.Full size imageWe see that the mean value and the variance of these populations are strongly affected by the presence of random fluctuations in the irradiance. Specifically, as the noise intensity increases the mean values of picophytoplankton and bacteria concentrations exhibit a shift. In particular, the results indicate that picophytoplankton is disavantaged by the presence of a noisy component in the irradiance, which indeed tends to inhibit its ability to absorbe the solar light, slowing down its growth. As a consequence, since phytoplankton and bacteria compete for the same resources, as the former declines the latter are favoured, with a compensation mechanism which allows their predators (zooplankton populations) to be almost not affected by the noisy behaviour of the irradiance. Further, we note that for intermediate values of the noise intensity ((D = 10^4 – 10^5)) a maximum of the variance occurs (the PDFs are clearly spread on a wider range of values). Such an effect indicates that the noisy behaviour of irradiance strongly influences the whole ecosystem dynamics. Moreover, the nonmonotonic behaviour of the variance (its PDFs become larger and then tighter again as the noise intensity increases) indicates that the noise pushes the ecosystem away from equilibrium, driving it towards a non-equilibrium steady state. Finally, we note that the nonmonotonic behaviour of CV as a function of the noise intensity remains also in the presence of seasonal driving.Figure 5Coefficient of variation ((CV=sigma / mu)) of nine planktonic populations resulting from numerical integration of model equations plotted versus the considered values of D; different curves correspond to different values of the correlation time (tau).Full size imageFigure 5 shows indeed the nonmonotonic response of the ecosystem to the change of D when the deterministic seasonal cycling of the four environmental parameters (temperature, irradiance, wind and salinity) is present. It is easy to observe that also in this instance the major noise-induced effect appears in nanoflagellates and diatoms with a percent standard deviation of 50(%) and 100(%), respectively. The coalescence of different curves (related to different values of (tau)), as D decreases, is due to the fact that for (D rightarrow 0) the impact of the noise is negligible and the evolution of the system practically resembles the deterministic one. On the contrary, for higher values of D remarkable differences arise and clear peaks of CV appear in the considered range of variation.These plots show that, for a fixed value of (tau), there exists a value of the noise intensity for which the planktonic concentrations are maximally spread around their mean values (corresponding to the maximum value of CV and then of the variance). Moreover, such a nonmonotonic behaviour suggests the presence of a resonance, which can be interpreted as the effect of the interplay between the nonlinearity of the system and the environmental random fluctuations.Also in this case, the interplay between the two parameters D and (tau) in determining and characterizing the dynamics of the ecosystem transparently emerges. The value of D corresponding to the maximum value of CV, indeed, basically depends on the specific value of (tau).Finally, we point out that the different dynamic scenarios identified by the D-(tau) couples can be experienced by the system during the year, since the two parameters may seasonally vary depending on the different weather conditions. In other words, a seasonally varying noise (see Fig. 1c) may cause the nine populations explore different regions of the D-(tau) space during the year. Therefore, the results reported in this paper can highlight the detectable yearly variability of a marine ecosystem which does not stem from the deterministic seasonal variation of environmental parameters.Effects on the organic carbonIn this subsection the effects of the irradiance noise on the biogechemistry are analysed. In Fig. 6 the dependence on (tau) of both the CV [panel (a)] and the mean value concentration [panel (b)] of detritus, labile dissolved organic carbon (L-DOC), semi-labile dissolved organic carbon (SL-DOC) and gross primary production (GPP) are shown. All these biogeochemical properties are correlated with carbon cycling. Gross primary production is related to the amount of carbon entering in the ecosystem, and is related to the maximum energy available in the ecosystem progressively dissipated in the trophic web. Gross primary production is directly affected by light fluctuation and its CV shape is very similar to that of the irradiance, Fig. 2. We selected also detritus and DOC because they are important indicators for the carbon cycling dynamics and are related to the cycling of chemicals like heavy metals62. The different curves, related to different values of D, approach the same (vanishing) value for large (tau). As previously discussed for the CV [Fig. 6(a)] of biomass concentrations, this circumstance is due to the fact that, in this case, the system dynamics tends to the deterministic case, characterized by a unique possible realization implying a vanishing standard deviation. For high correlation times thus the system is insensitive to the noise intensity. On the contrary, for small values of (tau), different values of D lead to significant differences of the variance. In particular, detritus, L-DOC and SL-DOC exhibit a clear non-monotonic behaviour whose maximum value depends on the combined values of D-(tau). Only the GPP presents a decreasing monotonic behaviour.The dependence of the mean value concentration on (tau), instead, is qualitatively the same for all the four parameters. Also in this case we can note a diversification with respect to D occurring at small (tau) and a (deterministic) constant value arising for low (high) values of D ((tau)).These results manifest that not only the population dynamics, but also all the biogeochemical processes are profoundly affected by the presence of stochastic environmental variables. The values and the behaviour of the examined quantities are indeed determined by the intimate interplay between the intensity and the time correlation of the noise fluctuations.Figure 6(a) Coefficient of variation ((CV=sigma / mu)) and (b) mean value concentration ((mu)) of detritus, labile dissolved organic carbon (L-DOC), semi-labile dissolved organic carbon (SL-DOC) and gross primary production (GPP) resulting from numerical integration of model equations plotted versus the considered values of (tau); the different curves are related to different values of the correlation time D.Full size image More

  • in

    Human recreation impacts seasonal activity and occupancy of American black bears (Ursus americanus) across the anthropogenic-wildland interface

    Chapron, G. et al. Recovery of large carnivores in Europe’s modern human-dominated landscapes. Science 346, 1517–1519 (2014).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Lute, M. L., Carter, N. H., López-Bao, J. V. & Linnell, J. D. C. Conservation professionals’ views on governing for coexistence with large carnivores. Biol. Cons. 248, 108668 (2020).Article 

    Google Scholar 
    Gantchoff, M. G. & Belant, J. L. Regional connectivity for recolonizing American black bears (Ursus americanus) in southcentral USA. Biol. Cons. 214, 66–75 (2017).Article 

    Google Scholar 
    Ripple, W. J. et al. Status and ecological effects of the world’s largest carnivores. Science 343, 25 (2014).Article 
    CAS 

    Google Scholar 
    Kays, R. et al. Does hunting or hiking affect wildlife communities in protected areas?. J. Appl. Ecol. 54, 242–252 (2017).Article 

    Google Scholar 
    Schipper, J. et al. The status of the world’s land and marine mammals: diversity, threat, and knowledge. Science 322, 225–230 (2008).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Smith, J. A., Wang, Y. & Wilmers, C. C. Top carnivores increase their kill rates on prey as a response to human-induced fear. Proc. R. Soc. B Biol. Sci. 282, 20142711 (2015).Article 

    Google Scholar 
    Stillfried, M., Belant, J. L., Svoboda, N. J., Beyer, D. E. & Kramer-Schadt, S. When top predators become prey: Black bears alter movement behaviour in response to hunting pressure. Behav. Proc. 120, 30–39 (2015).Article 

    Google Scholar 
    Støen, O.-G. et al. Physiological evidence for a human-induced landscape of fear in brown bears (Ursus arctos). Physiol. Behav. 152, 244–248 (2015).PubMed 
    Article 
    CAS 

    Google Scholar 
    Evans, M. J., Rittenhouse, T. A. G., Hawley, J. E. & Rego, P. W. Black bear recolonization patterns in a human-dominated landscape vary based on housing: New insights from spatially explicit density models. Landsc. Urban Plan. 162, 13–24 (2017).Article 

    Google Scholar 
    LaRue, M. A. et al. Cougars are recolonizing the midwest: Analysis of cougar confirmations during 1990–2008. J. Wildl. Manag. 76, 1364–1369 (2012).Article 

    Google Scholar 
    Cove, M. V., Fergus, C., Lacher, I., Akre, T. & McShea, W. J. Projecting mammal distributions in response to future alternative landscapes in a rapidly transitioning region. Remote Sens. 11, 2482 (2019).ADS 
    Article 

    Google Scholar 
    Frid, A. & Dill, L. Human-caused disturbance stimuli as a form of predation risk. Conserv. Ecol. 6, 25 (2002).
    Google Scholar 
    Clinchy, M. et al. Fear of the human “super predator” far exceeds the fear of large carnivores in a model mesocarnivore. Behav. Ecol. 27, 1826–1832 (2016).
    Google Scholar 
    Suraci, J. P., Clinchy, M., Zanette, L. Y. & Wilmers, C. C. Fear of humans as apex predators has landscape-scale impacts from mountain lions to mice. Ecol. Lett. 22, 1578–1586 (2019).PubMed 
    Article 

    Google Scholar 
    Gaynor, K. M., Hojnowski, C. E., Carter, N. H. & Brashares, J. S. The influence of human disturbance on wildlife nocturnality. Science 360, 1232–1235 (2018).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Smith, J. A., Thomas, A. C., Levi, T., Wang, Y. & Wilmers, C. C. Human activity reduces niche partitioning among three widespread mesocarnivores. Oikos 127, 890–901 (2018).Article 

    Google Scholar 
    Tucker, M. A. et al. Moving in the Anthropocene: Global reductions in terrestrial mammalian movements. Science 359, 466–469 (2018).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Carter, N. H., Brown, D. G., Etter, D. R. & Visser, L. G. American black bear habitat selection in northern Lower Peninsula, Michigan, USA, using discrete-choice modeling. Ursus 21, 57–71 (2010).Article 

    Google Scholar 
    Naidoo, R. & Burton, A. C. Relative effects of recreational activities on a temperate terrestrial wildlife assemblage. Conserv. Sci. Pract. 2, e271 (2020).
    Google Scholar 
    Geffroy, B., Samia, D. S. M., Bessa, E. & Blumstein, D. T. How nature-based tourism might increase prey vulnerability to predators. Trends Ecol. Evol. 30, 755–765 (2015).PubMed 
    Article 

    Google Scholar 
    Geffroy, B. et al. Evolutionary dynamics in the Anthropocene: Life history and intensity of human contact shape antipredator responses. PLoS Biol. 18, e3000818 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Beeco, J. A., Hallo, J. C. & Brownlee, M. T. J. GPS visitor tracking and recreation suitability mapping: tools for understanding and managing visitor use. Landsc. Urban Plan. 127, 136–145 (2014).Article 

    Google Scholar 
    Thorsen, N. H. et al. Smartphone app reveals that lynx avoid human recreationists on local scale, but not home range scale. Sci. Rep. 12, 1–13 (2022).Article 
    CAS 

    Google Scholar 
    Evans, M. J., Hawley, J. E., Rego, P. W. & Rittenhouse, T. A. G. Hourly movement decisions indicate how a large carnivore inhabits developed landscapes. Oecologia 190, 11–23 (2019).ADS 
    PubMed 
    Article 

    Google Scholar 
    Carlos, A. W. D., Bright, A. D., Teel, T. L. & Vaske, J. J. Human-black bear conflict in urban areas: an integrated approach to management response. Hum. Dimens. Wildl. 14, 174–184 (2009).Article 

    Google Scholar 
    Johnson, H. E. et al. Human development and climate affect hibernation in a large carnivore with implications for human–carnivore conflicts. J. Appl. Ecol. 55, 663–672 (2018).Article 

    Google Scholar 
    Gould, N. P., Powell, R., Olfenbuttel, C. & DePerno, C. S. Growth and reproduction by young urban and rural black bears. J. Mammal. 102, 1165–1173 (2021).Article 

    Google Scholar 
    Ditmer, M. A., Noyce, K. V., Fieberg, J. R. & Garshelis, D. L. Delineating the ecological and geographic edge of an opportunist: The American black bear exploiting an agricultural landscape. Ecol. Model. 387, 205–219 (2018).Article 

    Google Scholar 
    McFadden-Hiller, J. E. Jr. & Belant, J. L. Spatial distribution of black bear incident reports in michigan. PLoS One 11, e0154474 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Ladle, A., Steenweg, R., Shepherd, B. & Boyce, M. S. The role of human outdoor recreation in shaping patterns of grizzly bear-black bear co-occurrence. PLoS One 13, e0191730 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wilbur, R. C., Lischka, S. A., Young, J. R. & Johnson, H. E. Experience, attitudes, and demographic factors influence the probability of reporting human–black bear interactions. Wildl. Soc. Bull. 42, 22–31 (2018).Article 

    Google Scholar 
    Lustig, E. J., Lyda, S. B., Leslie, D. M., Luttbeg, B. & Fairbanks, W. S. Resource selection by recolonizing American Black Bears. J. Wildl. Manage. 85, 531–542 (2021).Article 

    Google Scholar 
    Sun, C. C., Fuller, A. K., Hare, M. P. & Hurst, J. E. Evaluating population expansion of black bears using spatial capture-recapture. J. Wildl. Manage. 81, 814–823 (2017).Article 

    Google Scholar 
    Kautz, T. M. et al. Large carnivore response to human road use suggests a landscape of coexistence. Glob. Ecol. Conserv. 30, e01772 (2021).Article 

    Google Scholar 
    Michigan Department of Natural Resources (MIDNR) (2021).Blount, J. D., Chynoweth, M. W., Green, A. M. & Şekercioğlu, Ç. H. Review: COVID-19 highlights the importance of camera traps for wildlife conservation research and management. Biol. Cons. 256, 108984 (2021).Article 

    Google Scholar 
    Weather Atlas. https://www.weather-atlas.com/enEvans, J. S. Spatial Analysis and Modelling Utilities. Package ‘spatialEco’. https://cran.r-project.org/web/packages/spatialEco/spatialEco.pdf (2021).Díaz-Ruiz, F., Caro, J., Delibes-Mateos, M., Arroyo, B. & Ferreras, P. Drivers of red fox (Vulpes vulpes) daily activity: prey availability, human disturbance or habitat structure?. J. Zool. 298, 128–138 (2016).Article 

    Google Scholar 
    Moore, J. F. et al. Comparison of species richness and detection between line transects, ground camera traps, and arboreal camera traps. Anim. Conserv. 23, 561–572 (2020).Article 

    Google Scholar 
    Parsons, A. W. et al. Urbanization focuses carnivore activity in remaining natural habitats, increasing species interactions. J. Appl. Ecol. 56, 1894–1904 (2019).Article 

    Google Scholar 
    Allen, M. L., Sibarani, M. C., Utoyo, L. & Krofel, M. Terrestrial mammal community richness and temporal overlap between tigers and other carnivores in Bukit Barisan Selatan National Park, Sumatra. Anim. Biodiv. Conserv. 43(1), 97–107 (2020).Article 

    Google Scholar 
    Tian, C. et al. Temporal niche patterns of large mammals in Wanglang National Nature Reserve, China. Glob. Ecol. Conserv. 22, e01015 (2020).Article 

    Google Scholar 
    Meredith, M. & Ridout, M. Estimates of coefficient of overlapping for animal activity patterns. Package ‘overlap’. https://cran.r-project.org/web/packages/overlap/overlap.pdf (2020).RStudio Team. RStudio: Integrated Development for R. RStudio, PBC, Boston, MA. http://www.rstudio.com/ (2021).Ridout, M. S. & Linkie, M. Estimating overlap of daily activity patterns from camera trap data. JABES 14, 322–337 (2009).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    Lashley, M. A. et al. Estimating wildlife activity curves: comparison of methods and sample size. Sci. Rep. 8, 1–11 (2018).CAS 
    Article 

    Google Scholar 
    Rowcliffe, M. Animal Activity Statistics. Package ‘activity’. https://cran.r-project.org/web/packages/activity/activity.pdf (2021).MacKenzie, D. I., Nichols, J. D., Hines, J. E., Knutson, M. G. & Franklin, A. B. Estimating site occupancy, colonization, and local extinction when a species is detected imperfectly. Ecology 84, 2200–2207 (2003).Article 

    Google Scholar 
    Wei, T., & Simko, V. Visualization of a Correlation Matrix. Package ‘corrplot’. https://cran.r-project.org/web/packages/corrplot/corrplot.pdf (2017).Norton, D. C. et al. Female American black bears do not alter space use or movements to reduce infanticide risk. PLoS One 13, e0203651 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Ditmer, M. A. et al. Behavioral and physiological responses of American black bears to landscape features within an agricultural region. Ecosphere 6, 1–21 (2015).Article 

    Google Scholar 
    Clark, D. et al. Using machine learning methods to predict the movement trajectories of the Louisiana black bear. SMU Data Sci. Rev. 5, 25 (2021).
    Google Scholar  More

  • in

    Climate variability and multi-decadal diatom abundance in the Northeast Atlantic

    Armbrust, E. V. The life of diatoms in the world’s oceans. Nature 459, 185–192 (2009).CAS 
    Article 

    Google Scholar 
    Mann, D. G. The species concept in diatoms. Phycologia 38, 437–495 (1999).Article 

    Google Scholar 
    Smetacek, V. Diatoms and the ocean carbon cycle. Protist 150, 25–32 (1999).CAS 
    Article 

    Google Scholar 
    Rynearson, T. A. et al. Major contribution of diatom resting spores to vertical flux in the sub-polar North Atlantic. Deep. Res. Part I Oceanogr. Res. Pap. 82, 60–71 (2013).CAS 
    Article 

    Google Scholar 
    Allen, J. T. et al. Diatom carbon export enhanced by silicate upwelling in the northeast Atlantic. Nature 437, 728–732 (2005).CAS 
    Article 

    Google Scholar 
    Boyd, P. W., Strzepek, R., Fu, F. & Hutchins, D. A. Environmental control of open-ocean phytoplankton groups: Now and in the future. Limnol. Oceanogr. 55, 1353–1376 (2010).CAS 
    Article 

    Google Scholar 
    Hátún, H., Somavilla, R., Rey, F., Johnson, C. & Mathis, M. The subpolar gyre regulates silicate concentrations in the North Atlantic. Sci. Rep. 1–9 https://doi.org/10.1038/s41598-017-14837-4 (2017).Bopp, L. Response of diatoms distribution to global warming and potential implications: a global model study. Geophys. Res. Lett. 32, 2–5 (2005).Article 
    CAS 

    Google Scholar 
    Warner, A. J. & Hays, G. C. Sampling by the Continuous Plankton Recorder survey. Prog. Oceanogr. 6611, 237–256 (1994).Article 

    Google Scholar 
    Edwards, M., Beaugrand, G., Reid, P. C., Rowden, A. A. & Jones, M. B. Ocean climate anomalies and the ecology of the North Sea. Mar. Ecol. Prog. Ser. 239, 1–10 (2002).Article 

    Google Scholar 
    Allen, S. et al. Interannual stability of phytoplankton community composition in the North-East atlantic. Mar. Ecol. Prog. Ser. 655, 43–57 (2020).Article 

    Google Scholar 
    Edwards, M. & Richardson, A. J. Impact of climate change on marine pelagic phenology and trophic mismatch. Nature 430, 881–884 (2004).CAS 
    Article 

    Google Scholar 
    Wihsgott, J. U. et al. Observations of vertical mixing in autumn and its effect on the autumn phytoplankton bloom. Prog. Oceanogr. 177, 1157–1165 (2019).Article 

    Google Scholar 
    Kamykowski, D. & Zentara, S. J. Predicting plant nutrient concentrations from temperature and sigma-T in the upper kilometer of the world ocean. Deep. Res. Part A-Oceanogr. Res. Pap. 33, 89–105 (1986).CAS 
    Article 

    Google Scholar 
    Kamykowski, D. & Zentara, S. J. Changes in world ocean nitrate availability through the 20th century. Deep. Res. Part I-Oceanogr. Res. Pap. 52, 1719–1744 (2005).Article 

    Google Scholar 
    Behrenfeld, M. J. et al. Climate-driven trends in contemporary ocean productivity. Nature 444, 752–755 (2006).CAS 
    Article 

    Google Scholar 
    López-Urrutia, A., San Martin, E., Harris, R. P. & Irigoien, X. Scaling the metabolic balance of the oceans. Proc. Natl. Acad. Sci. USA 103, 8739–8744 (2006).Article 
    CAS 

    Google Scholar 
    Richardson, A. J. & Schoeman, D. S. Climate impact on plankton ecosystems in the Northeast Atlantic. Science 305, 1609–1612 (2004).CAS 
    Article 

    Google Scholar 
    Edwards, M., Johns, D., Leterme, S., Svendsen, E. & Richardson, A. Regional climate change and harmful algal blooms in the northeast Atlantic. Limnol. Oceanogr. 51, 820–829 (2006).Article 

    Google Scholar 
    Hinder, S. L. et al. Changes in marine dinoflagellate and diatom abundance under climate change. Nat. Clim. Chang. 2, 271–275 (2012).Article 

    Google Scholar 
    Batten, S. et al. CPR sampling: the technical background, materials and methods, consistency and comparability. Prog. Oceanogr. 58, 193–215 (2003).Article 

    Google Scholar 
    Reid, P. C. et al. The Continuous Plankton Recorder: concepts and history, from plankton indicator to undulating recorders. Prog. Oceanogr. 58, 117–173 (2003).Article 

    Google Scholar 
    Edwards, M., Beaugrand, G., Hays, G. C., Koslow, J. A. & Richardson, A. J. Multi-decadal oceanic ecological datasets and their application in marine policy and management. Trends Ecol. Evol. 25, 602–610 (2010).Article 

    Google Scholar 
    Edwards, M. et al. North Atlantic warming over six decades drives decreases in krill abundance with no associated range shift. Commun. Biol. 4, 1–10 (2021).Article 

    Google Scholar 
    Richardson, A. J. et al. Using continuous plankton recorder data. Prog. Oceanogr. 68, 27–74 (2006).Article 

    Google Scholar 
    Hélaouët, P., Beaugrand, G. & Reygondeau, G. Reliability of spatial and temporal patterns of C. finmarchicus inferred from the CPR survey. J. Mar. Syst. 153, 18–24 (2016).Article 

    Google Scholar 
    Owens, N. J. P. et al. All plankton sampling systems underestimate abundance: response to “Continuous plankton recorder underestimates zooplankton abundance” by J.W. Dippner and M. Krause. J. Mar. Syst. 128, 240–242 (2013).Article 

    Google Scholar 
    Jonas, T. D., Walne, A., Beaugrand, G., Gregory, L. & Hays, G. C. The volume of water filtered by a Continuous Plankton Recorder sample: the effect of ship speed. J. Plankton Res. 26, 1499–1506 (2004).Article 

    Google Scholar 
    O’Reilly, C. H., Zanna, L. & Woollings, T. Assessing external and internal sources of Atlantic multidecadal variability using models, proxy data, and early instrumental indices. J. Clim 32, 7727–7745 (2019).Article 

    Google Scholar 
    Qin, M., Dai, A. & Hua, W. Quantifying contributions of internal variability and external forcing to atlantic multidecadal variability since 1870. Geophys. Res. Lett. 47, 1–11 (2020).
    Google Scholar 
    Mann, M. E., Steinman, B. A. & Miller, S. K. Absence of internal multidecadal and interdecadal oscillations in climate model simulations. Nat. Commun. 11, 1–9 (2020).Article 
    CAS 

    Google Scholar 
    Enfield, D. B., Mestas-Nuñez, A. M. & Trimble, P. J. The Atlantic Multidecadal Oscillation and its relation to rainfall and river flows in the continental U.S. Geophys. Res. Lett. 28, 2077–2080 (2001).Article 

    Google Scholar 
    Gray, S. T., Graumlich, L. J., Betancourt, J. L. & Pederson, G. T. A tree-ring based reconstruction of the Atlantic Multidecadal Oscillation since 1567 A.D. Geophys. Res. Lett. 31, 2–5 (2004).Article 

    Google Scholar 
    Edwards, M., Beaugrand, G., Helaouët, P., Alheit, J. & Coombs, S. Marine ecosystem response to the Atlantic Multidecadal Oscillation. PLoS One 8, e57212 (2013).CAS 
    Article 

    Google Scholar 
    Jones, P. D., New, M., Parker, D. E. & Martin, S. & Rigor, I. G. Surface air temperature and its changes over the past 150 years. Rev. Geophys. 37, 173–199 (1999).Article 

    Google Scholar 
    Beaugrand, G. et al. Reorganization of North Atlantic marine copepod biodiversity and climate. Science296, 1692–1694 (2002).CAS 
    Article 

    Google Scholar 
    Alvain, S., Moulin, C., Dandonneau, Y. & Bréon, F. M. Remote sensing of phytoplankton groups in case 1 waters from global SeaWiFS imagery. Deep Sea Res. Part I Oceanogr. Res. Pap. 52, 1989–2004 (2005).Article 

    Google Scholar 
    Kalnay, E. et al. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc. 77, 437–471 (1996).Article 

    Google Scholar 
    Huang, B. et al. Extended reconstructed sea surface temperature, version 5 (ERSSTv5): upgrades, validations, and intercomparisons. J. Clim. 30, 8179–8205 (2017).Article 

    Google Scholar 
    Lam, N. S. N. Spatial interpolation methods: a review. Am. Cartogr. 10, 129–150 (1983).Article 

    Google Scholar 
    Beaugrand, G., McQuatters-Gollop, A., Edwards, M. & Goberville, E. Long-term responses of North Atlantic calcifying plankton to climate change. Nat. Clim. Chang. 3, 263–267 (2012).Article 
    CAS 

    Google Scholar 
    Beaugrand, G. et al. Prediction of unprecedented biological shifts in the global ocean. Nat. Clim. Chang. 9, 237–243 (2019).Article 

    Google Scholar  More

  • in

    Organic and in-organic fertilizers effects on the performance of tomato (Solanum lycopersicum) and cucumber (Cucumis sativus) grown on soilless medium

    Growth conditions and plant materialsTwo experiments were conducted concurrently (sites A and B) in the same screen house in 2019 between the months of May and July at the Landmark University Greenhouse and Hydroponic Technology Center, a section of the Teaching and Research Farm of the University in Omu-Aran, Kwara State Nigeria. Experiment at site B was conducted simultaneously as A so as to validate the results of experiment A. Landmark University lies within Latitude 8° 7′ 26.21388″ and 5° 5′ 0.1788″. Both experiments (A & B) involved tomato (Solanum lycopersicum L. variety cherry) and cucumber (Cucumis sativus L. variety marketer) crops. For each crop, seeds were sown into a separate seed tray filled with coco peat (Coco peat, SRIMATHI EXPORT, INDIA). Cocopeat is the mesocarp tissue or husk after the grinding of coconut fruit. It has a lightweight and high water and nutrient holding capacities, it has an acceptable pH, electrical conductivity, and other chemical attributes27. Rice husk is the by-product of rice after milling. The rice husk used was collected from the rice processing mill of Landmark University. Rice husk is a highly porous and light weighted material with a very high specific area28.Two sets of seed trays (one for organic and another for inorganic fertilizers) were used each for tomato and cucumber crops in the nursery. Both were raised in the nursery for two weeks before transplanting. Black grow bags (30 × 17 cm) filled with a coco peat/rice husk (1:4 ratio by volume) mixture with a weight of about 10 kg were arranged in a screen house. Both the nursery and establishment of crop proper take place in a screen house. The screen house has a galvanized iron as the frame, a UV covering on top, side net for screening insect pests the floor fairly covered with granite. Temperature and relative humidity within the screen house during the period of the experiment was monitored using a Thermograph and a Barograph, and they were at an average of 31 °C and 75%, respectively.The grow bags were randomly placed in the screen house for the unbiased application of amendments. For both tomato and cucumber crops, the treatment comprised of six (6) levels of liquid organic fertilizer (5, 15, 25, 35, 45, 55 mL), in-organic fertilizer, and a control (ordinary borehole water). Levels of organic fertilizers were selected based on the recommendation of 20 mL of liquid organic fertilizer by29. The eight (8) treatments both for tomato and cucumber were arranged in a Completely Randomized Design replicated three times. One healthy plant was maintained per grow bag and four grow bags represent a treatment and there were 32 plants per block each for tomato and cucumber. For both crops, the experiment lasted for 90 days.Organic and in-organic nutrient solutionsThe liquid organic fertilizer used was obtained from the biomass of Mexican sunflower (Tithonia diversifolia). Fresh biomass (mainly leaves and stems) of the plant was collected from the Teaching and Research Farms of Landmark University, Nigeria. After rinsing, they were cut with a sterile knife into pieces of ≤ 1 cm size. A sample was taken for initial physicochemical analyses by grinding in a sterile mortal, diluted with sterile water and analyzed. The biomass was then soaked in sterile water inside a clean container, and allowed to ferment spontaneously for a period of 14 days. During the fermentation, samples were taken every 4 days for microbial analyses of the major players during the fermentation. At the end of fermentation, the mixture was separated using a sieve of mesh size ≤ 2 mm. The liquid portion was then refrigerated prior to the planting regime while another sample was taken to ascertain the physicochemical and microbial qualities of the produced liquid fertilizer. The chemical analysis is presented in Table 4. For inorganic fertilizer, Water soluble fertilizers employed in hydroponics were used (Hydroponics fertilizer, Anmol chemicals, India); calcium nitrate 650 mg L−1, potassium nitrate 450 mg L−1, magnesium 400 mg L−1, chelate 20 mg L−1, mono-ammonium phosphate 400 mg L−1. The electrical conductivity (EC) of the solution was 1.9 dS m-1.Irrigation and fertigationThe tomato and cucumber plants were fertigated morning and evening daily for one hour on each occasion according to the treatments. Preparation of the nutrient solution was with borehole water and was supplied to plants by an online pressure drip irrigation system set at 2.0 L h-1 using an arrowhead on each tomato and cucumber plant. Different tanks (250 L) were installed according to the various treatments making a total of 8 tanks. The organic fertilizer was diluted according to the various treatments equivalent to 1.25, 3.75, 6.25, 8.75, 11.25, and 13.75 L per 250 L of water respectively for 5, 15, 25, 35, 45, and 55 mL treatments. The nutrient solutions were refilled when the consumption is less than 20% of the initial volume (250 L) in the tank. One day per week, crops were irrigated with ordinary water to wash out pipes and prevent deposits of salts. The same concentration of nutrient was used from transplanting to the termination of the study for both tomato and cucumber crops, however, at the flowering of the crops, the volume of fertigation was increased to 3.0 L h-1 to be able to cope with the size of the plants.Trellising, pest and diseases controlFor both tomato and cucumber crops, plant vines were supported by twisting them around a wire that is- attached to the roof of the screen house and 2 m from the ground. Lateral outgrowths were cut off every week to ensure a sturdy single stem. Pests and diseases were scouted every day. Whiteflies, aphids, and other insects were controlled with orizon (Producer, location of producer) (active ingredient, acetamiprid, and abamectin) using 0.133% v/v. Fungi were controlled using ridomil gold (Producer, Location of producer) at 2% w/v.Determination of growth and yield of tomato and cucumberThree tomato and cucumber plants were randomly selected for each treatment for the determination of growth parameters (plant height, leaf area, number of leaves per plant, and stem diameter) at mid the flowering stage of tomato and cucumber plants.The leaf area of tomato was calculated using the model (A = KL2) developed by Lyon30, where L = Length of tomato leaf, K = constant which is 0.1551, and A = leaf area of tomato. Similarly, the leaf area of cucumber was calculated using A = 0.88LW – 4.27, where L = cucumber leaf length and W = cucumber leaf width, A = leaf area of cucumber31.Tomato fruits were ready for harvest from 65 days after transplanting, harvestings were done twice every week (Mondays and Fridays) for up to 85 days after transplanting. Similarly, harvesting of cucumber fruits started 35 days after transplanting and harvestings were also done twice a week (Mondays and Fridays), harvesting was carried out till 60 days after transplanting. Tomato and cucumber fruit yields were counted and weighed at each harvest.Analysis of tomato and cucumber leaves and fruitsAt the 50% flowering stage of tomato and cucumber plants, ten leaf samples were collected from each treatment. The leaf samples were oven-dried at 75 °C for 24 h and thereafter grounded. The grounded samples were later analyzed for nitrogen (N), phosphorous (P), potassium (K), calcium (Ca), and magnesium (Mg) content using the method of described by32. At harvest, four matured tomato and cucumber fruits of uniform size were selected per treatment, and their nutrient compositions were determined using the method of33.Statistical analysisAll data collected on the growth, yield, leaf, and fruit nutrient contents of tomato and cucumber were subjected to analysis of variance (ANOVA). The SPSS V 21.0 (New York, USA) software was used to perform ANOVA and Duncan’s multiple range test (DMRT) was used to compare means at a 5% probability level.
    Ethical approvalI confirm that all the research meets ethical guidelines and adheres to the legal requirements of the study country.Compliance with international, national and/or institutional guidelinesExperimental research (either cultivated or wild), comply with relevant institutional, national, and international guidelines and legislation. Experimental studies were carried out in accordance with relevant institutional, national or international guidelines or regulation. More