More stories

  • in

    Fluctuation spectra of large random dynamical systems reveal hidden structure in ecological networks

    Power spectral density for a general Ornstein–Uhlenbeck processIn the following we develop a method to compute the power spectral density of N-dimensional Ornstein–Uhlenbeck processes,$$frac{d{boldsymbol{xi }}}{dt}={boldsymbol{A}}{boldsymbol{xi }}+{boldsymbol{zeta }}(t),$$
    (15)
    where ζ(t) is an N-vector of Gaussian white noise with correlations ({mathbb{E}}[{boldsymbol{zeta }}(t){boldsymbol{zeta }}{(t^{prime} )}^{T}]=delta (t-t^{prime} ){boldsymbol{B}}). The matrix A determines the mean behaviour of ξ and is considered to be locally stable, i.e. all eigenvalues of A have negative real part. Using the matrices A and B one can fully determine the power spectral density of fluctuations for the Ornstein-Uhlenbeck process.We are interested in the case that the coefficients Aij and Bij are derived from a complex network of interactions with weights drawn at random, possibly with correlations. This framework encompasses a very general class of models with a wealth of real-world applications including but not limited to the ecological focus we have here. The method we describe exploits the underlying network structure of A and B to deduce a self-consistent scheme of equations whose solution contains information on the power spectral density.We start with the definition of the power spectral density Φ(ω) as the Fourier transform of the covariance ({mathbb{E}}[{boldsymbol{xi }}(t){boldsymbol{xi }}{(t+tau )}^{T}]) at equilibrium,$${mathbf{Phi }}(omega )=int_{-infty }^{infty }{{rm{e}}}^{-{rm{i}}omega tau }{mathbb{E}}[{boldsymbol{xi }}(t){boldsymbol{xi }}(t+tau )]dtau .$$
    (16)
    From ref. 33 on multivariate Ornstein–Uhlenbeck processes, we know that the power spectral density can also be written in the form of the matrix equation,$${mathbf{Phi }}(omega )={({boldsymbol{A}}-iomega {boldsymbol{I}})}^{-1}{boldsymbol{B}}{({{boldsymbol{A}}}^{T}+iomega {boldsymbol{I}})}^{-1}.$$
    (17)
    In practice, this equation is difficult to use for large systems as large matrix inversion is analytically intractable and numerical schemes are slow and sometimes unstable. We take an alternative route by recasting Eq. (17) as a complex Gaussian integral reminiscent of problems appearing in the statistical physics of disordered systems. Our approach in the following is to treat ω as a fixed parameter and drop the explicit dependence from our notation. We begin by writing$${mathbf{Phi }}(omega )=frac{| {boldsymbol{A}}-iomega {boldsymbol{I}}{| }^{2}}{{pi }^{N}| {boldsymbol{B}}| }int_{{mathbb{C}}}{e}^{-{{boldsymbol{u}}}^{dagger }{{boldsymbol{Phi }}}^{-1}{boldsymbol{u}}}{boldsymbol{u}}{{boldsymbol{u}}}^{dagger }mathop{prod }limits_{i=1}^{N}d{u}_{i} .$$
    (18)
    Simplification of the integrand is achieved by unpicking the matrix inversion in the exponent via a Hubbard-Stratonovich transformation46,47. To this end we recast the system in the language of statistical mechanics by introducing N complex-valued ‘spins’ ui and N auxiliary variables vi, with the ‘Hamiltonian’$${mathcal{H}}({boldsymbol{u}},{boldsymbol{v}})=-{{boldsymbol{u}}}^{dagger }({boldsymbol{A}}-{rm{i}}omega ){boldsymbol{v}}+{{boldsymbol{v}}}^{dagger }{({boldsymbol{A}}-{rm{i}}omega )}^{dagger }{boldsymbol{u}}+{{boldsymbol{v}}}^{dagger }{boldsymbol{B}}{boldsymbol{v}} .$$
    (19)
    Introducing a bracket operator$$langle cdots rangle :=frac{{int}_{{mathbb{C}}}{e}^{-{mathcal{H}}({boldsymbol{u}},{boldsymbol{v}})}(cdots )d{boldsymbol{u}}d{boldsymbol{v}}}{{int}_{{mathbb{C}}}{e}^{-{mathcal{H}}({boldsymbol{u}},{boldsymbol{v}})}d{boldsymbol{u}}d{boldsymbol{v}}} ,$$
    (20)
    we can obtain succinct expressions for the power spectral density Φ = 〈uu†〉 as well as the resolvent matrix ({boldsymbol{{mathcal{R}}}}={({rm{i}}omega -{boldsymbol{A}})}^{-1}=langle {boldsymbol{u}}{{boldsymbol{v}}}^{dagger }rangle). Thus we may write,$${mathbf{Phi }}=frac{1}{{mathcal{Z}}}{int}_{{mathbb{C}}}{e}^{-{mathcal{H}}({boldsymbol{u}},{boldsymbol{v}})}{boldsymbol{u}}{{boldsymbol{u}}}^{dagger }mathop{prod }limits_{i=1}^{N}d{u}_{i}d{v}_{i} ,$$
    (21)
    where ({mathcal{Z}}=| {boldsymbol{A}}-iomega {boldsymbol{I}}{| }^{2}/{pi }^{2N}).This construction may seem laborious at first, but it unlocks a powerful collection of statistical mechanics tools, including the ‘cavity method’. Originally, the cavity method has been introduced in order to analyse a model for spin glass systems48,49. Further applications of the method include the analysis of the eigenvalue distribution in sparse matrices50,51,52. We will exploit the network structure in a similar fashion in order to compute the power spectral density.In our analysis, we find that it is convenient to split the Hamiltonian in Eq. (19) into the sum of its local contributions at sites i, ({{mathcal{H}}}_{i}), and contributions from interactions between i and j, ({{mathcal{H}}}_{ij}),$${mathcal{H}}=mathop{sum}limits_{i}{{mathcal{H}}}_{i}+mathop{sum}limits_{i sim j}{{mathcal{H}}}_{ij} .$$
    (22)
    These terms can be decomposed as ({{mathcal{H}}}_{i}={{boldsymbol{w}}}_{i}^{dagger }{{boldsymbol{chi }}}_{i}{{boldsymbol{w}}}_{i}) and ({{mathcal{H}}}_{ij}={{boldsymbol{w}}}_{i}^{dagger }{{boldsymbol{chi }}}_{ij}{{boldsymbol{w}}}_{j}), where we introduce the compound spins ({{boldsymbol{w}}}_{i}={({u}_{i},{v}_{i})}^{T}) and transfer matrices,$${{boldsymbol{chi }}}_{i} = , left(begin{array}{ll}0&{A}_{ii}+iomega \ -{A}_{ii}+iomega &{B}_{ii}end{array}right) ,\ {{boldsymbol{chi }}}_{ij} = , left(begin{array}{ll}0&{A}_{ji}\ -{A}_{ij}&{B}_{ij}end{array}right) .$$
    (23)
    Let us focus on the power spectral density of a particular variable ξi, obtained from the diagonal element ϕi = Φii. For this we compute the single-site marginal fi by integrating over all other variables,$${f}_{i}({{boldsymbol{w}}}_{i})=frac{1}{{mathcal{Z}}}{int}_{{mathbb{C}}}{e}^{-{mathcal{H}}}mathop{prod}limits_{jne i}d{{boldsymbol{w}}}_{j}.$$
    (24)
    Alternatively, ϕi can be obtained as the top left entry of the covariance matrix ({{mathbf{Psi }}}_{i}=langle {{boldsymbol{w}}}_{i}{{boldsymbol{w}}}_{i}^{dagger }rangle). We write the covariance matrix as the integral,$${{mathbf{Psi }}}_{i}={int}_{{mathbb{C}}}{f}_{i}({{boldsymbol{w}}}_{i}){{boldsymbol{w}}}_{i}{{boldsymbol{w}}}_{i}^{dagger }d{{boldsymbol{w}}}_{i} ,$$
    (25)
    which could also be expressed in terms of a Gaussian integral,$${{mathbf{Psi }}}_{i}=frac{1}{{pi }^{2}| {{mathbf{Psi }}}_{i}| }{int}_{{mathbb{C}}}{e}^{-{{boldsymbol{w}}}_{i}^{dagger }{{mathbf{Psi }}}_{i}^{-1}{{boldsymbol{w}}}_{i}}{{boldsymbol{w}}}_{i}{{boldsymbol{w}}}_{i}^{dagger }d{{boldsymbol{w}}}_{i} .$$
    (26)
    By comparing Eqs. (25) and (26) we find that$${f}_{i}({{boldsymbol{w}}}_{i})=frac{1}{{pi }^{2}| {{mathbf{Psi }}}_{i}| }{e}^{-{{boldsymbol{w}}}_{i}^{dagger }{{mathbf{Psi }}}_{i}^{-1}{{boldsymbol{w}}}_{i}} .$$
    (27)
    We now insert Eq. (22) into Eq. (24) and obtain,$${f}_{i}({{boldsymbol{w}}}_{i})=frac{1}{{pi }^{2}| {{mathbf{Psi }}}_{i}| }{e}^{-{{mathcal{H}}}_{i}}{int}_{{mathbb{C}}}mathop{prod}limits_{i sim j}left({e}^{-{{mathcal{H}}}_{ij}-{{mathcal{H}}}_{ji}}{f}_{j}^{(i)}d{{boldsymbol{w}}}_{j}right) ,$$
    (28)
    where we write ({f}_{j}^{(i)}) for the ‘cavity marginals’,$${f}_{j}^{(i)}({{boldsymbol{w}}}_{j})=frac{1}{{{mathcal{Z}}}^{(i)}}{int}_{{mathbb{C}}}{e}^{-{{mathcal{H}}}^{(i)}}mathop{prod}limits_{kne i,j}d{{boldsymbol{w}}}_{k} .$$
    (29)
    In essence, the above discussion amounts to organising the 2N integrals in Eq. (21) in a convenient way, with the advantage of providing a simple intuition for the role of the underlying network. The superscript (i) is used to indicate that the quantity corresponds to the cavity network where node i has been removed. We will further use this notation for the ‘cavity covariance matrix’ ({{mathbf{Psi }}}_{jl}^{(i)}) introduced in the following.Next we perform the integration in Eq. (28) and compare to the form in Eq. (27). We thus obtain a recursion formula for the covariance matrix Ψi and the cavity covariance matrices ({{mathbf{Psi }}}_{jl}^{(i)}),$${{mathbf{Psi }}}_{i}={left({{boldsymbol{chi }}}_{i}-mathop{sum}limits_{{{i !sim! j}atop {i! sim! l}}}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}_{jl}^{(i)}{{boldsymbol{chi }}}_{li}right)}^{-1},$$
    (30)
    where the notation i ~ j indicates that we sum over nodes j connected to node i. Unless there is some specific structure underlying the network, we assume that most real world cases have a ‘tree-like’ structure from the local view point of a single node i. Hence, it is highly unlikely that the nodes j and l are nearby in the cavity network where node i is removed, and thus ({{mathbf{Psi }}}_{jl}^{(i)}) only gives non-zero contributions if j = l. We therefore reduce Eq. (30) and obtain for the covariance matrix,$${{mathbf{Psi }}}_{i}={left({{boldsymbol{chi }}}_{i}-mathop{sum}limits_{i sim j}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}_{j}^{(i)}{{boldsymbol{chi }}}_{ji}right)}^{-1}.$$
    (31)
    Similarly, the cavity covariance matrix obeys the equation,$${{mathbf{Psi }}}_{j}^{(i)}={left({{boldsymbol{chi }}}_{j}-mathop{sum}limits_{j sim k,kne i}{{boldsymbol{chi }}}_{jk}{{mathbf{Psi }}}_{k}^{(j)}{{boldsymbol{chi }}}_{kj}right)}^{-1}.$$
    (32)
    Here we use that Ψ(i, j) = Ψ(j) when the nodes i and k are not connected. In other words, removing node j from the cavity network where node i is missing, has the same effect as removing it from the full network. The system in Eq. (31) describes a collection of nonlinear matrix equations that must be solved self-consistently.For networks with high enough connectivity (and to good approximation even with modest connectivity), the removal of a single node does not affect the rest of the network, as its contribution is negligible compared to the full system. Hence the system in Eq. (31) can be reduced to a smaller set of equations approximately satisfied by the matrices Ψi:$${{mathbf{Psi }}}_{i}approx {left({{boldsymbol{chi }}}_{i}-mathop{sum}limits_{i sim j}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}_{j}{{boldsymbol{chi }}}_{ji}right)}^{-1}.$$
    (33)
    The power spectral density ϕi can be obtained as the top left entry of Ψi.In order to progress further, we now consider specific approximations that help us compute the power spectral density. First, we take a mean-field approach in order to obtain the mean power spectral density for all nodes part of the network; we then use the result for the mean-field in order to compute a close approximation to the local power spectral density of a single node. Later, we adapt the method to partitioned networks where nodes belong to different types of connected groups.Mean fieldFor the following, we assume that all agents in the system behave the same on average. In practice, the terms governed by self-interactions Aii are drawn from the same distribution for all agents. Similarly, the terms including Bii are governed by one distribution. Interaction strengths and connections with other nodes in the network are also sampled equally for all agents (we have explored a large Lotka-Volterra ecosystem as an example of such a network). In the mean-field (MF) formulation we assume that the mean degree and excess degree are approximately equal, and replace all quantities in Eqs. (31) and (32) with their average. Ψi = ΨMF ∀ i. We then obtain the following recursion equation,$${{mathbf{Psi }}}^{{rm{MF}}}={left[{mathbb{E}}[{{boldsymbol{chi }}}_{i}]-{mathbb{E}}left(mathop{sum}limits_{i sim j}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}^{{rm{MF}}}{{boldsymbol{chi }}}_{ji}right)right]}^{-1}.$$
    (34)
    In order to solve this equation, we parameterise,$${{mathbf{Psi }}}^{{rm{MF}}}=left(begin{array}{ll}phi &r\ -bar{r}&0end{array}right),$$
    (35)
    where the top left entry ϕ corresponds to the mean power spectral density, and we introduce r as the mean diagonal element of the resolvent matrix ({boldsymbol{{mathcal{R}}}}). Finally by inserting the ansatz of Eq. (35) into Eq. (34) we obtain,$$left(begin{array}{ll}phi &r\ -bar{r}&0end{array}right)^{-1}= , left(begin{array}{ll}0&{mathbb{E}}[{A}_{ii}]+iomega \ -{mathbb{E}}[{A}_{ii}]+iomega &{mathbb{E}}[{B}_{ii}]end{array}right)\ , +cleft(begin{array}{ll}0&bar{r}{mathbb{E}}[{A}_{ij}{A}_{ji}]\ -r{mathbb{E}}[{A}_{ij}{A}_{ji}]&phi {mathbb{E}}[{A}_{ij}^{2}]+(r+bar{r}){mathbb{E}}[{A}_{ij}{B}_{ij}]end{array}right),$$
    (36)
    where c is the average degree (i.e. number of connections) per node. Moreover, the expectations in the second term are to be taken over connected nodes i ~ j (i.e. non-zero matrix entries).From Eq. (36) above, we obtain the equations,$$frac{phi }{| r{| }^{2}} = , {mathbb{E}}[{B}_{ii}]+cleft(phi {mathbb{E}}[{A}_{ij}^{2}]+2{rm{Re}}(r){mathbb{E}}[{A}_{ij}{B}_{ij}]right),\ frac{bar{r}}{| r{| }^{2}} = , -!{mathbb{E}}[{A}_{ii}]+iomega -cr{mathbb{E}}[{A}_{ij}{A}_{ji}].$$
    (37)
    We solve the second equation in Eq. (37) for r and write the mean power spectral density in terms of r,$$phi = , | r{| }^{2}frac{{mathbb{E}}[{B}_{ii}]+2c{rm{Re}}(r){mathbb{E}}[{A}_{ij}{B}_{ij}]}{1-c| r{| }^{2}{mathbb{E}}[{A}_{ij}^{2}]},\ r = , frac{1}{2c{mathbb{E}}[{A}_{ij}{A}_{ji}]}left[-{mathbb{E}}[{A}_{ii}]+iomega right.\ , left.-sqrt{{(-{mathbb{E}}[{A}_{ii}]+iomega )}^{2}-4c{mathbb{E}}[{A}_{ij}{A}_{ji}]}right]$$
    (38)
    This equation informs the first part of the results presented in the main text.Single defect approximationThe single defect approximation (SDA) makes use of the mean-field approximation for the cavity fields, but retains local information about individual nodes. We parameterise similarly to Eq. (35) for a single individual. Moreover, we replace all other quantities with the respective mean-field approximation. Specifically, we obtain$${left(begin{array}{ll}{phi }_{i}^{{rm{SDA}}}&{r}_{i}^{{rm{SDA}}}\ -{bar{r}}_{i}^{{rm{SDA}}}&0end{array}right)}^{-1}= ,left(begin{array}{ll}0&{A}_{ii}+iomega \ -{A}_{ii}+iomega &{B}_{ii}end{array}right)\ , +mathop{sum}limits_{i sim j}left(begin{array}{ll}0&{bar{r}}^{{rm{MF}}}{A}_{ij}{A}_{ji}\ -{r}^{{rm{MF}}}{A}_{ij}{A}_{ji}&{phi }^{{rm{MF}}}{A}_{ij}^{2}+({r}^{{rm{MF}}}+{bar{r}}^{{rm{MF}}}){A}_{ij}{B}_{ij}end{array}right).$$
    (39)
    We solve this equation for ({phi }_{i}^{{rm{SDA}}},{r}_{i}^{{rm{SDA}}}), which delivers$$frac{{phi }_{i}^{{rm{SDA}}}}{| {r}_{i}^{{rm{SDA}}}{| }^{2}} = , {phi }^{{rm{MF}}}mathop{sum}limits_{i sim j}{A}_{ij}^{2}+2{rm{Re}}({r}^{{rm{MF}}})mathop{sum}limits_{i sim j}{A}_{ij}{B}_{ij}+{B}_{ii} ,\ {r}_{i}^{{rm{SDA}}} = , {left({A}_{ii}+iomega +{bar{r}}^{{rm{MF}}}mathop{sum}limits_{i sim j}{A}_{ij}{A}_{ji}right)}^{-1}.$$
    (40)
    Partitioned networkPreviously we assumed that all nodes in a network are interchangeable in distribution. However, many real-world applications feature agents with different properties, imposing a high-level structure on the network. We realise this by partitioning nodes into distinct groups that interact with each other (see the section Trophic structure model for a simple example).In order to handle different connected groups we make use of the cavity method as in Eqs. (31) and (32). In particular, we split the sum in the second term on the right-hand side of these equations into contributions from each group in the partitioned network. Let M denote the number of subgroups Vm in a partitioned network then we write,$${{mathbf{Psi }}}_{i}= , {left({{boldsymbol{chi }}}_{i}-mathop{sum }limits_{m}^{M}mathop{sum}limits_{{{i !sim! j}atop {j!in! {V}_{m}}}}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}_{j}^{(i)}{{boldsymbol{chi }}}_{ji}right)}^{-1} ,\ {{mathbf{Psi }}}_{j}^{(i)} = , {left({{boldsymbol{chi }}}_{j}-mathop{sum }limits_{m}^{M}mathop{sum}limits_{{{j !sim! k}atop {k!in !{V}_{m}}}}{{boldsymbol{chi }}}_{jk}{{mathbf{Psi }}}_{k}^{(j)}{{boldsymbol{chi }}}_{kj}right)}^{-1}.$$
    (41)
    Similar to the previous sections we replace all quantities with a mean-field average ({{mathbf{Psi }}}_{m}^{{rm{MF}}}), but for each group separately. Hence we obtain M equations of the form$${{mathbf{Psi }}}_{i}^{{rm{MF}}}={left[{mathbb{E}}[{{boldsymbol{chi }}}_{i}]-{mathbb{E}}left(mathop{sum }limits_{m}^{M}mathop{sum}limits_{{{i !sim! j}atop {j!in !{V}_{m}}}}{{boldsymbol{chi }}}_{ij}{{mathbf{Psi }}}_{m}^{{rm{MF}}}{{boldsymbol{chi }}}_{ji}right)right]}^{-1}.$$
    (42)
    In order to compute the mean power spectral density for different groups separately, we use a parameterisation as in Eq. (35) for each group. Therefore we have,$${{mathbf{Psi }}}_{m}^{{rm{MF}}}=left(begin{array}{ll}{phi }_{m}&{r}_{m}\ -{bar{r}}_{m}&0end{array}right),$$
    (43)
    for all m = 1, …, M. This delivers 2M equations to solve for all rm and ϕm. Numerically this is straightforward, although algebraically long-winded for the general case. However, the equations simplify for special cases. In the section Trophic structure model we demonstrate this method for a bipartite network where a lack of intra-group interactions simplifies the analysis.Large Lotka-Volterra ecosystemModel descriptionFirst, we define the framework for a general Lotka-Volterra ecosystem with N species and a large but finite system size V ≫ 1. Note that this parameter can be interpreted as a scaling factor for the fluctuation amplitude and thus, larger systems exhibit higher stability and quantitative reliability for our analytic results. Let Xi denote the number of individuals and xi = Xi/V the density of species i = 1, …, N. We start from the following set of reactions that define the underlying stochastic dynamics of the system:$$ , {X}_{i},mathop{to }limits^{{b}_{i}},2{X}_{i} ({rm{birth}})\ , 2{X}_{i},mathop{to }limits^{{R}_{ii}},{X}_{i} ({rm{death}})\ , {X}_{i}+{X}_{j},mathop{to }limits^{{R}_{ij}},left{begin{array}{ll}2{X}_{i}+{X}_{j}&({rm{mutualism}}),hfill\ {X}_{i}&({rm{competition}}),\ 2{X}_{i}&({rm{predation}}).hfillend{array}right.$$
    (44)
    The self-interactions are governed by the birth rate bi  > 0 and density-dependent mortality rate Rii  > 0. Furthermore, we define three interaction types between species i and j, namely mutualism, competition and predation. In the case of mutualistic interactions, both species benefit from each other, whereas competition means that both species have a higher mortality rate, depending on the density of the other species. For predator-prey pairs, one predator species benefits from the death of a prey species. The predator and prey species are chosen randomly, such that species i is equally likely to be a predator or prey of species j.With probability Pc we assign an interaction rate Rij  > 0 to the species pair (i, j), and with probability 1 − Pc there is no interaction between species i and j (i.e. Rij = 0). In other words, each species has on average c = NPc interaction partners. The reaction rates are considered to be i.i.d. random variables drawn from a half-normal distribution (| {mathcal{N}}(0,{sigma }^{2})|), where we write for the mean reaction rate (mu ={mathbb{E}}[{R}_{ij}]=sigma sqrt{2/pi }) and raw second moment ({sigma }^{2}={mathbb{E}}[{R}_{ij}^{2}]). For each interaction pair, the interaction type is chosen such that the proportion of predator-prey pairs is p ∈ [0, 1], and all non-predator-prey interactions are equally distributed between mutualistic and competitive interactions (i.e. the overall proportion of mutualistic/competitive interactions is 1/2(1 − p)). Lastly, we define the symmetry parameter γ = 1 − 2p, where γ = −1 if all interactions are of predator-prey type (p = 1), and similarly γ = +1 if there are no predator-prey interactions (p = 0). In a mixed case where predator-prey and mutualistic/competitive interactions have equal proportion (p = 1/2), we have γ = 0. Later we will see that γ is equivalent to the correlation of signed interaction strengths.In the limit V → ∞, the dynamics of the species density xi obey the ordinary differential equations,$$frac{d{x}_{i}}{dt}={x}_{i}left({b}_{i}+mathop{sum }limits_{j}^{N}{alpha }_{ij}{x}_{j}right),$$
    (45)
    where αij are the interaction coefficients with ∣αij∣ = ∣αji∣ = Rij. The signs of the interaction coefficients are determined by the type of interaction between species i and j. For mutualistic interactions we have αij = αji  > 0, and αij = αji  More

  • in

    Organic fertilization improves soil aggregation through increases in abundance of eubacteria and products of arbuscular mycorrhizal fungi

    1.Lal, R. Carbon sequestration in dryland ecosystems. Environ. Manag. 33, 528–544 (2004).Article 

    Google Scholar 
    2.Six, J., Elliot, E. T. & Paustian, K. Soil microaggregate turnover and microaggregate formation: A mechanism for C organic under no-tillage agriculture. Soil Biol. Biochem. 32, 2099–2103 (2000).CAS 
    Article 

    Google Scholar 
    3.Costa, O. Y. A., Raaijmakers, J. M. & Kuramae, E. E. Microbial extracellular polymeric substances: Ecological function and impact on soil aggregation. Front. Microbiol. 9, 1636 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    4.Bronick, C. J. & Lal, R. Soil structure and management: A review. Geoderma 124, 3–22 (2005).ADS 
    CAS 
    Article 

    Google Scholar 
    5.Reganold, J. P., Glover, J. D., Andrews, P. K. & Hinman, H. R. Sustainability of three apple production systems. Nature 410, 926–930 (2001).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    6.Aziz, I., Mahmood, T. & Islam, K. R. Effect of long term no-till and conventional tillage practices on soil quality. Soil Till. Res. 131, 28–35 (2013).Article 

    Google Scholar 
    7.Gautam, A., Guzman, J., Kovacs, P. & Kumar, S. Manure and inorganic fertilization impacts on soil nutrients, aggregate stability, and organic carbon and nitrogen in different aggregate fractions. Arch. Agron. Soil Sci. https://doi.org/10.1080/03650340.2021.1887480 (2021).Article 

    Google Scholar 
    8.Lin, Y., Ye, G., Liu, D., Fan, J. & Ding, W. Long-term manure application increases soil organic matter and aggregation, and alters microbial community structure and keystone taxa. Soil Biol. Biochem. 134, 187–196 (2019).CAS 
    Article 

    Google Scholar 
    9.Six, J., Bossuyt, H., DeGryze, S. & Denef, K. A history of research on the link between (micro)aggregates, soil biota, and soil organic matter dynamics. Soil Till. Res. 79, 7–31 (2004).Article 

    Google Scholar 
    10.Kumar, R., Rawat, K. S., Singh, J., Singh, A. & Rai, A. Soil aggregation dynamics and carbon sequestration. J. Appl. Nat. Sci. 5, 250–267 (2013).CAS 
    Article 

    Google Scholar 
    11.Lehmann, A., Zheng, W. S. & Rillig, M. C. Soil biota contributions to soil aggregation. Nat. Ecol. Evol. 1, 1828–1835 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    12.Spatafora, J. W. et al. A phylum-level phylogenetic classification of zygomycete fungi based on genome-scale data. Mycologia 108(5), 1028–1046 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.Miller, R. M. & Jastrow, J. D. Mycorrhizal fungi influence soil structure. In Arbuscular Mycorrhizas: Molecular Biology and Physiology (eds Kapulnik, Y. & Douds, D. D.) 3–18 (Kluwer, 2000).Chapter 

    Google Scholar 
    14.Leifheit, E. E., Veresoglou, S. D., Lehmann, A., Morris, E. K. & Rillig, M. C. Multiple factors influence the role of arbuscular mycorrhizal fungi in soil aggregation—A meta-analysis. Plant Soil 374, 523–537 (2014).CAS 
    Article 

    Google Scholar 
    15.Lehmann, A., Leifheit, E. F. & Rillig, M. C. Mycorrhizas and Soil Aggregation (Elsevier, 2017).Book 

    Google Scholar 
    16.Bethlenfalvay, G. J., Cantrell, I. C., Mihara, K. L. & Schreiner, R. P. Relationships between soil aggregation and mycorrhizae as influenced by soil biota and nitrogen nutrition. Biol. Fertil. Soil. 28, 356–363 (1999).Article 

    Google Scholar 
    17.Rashid, M. I. et al. Bacteria and fungi can contribute to nutrients bioavailability and aggregate formation in degraded soils. Microbiol. Res. 183, 26–41 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    18.Velmourougane, K., Prasanna, R. & Saxena, A. K. Agriculturally important microbial biofilms: Present status and future prospects. J. Basic Microbiol. 57, 548–573 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    19.Bomfeti, C. A. et al. Exopolysaccharides produced by the symbiotic nitrogen fixing bacteria of leguminosae. Rev. Bras. Cienc. Sol. 35, 657–671 (2011).CAS 
    Article 

    Google Scholar 
    20.Daynes, C. N., Zhang, N., Saleeba, J. A. & McGee, P. A. Soil aggregates formed in vitro by saprotrophic Trichocomaceae have transient water-stability. Soil Biol. Biochem. 48, 151–161 (2012).CAS 
    Article 

    Google Scholar 
    21.Bystrianský, L. et al. Observations on two microbial life strategies in soil: Planktonic and biofilmforming microorganisms are separable. Soil Biol. Biochem. 136, 107535 (2019).Article 
    CAS 

    Google Scholar 
    22.Rillig, M. C. Arbuscular mycorrhizae, glomalin, and soil aggregation. Can. J. Soil Sci. 84, 355–363 (2004).Article 

    Google Scholar 
    23.Yang, Y., Chuangjun, H., Huang, L., Ban, Y. & Tang, M. The effects of arbuscular mycorrhizal fungi on glomalin-related soil protein distribution, aggregate stability and their relationships with soil properties at different soil depths in lead-zinc contaminated area. PLoS ONE 12, e0182264 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    24.Baldock, J. A. Interactions of organic materials and microorganisms with minerals in the stabilization of soil structure. In Interactions Between Soil Particles and Microorganisms (eds Huang, P. M. et al.) (Wiley, 2002).
    Google Scholar 
    25.Tisdall, J. M. & Oades, J. M. Organic matter and water-stable aggregates in soils. J. Soil Sci. 33, 141–163 (1982).CAS 
    Article 

    Google Scholar 
    26.Martens, D. A. Management and crop residue influence soil aggregate stability. J. Environ. Qual. 29, 723–727 (2000).CAS 
    Article 

    Google Scholar 
    27.Aiken, G. R., McKnight, D. M., Wershaw, R. L. & MacCarthy, P. Humic Substances in Soil, Sediment, and Water (Wiley, 1985).
    Google Scholar 
    28.Jandl, R. & Sletten, R. S. Mineralization of forest soil carbon: Interactions with metals. J. Plant Nutr. Soil Sci. 162, 623–629 (1999).CAS 
    Article 

    Google Scholar 
    29.Bipfubusa, M., Angers, D. A., N’Dayegamiye, A. & Antoun, H. Soil aggregation and biochemical properties following the application of fresh and composted organic amendments. Soil Sci. Soc. Am. J. 72, 160–166 (2008).ADS 
    CAS 
    Article 

    Google Scholar 
    30.Jiao, Y., Whalen, J. K. & Hendershot, W. H. No-tillage and manure applications increase aggregation and improve nutrient retention in a sandy-loam soil. Geoderma 134, 24–33 (2006).ADS 
    CAS 
    Article 

    Google Scholar 
    31.Gielnik, A. et al. Effect of digestate application on microbial respiration and bacterial communities’ diversity during bioremediation of weathered petroleum hydrocarbons contaminate. Sci. Total Environ. 670, 271–281 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    32.Erktan, A., Rillig, M., Carminati, A., Jousset, A. & Scheu, S. Fungal and bacterial predator-prey systems influence soil aggregate formation and stabilization. Geophys. Res. Abstracts 21, 1–1 (2019).
    Google Scholar 
    33.Rillig, M. C. et al. Plant root and mycorrhizal fungal traits for understanding soil aggregation. New Phytol. 205, 1385–1388 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    34.Kandeler, E. Aggregate stability. In Methods in Soil Biology (eds Schiner, F. et al.) (Springer, 1996).
    Google Scholar 
    35.Demyan, M. S. et al. Use of specific peaks obtained by diffuse reflectance Fourier transform mid-infrared spectroscopy to study the composition of organic matter in a Haplic Chernozem. Eur. J. Soil Sci. 63(2), 189–199 (2012).CAS 
    Article 

    Google Scholar 
    36.Körschens, M., Schulz, E. & Behm, R. Hot water extractable carbon and nitrogen of soils as a criterion for their ability of N-release. Zentralbl. Mikrobiol. 145, 305–311 (1990).Article 

    Google Scholar 
    37.Wright, S. & Upadhyaya, A. Extraction of an abundant and unusual protein from soil and comparison with hyphal protein of arbuscular mycorrhizal fungi. Soil Sci. 161, 575–586 (1996).ADS 
    CAS 
    Article 

    Google Scholar 
    38.Bradford, M. M. A rapid and senstive method for the quantification of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72, 248–254 (1976).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Thonar, C., Erb, A. & Jansa, J. Real-time PCR to quantify composition of arbuscular mycorrhizal fungal communities—marker design, verification, calibration and field validation. Mol. Ecol. Res. 12, 219–232 (2012).CAS 
    Article 

    Google Scholar 
    40.Simon, L. M., Lalonde, T. D. & Bruns, T. D. Specific amplification of 18S fungal ribosomal genes from vesicular arbuscular endomycorrhizal fungi colonizing roots. Appl. Environ. Microbiol. 58, 291–295 (1992).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    41.Lee, J., Lee, S. & Young, J. P. W. Improved PCR primers for the detection and identification of arbuscular mycorrhizal fungi. FEMS Microbiol. Ecol. 65, 339–349 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    42.Stoeck, T. et al. Multiple marker parallel tag environmental DNA sequencing reveals a highly complex eukaryotic community in marine anoxic water. Mol. Ecol. 19, 21–31 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    43.Ampe, F., Omar, N. B., Moizan, C., Wacher, C. & Guyot, J.-P. Polyphasic study of the spatial distribution of microorganisms in Mexican pozol, a fermented maize dough, demonstrates the need for cultivation independent methods to investigate traditional fermentations. Appl. Environ. Microbiol. 65, 5464–5473 (1999).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Muyzer, G., de Wall, E. C. & Uitterlinden, A. G. Profiling of complex microbial populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-amplified genes coding for 16S rRNA. Appl. Environ. Microbiol. 59, 695–700 (1993).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    45.Verhoeven, K. J. F., Simonsen, K. L. & McIntyre, L. M. Implementing false discovery rate control: Increasing your power. Oikos 108, 643–647 (2005).Article 

    Google Scholar 
    46.ter Braak, C. J. F. & Šmilauer, P. Canoco Reference Manual and User’s Guide: Software for Ordination (version 5.10). (Biometris, Wageningen University & Research, 2018). More

  • in

    Chemical cues that attract cannibalistic cane toad (Rhinella marina) larvae to vulnerable embryos

    Parotoid gland extract preparationAdult cane toads (obtained in south-eastern Queensland, December 2018) were killed humanely using the cool/freeze method8 and stored at − 20 °C. Parotoid glands (54 g) excised from 23 thawed toads were macerated in H2O (250 mL) with a commercial blender, and filtered through a bed of Celite 545. The filtrate was concentrated in vacuo at 40 °C, and was partitioned into ethyl acetate (EtOAc) and H2O solubles. The EtOAc extract (750 mg) containing mostly bufagenins (Fig. 1a) was used in the attractant assay without further purification.Figure 1Analytical HPLC (298 nm) chromatogram of extracts obtained from (a) frozen parotoid gland, (b) eggs, (c) early-development tadpole, (d) late-development tadpole and (e) fresh parotoid secretion of cane toads, Rhinella marina [HPLC condition: Agilent Zorbax C8 column, 5 μm, 4.6 × 150 mm, 1 mL/min flow rate 15 min gradient elution from 90% H2O in MeCN, to 100% MeCN, with a constant 0.01% TFA in MeCN modifier]. Highlights: light blue = unspecified bufagenins (MW 400–432); light pink = unspecified bufolipins (MW 630–700); blue = bufagenins 1–5; red = bufotoxins 6–10; pink = bufolipin 11. Structures for 1–11 are shown in Fig. 2, and were assigned on the basis of spectroscopic analysis and comparisons with authentic standards.Full size imageEgg extract preparationCane toad eggs obtained from two laboratory-laid clutches (see method below, Northern Territory, October 2010) were stored at − 20 °C until extraction. Frozen eggs were freeze-dried to yield dry material (1.5 g) that was extracted overnight at room temperature with 90:10 MeOH:H2O (100 mL). The resulting solvent extract was concentrated in vacuo at 40 °C to give a crude material (251 mg) which was partitioned into EtOAc and H2O solubles. The EtOAc (160 mg) extract containing mostly bufagenins and bufolipins (Fig. 1b) was used in the attractant assay without further purification.Early-development tadpole extract preparationEarly developmental stage cane toad tadpoles were collected live from the wild (Northern Territory, March 2010), and stored at − 18 °C until extraction. Frozen tadpoles were freeze-dried to yield dry material (2.6 g) that was extracted overnight at room temperature with 90:10 MeOH:H2O (100 mL). The resulting solvent extract was concentrated in vacuo at 40 °C to give a crude material (1062 mg), which was partitioned into n-BuOH and H2O solubles. The BuOH extract (582 mg) containing mostly bufagenins and bufolipins (Fig. 1c) was used in the attractant assay without further purification.Late-development tadpole extract preparationMid to late developmental stage tadpoles were collected live from the wild (Northern Territory, December 2010), and stored at − 18 °C until extraction. Frozen tadpoles were freeze-dried to give dry material (13.4 g) that was extracted overnight at room temperature with 90:10 MeOH:H2O (100 mL). The resulting solvent extract was concentrated in vacuo at 40 °C to give a crude material (6899 mg), which was partitioned into n-BuOH and H2O solubles. The BuOH extract (3885 mg) containing mostly bufagenins (Fig. 1d) was used in the attractant assay without further purification.Parotoid secretion extract preparationParotoid secretion was obtained from a live adult toad (Northern Territory, August 2011) by mechanical compression of the parotoid gland directly into MeOH, which following concentration in vacuo yielded a crude MeOH extract (26.2 mg). The crude MeOH extract containing mostly bufotoxins (Fig. 1e) was used in the attractant assay without further purification.Pure compounds preparationMarinobufagin (1), marinobufotoxin (6) and suberoyl-l-arginine (13) were obtained from our in-house pure compound library, and their purities were confirmed by LCMS, HRMS and NMR (see Supporting Information for 1H NMR spectra of the pure compounds). Plant cardenolides: digitoxigenin (14), ouabain (15) and digoxin (16) (Fig. 2) were purchased from Sigma Aldrich and were used in the attractant assay without further purification.Figure 2Compounds identified in different stages of cane toad (Rhinella marina) (1–13) and plant derived cardenolides (14–16).Full size imageChemical analysesAnalytical HPLC was performed using an Agilent 1100 series module equipped with a diode array detector on an Agilent Zorbax Stable Bond C8 column (4.6 × 150 mm, 5 μm), 1 mL/min flow rate, 15 min gradient elution from 90% H2O in MeCN to 100% MeCN with a constant 0.01% TFA in MeCN modifier. All analytes were prepared in MeOH stock solutions (1 mg/mL) and an aliquot (10 μL) used for each analysis. HPLC chromatograms were monitored at 298 nm (the α-pyrone chromophore common to all bufadienolides). Compounds 1–12 (Fig. 2) present in the extracts were identified by LC-DAD-ESIMS and comparison with authentic standards (see Supporting Information Table S1). LC-DAD-ESIMS (Liquid Chromatography coupled to Diode Array Detector and Electrospray Ionization Mass Spectra) was acquired using an Agilent 1100 Series LC/MSD mass detector in both positive and negative modes using Agilent Zorbax Stable Bond C8 column (4.6 × 150 mm, 5 μm) with 1 mL/min flow rate, 15 min gradient elution from 90% H2O in MeCN to 100% MeCN with a constant 0.05% formic acid in MeCN modifier.Bait preparationsStock solutions of all attractant extracts were prepared in MeOH (20, 2.0 and 0.20 mg/mL concentrations), with a fixed volume (0.5 mL) of each loaded onto porous ceramic rings (Majestic Aquariums, Sydney, NSW) to give a series of loadings per ceramic ring (10, 1.0 and 0.1 mg) per attractant extract preparation. Stock solutions were also prepared for all pure compound attractants in MeOH (5.0 and 0.5 mM) with a fixed volume (0.5 mL) of each loaded onto porous ceramic rings to give a series of loadings per ceramic ring (2.5 and 0.25 µmoles) per attractant pure compound preparation (marinobufagin, 1.00 and 0.10 mg; digitoxigenin, 0.94 and 0.094 mg; marinobufotoxin, 1.78 and 0.178 mg; ouabain octahydrate, 1.82 and 0.182 mg; digoxin, 1.95 and 0.195 mg; suberoyl-l-arginine, 0.825 and 0.0825 mg per ceramic ring, respectively). Negative controls were ceramic rings loaded with MeOH (0.5 mL/ring) only. All impregnated ceramic rings were left in the fume-hood overnight at room temperature to allow the MeOH to evaporate, and to fix the attractants to the ceramic matrix.Toad breedingAdult toads were collected from the Adelaide River floodplain, near the city of Darwin in tropical Australia, and the animals were held in outdoor enclosures at The University of Sydney Tropical Ecology Research Facility at Middle Point, Northern Territory (12°34.73′S, 131°18.85″E). Breeding was induced by injection of the synthetic gonadotrophin leuprorelin acetate (Lucrin, Abbott Australasia). Females were injected with 0.75 mL doses of 0.25 mg/mL, while males were injected with doses of 0.25 mL4,9. Toads were injected just prior to sunset, and the pairs were placed in 70 L plastic tubs set on an angle with 8 L water. The following morning, eggs were collected and placed in 18 L tanks holding 9 L aerated water. When eggs developed into free-swimming tadpoles (Gosner10 stage 25), tadpoles were transferred to outdoor 750 L mesocosms located in a shaded area. Tadpoles were fed algae wafers (Kyorin, Japan) ad libitum daily, with 50% of the water in mesocosms changed every 3 days. Tadpoles (stage 30–39) were haphazardly selected from mesocosms for use in attraction trials as required.Attraction trialsAttraction trials were conducted in a covered outdoor enclosure exposed to ambient temperature between 0930 and 1700 hours (maximum daily water temperature range over all trials: 26–32 °C). Each trial used plastic pools (1 m diameter) filled with 90 L of well water. Within each pool we placed two plastic traps (175 mm × 120 mm × 70 mm), each of which had a funnel (1 cm diameter) attached to one side. The traps were positioned in the centre of the pool 5 cm apart, with the funnels facing outward. Each pool was stocked with 50 tadpoles from a single clutch. Tadpoles were allowed to settle for 2 h, after which we randomly allocated treatments to traps (i.e., control or chemical). A single bait was added to each trap, and the number of tadpoles within each trap was counted hourly for 6 h. Water temperature was measured at hourly intervals using a hand-held thermometer.Attraction responses to 26 combinations of chemical/concentration were tested, using a total of nine tadpole clutches. Each concentration of each chemical was tested using 4–7 tadpole clutches. The tadpole clutches used for each trial were chosen randomly, with the proviso that they had not been previously tested with the same chemical concentration. Individual tadpoles and baits were used only once in trials.Statistical analysisWe analysed tadpole attraction as a binomial response (trap preference: chemical trap vs control trap) using logistic regression11 in R12, package MASS:glmmPQL). Models were based on the quasibinomial distribution to account for overdispersion of data, with Treatment (control vs. chemical) and Time (hourly intervals) as fixed effects. Random effects were accounted for by nesting trap within pool and responding tadpole clutch. We did not apply Bonferroni corrections to treatment p values due to the highly subjective nature of deciding when to apply such corrections13,14, see both papers for further problems with use of Bonferroni corrections). Rather, we provide unadjusted treatment p values in association with effect sizes (i.e., odds ratio of trap preference; this being a more meaningful indicator of biological significance) to interpret our attraction results13,14.Ethics approvalThis research was approved under permit 6033 from the University of Sydney Animal Care Committee. All methods were performed in accordance with the relevant guidelines and regulations, including ARRIVE guidelines.Consent for publicationAll authors agree to publication of this work. More

  • in

    Evidence for large microbial-mediated losses of soil carbon under anthropogenic warming

    1.Jackson, R. B. et al. The ecology of soil carbon: pools, vulnerabilities, and biotic and abiotic controls. Annu. Rev. Ecol. Evol. Syst. 48, 419–445 (2017).Article 

    Google Scholar 
    2.Köchy, M., Hiederer, R. & Freibauer, A. Global distribution of soil organic carbon – Part 1: Masses and frequency distributions of SOC stocks for the tropics, permafrost regions, wetlands, and the world. Soil 1, 351–365 (2015).Article 

    Google Scholar 
    3.Le Quéré, C. et al. Global carbon budget 2013. Earth Syst. Sci. Data 6, 235–263 (2014).Article 

    Google Scholar 
    4.Davidson, E. A. & Janssens, I. A. Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature 440, 165–173 (2006).Article 

    Google Scholar 
    5.Bond-Lamberty, B. & Thomson, A. Temperature-associated increases in the global soil respiration record. Nature 464, 579–582 (2010).Article 

    Google Scholar 
    6.Crowther, T. W. et al. Quantifying global soil carbon losses in response to warming. Nature 540, 104–108 (2016).Article 

    Google Scholar 
    7.Intergovernmental Panel on Climate Change (IPCC). Climate Change 2013: The Physical Science Basis. Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge Univ. Press, 2013).8.Wieder, W. R., Sulman, B. N., Hartman, M. D., Koven, C. D. & Bradford, M. A. Arctic soil governs whether climate change drives global losses or gains in soil carbon. Geophys. Res. Lett. 46, 14486–14495 (2019).Article 

    Google Scholar 
    9.Guo, X. et al. Gene-informed decomposition model predicts lower soil carbon loss due to persistent microbial adaptation to warming. Nat. Commun. 11, 4897 (2020).Article 

    Google Scholar 
    10.Friedlingstein, P. et al. Climate–carbon cycle feedback analysis: results from the C4MIP model intercomparison. J. Clim. 19, 3337–3353 (2006).Article 

    Google Scholar 
    11.Zhu, Z. et al. Greening of the Earth and its drivers. Nat. Clim. Chang. 6, 791–795 (2016).Article 

    Google Scholar 
    12.Keenan, T. F. & Riley, W. J. Greening of the land surface in the world’s cold regions consistent with recent warming. Nat. Clim. Chang. 8, 825–828 (2018).Article 

    Google Scholar 
    13.Euskirchen, E. S., Mcguire, A. D., Chapin, F. S., Yi, S. & Thompson, C. C. Changes in vegetation in northern Alaska under scenarios of climate change, 2003–2100: implications for climate feedbacks. Ecol. Appl. 19, 1022–1043 (2009).Article 

    Google Scholar 
    14.Conant, R. T. et al. Temperature and soil organic matter decomposition rates – synthesis of current knowledge and a way forward. Glob. Chang. Biol. 17, 3392–3404 (2011).Article 

    Google Scholar 
    15.Peñuelas, J. et al. Shifting from a fertilization-dominated to a warming-dominated period. Nat. Ecol. Evol. 1, 1438–1445 (2017).Article 

    Google Scholar 
    16.Naidu, D. G. & Bagchi, S. Greening of the earth does not compensate for rising soil heterotrophic respiration under climate change. Glob. Chang. Biol. 27, 2029–2038 (2021).Article 

    Google Scholar 
    17.Koven, C. D., Lawrence, D. M. & Riley, W. J. Permafrost carbon–climate feedback is sensitive to deep soil carbon decomposability but not deep soil nitrogen dynamics. Proc. Natl Acad. Sci. USA 112, 3752–3757 (2015).Article 

    Google Scholar 
    18.Xu, X. et al. Plant community structure regulates responses of prairie soil respiration to decadal experimental warming. Glob. Chang. Biol. 21, 3846–3853 (2015).Article 

    Google Scholar 
    19.Melillo, J. M. et al. Long-term pattern and magnitude of soil carbon feedback to the climate system in a warming world. Science 358, 101–105 (2017).Article 

    Google Scholar 
    20.Guo, X. et al. Climate warming leads to divergent succession of grassland microbial communities. Nat. Clim. Chang. 8, 813–818 (2018).Article 

    Google Scholar 
    21.Zhou, J. et al. Microbial mediation of carbon-cycle feedbacks to climate warming. Nat. Clim. Chang. 2, 106–110 (2012).Article 

    Google Scholar 
    22.Luo, Y., Wan, S., Hui, D. & Wallace, L. L. Acclimatization of soil respiration to warming in a tall grass prairie. Nature 413, 622–625 (2001).Article 

    Google Scholar 
    23.García-Palacios, P. et al. Pathways regulating decreased soil respiration with warming in a biocrust-dominated dryland. Glob. Chang. Biol. 24, 4645–4656 (2018).Article 

    Google Scholar 
    24.Karhu, K. et al. Temperature sensitivity of soil respiration rates enhanced by microbial community response. Nature 513, 81–84 (2014).Article 

    Google Scholar 
    25.Allison, S. D., Wallenstein, M. D. & Bradford, M. A. Soil-carbon response to warming dependent on microbial physiology. Nat. Geosci. 3, 336–340 (2010).Article 

    Google Scholar 
    26.Bradford, M. A. et al. Thermal adaptation of soil microbial respiration to elevated temperature. Ecol. Lett. 11, 1316–1327 (2008).Article 

    Google Scholar 
    27.Crowther, T. W. & Bradford, M. A. Thermal acclimation in widespread heterotrophic soil microbes. Ecol. Lett. 16, 469–477 (2013).Article 

    Google Scholar 
    28.Wieder, W. R., Bonan, G. B. & Allison, S. D. Global soil carbon projections are improved by modelling microbial processes. Nat. Clim. Chang. 3, 909–912 (2013).Article 

    Google Scholar 
    29.Mueller, C. W. et al. Large amounts of labile organic carbon in permafrost soils of northern Alaska. Glob. Chang. Biol. 21, 2804–2817 (2015).Article 

    Google Scholar 
    30.Poeplau, C., Kätterer, T., Leblans, N. I. W. & Sigurdsson, B. D. Sensitivity of soil carbon fractions and their specific stabilization mechanisms to extreme soil warming in a subarctic grassland. Glob. Chang. Biol. 23, 1316–1327 (2017).Article 

    Google Scholar 
    31.Wang, X. et al. Soil respiration under climate warming: Differential response of heterotrophic and autotrophic respiration. Glob. Chang. Biol. 20, 3229–3237 (2014).Article 

    Google Scholar 
    32.Carey, J. C. et al. Temperature response of soil respiration largely unaltered with experimental warming. Proc. Natl Acad. Sci. USA 113, 13797–13802 (2016).Article 

    Google Scholar 
    33.Crowther, T. W. et al. The global soil community and its influence on biogeochemistry. Science 365, eaav0550 (2019).Article 

    Google Scholar 
    34.Raich, J. W. & Schlesinger, W. The global carbon dioxide flux in soil respiration and its relationship to vegetation and climate. Tellus 44, 81–99 (1992).Article 

    Google Scholar 
    35.Jian, J. et al. A restructured and updated global soil respiration database (SRDB-V5). Earth Syst. Sci. Data 13, 255–267 (2021).Article 

    Google Scholar 
    36.Bond-Lamberty, B. & Thomson, A. A global database of soil respiration data. Biogeosciences 7, 1915–1926 (2010).Article 

    Google Scholar 
    37.Rustad, L. E., Huntington, T. G. & Boone, R. D. Controls on soil respiration: Implications for climate change. Biogeochemistry 48, 1–6 (2000).Article 

    Google Scholar 
    38.Hursh, A. et al. The sensitivity of soil respiration to soil temperature, moisture, and carbon supply at the global scale. Glob. Chang. Biol. 23, 2090–2103 (2017).Article 

    Google Scholar 
    39.Xue, K. et al. Tundra soil carbon is vulnerable to rapid microbial decomposition under climate warming. Nat. Clim. Chang. 6, 595–600 (2016).Article 

    Google Scholar 
    40.Hengl, T. et al. SoilGrids250m: Global gridded soil information based on machine learning. PLoS One 12, e0169748 (2017).Article 

    Google Scholar 
    41.Ballantyne, A. et al. Accelerating net terrestrial carbon uptake during the warming hiatus due to reduced respiration. Nat. Clim. Chang. 7, 148–152 (2017).Article 

    Google Scholar 
    42.Schuur, E. A. G. et al. Climate change and the permafrost carbon feedback. Nature 520, 171–179 (2015).Article 

    Google Scholar 
    43.Shi, Z. et al. The age distribution of global soil carbon inferred from radiocarbon measurements. Nat. Geosci. 13, 555–559 (2020).Article 

    Google Scholar 
    44.Raich, J. W., Russell, A. E., Kitayama, K., Parton, W. & Vitousek, P. M. Temperature influences carbon accumulation in moist tropical forests. Ecology 87, 76–87 (2006).Article 

    Google Scholar 
    45.Parton, W. J., Stewart, J. W. B. & Cole, C. V. Dynamics of C, N, P and S in grassland soils: a model. Biogeochemistry 5, 109–131 (1988).Article 

    Google Scholar 
    46.Plaza, C. et al. Direct observation of permafrost degradation and rapid soil carbon loss in tundra. Nat. Geosci. 12, 627–631 (2019).Article 

    Google Scholar 
    47.Koven, C. D., Hugelius, G., Lawrence, D. M. & Wieder, W. R. Higher climatological temperature sensitivity of soil carbon in cold than warm climates. Nat. Clim. Chang. 7, 817–822 (2017).Article 

    Google Scholar 
    48.Kirschbaum, M. U. F. The temperature dependence of soil organic matter decomposition, and the effect of global warming on soil organic C storage. Soil Biol. Biochem. 27, 753–760 (1995).Article 

    Google Scholar 
    49.Kirschbaum, M. U. F. The temperature dependence of organic-matter decomposition—still a topic of debate. Soil Biol. Biochem. 38, 2510–2518 (2006).Article 

    Google Scholar 
    50.Serreze, M. C. & Barry, R. G. Processes and impacts of Arctic amplification: A research synthesis. Glob. Planet. Change 77, 85–96 (2011).Article 

    Google Scholar 
    51.Serna-Chavez, H. M., Fierer, N. & Van Bodegom, P. M. Global drivers and patterns of microbial abundance in soil. Glob. Ecol. Biogeogr. 22, 1162–1172 (2013).Article 

    Google Scholar 
    52.Xu, X. et al. Global pattern and controls of soil microbial metabolic quotient. Ecol. Monogr. 87, 429–441 (2017).Article 

    Google Scholar 
    53.van den Hoogen, J. et al. Soil nematode abundance and functional group composition at a global scale. Nature 572, 194–198 (2019).Article 

    Google Scholar 
    54.Hashimoto, S. et al. Global spatiotemporal distribution of soil respiration modeled using a global database. Biogeosciences 12, 4121–4132 (2015).Article 

    Google Scholar 
    55.Buchkowski, R. W., Bradford, M. A., Grandy, A. S., Schmitz, O. J. & Wieder, W. R. Applying population and community ecology theory to advance understanding of belowground biogeochemistry. Ecol. Lett. 20, 231–245 (2017).Article 

    Google Scholar 
    56.Bradford, M. A. et al. Managing uncertainty in soil carbon feedbacks to climate change. Nat. Clim. Chang. 6, 751–758 (2016).Article 

    Google Scholar 
    57.Trivedi, P., Anderson, I. C. & Singh, B. K. Microbial modulators of soil carbon storage: Integrating genomic and metabolic knowledge for global prediction. Trends Microbiol. 21, 641–651 (2013).Article 

    Google Scholar 
    58.Cavicchioli, R. et al. Scientists’ warning to humanity: microorganisms and climate change. Nat. Rev. Microbiol. 17, 569–586 (2019).Article 

    Google Scholar 
    59.Hutchins, D. A. et al. Climate change microbiology — problems and perspectives. Nat. Rev. Microbiol. 17, 391–396 (2019).Article 

    Google Scholar 
    60.Hartley, I. P., Hopkins, D. W., Garnett, M. H., Sommerkorn, M. & Wookey, P. A. Soil microbial respiration in arctic soil does not acclimate to temperature. Ecol. Lett. 11, 1092–1100 (2008).Article 

    Google Scholar 
    61.Bradford, M. A. et al. Cross-biome patterns in soil microbial respiration predictable from evolutionary theory on thermal adaptation. Nat. Ecol. Evol. 3, 223–231 (2019).Article 

    Google Scholar 
    62.Dacal, M., Bradford, M. A., Plaza, C., Maestre, F. T. & García-Palacios, P. Soil microbial respiration adapts to ambient temperature in global drylands. Nat. Ecol. Evol. 3, 232–238 (2019).Article 

    Google Scholar 
    63.Sierra, C. A., Trumbore, S. E., Davidson, E. A., Vicca, S. & Janssens, I. Sensitivity of decomposition rates of soil organic matter with respect to simultaneous changes in temperature and moisture. J. Adv. Model. Earth Syst. 7, 335–356 (2015).Article 

    Google Scholar 
    64.Birgander, J., Reischke, S., Jones, D. L. & Rousk, J. Temperature adaptation of bacterial growth and 14C-glucose mineralisation in a laboratory study. Soil Biol. Biochem. 65, 294–303 (2013).Article 

    Google Scholar 
    65.Sokol, N. W., Sanderman, J. & Bradford, M. A. Pathways of mineral-associated soil organic matter formation: Integrating the role of plant carbon source, chemistry, and point of entry. Glob. Chang. Biol. 25, 12–24 (2019).Article 

    Google Scholar 
    66.Bradford, M. A. Thermal adaptation of decomposer communities in warming soils. Front. Microbiol. 4, 333 (2013).Article 

    Google Scholar 
    67.Walker, T. W. N. et al. Microbial temperature sensitivity and biomass change explain soil carbon loss with warming. Nat. Clim. Chang. 8, 885–889 (2018).Article 

    Google Scholar 
    68.Tucker, C. L., Bell, J., Pendall, E. & Ogle, K. Does declining carbon-use efficiency explain thermal acclimation of soil respiration with warming? Glob. Chang. Biol. 19, 252–263 (2013).Article 

    Google Scholar 
    69.Takriti, M. et al. Soil organic matter quality exerts a stronger control than stoichiometry on microbial substrate use efficiency along a latitudinal transect. Soil Biol. Biochem. 121, 212–220 (2018).Article 

    Google Scholar 
    70.Frey, S. D., Lee, J., Melillo, J. M. & Six, J. The temperature response of soil microbial efficiency and its feedback to climate. Nat. Clim. Chang. 3, 395–398 (2013).Article 

    Google Scholar 
    71.Alster, C. J., von Fischer, J. C., Allison, S. D. & Treseder, K. K. Embracing a new paradigm for temperature sensitivity of soil microbes. Glob. Chang. Biol. 26, 3221–3229 (2020).Article 

    Google Scholar 
    72.Bååth, E. Temperature sensitivity of soil microbial activity modeled by the square root equation as a unifying model to differentiate between direct temperature effects and microbial community adaptation. Glob. Chang. Biol. 24, 2850–2861 (2018).Article 

    Google Scholar 
    73.Ratkowsky, D. A., Olley, J., McMeekin, T. A. & Ball, A. Relationship between temperature and growth rate of bacterial cultures. J. Bacteriol. 149, 1–5 (1982).Article 

    Google Scholar 
    74.Rinnan, R., Rousk, J., Yergeau, E., Kowalchuk, G. A. & Bååth, E. Temperature adaptation of soil bacterial communities along an Antarctic climate gradient: predicting responses to climate warming. Glob. Chang. Biol. 15, 2615–2625 (2009).Article 

    Google Scholar 
    75.Roller, B. R. K. & Schmidt, T. M. The physiology and ecological implications of efficient growth. ISME J. 9, 1481–1487 (2015).Article 

    Google Scholar 
    76.Rousk, J., Frey, S. D. & Bååth, E. Temperature adaptation of bacterial communities in experimentally warmed forest soils. Glob. Chang. Biol. 18, 3252–3258 (2012).Article 

    Google Scholar 
    77.Coleman, K. et al. Simulating trends in soil organic carbon in long-term experiments using RothC-26.3. Geoderma 81, 29–44 (1997).Article 

    Google Scholar 
    78.Treseder, K. K. et al. Integrating microbial ecology into ecosystem models: Challenges and priorities. Biogeochemistry 109, 7–18 (2012).Article 

    Google Scholar 
    79.Hagerty, S. B. et al. Accelerated microbial turnover but constant growth efficiency with warming in soil. Nat. Clim. Chang. 4, 903–906 (2014).Article 

    Google Scholar 
    80.Steinweg, J. M., Plante, A. F., Conant, R. T., Paul, E. A. & Tanaka, D. L. Patterns of substrate utilization during long-term incubations at different temperatures. Soil Biol. Biochem. 40, 2722–2728 (2008).Article 

    Google Scholar 
    81.Sinsabaugh, R. L. et al. Stoichiometry of microbial carbon use efficiency in soils. Ecol. Monogr. 86, 172–189 (2016).Article 

    Google Scholar 
    82.Ye, J.-S., Bradford, M. A., Dacal, M., Maestre, F. T. & García-Palacios, P. Increasing microbial carbon use efficiency with warming predicts soil heterotrophic respiration globally. Glob. Chang. Biol. 25, 3354–3364 (2019).Article 

    Google Scholar 
    83.Ye, J. S., Bradford, M. A., Maestre, F. T., Li, F. M. & García-Palacios, P. Compensatory thermal adaptation of soil microbial respiration rates in global croplands. Glob. Biogeochem. Cycles 34, e2019GB006507 (2020).
    Google Scholar 
    84.Schmidt, M. W. I. et al. Persistence of soil organic matter as an ecosystem property. Nature 478, 49–56 (2011).Article 

    Google Scholar 
    85.Hartley, I. P., Heinemeyer, A. & Ineson, P. Effects of three years of soil warming and shading on the rate of soil respiration: Substrate availability and not thermal acclimation mediates observed response. Glob. Chang. Biol. 13, 1761–1770 (2007).Article 

    Google Scholar 
    86.Hopkins, F. M., Torn, M. S. & Trumbore, S. E. Warming accelerates decomposition of decades-old carbon in forest soils. Proc. Natl Acad. Sci. USA 109, 1753–1751 (2012).Article 

    Google Scholar 
    87.Feng, W. et al. Enhanced decomposition of stable soil organic carbon and microbial catabolic potentials by long-term field warming. Glob. Chang. Biol. 23, 4765–4776 (2017).Article 

    Google Scholar 
    88.Chen, J. et al. Differential responses of carbon-degrading enzyme activities to warming: Implications for soil respiration. Glob. Chang. Biol. 24, 4816–4826 (2018).Article 

    Google Scholar 
    89.Anthony, M. A., Crowther, T. W., Maynard, D. S., van den Hoogen, J. & Averill, C. Distinct assembly processes and microbial communities constrain soil organic carbon formation. One Earth 2, 349–360 (2020).Article 

    Google Scholar 
    90.Lehmann, J. & Kleber, M. The contentious nature of soil organic matter. Nature 528, 60–68 (2015).Article 

    Google Scholar 
    91.Cotrufo, M. F., Ranalli, M. G., Haddix, M. L., Six, J. & Lugato, E. Soil carbon storage informed by particulate and mineral-associated organic matter. Nat. Geosci. 12, 989–994 (2019).Article 

    Google Scholar 
    92.van Gestel, N. et al. Predicting soil carbon loss with warming. Nature 554, E4–E5 (2018).Article 

    Google Scholar 
    93.Hawkes, C. V., Waring, B. G., Rocca, J. D. & Kivlin, S. N. Historical climate controls soil respiration responses to current soil moisture. Proc. Natl Acad. Sci. USA 114, 6322–6327 (2017).Article 

    Google Scholar 
    94.Ning, D. et al. A quantitative framework reveals ecological drivers of grassland microbial community assembly in response to warming. Nat. Commun. 11, 4717 (2020).Article 

    Google Scholar 
    95.Karger, D. N. et al. Climatologies at high resolution for the earth’s land surface areas. Sci. Data 4, 170122 (2017).Article 

    Google Scholar 
    96.Tifafi, M., Guenet, B. & Hatté, C. Large differences in global and regional total soil carbon stock estimates based on SoilGrids, HWSD, and NCSCD: Intercomparison and evaluation based on field data from USA, England, Wales, and France. Glob. Biogeochem. Cycles 32, 42–56 (2018).Article 

    Google Scholar 
    97.Manzoni, S., Taylor, P., Richter, A., Porporato, A. & Ågren, G. I. Environmental and stoichiometric controls on microbial carbon-use efficiency in soils. New Phytol. 196, 79–91 (2012).Article 

    Google Scholar  More

  • in

    Integrating spatial analysis and questionnaire survey to better understand human-onager conflict in Southern Iran

    Study areaQatruiyeh National Park, established in 2008, is a core zone in the Bahram-e-Goor Protected Area (established in 1972) at the border of Fars and Kerman provinces in southern Iran (Fig. 1). It covers 310 km2 and is part of the Zagros Mountains. It is a semi-desert with temperate arid climate, vegetated mainly with Zygophyllum eurypterum and Artemisia sieberi20. There are seven villages in the vicinity of the protected area, where pastoralism is the main source of livelihood21.Figure 1Location of the study area. The software ArcGIS. Version 10.2. was used to generate figure. DEM map was downloaded from the WorldClim database (http://www.worldclim.org).Full size imageOne of the major threats for the Persian onager populations in this area is increasing construction of new roads and increasing road traffic. The Sirjan-Yazd (Hassan Abad-Meshkaan) asphalt road, which passes through the Bahram-e-Goor Protected Area, was recently converted into a highway and represents a substantial threat to Onagers (Fig. 1). This road has two lanes in each direction. The day-time speed limit on this road is 110 km/h and 90 km/h at night. Most vehicles on this road are heavy trucks, which pass at high speed (more than 90 km/h), with high traffic volumes at night. During winter, late autumn and summer of drought years, when fodder is scarce, onagers frequently cross the road to access gardens and agricultural fields, which causes high onager mortality due to vehicle collisions. In this research, we used spatial randomization of vehicle collisions and crossing locations to test the predictive ability of resistant kernel and factorial least-cost path predictions of movement18. We also conducted questionnaires with residents from local communities to determine the most important factors influencing human-onager conflicts in the Bahram-e-Goor Protected Area.Human-onager conflict assessmentQualitative data collectionWe administered a questionnaire through a personal interview to 200 randomly chosen farmers residing near onager populations in the Bahram-e-Goor Protected Area in Fars province. Data were collected through a questionnaire between May and August 2018 (Table S1). Ethical clearance was obtained from the DOE (under permit 32–239). All participants were given a printed descriptive summary of the research (if participants were illiterate, the document was read to them). Prior informed consent was obtained orally from all participants. In this research, we followed legal requirements of ethical issues.We calculated the sample size needed by using the family size in rural areas around Bahram-e-Goor Protected Area using the Daniel method22 (Table S1) as described below (Eq. 1):We randomly conducted 200 questionnaires in total.$$N=frac{ {Z}^{2 }P (1-P) }{{d}^{2}}$$
    (1)
    In this equation, Z is the Z statistic for a level of confidence, P is expected prevalence or proportion (if the expected prevalence is 20%, then P = 0.2), and d is precision (if the precision is 5%, then d = 0.05). In this research, we used d = 0.5 and p was selected according to family sizes in each district of rural areas22.All interviewees were adult males. We collected information on interviewees’ demographic and socioeconomic background (occupation, property, age, and income) as well as their knowledge and opinion on how to prevent onager crop-raiding.We used logistic regression to analyze the significance of sociological factors related to crop damage by onagers. Our dependent variable was “Have you had any of your crop raided by onager during the last year? (Binary response: 1 = Yes, 0 = No)”. Our independent variables included: (1) traditional solutions for reducing Persian onager damages (Response: 1 guarding dogs, 2: fencing around agricultural land, 3: use of traditional barriers (a plastic cuff with a bell on it), 4: scarecrow, 5: turn on the lights at night , 6: Bird-Scarer (Kalaghparan in Persian); (2) which of these solutions could be effective in reducing Persian onager damages (Responses included: 1: fencing around Persian onager habitat, 2: fencing around farmland, 3: give fodder and provide water for Persian onager, 4: buying fodder from local people by DoE, 4: capturing and relocating Persian onager); (3): do you agree with Persian onager hunting? (Binary response: 1 = Yes, 0= No); (4): what is the role of the Persian onager in the wild? (Response 1: distributing seed of plants, the rangelands are restored, 2: it attracts tourists in the region, 3: beauty of nature: God’s creature with a right to live (Intrinsic value), 4: none) (5): age (response: 1:  50 Years), (6): education (response: 1: Incomplete Elementary (lower than 5th grade of elementary), 2: Complete Elementary (5th grade of elementary), 3: Incomplete High school, 4: Associate Degree, 5: Bachelor of Science (BSc), 5: Master of Science (MSc) or Higher), (7) Experience of Persian onager observation in nature: Have you ever seen a Persian onager in the wild? (Response scale: 1 = Yes, frequently, 2 = Yes, several times, 3: Yes, a few times 4: No, never, 5: only seen the Asiatic wild ass carcass), (8) the presence of a Persian onager around your village damages your farms and gardens. How do you feel about this statement? (Response scale) 1: completely disagree, 2: Somewhat disagree, 3: I do not agree or disagree, 4: I agree somewhat, 5: completely agree.All statistical tests were conducted in IBM SPSS Statistics (V. 23.0). Independent variables in the logistic regression analysis were coded as showed in Table S1.Naïve Bayes classificationNaïve Bayes Classification uses a group of simple classifiers based on probabilities, which are applicable to the types of random independent variables in our study. This approach is a supervised machine learning algorithm based on the Bayes Theorem that is used to solve classification problems by following a probabilistic approach. We used the e1071 library23 in R version 3.5.324 for Naïve Bayes classification of onager crop-raiding under this scheme. We considered: Yes (local communities with experience of crop-raid damages), or No (local communities without experience of crop-raid damages during the last one year) as a dependent variable, as a function of the independent variables described in logistic regression section, except we also included farm land area (1:  5 ha) as an additional variable.We categorized data into two groups (testing and training) to determine whether the model performed correctly based on training data. Subsequently, 70% of the data were used to test and run the model along with training confirmation. The Naïve Bayes Classifier was trained to anticipate each attitude in the test data. We calculated the randomness of our results using the Mclust library25 in R version 3.5.324.Onager vehicle collisionsA 25-km section of the 99-km Hassan Abad-Meshkaan road (the area with the highest wildlife-vehicle collision reports) was monitored by motorcycling and walking daily from August to October 2017 (3 weeks). Every morning, we inspected for mammal roadkill within a 30-m buffer on each side of the road, and all carcasses of mammals were recorded using a handheld GPS (Garmin GPS Map 62S). To avoid double-counting, we removed the carcasses after recording. We also obtained collision location data during 2004–2018 from the DoE.The crossing data for onager were obtained from a variety of sources including opportunistic direct observation, environmental guard’s information, and monitoring by LED portable flashlight at night (summer and autumn seasons of 2017 and 2018).Habitat connectivity analysisHabitat suitability modelingA total of 103 presence points were obtained from DoE (2015) in the study area, including Bahram-e-Goor Protected Area, as well as nearby surroundings. To minimize spatial autocorrelation, a 1-km radius was used to eliminate points around each presence location using the SDM toolbox26. The remaining 90 presence points were used in the modeling.A habitat suitability map for onager was developed using MaxEnt software version 3.3.3k27 to create a resistance map for connectivity modeling28. We used 10,000 pseudo-absence points29. For the training data set, 75% of the presence points were randomly chosen to train and the remaining 25% were used to test the model30. We used the area under the ROC curve (AUC) to evaluate model performance. MaxEnt models were completed with 10 bootstrapped replicates.Environmental layers included in MaxEnt modeling included (1) elevation (digital elevation model [DEM]), (2) slope, (3) land cover, (4) distance from agricultural lands, (5) distance from roads and (6) distance from villages. All layers had a 30 m × 30 m resolution (Table 1).Table 1 Environmental variables used for habitat modeling of the Persian onager in the study area.Full size tableSlope was calculated from the DEM layer. Land cover for 27 vegetation classes in the study area was reclassified to 10 classes based on similarities between classes in the original landcover map and due to the importance of agricultural lands (5% of the study area) to onagers. Distance from agricultural lands, roads and villages were included as predictor variables, and were calculated with the Euclidean distance tool in the Spatial Analyst extension of ArcGIS 10.2. We checked for multi-collinearity among variables and correlation was  3 were used as a threshold to exclude variables32. VIF ranged from 1.2 to 1.8 for all variables. Therefore, all variables were retained for habitat modeling.Resistance surface for connectivity analysisTo estimate landscape resistance, we converted the habitat suitability maps to resistance maps using a negative exponential function (R = 1000(−1×HS)) where R represents the cost resistance value assigned to each pixel and HS represents the predicted habitat suitability derived from the suitability models described above33. We used 1000 as the base of our exponential decay function such that areas with  > 0.3 habitat suitability would have low-cost resistance. We rescaled the resistance values to a range between 1 and 100 by linear interpolation, such that minimum resistance (Rmin) was 1 when HS was 1, and maximum resistance (Rmax) was 100 when HS was 033.Connectivity corridor network simulationWe used the universal corridor network simulator (UNICOR)34 to predict movement core areas and corridors for Onagers. UNICOR’s key features include a driver-module framework, connectivity mapping with thresholding and buffering, and graph theory metrics. UNICOR produces two kinds of connectivity predictions: (1) resistant kernels16 and (2) factorial least-cost paths15. The factorial least-cost path analysis implanted in UNICOR simulator uses Dijkstra’s algorithm34 to solve the single-source shortest path problem from every mapped species occurrence location on a landscape to every other occurrence location34. The analysis produces predicted least-cost path routes from each source point to each destination point. The resistant kernel algorithm calculates the resistance cost weighted dispersal kernel around each source point up to a user-defined dispersal threshold, and then sums these, producing an incidence function of the rate of organism movement through every pixel in the landscape as a function of the number and density of source points, the dispersal ability of the species, and the resistance of the landscape.According to observation and reports of experts in the DoE, the maximum dispersal of threshold for movement of Onagers is about 100 km. We thus specified a dispersal threshold of 100,000 cost units for the resistant kernel analysis35. We calculated the factorial least-cost path network without dispersal the threshold35 to provide a broad-scale assessment of the regional pattern of potential linkage and to map corridors. The buffered least-cost paths were then combined through summation15 to produce maps of connectivity among all pairs of presence points.Evaluating congruence between crossing points and predicted connectivityWe used a spatial randomization testing procedure to evaluate congruence between the locations where onagers were observed crossing the road and resistant kernel values of predicted connectivity18. Spatial randomization testing of this kind is recommended in cases where there is spatial dependence among observations, and produces an unbiased estimate of the probability of the observed outcome given the data18.We compared the median value of predicted connectivity (resistant kernel) for the 104 actual onager crossing locations with the distribution of median values of 1 × 107 random samples of 104 locations along the highway within the study area. For each combination of resistance surface and connectivity modeling approach, we calculated the ranking of the median of observed values within the distribution of the medians of the 1 × 107 random samples. More

  • in

    Major ocean currents may shape the microbiome of the topshell Phorcus sauciatus in the NE Atlantic Ocean

    1.Rees, H. C. et al. The detection of aquatic animal species using environmental DNA – a review of eDNA as a survey tool in ecology. J. Appl. Ecol. 51, 1450–1459 (2014).CAS 
    Article 

    Google Scholar 
    2.Palumbi, S. R. What can molecular genetics contribute to marine biogeography? An urchin’s tale. J. Exp. Mar. Biol. Ecol. 203, 75–92 (1996).CAS 
    Article 

    Google Scholar 
    3.Vucetich, J. A. & Waite, T. A. Spatial patterns of demography and genetic processes across the species’ range: null hypotheses for landscape conservation genetics. Conserv. Genet. 4(5), 639–645 (2003).Article 

    Google Scholar 
    4.Krishnamurthy, P. K. & Francis, R. A. A critical review on the utility of DNA barcoding in biodiversity conservation. Biodivers. Conserv. 21(8), 1901–1919 (2012).Article 

    Google Scholar 
    5.Metzker, M. L. Emerging technologies in DNA sequencing. Genome Res. 15(12), 1767–1776 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    6.Gloor, G. B. et al. Microbiome Profiling by Illumina Sequencing of Combinatorial Sequence-Tagged PCR Products. PLoS ONE 5(10), e15406 (2010).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    7.Caporaso, J. G. et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc. Natl. Acad. Sci. 108(1), 4516–4522 (2011).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    8.Douglas, A. E. Symbiosis as a general principle in eukaryotic evolution. Cold Spring Harb. Perspect. Biol. 6(2), a016113 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    9.Bolhuis, H. & Cretoiu, M. S. What is so special about marine microorganisms? Introduction to the marine microbiome—from diversity to biotechnological potential. In the Marine Microbiome (eds Stal, L. & Cretoiu, M.) (Springer, Cham, 2016).
    Google Scholar 
    10.Stal, L. & Cretoiu, M. S. (eds) The Marine Microbiome (Springer, 2016).
    Google Scholar 
    11.Huse, S. M. et al. Exploring microbial diversity and taxonomy using SSU rRNA hypervariable tag sequencing. PLoS Genet. 4(11), e1000255 (2008).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    12.Pedrós-Alió, C. The rare bacterial biosphere. Ann. Rev. Mar. Sci. 4, 449–466 (2012).PubMed 
    Article 

    Google Scholar 
    13.Thomas, T. et al. Diversity, structure and convergent evolution of the global sponge microbiome. Nat. Commun. 7, 11870 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    14.Taylor, M. W., Radax, R., Steger, D. & Wagner, M. Sponge-associated microorganisms: evolution, ecology, and biotechnological potential. Microbiol. Mol. Biol. Rev. 71(2), 295–347 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    15.Hentschel, U., Piel, J., Degnan, S. M. & Taylor, M. W. Genomic insights into the marine sponge microbiome. Nat. Rev. Microbiol. 10(9), 641–654 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    16.Cleary, D. F. R. et al. The sponge microbiome within the greater coral reef microbial metacommunity. Nat. Commun. 10, 1644 (2019).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    17.León-Palmero, E. et al. Diversity and antimicrobial potential in sea anemone and holothurian microbiomes. PLoS One 13(5), e0196178 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    18.Blasiak, L. C., Zinder, S. H., Buckley, D. H. & Russell, T. H. Bacterial diversity associated with the tunic of the model chordate Ciona intestinalis. ISME J. 8(2), 309–320 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    19.Givens, C. E. et al. Comparison of the gut microbiome of 12 bony fish and 3 shark species. Mar. Ecol. Prog. Ser. 518, 209–223 (2015).Article 
    ADS 

    Google Scholar 
    20.Neu, A. T., Allen, E. E. & Roy, K. Diversity and composition of intertidal gastropod microbiomes across a major marine biogeographic boundary. Environ. Microbiol. Rep. 11(3), 434–447 (2019).PubMed 
    Article 

    Google Scholar 
    21.Salazar, G. & Sunagawa, S. Marine microbial diversity. Curr. Biol. 27(11), 489–494 (2017).Article 
    CAS 

    Google Scholar 
    22.Dick, G. J. The microbiomes of deep-sea hydrothermal vents: distributed globally, shaped locally. Nat. Rev. Microbiol. 17(5), 271–283 (2019).CAS 
    PubMed 

    Google Scholar 
    23.Cavicchioli, R. Microbial ecology of Antarctic aquatic systems. Nat. Rev. Microbiol. 13, 691–706 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    24.Zinger, L. et al. Global patterns of bacterial beta-diversity in seafloor and seawater ecosystems. PLoS ONE 6(9), e24570 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    25.Palumbi, S. & Hedgecock, D. The life of the sea: implications of marine population biology to conservation policy. In Marine conservation biology: the science of maintaining the sea’s biodiversity (eds Norse, E. A. & Crowder, L. B.) (Island Press, Washington, 2005).
    Google Scholar 
    26.Sousa, R., Delgado, J., González, J. A., Freitas, M. & Henriques, P. Marine snails of the genus Phorcus: Biology and ecology of sentinel species for human impacts on the rocky shores. In Biological Resources of Water (ed. Ray, S.) 141–167 (Croatia, IntechOpen., 2018).
    Google Scholar 
    27.Sousa, R. et al. Filling biological information gaps of the marine topshell Phorcus sauciatus (Gastropoda: Trochidae) to ensure its sustainable exploitation. J. Mar. Biol. Assoc. U.K. 99(4), 841–849 (2019).Article 

    Google Scholar 
    28.Sousa, R. et al. Disentangling exploitation of the intertidal grazer Phorcus sauciatus (Gastropoda: Trochidae) in an ocean archipelago: Implications for conservation. Mar. Ecol. 40(2), e12540 (2019).Article 
    ADS 

    Google Scholar 
    29.Donald, K. M. et al. Phylogenetic relationships elucidate colonization patterns in the intertidal grazers Osilinus Philippi, 1847 and Phorcus Risso, 1826 (Gastropoda: Trochidae) in the northeastern Atlantic Ocean and Mediterranean Sea. Mol. Phylogenet. Evol. 62(1), 35–45 (2012).PubMed 
    Article 

    Google Scholar 
    30.Takahashi, et al. Development of prokaryotic universal primer for simultaneous analysis of Bacteria and Archaea using next-generation sequencing. PLoS ONE 9(8), e105592 (2014).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    31.Herlemann, D. P. et al. Transitions in bacterial communities along the 2000 km salinity gradient of the Baltic Sea. ISME J. 5(10), 1571–1579 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    32.QGIS Development Team. QGIS Geographic Information System (QGIS Association, 2020).33.Edgar, R. C., Haas, B. J., Clemente, J. C., Quince, C. & Knight, R. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27, 2194–2200 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    34.Weiss, S. J. et al. Effects of library size variance, sparsity, and compositionality on the analysis of microbiome data. PeerJ PrePrints 3, e1157v1 (2015).
    Google Scholar 
    35.Staley, J. T. & Bauld, J. The genus Planctomyces. In The Prokaryotes (eds Starr, M. P. et al.) (Springer, Berlin, 1981).
    Google Scholar 
    36.Nelson, L. et al. Molecular analysis of gut microflora in captive-raised sea urchins (Lytechinus variegatus). J. World Aquaculture Soc. 41(5), 807–815 (2010).Article 

    Google Scholar 
    37.Hakim, J. A. et al. An abundance of Epsilonproteobacteria revealed in the gut microbiome of the laboratory cultured sea urchin, Lytechinus variegatus. Front. Microbiol. 6, 1047 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Moisander, P. H., Sexton, A. D. & Daley, M. C. Stable associations masked by temporal variability in the marine copepod microbiome. PLoS ONE 10(9), e0138967 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    39.Motarjemi, Y., Todd, E. & Moy, G. Encyclopedia of Food Safety (Elsevier, 2014).
    Google Scholar 
    40.Zilber-Rosenberg, I. & Rosenberg, E. Role of microorganisms in the evolution of animals and plants: the hologenome theory of evolution. FEMS Microbiol. Rev. 32(5), 723–735 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    41.Lokmer, A. & Wegner, K. M. Hemolymph microbiome of Pacific oysters in response to temperature, temperature stress and infection. ISME J. 9(3), 670–682 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    42.Lokmer, A. et al. Spatial and temporal dynamics of Pacific oyster hemolymph microbiota across multiple scales. Front. Microbiol. 7, 1367 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Kellogg, C. A., Goldsmith, D. B. & Gray, M. A. Biogeographic comparison of Lophelia-associated bacterial communities in the western Atlantic reveals conserved core microbiome. Front. Microbiol. 8, 796 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Burgsdorf, I. et al. Biogeography rather than association with cyanobacteria structures symbiotic microbial communities in the marine sponge Petrosia ficiformis. Front. Microbiol. 5, 1–11 (2014).Article 

    Google Scholar 
    45.Kellogg, C. A. Microbiomes of stony and soft deep-sea corals share rare core bacteria. Microbiome 7, 90 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Hamdan, L. et al. Ocean currents shape the microbiome of Arctic marine sediments. ISME J. 7, 685–696 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    47.Donaldson, G. P., Lee, S. M. & Mazmanian, S. K. Gut biogeography of the bacterial microbiota. Nat. Rev. Microbiol. 14(1), 20–32 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    48.Xiao, Y. et al. Comparative biogeography of the gut microbiome betweeen Jinhua and Landrace pigs. Sci. Rep. 8, 5985 (2018).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    49.Morelan, et al. Microbiome variation in an intertidal sea anemone across latitudes and symbiotic states. Front. Mar. Sci. 6, 7 (2019).Article 
    ADS 

    Google Scholar 
    50.Barton, E. D. Canary and Portugal currents. In Encyclopedia of Ocean Sciences (eds Steele, J. et al.) 380–389 (Academic Press, 2001).Chapter 

    Google Scholar 
    51.Spalding, M. D. et al. Marine ecoregions of the world: a bioregionalization of coast and shelf areas. BioScience 57, 573–583 (2007).Article 

    Google Scholar 
    52.Tuya, F. et al. Phytogeography of Lusitanian Macaronesia: biogeographic affinities in species richness and assemblage composition. Eur. J. Phycol. 44(3), 405–413 (2009).Article 

    Google Scholar 
    53.Freitas, R. et al. Restructuring of the “Macaronesia” biogeographic unit: A marine multi-taxon biogeographic approach. Sci. Rep. 9, 15792 (2019).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    54.Faria, J. et al. Disentangling the genetic and morphological structure of Patella candei complex in Macaronesia (NE Atlantic). Ecol. Evol. 7(16), 6125–6140 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    55.Stramma, L. & Siedler, G. Seasonal changes in the North Atlantic subtropical gyre. J. Geophys. Res. Oceans 93(C7), 8111–8118 (1988).Article 
    ADS 

    Google Scholar 
    56.Mason, E. et al. Seasonal variability of the Canary current: a numerical study. J. Geophys. Res. 116(6), C06001 (2011).ADS 

    Google Scholar 
    57.Follows, M. J., Dutkiewicz, S., Grant, S. & Chisholm, S. W. Emergent biogeography of microbial communities in a model ocean. Science 315, 1843–1846 (2007).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    58.Galand, P. E., Potvin, M., Casamayor, E. O. & Lovejoy, C. Hydrography shapes bacterial biogeography of the deep Arctic Ocean. ISME J. 4, 564–576 (2010).PubMed 
    Article 

    Google Scholar 
    59.Cicala, F., Cisterna-Céliz, J. A., Moore, J. D. & Rocha-Olivares, A. Structure, dynamics and predicted functional role of the gut microbiota of the blue (Haliotis fulgens) and yellow (H. corrugata) abalone from Baja California Sur. Mexico. PeerJ. 6, e5830 (2018).PubMed 
    Article 
    CAS 

    Google Scholar 
    60.Gutiérrez-Díaz, I., Fernández-Navarro, T., Sánchez, B., Margolles, A. & González, S. Mediterranean diet and faecal microbiota: a transversal study. Food Funct. 7(5), 2347–2356 (2016).PubMed 
    Article 
    CAS 

    Google Scholar 
    61.Mueller, U. G. & Sachs, J. L. Engineering microbiomes to improve plant and animal health. Trends Microbiol. 23(10), 606–617 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    62.Jost, L., Chao, A. & Chazdon, R. L. Compositional similarity and beta diversity. In Biological Diversity Frontiers in Measurement and Assessment (eds Magurran, A. E. & McGill, B. J.) (Oxford University Press, Oxford, 2011).
    Google Scholar  More

  • in

    Tapping local knowledge to save a Papua New Guinea forest

    Download PDF

    It takes collaboration to get the full picture of a forest. Here, I’m teaming up with Sammy, a local senior-school student, to count and identify ants in a dense fragment of lowland rainforest. It’s November 2019, when I was a research technician with the New Guinea Binatang Research Centre in Madang, and we’re near the village of Boredoa on the southern coast of Papua New Guinea.I’m impressed by the locals’ knowledge of the forest and its inhabitants. It’s important to get more villagers involved in forest surveys and other conservation efforts so that they can work to protect them.Papua New Guinea — a country that makes up the eastern half of the island of New Guinea — is home to one of the world’s largest and most biodiverse rainforests, but mining and timber companies are taking a terrible toll. The areas beyond these trees have been heavily logged, and we’re checking to see how life in this remaining forest is faring, from the ants to the trees.I grew up in the northern city of Lae, a place very different from this forest. I have formal training in forestry and entomology, but, unlike Sammy and other villagers, I don’t have the experiences and insights that come from a lifetime of living on the land.Our ant survey was part of the National Forest Inventory, a project of the United Nations Programme on Reducing Emissions from Deforestation and Forest Degradation. There are many places still waiting to be studied.This forest is a hot, wet, challenging place to work. Villagers helped us to find relatively dry places to set up our tents. We had to wait for a break in the rain to set out our ant traps of tuna and fruit-flavoured drinks. In this particular sample, we identified six species, all native to the area. Introduced species such as fire ants and army ants have been taking over elsewhere in Papua New Guinea, but the local ants here have managed to hold on to their territory. For now.

    Nature 594, 466 (2021)
    doi: https://doi.org/10.1038/d41586-021-01587-7

    Latest on:

    Careers

    The authorship rows that sour scientific collaborations
    Career Feature 14 JUN 21

    Six mentoring tips as we enter year two of COVID
    Career Column 11 JUN 21

    Researchers’ career insecurity needs attention and reform now, says international coalition
    Career News 08 JUN 21

    Ecology

    Indigenous lands: make Brazil stop mining to secure US deal
    Correspondence 08 JUN 21

    Widespread deoxygenation of temperate lakes
    Article 02 JUN 21

    French vote for river barriers defies biodiversity strategy
    Correspondence 01 JUN 21

    Jobs from Nature Careers

    All jobs

    POST-DOCTORAL ASSOCIATE BIOCHEMISTRY/CHEMICAL BIOLOGY
    University of Illinois at Urbana-Champaign (UIUC)
    Urbana, IL, United States

    JOB POST

    Postdoctoral Training Fellow
    Institute of Cancer Research (ICR)
    London, United Kingdom

    JOB POST

    Associate Editor or Senior Editor, BMC Series
    Springer Nature
    London, United Kingdom

    JOB POST

    Director of the Biomedical Computer Vision Facility
    Harvard Medical School (HMS)
    Boston, MA, United States

    JOB POST

    Nature Briefing
    An essential round-up of science news, opinion and analysis, delivered to your inbox every weekday.

    Email address

    Yes! Sign me up to receive the daily Nature Briefing email. I agree my information will be processed in accordance with the Nature and Springer Nature Limited Privacy Policy.

    Sign up More

  • in

    Social-ecological filters drive the functional diversity of beetles in homegardens of campesinos and migrants in the southern Andes

    1.Berkes, F. & Folke, C. Linking social and ecological systems for resilience and sustainability. in Linking social and ecological systems: management practices and social mechanisms for building resilience (eds. Berkes, F. & Folke, C.) 1–25 (Cambridge University Press, 2002).2.Pretty, J. et al. The intersections of biological diversity and cultural diversity: Towards integration. Conserv. Soc. 7, 100 (2009).Article 

    Google Scholar 
    3.IPBES. The regional assessment report on biodiversity and ecosystem services for the Americas (Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, 2018).4.Galluzzi, G., Eyzaguirre, P. & Negri, V. Home gardens: Neglected hotspots of agro-biodiversity and cultural diversity. Biodivers. Conserv. 19, 3635–3654 (2010).Article 

    Google Scholar 
    5.Fernandes, E. C. M. & Nair, P. K. R. An evaluation of the structure and function of tropical homegardens. Agric. Syst. 21, 279–310 (1986).Article 

    Google Scholar 
    6.Ibarra, J. T., Caviedes, J., Barreau, A. & Pessa, N. Huertas familiares y comunitarias: cultivando soberanía alimentaria (Ediciones Universidad Católica de Chile, 2019).7.Eyzaguirre, P. B. & Linares, O. F. Home Gardens and Agrobiodiversity (Smithsonian Institution Press, 2010).8.Timsuksai, P. & Rambo, A. T. The influence of culture on agroecosystem structure: a comparison of the spatial patterns of homegardens of different ethnic groups in Thailand and Vietnam. PLoS ONE 11, e0146118 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    9.Lemessa, D., Hambäck, P. A. & Hylander, K. The effect of local and landscape level land-use composition on predatory arthropods in a tropical agricultural landscape. Landsc. Ecol. 30, 167–180 (2015).Article 

    Google Scholar 
    10.Mattsson, E., Ostwald, M., Nissanka, S. P. & Pushpakumara, D. K. N. G. Quantification of carbon stock and tree diversity of homegardens in a dry zone area of Moneragala district, Sri Lanka. Agrofor. Syst. 89, 435–445 (2015).Article 

    Google Scholar 
    11.Mohri, H. et al. Assessment of ecosystem services in homegarden systems in Indonesia, Sri Lanka, and Vietnam. Ecosyst. Serv. 5, 124–136 (2013).Article 

    Google Scholar 
    12.Pakeman, R. J. & Stockan, J. A. Drivers of carabid functional diversity: abiotic environment, plant functional traits, or plant functional diversity?. Ecology 95, 1213–1224 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Altieri, M. A. Agroecology: The Science of Sustainable Agriculture (Westview Press, 1995).14.Ellis, E. C. & Ramankutty, N. Putting people in the map: Anthropogenic biomes of the world. Front. Ecol. Environ. 6, 439–447 (2008).Article 

    Google Scholar 
    15.Piccini, I. et al. Dung beetles as drivers of ecosystem multifunctionality: Are response and effect traits interwoven?. Sci. Total Environ. 616–617, 1440–1448 (2018).ADS 
    PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    16.Boonstra, W. J., Björkvik, E., Haider, L. J. & Masterson, V. Human responses to social-ecological traps. Sustain. Sci. 11, 877–889 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Cadotte, M. W., Carscadden, K. & Mirotchnick, N. Beyond species: functional diversity and the maintenance of ecological processes and services. J. Appl. Ecol. 48, 1079–1087 (2011).Article 

    Google Scholar 
    18.Kraft, N. J. B. et al. Community assembly, coexistence and the environmental filtering metaphor. Funct. Ecol. 29, 592–599 (2015).Article 

    Google Scholar 
    19.Toledo-Hernández, M., Denmead, L. H., Clough, Y., Raffiudin, R. & Tscharntke, T. Cultural homegarden management practices mediate arthropod communities in Indonesia. J. Insect Conserv. 20, 373–382 (2016).Article 

    Google Scholar 
    20.Jaganmohan, M., Vailshery, L. S. & Nagendra, H. Patterns of insect abundance and distribution in urban domestic gardens in Bangalore, India. Diversity 5, 767–778 (2013).Article 

    Google Scholar 
    21.Huerta, E. & Van der Wal, H. Soil macroinvertebrates’ abundance and diversity in home gardens in Tabasco, Mexico, vary with soil texture, organic matter and vegetation cover. Eur. J. Soil Biol. 50, 68–75 (2012).Article 

    Google Scholar 
    22.Pizzolotto, R. et al. Ground beetles in Mediterranean olive agroecosystems: their significance and functional role as bioindicators (Coleoptera, Carabidae). PLoS ONE 13, e0194551 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    23.Grez, A. A., Zaviezo, T., Casanoves, F., Oberti, R. & Pliscoff, P. The positive association between natural vegetation, native coccinellids and functional diversity of aphidophagous coccinellid communities in alfalfa. Insect Conserv. Divers. https://doi.org/10.1111/icad.12473 (2021).Article 

    Google Scholar 
    24.Villéger, S., Mason, N. W. H. & Mouillot, D. New multidimensional functional diversity indices for a multifaceted framework in functional ecology. Ecology 89, 2290–2301 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Guerrero, I., Carmona, C. P., Morales, M. B., Oñate, J. J. & Peco, B. Non-linear responses of functional diversity and redundancy to agricultural intensification at the field scale in Mediterranean arable plant communities. Agric. Ecosyst. Environ. 195, 36–43 (2014).Article 

    Google Scholar 
    26.Mayfield, M. M. et al. What does species richness tell us about functional trait diversity? Predictions and evidence for responses of species and functional trait diversity to land-use change. Glob. Ecol. Biogeogr. 19, 423–431 (2010).
    Google Scholar 
    27.Upreti, B. R. & Upreti, Y. G. Factors leading to agro-biodiversity loss in developing countries: the case of Nepal. Biodivers. Conserv. 11, 1607–1621 (2002).Article 

    Google Scholar 
    28.Reyes-García, V. et al. Resilience of traditional knowledge systems: The case of agricultural knowledge in home gardens of the Iberian Peninsula. Glob. Environ. Chang. 24, 223–231 (2014).Article 

    Google Scholar 
    29.Kawa, N. C. How religion, race, and the weedy agency of plants shape Amazonian home gardens. Cult. Agric. Food Environ. 38, 84–93 (2016).Article 

    Google Scholar 
    30.Brondizio, E. S. et al. Re-conceptualizing the Anthropocene: A call for collaboration. Glob. Environ. Chang. 39, 318–327 (2016).Article 

    Google Scholar 
    31.Benson, M. & O’Reilly, K. Lifestyle Migration: Expectations, Aspirations, and Experiences (Ashgate Publishing, 2009).32.Marchant, C. Lifestyle migration and the nascent agroecological movement in the Andean Araucanía, Chile: Is it promoting sustainable local development?. Mt. Res. Dev. 37, 406–414 (2017).Article 

    Google Scholar 
    33.Ibarra, J. T., Barreau, A., Caviedes, J., Pessa, N. & Urra, R. Huertas familiares tradicionales y emergentes: cultivando biodiversidad, aprendizaje y soberanía desde la interculturalidad. in Huertas familiares y comunitarias: cultivando soberanía alimentaria (eds. Ibarra, J. T., Caviedes, J., Barreau, A. & Pessa, N.) 138–165 (Ediciones Universidad Católica de Chile, 2019).34.Myers, N., Mittermeier, R. A., Mittermeier, C. G., da Fonseca, G. A. B. & Kent, J. Biodiversity hotspots for conservation priorities. Nature 403, 853–858 (2000).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Arroyo, M. T. K. et al. El hotspot chileno, prioridad mundial para la conservación. in Diversidad de Chile: patrimonios y desafíos (ed. Mnisterio del Medio Ambiente, G. de C.) 90–95 (Ocho Libros Editores, 2006).36.Farias, A. A. & Jaksic, F. M. Low functional richness and redundancy of a predator assemblage in native forest fragments of Chiloe Island Chile. J. Anim. Ecol. 80, 809–817 (2011).PubMed 
    Article 

    Google Scholar 
    37.Ibarra, J. T. & Martin, K. Biotic homogenization: loss of avian functional richness and habitat specialists in disturbed Andean temperate forests. Biol. Conserv. 192, 418–427 (2015).Article 

    Google Scholar 
    38.Lavelle, P. et al. Soil function in a changing world: The role of invertebrate ecosystem engineers. Eur. J. Soil Biol. 33, 159–193 (1997).CAS 

    Google Scholar 
    39.Cole, L. J. et al. Relationships between agricultural management and ecological groups of ground beetles (Coleoptera: Carabidae) on Scottish farmland. Agric. Ecosyst. Environ. 93, 323–336 (2002).Article 

    Google Scholar 
    40.Van Klink, R. et al. Meta-analysis reveals declines in terrestrial but increases in freshwater insect abundances. Science 368, 417–420 (2020).ADS 
    PubMed 
    Article 
    CAS 

    Google Scholar 
    41.Sánchez-Bayo, F. & Wyckhuys, K. A. G. Worldwide decline of the entomofauna: a review of its drivers. Biol. Conserv. 232, 8–27 (2019).Article 

    Google Scholar 
    42.Lencinas, M. V., Sola, F. J., Cellini, J. M., Peri, P. L. & Martínez Pastur, G. Land sharing in South Patagonia: Conservation of above-ground beetle diversity in forests and non-forest ecosystems. Sci. Total Environ. 690, 132–139 (2019).43.Roig-Juñent, S. & Domínguez, M. C. Diversidad de la familia Carabidae (Coleoptera) en Chile. Rev. Chil. Hist. Nat. 74, 549–571 (2001).Article 

    Google Scholar 
    44.Grez, A. A., Moreno, P. & Elgueta, M. Coleópteros (Insecta: Coleoptera) epígeos asociados al bosque maulino y plantaciones de pino aledañas. Rev. Chil. Entomol. 29, 9–18 (2003).
    Google Scholar 
    45.Richardson, B. J. & Arias-Bohart, E. T. Why so many apparently rare beetles in Chilean temperate rainforests?. Rev. Chil. Hist. Nat. 84, 419–432 (2011).Article 

    Google Scholar 
    46.Cifuentes-Croquevielle, C., Stanton, D. E. & Armesto, J. J. Soil invertebrate diversity loss and functional changes in temperate forest soils replaced by exotic pine plantations. Sci. Rep. 10, 7762 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Fonseca, C. R. & Ganade, G. Species functional redundancy, random extinctions and the stability of ecosystems. J. Ecol. 89, 118–125 (2001).Article 

    Google Scholar 
    48.Rosenfeld, J. S. Functional redundancy in ecology and conservation. Oikos 98, 156–162 (2013).Article 

    Google Scholar 
    49.Petchey, O. L. & Gaston, K. J. Functional diversity (FD), species richness and community composition. Ecol. Lett. 5, 402–411 (2002).Article 

    Google Scholar 
    50.Mori, A. S. Resilience in the studies of biodiversity-ecosystem functioning. Trends Ecol. Evol. 31, 87–89 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.Ibarra, J. T. et al. Nurturing resilient forest biodiversity: nest webs as complex adaptive systems. Ecol. Soc. 25, 27 (2020).Article 

    Google Scholar 
    52.Ibarra, J. T., Martin, M., Cockle, K. L. & Martin, K. Maintaining ecosystem resilience: functional responses of tree cavity nesters to logging in temperate forests of the Americas. Sci. Rep. 7, 4467 (2017).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    53.Elgueta, M. & Arriagada, G. Estado actual del conocimiento de los coleópteros de Chile (Insecta: Coleoptera). Rev. Chil. Entomol. 17, 5–60 (1989).
    Google Scholar 
    54.Díaz, S. & Cabido, M. Vive la différence: plant functional diversity matters to ecosystem processes. Trends Ecol. Evol. 16, 646–655 (2001).Article 

    Google Scholar 
    55.Petchey, O. L., Evans, K. L., Fishburn, I. S. & Gaston, K. J. Low functional diversity and no redundancy in British avian assemblages. J. Anim. Ecol. 76, 977–985 (2007).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    56.Villagrán, C. & Hinojosa, L. F. Historia de los bosques del sur de Sudamérica, II : análisis fitogeográfico. Rev. Chil. Hist. Nat. 70, 241–267 (1997).
    Google Scholar 
    57.Vuilleumier, F. & Simpson, B. Pleistocene changes in the fauna and flora of South America. Science 173, 771–780 (1971).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Niemelä, J. Habitat distribution of carabid beetles in Tierra del Fuego South America. Entomol. Fenn. 29, 3–16 (1990).Article 

    Google Scholar 
    59.O’Brien, C. The biogeography of Chile through entomofaunal regions. Entomol. News 82, 197–202 (1971).
    Google Scholar 
    60.Vergara, O. E., Jerez, V. & Parra, L. E. Diversidad y patrones de distribución de coleópteros en la Región del Biobío, Chile : una aproximación preliminar para la conservación de la diversidad. Rev. Chil. Hist. Nat. 79, 369–388 (2006).Article 

    Google Scholar 
    61.Mason, N. W. H., Irz, P., Lanoiselée, C., Mouillot, D. & Argillier, C. Evidence that niche specialization explains species-energy relationships in lake fish communities. J. Anim. Ecol. 77, 285–296 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    62.Clavel, J., Julliard, R. & Devictor, V. Worldwide decline of specialist species: toward a global functional homogenization?. Front. Ecol. Environ. 9, 222–228 (2011).Article 

    Google Scholar 
    63.Devictor, V. et al. Spatial mismatch and congruence between taxonomic, phylogenetic and functional diversity: the need for integrative conservation strategies in a changing world. Ecol. Lett. 13, 1030–1040 (2010).PubMed 
    PubMed Central 

    Google Scholar 
    64.Trinh, L. N. et al. Agrobiodiversity conservation and development in Vietnamese home gardens. Agric. Ecosyst. Environ. 97, 317–344 (2003).Article 

    Google Scholar 
    65.MacArthur, R. H. & Wilson, E. O. The Theory of Island Biogeography (Princeton University Press, 1967).66.Serge, M. M. P., Giovani, E. T. & Mony, R. Household and home garden infesting arthropods (Ants and Myriapods) in the city of Yaoundé, Cameroon. J. Entomol. Zool. Stud. 7, 1030–1037 (2019).
    Google Scholar 
    67.Jacquet, C., Mouillot, D., Kulbicki, M. & Gravel, D. Extensions of island biogeography theory predict the scaling of functional trait composition with habitat area and isolation. Ecol. Lett. 20, 135–146 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    68.Gravel, D., Massol, F., Canard, E., Mouillot, D. & Mouquet, N. Trophic theory of island biogeography. Ecol. Lett. 14, 1010–1016 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    69.Regman, T. P. et al. Species interactions regulate the collapse of biodiversity and ecosystem function in tropical forest fragments. Ecology 96, 2692–2704 (2015).Article 

    Google Scholar 
    70.Bolger, D. T., Suarez, A. V., Crooks, K. R., Morrison, S. A. & Case, T. J. Arthropods in urban habitat fragments in southern California: area, age and edge effects. Ecol. Appl. 10, 1230–1248 (2000).Article 

    Google Scholar 
    71.Barreau, A., Ibarra, J. T., Wyndham, F. S. & Kozak, R. A. Shifts in Mapuche food systems in southern Andean forest landscapes: historical processes and current trends of biocultural homogenization. Mt. Res. Dev. 39, 12–23 (2019).Article 

    Google Scholar 
    72.Caviedes, J. & Ibarra, J. T. Influence of anthropogenic disturbances on stand structural complexity in Andean temperate forests: implications for managing key habitat for biodiversity. PLoS ONE 12, e0169450 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    73.Altieri, M. A. & Nicholls, C. I. The adaptation and mitigation potential of traditional agriculture in a changing climate. Clim. Change 140, 33–45 (2017).ADS 
    Article 

    Google Scholar 
    74.Sánchez-Bayo, F. Impacts of agricultural pesticides on terrestrial ecosystems. in Ecological Impacts of Toxic Chemicals (eds. Sánchez-Bayo, F., Van den Brink, P. J. & Mann, R.) 63–87 (Bentham Science Publishers, 2011).75.Geiger, F. et al. Persistent negative effects of pesticides on biodiversity and biological control potential on European farmland. Basic Appl. Ecol. 11, 97–105 (2010).CAS 
    Article 

    Google Scholar 
    76.Barreau, A., Ibarra, J. T., Wyndham, F. S., Rojas, A. & Kozak, R. A. How can we teach our children if we cannot access the forest? Generational change in Mapuche knowledge of wild edible plants in Andean temperate ecosystems of Chile. J. Ethnobiol. 36, 412–432 (2016).Article 

    Google Scholar 
    77.Newing, H. Conducting research in conservation: a social science perspective. (Routledge, 2011). https://doi.org/10.1007/s13398-014-0173-7.278.Caballero-Serrano, V. et al. Plant diversity and ecosystem services in Amazonian homegardens of Ecuador. Agric. Ecosyst. Environ. 225, 116–125 (2016).Article 

    Google Scholar 
    79.Schneider, J. Toward an analysis of home-garden cultures: on the use of socio-cultural variables in home garden studies. in Home gardens and agrobiodiversity (eds. Eyzaguirre, P. B. & Linares, O. F.) 41–55 (Smithsonian Books, 2010).80.Rohr, J. R., Mahan, C. G. & Kim, K. C. Developing a monitoring program for invertebrates: guidelines and a case study. Conserv. Biol. 21, 422–433 (2007).PubMed 
    Article 

    Google Scholar 
    81.Gotelli, N. J. & Colwell, R. K. Quantifying biodiversity: Procedures and pitfalls in the measurement and comparison of species richness. Ecol. Lett. 4, 379–391 (2001).Article 

    Google Scholar 
    82.Iida, T., Soga, M., Hiura, T. & Koike, S. Life history traits predict insect species responses to large herbivore overabundance: a multitaxonomic approach. J. Insect Conserv. 20, 295–304 (2016).Article 

    Google Scholar 
    83.Vanderwel, M. C., Malcolm, J. R., Smith, S. M. & Islam, N. Insect community composition and trophic guild structure in decaying logs from eastern Canadian pine-dominated forests. For. Ecol. Manage. 225, 190–199 (2006).Article 

    Google Scholar 
    84.Zarazaga, M. A. Clase Insecta Orden Coleoptera. Rev. IDE-SEA 56, 1–18 (2015).
    Google Scholar 
    85.Lazo, W. Insectos de Chile: atlas entomológico. (Universidad de Chile, 2015).86.Briones, R., Gárate-Flores, F. & Jerez, V. Insectos de Chile. Nativos, introducidos y con problemas de conservacion. (Corporación Chilena de la Madera, 2012).87.Elgueta, M. & Arriagada, G. Estado actual del conocimiento de los coleópteros de Chile (Insecta: Coleoptera). Rev. Chil. Entomol. 17, 05–60 (1989).
    Google Scholar 
    88.Elgueta, M. & Marvaldi, A. E. Lista sistemática de las especies de curculionoidea (insecta: coleoptera) presentes en Chile, con su sinonimia. Boletín del Mus. Nac. Hist. Nat. 55, 113–153 (2006).
    Google Scholar 
    89.Moore, T. & Vidal, P. Los Bupréstidos de Chile. (Ediciones UC, 2013).90.Roig-Juñent, S. & Domínguez, M. C. Diversity of the family Carabidae (Coleoptera) in Chile. Rev. Chil. Hist. Nat. 74, 549–571 (2001).Article 

    Google Scholar 
    91.Arriagada, G. Histéridos chilenos (Coleoptera: Histeridae). Rev. Chil. Entomol. 14, 71–80 (1986).
    Google Scholar 
    92.González, G. Lista y distribución geográfica de especies de Coccinelidae (Insecta: Coleoptera) presentes en Chile. Boletín del Mus. Nac. Hist. Nat. 57, 77–107 (2008).
    Google Scholar 
    93.Lister, B. C. & Garcia, A. Climate-driven declines in arthropod abundance restructure a rainforest food web. Proc. Natl. Acad. Sci. 115, E10397–E10406 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    94.Johnson, M. D. & Strong, A. M. Length-weight relationships of Jamaican arthropods. Entomol. News 111, 270–281 (2000).
    Google Scholar 
    95.Laliberté, E., Legendre, P. & Shipley, B. FD: measuring functional diversity (FD) from multiple traits, and other tools for functional ecology. (2011).96.Zuur, A., Leno, E. N., Walker, N. J., Saveliev, A. A. & Smith, G. M. Mixed effects models and extensions in ecology with R. Statistics for Biology and Health 36, (Springer, 2009).97.Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1:48 (2015).98.Mazerolle, M. J. AICcmodavg: model selection and multimodel inference based on (Q)AIC(c). R Packag. version 2.1–1 (2017).99.R Core Team. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/. (2021).100.Burnham, K. P. & Anderson, D. R. Model selection and inference: a practical information-theoretic approach. (Springer-Verlag, 2002).101.Oliver, M. A. & Webster, R. Kriging: a method of interpolation for geographical information systems. Int. J. Geogr. Inf. Syst. 4, 313–332 (1990).Article 

    Google Scholar  More