Zum Hauptinhalt springen

Biodiversity and ecosystem functions depend on environmental conditions and resources rather than the geodiversity of a tropical biodiversity hotspot.

Wallis, CIB ; Tiede, YC ; et al.
In: Scientific reports, Jg. 11 (2021-12-31), Heft 1, S. 24530
Online academicJournal

Biodiversity and ecosystem functions depend on environmental conditions and resources rather than the geodiversity of a tropical biodiversity hotspot 

Biodiversity and ecosystem functions are highly threatened by global change. It has been proposed that geodiversity can be used as an easy-to-measure surrogate of biodiversity to guide conservation management. However, so far, there is mixed evidence to what extent geodiversity can predict biodiversity and ecosystem functions at the regional scale relevant for conservation planning. Here, we analyse how geodiversity computed as a compound index is suited to predict the diversity of four taxa and associated ecosystem functions in a tropical mountain hotspot of biodiversity and compare the results with the predictive power of environmental conditions and resources (climate, habitat, soil). We show that combinations of these environmental variables better explain species diversity and ecosystem functions than a geodiversity index and identified climate variables as more important predictors than habitat and soil variables, although the best predictors differ between taxa and functions. We conclude that a compound geodiversity index cannot be used as a single surrogate predictor for species diversity and ecosystem functions in tropical mountain rain forest ecosystems and is thus little suited to facilitate conservation management at the regional scale. Instead, both the selection and the combination of environmental variables are essential to guide conservation efforts to safeguard biodiversity and ecosystem functions.

These authors contributed equally: Christine I. B. Wallis and Yvonne C. Tiede.

Introduction

Safeguarding biodiversity[1]–[3] and ecosystem functions under global change[4],[5] is of key importance for the provision of essential services to human societies[6]. Geodiversity indices (i.e., the diversity of earth system's abiotic and structural features) are increasingly used to predict biodiversity and ecosystem functions[1],[2],[7]. A comprehensive review on how the term geodiversity evolved and diversified over time is given in ref.[8]. Early concepts of geodiversity proposed the conservation of a broad range of environments to maintain biological diversity[9], in particular, considering geological features of a region (Supplementary Table S1). These concepts were later extended by a variety of geomorphological, hydrological and pedological features (Supplementary Table S1, see also ref.[10]). To describe geodiversity, simple feature numbers were initially used but later, inspired by biology, replaced by metrics such as the Shannon diversity index[11]–[13].

Using geodiversity as a surrogate for biodiversity burgeoned around the new millennium[13], mainly to identify areas for conservation or natural heritage or to manage the exploitation of ecosystem services[12],[14]–[16]. This endeavour was based on the hypothesis that specific geo-sites should support unique biota so that high geodiversity is coupled with high biodiversity[17]–[19]. This concept has been particularly addressed by the Conserving Nature's Stage (CNS) approach which focusses on the use of geodiversity in conservation planning[9],[18],[19]. Because geodiversity was expected to be more resilient against environmental changes than biodiversity, it was claimed that preserving geodiversity should safeguard biodiversity and ecosystem functions and services[20]–[22]. Consequently, geodiversity data, in parts easily available through remote sensing, were proposed as suitable planning tools to predict biodiversity and ecosystem functions, particularly in the frequent case of biodiversity data unavailability[23]–[25].

To date, different types of geodiversity indices are used to predict biodiversity pattern[8]. They range from simple feature-based to complex multivariate indices such as geodiversity indices that consider climate, hydrology, soil, topography and habitat, and their variability in space and time (Supplementary Table S1, see also ref.[23]). Our understanding of the concept of geodiversity is in line with the resource-based measure of geodiversity explained in ref.[23] (also discussed in ref.[8]). In the following, we will refer to geodiversity as the spatial variation of environmental variables. The environmental variables (also elsewhere termed as abiotic conditions, abiotic variables or abiotic stages[26]) used in this study consider conditions and resources related to climate, habitat and soil which are requirements for the establishment and survival of organisms and have been used as meaningful predictors of species diversity and ecosystem functions (Supplementary Methods). These variables are here used as a direct (the value measured at a plot) or indirect (the diversity of multiple variables at a plot and its surrounding; hereafter the compound geodiversity index) surrogate for species diversity and ecosystem functions (see Methods and Supplementary Methods). Our view on using the environmental variables to calculate a geodiversity compound index is supported by a review on the CNS approach that recommends to consider a broader spectrum of environmental variables at a finer thematic resolution[27]. At the same time, a harmonization of geodiversity approaches has recently been claimed to reach intercomparability of studies focussing on the geodiversity-biodiversity relationship, and further test this relationship for underrepresented ecosystems (such as tropical mountain rain forests)[7].

While biodiversity is defined by the wealth and variety of life, depending on the scale and the level of complexity, a range of measures can be applied to surrogate it, i.e., the diversity of ecosystems, communities, or taxa[28],[29]. So far, most biodiversity studies used measures of species diversity such as species richness, or abundance-weighted diversity metrics, e.g. the Shannon index[30]. This is often due to the simplicity of these indices and the ease with which they can be compared within and across ecosystems. As taxonomic species diversity alone, however, misses much of the overall diversity of an ecosystem[31], measures of ecosystem functions can be additionally addressed to describe the state of an ecosystem. Ecosystem functions are often defined as the biological, physical, and geochemical processes that occur within an ecosystem and consider the sizes of compartments (such as carbon storage) as well as the rates of their changes (such as rates of carbon sequestration)[32]. They can provide or regulate ecosystem services[32] and are, therefore, an integral part of the biodiversity of ecosystems[32].

Using the Shannon index for the calculation of geodiversity and species diversity could thus facilitate the direct comparisons of geodiversity and species diversity. But research on geodiversity-biodiversity relations showed different performance depending on selected ecosystem, taxa, and scale, where geodiversity predictors seemed to generally work better for intermediate to large scales[33]–[35]. CNS research activities, however, revealed best performance of geo-biodiversity at landscape to regional scales that are most important for conservation planning[26]. Also challenging is often the match of geo- and biodiversity scales. It is important that the environmental variables used for the calculation of a geodiversity compound index are drivers of the species diversity of the taxon under investigation. Since the strength of environmental variables as predictors differs between different taxa, it is unlikely that there is a geodiversity index which can be used as a surrogate of species diversity across multiple taxa. Therefore, the choice of environmental variables to calculate taxon-specific geodiversity indices is of great importance. Not only the environmental variables but also the scales used to calculate the geodiversity compound index should have ecological significance for the taxa and ecosystem functions studied[35].

Here, we test the hypothesis that geodiversity compound indices are powerful predictors of species diversity and ecosystem functions in a tropical mountain rainforest hotspot for biodiversity. Specifically, we compare predictive models of species diversity and ecosystem functions based on a geodiversity index with those based on environmental variables. In this study we use the following definitions: (i) Environmental variables consider the conditions and resources at each plot. (ii) The geodiversity index is assembled as the sum of the spatial diversity of multiple environmental variables measured within each plot and its surrounding (see Methods and Supplementary Fig. S1). For both, (i) and (ii) environmental variables belong to the three groups climate, habitat, and soil, each comprising a set of different environmental conditions and resources available as spatial grids (Table 1). We consider one condition or resource from each group and subsequently use the selected three environmental variables as predictors of species diversity and ecosystem functions (Extended Data Table E3). For the geodiversity compound index, all spatial grids are classified, and the central pixel of each plot and its adjacent pixels are used to calculate their spatial diversity analogous to the calculation of Shannon diversity (Supplementary Fig. S1). The geodiversity index is subsequently calculated by summing up the diversities of the selected environmental variables indicating the degree of variation across space. Comparison of the geodiversity index models with models of the environmental variables has been conducted for the following taxa: trees, testate amoebae, ants and birds, and their associated ecosystem functions: carbon sequestration, decomposition, predation, and seed dispersal, along an elevational gradient ranging from 1000 to 3000 m a.s.l.

Table 1 Potential predictors considering environmental conditions and resources within climate, habitat, and soil variables.

Environmental variables

Raster source layer/spatial resolution

Calculation/references

Climate

Based on averaged monthly air temperature in C°: annual mean (mean), annual maximum (max), annual standard deviation (sd)

Landsat-8 scene, 30 m spatial resolution

After ref.80

Humidity in %: mean annual

Landsat-8 scene, 30 m spatial resolution

After ref.81

Habitat

Normalized difference vegetation index (NDVI)

Landsat-8 scene, 30 m spatial resolution

(near infrared − red)/(near infrared + red)

Image textural metric: NDVI correlation

Landsat-8 scene, 30 m spatial resolution

'glcm' package in R; using all directions82,83

Red-Blue normalized difference vegetation index (RBVI)

Landsat-8 scene, 30 m spatial resolution

(red − blue)/(red + blue)

Forest cover in %

Landsat-8 scene, 30 m spatial resolution

Spectral unmixing in IDRISI Andes after92

Topographical position index (TPI)

Sigtierras image, 6 m spatial resolution

After ref.93 with a surrounding of 17 pixels

Leaf area index (LAI)

Sentinel-2 scene, 10 m spatial resolution

S2 SNAP (Sentinel Application Platform) Toolbox Biophysical Processor94

Soil

pH in Bv-Horizon

30 m spatial resolution

Model prediction; Extended Data Tables E1, E2

Total phosphorus soil in kg/ha in Ah-Horizon (phosphorus)

30 m spatial resolution

Model prediction; Extended Data Tables E1, E2

Organic layer depth in cm

30 m spatial resolution

Model prediction; Extended Data Tables E1, E2

Raster source means the underlying raster data that were used to determine the environmental variables in space.

Results and discussion

Geodiversity and environmental variables

First, we compared the extent to which the geodiversity index and environmental variables were associated with species diversity and ecosystem functions using GAMs (generalized additive model, Fig. 1, see also Extended Data Figs. E1, E2, Extended Data Tables E3, E4, E5). Contrary to the expectation, we found that the geodiversity index explains only a very limited amount of the variation of species diversity and ecosystem functions along the elevational gradient. In contrast, models based on selected environmental conditions and resources explain most of the variation in species diversity and associated ecosystem functions (Fig. 1).

Graph: Figure 1 Explained deviance of species diversity (left) and ecosystem functions (right) by a geodiversity index as well as environmental variables. Models using environmental variables are based on three predictors accounting for conditions and resources of the three groups climate, habitat, and soil, while models using the geodiversity index consider the summed spatial diversities of the same three predictors. The predictor sets are selected individually for each response (depicted in Fig. 2) and are hold equal for both models based on the geodiversity index as well as environmental variables (see also Extended Data Figs. E1 and E2).

Our findings contradict the results of previous studies which successfully linked geo- and biodiversity[35],[36]. On closer inspection however, only a few of the previous studies rely on geodiversity indices only (Supplementary Table S1). Most of them used mixed sets of predictors comprising both, geodiversity features and indices, and environmental variables considering different conditions and resources (e.g., climate, hydrology, soil, habitat, topography). For instance, ref.[10] combined geodiversity indices with levels of potential radiation (climate) and NDVI (Normalized Difference Vegetation Index; habitat) to explain biodiversity. In many studies that used such a mixed set of predictors, environmental conditions and resources had higher predictive power than geodiversity indices to explain biodiversity across taxa and scales[5],[34],[37].

A reason for the relatively poor performance of geodiversity indices might be the information loss by aggregating fine-grain resource data into lower resolution compound metrics. Different predictors with reversed patterns might cancel each other out in geodiversity compound indices[1]. Secondly, scale mismatches might impact the predictive performance of geodiversity. In general, the combination of environmental variables should match the targeted measurement scales at which the diversities of taxa and ecosystem functions have been sampled. In our study, these measurement scales varied depending on the different taxonomic groups studied and the associated ecosystem functions due to differences in activity ranges among the taxa (Methods). On the other hand, the used environmental data varied in their spatial resolution, mainly due to the fact that we used mostly freely available data at a certain spatial resolution (Table 1) which is a common practice in geo-biodiversity studies[38]. While there is only a small mismatch in the case of models using environmental variables (e.g., comparing tree Shannon diversity measured within an area of 400 m2 with the values of environmental predictors at 30 m spatial resolution), the mismatch increases when calculating spatial diversity measures from these environmental variables. In detail, the spatial scale of the environmental data in our study is in a resolution of 6–30 m and for the calculation of geodiversity the diversity measures are calculated in a 3 pixel × 3 pixel environment (320–8100 m2) which differ from the measurement scales of the investigated taxa and ecosystem functions (400–10,000 m2). In addition, this limitation resulted in geodiversity indices that in some cases combined the spatial variation of environmental variables calculated within different areas. This is a limitation of our study and needs to be addressed in future research as it is a common practice in geodiversity studies to combine remote sensing products such as digital elevation models at only moderate spatial resolution with local field data at finer resolution[35],[39]. But as stated in ref.[26] smaller mismatches between geo- and biodiversity can be tolerated and, furthermore, the knowledge that the plots have been chosen to be representative of their surrounding environment also counteracts this limitation.

Despite these limitations and recommendations, our findings suggest that geodiversity computed as a single compound index is not able to model species diversity and ecosystem functions in the here studied highly diverse tropical mountain ecosystem. This is in line with the findings in ref.[38], that geodiversity indices better work as a surrogate for common than rare species, which often occur in biodiversity hotspot areas. By contrast, we showed that environmental variables are fundamental to conservation planning, as they imply that the availability of specific environmental conditions and resources, rather than surrogate geodiversity indices, are required to accurately predict species diversity and ecosystem functions.

Relationships of environmental variables with species diversity and ecosystem functions

The selected environmental predictors accounting for climate, habitat, and soil variables revealed broad-scale relationships with species diversity for most taxa (Fig. 2a–d, lower panels) and ecosystem functions (Fig. 2e–h, lower panels). In the univariate models, climate conditions explained the patterns in species diversity and ecosystem functions on average better than habitat and soil predictors. In addition, variation partitioning of multivariate models showed that the most important predictors considering climate, habitat and soil differed among taxa and ecosystem functions (Fig. 2, upper panels). Conditions of air temperature were the main driver of animal diversity and ecosystem functions, whereas habitat conditions gained in relevance for plants and soil-associated taxa and functions. While tree diversity was best explained by habitat but with little explanatory support, C-sequestration was strongly related to climate, but at high shared explained variance with soil and habitat predictors. The diversity of the least mobile taxon, testate amoebae, was best explained by habitat. An increase in the diversity of testate amoebae was positively related to an increasing forest coverage which is associated with lower climatic stress[40]. Less important predictors for the diversity of testate amoebae were soil P as a key food resource[40], and humidity[41]. Litter decomposition was best explained by climate, and in particular an increase in mean annual temperature (MAT). Litter decomposition is known to be related to the soil microbiome[42] whose activity is positively related to MAT[43]. An increase in ant diversity and ant predation was best explained by climate, i.e., a low seasonal variability in air temperature. Ant diversity is generally positively related to enhanced temperatures[44],[45]. Because ants with specifically narrow optimum temperature ranges are mostly occurring in the tropics, they particularly profit from low temperature fluctuations[46]. Bird diversity and seed dispersal by birds were positively related to MAT. This relationship corroborates findings on the global scale[47]. The importance of MAT for seed dispersal is similar to the effect on bird diversity itself, which points to more dispersal of seeds under increased temperatures. Furthermore, bird diversity showed a hump-shaped relationship to habitat, i.e., a spatial texture metric of the NDVI, confirming that species diversity is positively related to the number of niches in heterogeneous habitats[48]. The relationships between species richness and ecosystem functions with environmental conditions and resources found in this study are largely consistent with the patterns found in other tropical mountain systems[49],[50] such as Kilimanjaro[51].

Graph: Figure 2 Relative importance of the three environmental variables accounting for conditions and resources from climate, habitat, and soil in explaining species diversity of four taxonomic groups (a–d) and related ecosystem functions (e–h). Upper panels: We conducted variation partitioning (VP) based on generalized additive models (GAM). Venn diagrams show the pure proportions explained exclusively by each predictor, and their share of explained deviance (expl. dev.) due to overlapping effects of pairs of predictors or all three predictors. A negative value indicates that the explained deviance by the predictor overlap is less than that explained separately by the predictors[62]. Lower panels: Dots present the response of the three selected predictors (same as in the upper panel) against Shannon taxon diversity as a measure of species diversity and the associated ecosystem functions. Lines depict the respective univariate GAM smooth functions with the corresponding explained deviance indicated. NDVI: Normalized Difference Vegetation Index, sd: standard deviation of air temperature and air humidity over time. The reader may refer to the Methods section and the Supplementary Methods for the individual selection procedure of the three predictors for each response. The plant and animal images used in the figure were provided by Cordula Mann and Christiane Enderle.

Relevance of scale, identity, and complementarity of geodiversity for biodiversity and ecosys...

So far, most evidence for a general relationship between geodiversity, biodiversity and ecosystem functions is based on global analyses[2] (but see also refs.[34],[37]). Based on these studies, geodiversity has been proposed to be in principle a useful surrogate of biodiversity for conservation planning on the global scale[4],[37]. Yet, our results revealed that species diversity and ecosystem functioning cannot be inferred from a geodiversity index on the regional spatial scale within a biodiversity hotspot. Instead, we found that the combination of individually selected environmental conditions and resources was much more indicative and thus, is needed to reveal associations between biodiversity and its environment on regional spatial scales. Geodiversity approaches that work well on a global scale are apparently of relatively little value for conservation management that must be implemented on the regional scale.

Our findings are analogous to those from biodiversity-ecosystem function (BEF) research where trait values are often more relevant than trait diversity to understand the role of biodiversity for ecosystem functioning at regional scales[52]–[54]. Similarly, a geodiversity index does not hold as an all-in-one solution to predict species diversity and ecosystem functioning. Instead, the individual environmental conditions and resources matter—and they differ between the different components of biodiversity and ecosystem functions. Both, trait identity and complementarity are needed to predict ecosystem functions from biodiversity[52]–[54]. This analogy suggests that research linking geo- with biodiversity and ecosystem functions might go through a similar progression of conceptual and methodological refinement as BEF research. Consequently, it must be also acknowledged, that from a conservation perspective, the here targeted species diversity is always a simplification of overall species composition. Species composition, and thus changes in species identity, are therefore often more informative and thus relevant to the conservation of biodiversity and functions[55]–[58]. Since the Shannon index is limited in surrogating the biotic and abiotic composition of ecosystems, we suggest that future studies should test geodiversity-biodiversity relationships focusing on the biotic and abiotic composition of environments instead of their diversity. This could be realised, i.e. using techniques of ordination to link the composition of taxa and their environment in a distance-based framework[58]–[61].

Conclusion

Our study identifies four key research gaps that future geo-biodiversity approaches should address and test with comparative studies. First, geodiversity indices seem to be better predictors of biodiversity in simple than in complex ecosystems given the selection and availability of key conditions or resources could be more important than the composition of different conditions and resources for biodiversity and functionality in these systems[34],[37]. Second, the relevance of spatial scale, i.e., at larger spatial and temporal scales geodiversity indices might be more relevant than the identity of specific environmental variables given that it provides higher flexibility depending on the components of biodiversity and functions that are studied[63]. Third, the need to consider species identity in geodiversity studies, e.g., by testing the relationship between species and environmental composition instead of diversity. Fourth, the tailored selection of environmental conditions and resources is important as these affect the extent to which biodiversity and ecosystem functions can be predicted[34],[37]. Our study provides new insights into how combinations of environmental variables, rather than geodiversity indices, shape species diversity and ecosystem functions at regional scales. At such spatial scales, both the selection and the combination of environmental variables can increase our mechanistic understanding and support conservation management of biodiversity and ecosystem functions.

Methods

Study area

We exemplify our analysis on the mountain rainforest of south-eastern Ecuador that belongs to the biodiversity hotspot of the tropical Andes which is globally most important for conservation support[64]. We use data of a unique ecosystem, the mountain rainforest of the eastern Andean slopes[65]. This is essential to avoid confounding effects from multiple biogeographic units, as it is the case in most other studies on the relation of geo- and biodiversity. The field data were collected between 2007 and 2017 along an elevational gradient between 1000 and 3000 m a.s.l. (Fig. 3) with study sites equally distributed along the gradient[66]–[68].

Graph: Figure 3 Location of the study area in the southern part of Ecuador (left) and its three main elevational levels from 3000 to 1000 m a.s.l. (white boxes from left to right). Map created in QGIS 3.14.16-Pi (https://qgis.org/). Background imagery (right): Landsat-8 true-color composite (in courtesy of USGS, https://earthexplorer.usgs.gov/).

Species diversity and ecosystem functions

We assessed the species diversity of four taxa and four associated ecosystem functions using standardized methods. The details of the sampling procedure of each taxon and function have mostly been described in their original publications. We sampled species diversity data of trees[40],[69], testate amoebae[70],[71], ants[72] and birds[73]. For ants, species abundance data from two field campaigns were pooled and for birds, nine point count records per plot were pooled to obtain the bird species abundance of all species and individuals per plot. Species abundance for trees and testate amoebae was measured based on distinct samples. To compare the Shannon diversity of the four taxa in a standardized way, we used a rarefaction and extrapolation framework based on sample coverage, implemented in the R package iNEXT[74],[75]. In this framework, it is statistically valid to extrapolate the observed Shannon diversity values to the maximal sample coverage (the asymptote of the rarefaction and extrapolation curve). We therefore set the extrapolation ratio for the calculation of the standardized species diversity values to one, corresponding to a complete sample coverage for all taxa[76].

Four important ecosystem functions that are associated with the four taxa were measured in the study area. C-sequestration was observed as the aboveground net primary production[77], deduced from annual wood production and annual fine litter production (tree diversity/C-sequestration, partial overlap of plots, Pearson r = 0.35, p < 0.01, n = 52). The decomposition of soil organic matter was measured using a Tea Bag experiment[78] reflecting the weight loss of tea bags after 21 days (testate amoebae diversity/decomposition, distinct plots). Predation was assessed using the number of bite marks of ants on artificial caterpillars[72] (ant diversity/predation, Pearson r = 0.40 p < 0.05, n = 27). Seed dispersal was measured as the number of observed seed removal events of frugivorous birds on fleshy fruited plants[79] (bird diversity/seed dispersal, Pearson r = 0.62, p < 0.01, n = 15).

Environmental variables

In this study we investigated a set of 13 environmental variables based on conditions and resources within the three groups climate, habitat and soil (Table 1). We did not include geology which has often been used in other geodiversity research studies due to the geological homogeneity of our study area[66]. The predictors of the three groups were available as spatial datasets or were spatially predicted with varying spatial resolution (see Table 1). The following predictors were considered for the three groups:

Climate: We used climate stations in combination with a Landsat-8 image classification (recorded on 20/11/2016, see also ref.[77]) to derive gridded forest temperature and humidity data at a spatial resolution of 30 m following previous approaches[80],[81]. We calculated four climatic predictors: the annual mean, maximum and standard deviation of mean monthly air temperatures over the year as well as the mean annual relative air humidity.

Habitat: Based on the pre-processed Landsat-8 scene, we calculated the Normalized Difference Vegetation Index (NDVI) and its textural metric 'correlation' using all directions in the 'glcm' package[82] in R as well as the Red-Blue Difference Vegetation index (RBVI, see refs.[77],[83]) and the forest cover per pixel in percentage. These habitat variables are known to be related to habitat structure and productivity in the study area[77],[83]. In addition, we used the Sentinel-2 product Leaf Area Index (LAI) available at 10 m spatial resolution. To characterize topography, we calculated the Topographical Position Index (TPI) based on an airborne high-spatial-resolution digital elevation model from the SIGTIERRAS campaign[83].

Soil: Due to the lack of spatial soil data products at a moderate spatial resolution, we modeled and spatially predicted a suite of soil variables sampled in the study area[84]. These soil variables were related to phosphorus and nitrogen availability and pH (in H2O) in the mineral soil horizon. We used multi-linear models with a stepwise predictor selection (forward and backward) using spatial predictors (Extended Data Table E1). Only three models explained more than 45% in organic layer depth, Phosphorus content in Ah-horizon and pH in Bv-horizon. The explained variance of their models ranged between 45 and 69% (Extended Data Table E2). Their spatial predictions were subsequently used as soil predictors.

Environmental variables and geodiversity indices

Our study targets to compare models of species diversity and ecosystem functions using combinations of environmental predictors grouped within climate, habitat, and soil variables with models using a single geodiversity compound index which is defined as the summed spatial diversity of the same selected environmental variables. For this purpose, from each group (climate, habitat, and soil) one condition or resource has been selected as predictor. The choice of a predictor within each group (e.g., temperature, rainfall etc. out of climate) was guided by the findings of prior studies (Extended Data Table E3, Supplementary Methods) and the resulting combination of environmental variables, thus, comprises the requirements that the corresponding organisms need for their establishment and survival. A second criterion was the collinearity present in the predictor space. We changed the predictor combination if the correlation between two or more predictors were greater than r = 0.6. Whenever present, different statistics such as mean, maximum, and standard deviation of the climate group have been compared and the environmental variables with the best performance concerning the Akaike information criterion of the corresponding model fit was chosen. For all response variables, the pixel values of the selected environmental variables were used that overlapped with either the point or shape geometries of the corresponding plots. The resulting predictor set, thus, assess the central tendency of three environmental factors at each plot.

The geodiversity index, in contrast, is computed as the summed spatial diversity of the same three environmental factors measured within each plot and its surrounding and follows the recent plea to consistently use diversity indices commonly applied in biodiversity studies[7]. For this purpose, we classified the spatial grids of the three selected predictors into five classes using intervals following the Fisher style[85] in the 'classInt' R package[86]. After the classification, the central pixel of each plot and in addition all adjacent pixels were extracted. This resulted in nine pixels per plot. For these pixels, the frequency of all present classes was calculated. From an ecological perspective each class was used analogous to a present species at the corresponding plot and their frequency was used as their abundance. For the resulting abundance matrix of the three predictors, Shannon diversity was calculated using the diversity function of the 'vegan' package[87],[88]. This results in three predictors addressing the spatial variation of the chosen environmental variables. As in some ecological studies where the diversity of multiple taxa has been summed or averaged to study multi-taxa diversity[51], we here treated each of the three predictors like a single taxon. Thus, analogously to multi-taxa approaches, we calculated the geodiversity index by summing up the spatial diversity of the three environmental variables (for further explanation see also Supplementary Fig. S1).

Statistical analyses

We used generalized additive regression models (GAMs) to model species diversity, and ecosystem functions. To address the (non-)linear relationships between the index of geodiversity, species diversity and ecosystem functions, univariate GAMs were used. Multivariate GAMs were used to address the relationships to the three environmental variables considering conditions and resources of the three groups climate, habitat, and soil. For all models, GAMs were calibrated using a Gaussian error distribution in the model fitting based on the exploration of residuals and the generalized cross-validation score of the models (GCVS). Non-linear relationships were considered using cubic regression splines with a degree of freedom permitted to vary between one and three. GAMs however maintained the flexibility to model linear relationships wherever present. We performed all GAM analyses using the 'mgcv' package in R[89]. The model performance of combined environmental variables and the geodiversity index among all models was compared (Fig. 1).

For the combination of environmental variables, we further explored the driving conditions and resources within climate, habitat and soil variables considering their pure and shared proportions in explaining species diversity and ecosystem functions. For that purpose, we used Variance Partitioning (VP) implemented in the 'modEvA' R package[90] in combination with GAMs. All environmental variable combinations from single to three predictors were considered, resulting in 7 different GAMs (3 models with single predictors, three models with two predictors and one model with three predictors). Model parameters have been fixed as in the model using all three climate, habitat, and soil predictors to enhance model comparability.

Extended data

See Extended Data Figs. E1, E2 and Tables E1, E2, E3, E4 and E5.

Graph: Extended Data Figure E1 Effects of the geodiversity index on species diversity and ecosystem functions. The index of geodiversity is defined as the sum of Shannon diversity of three selected environmental predictors considering climate, habitat, and soil (see Supplementary Fig. S1). For each sampling plot, fitted (x-axis) versus observed (y-axis) values of species diversity of four taxa (trees, testate amoebae, ants, and birds; top) and related ecosystems functions (C-sequestration, decomposition, predation, and seed dispersal; bottom) are represented derived from generalized additive models (GAMs). The dashed line indicates the 1:1 line. The goodness of fit is depicted by the adjusted R2 and the explained deviance. Legend is shown in Fig. 2. The plant and animal images used in the figure were provided by Cordula Mann and Christiane Enderle.

Graph: Extended Data Figure E2 Effects of three selected environmental variables on taxon Shannon diversity and ecosystem functions. The three predictors account for environmental conditions and resources of the three groups climate, habitat, and soil. Further explanation is given in Extended Data Fig. E1 and Fig. 2. The plant and animal images used in the figure were provided by Cordula Mann and Christiane Enderle.

Extended Data Table E1 Spatial predictors to derive the three soil variables (shown in Table 1).

Predictor of soil variables

Source

Calculation/references

Digital elevation model (DEM)

Aster image

USGS

Slope

Aster image

Based on DEM, in degree

Topographical position index (TPI)

Aster image

After ref.93, with a surrounding of 11 pixels

Mean temperature

Weather stations, Landsat classification

After ref.44

Canopy water content (Canopy water)

Sentinel 2

S2 SNAP (Sentinel Application Platform) Toolbox Biophysical Processor54

Foliar N content (Foliar N)

Landsat-8

From ref.43

The spatial predictors were used to model the soil variables and derive prediction maps. Equations shown in Extended Data Table E2.

Extended Data Table E2 Multi-linear model equations with stepwise predictor selection and model fit to derive the three soil variables of Table 1.

Predicted soil variables

Model equation

Adjusted R2

Organic layer depth in cm (organic layer depth, n = 53)

OLD ~ FoliarN + Canopy water + Slope + TPI

0.45

Phosphorus content in Ah-Horizon in kg/ha (Phosphorus; n = 43)

Phosphorus ~ FoliarN + Canopy water + DEM

0.69

pH in Bv-Horizon (pH; n = 50)

pH ~ Mean temperature + TPI

0.67

All initial predictors are shown in Extended Data Table E1. The number of soil samples is given by n in brackets. All predictors were resampled to 30 m spatial resolution.

Extended Data Table E3 Selected environmental variables used as predictors in generalized additive models for modeling species diversity for four taxa and associated ecosystem functions.

Climate

Habitat

Soil

References

Predictor

Mean

Sd

Predictor

Mean

Sd

Predictor

Mean

Sd

Shannon diversity

Trees (n = 67)

Max temperature

19.30

3.07

LAI

2.38

0.34

Phosphorus

137.98

72.59

95

Testate amoebae (n = 24)

Sd humidity

1.43

1.59

Forest cover/pixel

0.94

0.04

Phosphorus

106.47

74.9

96

Ants (n = 27)

Sd temperature

0.60

0.11

TPI

0.31

0.71

pH

4.24

0.52

45,83

Birds (n = 15)

Mean temperature

15.89

3.94

NDVI correlation

34.17

23.91

Phosphorus

147.92

75.58

73,83

Ecosystem functions

C-sequestration (n = 54)

Max temperature

0.63

0.14

LAI

2.43

0.36

Phosphorus

135.45

71.86

40

Decomposition (n = 27)

Mean temperature

15.48

3.19

NDVI

0.84

0.04

Organic layer depth

10.55

6.82

43

Predation (n = 27)

Sd temperature

0.60

0.11

TPI

0.31

0.71

Phosphorus

132.94

70.05

72

Seed dispersal (n = 15)

Mean temperature

15.89

3.94

RBVI

0.20

0.07

Phosphorus

147.92

75.58

79

Predictors were grouped within climate, habitat, and soil conditions and resources, respectively. References highlighting their use as predictors of species diversity are indicated (see also Supplementary Methods). The number of samples of each response is given by n in brackets.

Extended Data Table E4 Generalized additive model results using a geodiversity index as a single predictor.

Response

Parametric coefficients (Intercept)

Approximate significance of smooth terms

Model fit

Estimate

Std. Error

t value

Pr(>|t|)

edf

Ref.df

F

p-value

EDF total

adjR2

Exp. Dev

Shannon diversity

Trees

33.42

3.1

10.77

0

1

1

1.88

0.18

2

0.01

0.03

Testate amoebae

26.07

1.07

24.33

0

1

1

2.27

0.15

2

0.05

0.09

Ants

12.23

1.81

6.74

0

1

1

1.95

0.17

2

0.04

0.07

Birds

33.17

1.88

17.65

0

1.4

1.64

0.34

0.72

2.4

-0.02

0.08

Ecosystem functions

C-sequestration

7.29

0.35

20.52

0

1.85

1.98

2.75

0.07

2.85

0.08

0.11

Decomposition

52.07

1.13

46.11

0

1

1

4.91

0.04

2

0.13

0.16

Predation

0.10

0.01

8.64

0

1

1

0.67

0.42

2

-0.01

0.03

Seed dispersal

359.40

93.99

3.82

0

1

1

1.75

0.21

2

0.05

0.12

A Gaussian family was used for all models. Exp. Dev. = explained deviance.

Extended Data Table E5 Generalized additive model results using selected environmental variables as predictors.

Response

Parametric coefficients (intercept)

Approximate significance of smooth terms (predictors)

Model fit

Est

Std. error

t value

Pr (>|t|)

Climate

P-value climate

Habitat

P-value habitat

Soil

P-value soil

EDF total

adjR2

Exp. Dev

Shannon diversity

Trees

33.42

2.90

11.52

0

Max temperature

0.20

LAI

0.02

Phosphorus

0.20

4.27

0.14

0.18

Testate amoebae

26.07

0.73

35.69

0

Sd humidity

0.03

Forest cover/pixel

0.00

Phosphorus

0.16

5.78

0.56

0.65

Ants

12.23

0.91

13.46

0

Sd temperature

0.00

TPI

0.02

pH

0.01

4.86

0.76

0.79

Birds

33.17

0.85

39.19

0

Mean temperature

0.00

NDVI correlation

0.03

Phosphorus

0.44

5.59

0.79

0.86

Ecosystem functions

C-sequestration

7.29

0.21

34.74

0

Max temperature

0.00

LAI

0.53

Phosphorus

0.00

4.67

0.68

0.70

Decomposition

52.07

0.54

96.12

0

Mean temperature

0.00

NDVI

0.08

Organic layer depth

0.00

5.74

0.80

0.84

Predation

0.10

0.01

11.86

0

Sd temperature

0.00

TPI

0.02

Phosphorus

0.32

4.02

0.46

0.53

Seed dispersal

359.40

33.14

10.84

0

Mean temperature

0.00

RBVI

0.01

Phosphorus

0.00

6.94

0.88

0.93

A Gaussian family was used for all models. Est. = Estimate, Exp. Dev. = explained deviance.

Acknowledgements

We thank V. Krashevska, M. Quitián, V. Santillán, J. Schlautmann, N. Cumbicus and numerous interns for field sampling and Mark Vellend for constructive criticism on the manuscript. We also thank two anonymous reviewers for their thoughtful comments and efforts towards improving our manuscript. This work was funded by the German Research Foundation in the scope of the research units FOR402, 816 and 2730 as well as the knowledge transfer program PAK823-825. We thank the Ministerio del Ambiente y Aqua and the Instituto Nacional de Biodiversidad de Ecuador (INABIO) for granting research permits and the foundation "Nature and Culture International" for providing research facilities.

Author contributions

J.B., N.F., Y.T. and C.W. conceptualized the idea of this manuscript. J.H., K.M., Y.T., C.W. collected data. Y.T. and C.W. prepared the ecological data and C.W. prepared the environmental data and ran the models and analyses with notable contributions of Y.T., N.F. and J.B. J.B., N.F., Y.T. and C.W. wrote the first manuscript draft and finalized the manuscript. J.B., N.F., K.B.-G., R.B., D.D., A.F., J.H., C.L., M.M., K.M, E.L.N., S.S., M.S., B.T. made substantial contributions collecting the underlaying data. E.B., K.B.-G., R.B., J.H., E.L.N., C.L., M.S., S.S. contributed to the conceptual development of the work. C.E., D.I., J.S. and all other authors commented on the manuscript.

Funding

Open Access funding enabled and organized by Projekt DEAL.

Data availability

Data on species abundance of taxonomic groups and ecosystem functions will be made available in the data warehouse[91] under http://www.TropicalMountainForest.org.

Code availability

The R code and the datasets generated and analysed within the current study are available from the corresponding authors on reasonable request.

Competing interests

The authors declare no competing interests.

Supplementary Information

Graph: Supplementary Information.

Supplementary Information

The online version contains supplementary material available at https://doi.org/10.1038/s41598-021-03488-1.

Publisher's note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References 1 Muellner-Riehl AN. Origins of global mountain plant biodiversity: Testing the 'mountain-geobiodiversity hypothesis'. J. Biogeogr. 2019; 46: 2826-2838 2 Antonelli A. Geological and climatic influences on mountain biodiversity. Nat. Geosci. 2018; 11: 718-725. 2018NatGe.11.718A. 1:CAS:528:DC%2BC1cXhvVGntbbI 3 Schrodt F. Opinion: To advance sustainable stewardship, we must document not only biodiversity but geodiversity. Proc. Natl. Acad. Sci. 2019; 116: 16155-16158. 2019PNAS.11616155S. 1:CAS:528:DC%2BC1MXhsFylsLnE. 31409742. 6697798 4 Alahuhta J. The role of geodiversity in providing ecosystem services at broad scales. Ecol. Indic. 2018; 91: 47-56 5 Read QD. Beyond counts and averages: Relating geodiversity to dimensions of biodiversity. Glob. Ecol. Biogeogr. 2020; 29: 696-710 6 Díaz S. Pervasive human-driven decline of life on Earth points to the need for transformative change. Science. 2019; 366: eaax3100. 31831642 7 Alahuhta J, Toivanen M, Hjort J. Geodiversity–biodiversity relationship needs more empirical evidence. Nat. Ecol. Evol. 2020; 4: 2-3. 31745262 8 Boothroyd A, McHenry M. Old processes, new movements: the inclusion of geodiversity in biological and ecological discourse. Diversity. 2019; 11: 216 9 Hunter ML, Jacobson GL, Webb T. Paleoecology and the coarse-filter approach to maintaining biological diversity. Conserv. Biol. 1988; 2: 375-385 Hjort J, Luoto M. Can geodiversity be predicted from space?. Geomorphology. 2012; 153–154: 74-80. 2012Geomo.153.74H Benito-Calvo A, Pérez-González A, Magri O, Meza P. Assessing regional geodiversity: the Iberian Peninsula. Earth Surf. Process. Landf. 2009; 34: 1433-1445. 2009ESPL.34.1433B dos Santos FM, de La Corte Bacci D, Saad AR, da Silva Ferreira AT. Geodiversity index weighted by multivariate statistical analysis. Appl. Geomat. 2020; 12: 361-370 Crisp JR, Ellison JC, Fischer A. Current trends and future directions in quantitative geodiversity assessment. Prog. Phys. Geogr. Earth Environ. 2020. 10.1177/0309133320967219 Pereira DI, Pereira P, Brilha J, Santos L. Geodiversity assessment of Paraná State (Brazil): An innovative approach. Environ. Manag. 2013; 52: 541-552. 2013EnMan.52.541P Gray M. Geodiversity and geoconservation: What, why, and how?. George Wright Forum. 2005; 22: 4-12 Ruban DA. Quantification of geodiversity and its loss. Proc. Geol. Assoc. 2010; 121: 326-333 Hjort J, Gordon JE, Gray M, Hunter ML. Why geodiversity matters in valuing nature's stage: Why geodiversity matters. Conserv. Biol. 2015; 29: 630-639. 25923307 Beier P, Brost B. Use of land facets to plan for climate change: Conserving the arenas, not the actors. Conserv. Biol. J. Soc. Conserv. Biol. 2010; 24: 701-710 Anderson MG, Ferree CE. Conserving the stage: Climate change and the geophysical underpinnings of species diversity. PLoS ONE. 2010; 5: e11554. 2010PLoSO.511554A. 20644646. 2904386 Knudson C, Kay K, Fisher S. Appraising geodiversity and cultural diversity approaches to building resilience through conservation. Nat. Clim. Change. 2018; 8: 678-685. 2018NatCC.8.678K Turner JAvon Haaren C, Lovett AA, Albert C. Geodiversity: The natural support system of ecosystems. Landscape Planning with Ecosystem Services: Theories and Methods for Application in Europe 253–265. 2019; Springer. 10.1007/978-94-024-1681-7_16 Fox N, Graham LJ, Eigenbrod F, Bullock JM, Parks KE. Incorporating geodiversity in ecosystem service decisions. Ecosyst. People. 2020; 16: 151-159 Parks KE, Mulligan M. On the relationship between a resource based measure of geodiversity and broad scale biodiversity patterns. Biodivers. Conserv. 2010; 19: 2751-2766 Comer PJ. Incorporating geodiversity into conservation decisions: Geodiversity and conservation decisions. Conserv. Biol. 2015; 29: 692-701. 25923052 Chakraborty A, Gray M. A call for mainstreaming geodiversity in nature conservation research and praxis. J. Nat. Conserv. 2020; 56: 125862 Lawler J. The theory behind, and the challenges of, conserving nature's stage in a time of rapid change. Conserv. Biol. 2015; 29: 618-629. 25922899 Beier P. A review of selection-based tests of abiotic surrogates for species representation. Conserv. Biol. J. Soc. Conserv. Biol. 2015; 29: 668-679 Purvis A, Hector A. Getting the Measure of Biodiversity. Nature. 2000; 405: 212-219. 1:CAS:528:DC%2BD3cXjsFyjs7w%3D. 10821281 Moreno C. Measuring biodiversity in the Anthropocene: A simple guide to helpful methods. Biodivers. Conserv. 2017; 26: 2993-2998 Roswell M, Dushoff J, Winfree R. A conceptual guide to measuring species diversity. Oikos. 2021; 130: 321-338 Chiarucci A, Bacaro G, Scheiner SM. Old and new challenges in using species diversity for assessing biodiversity. Philos. Trans. R. Soc. B Biol. Sci. 2011; 366: 2426-2437 Hooper DU. Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecol. Monogr. 2005; 75: 3-35 Hjort J, Heikkinen RK, Luoto M. Inclusion of explicit measures of geodiversity improve biodiversity models in a boreal landscape. Biodivers. Conserv. 2012; 21: 3487-3506 Bailey JJ, Boyd DS, Hjort J, Lavers CP, Field R. Modelling native and alien vascular plant species richness: At which scales is geodiversity most relevant?. Glob. Ecol. Biogeogr. 2017; 26: 763-776 Zarnetske PL. Towards connecting biodiversity and geodiversity across scales with satellite remote sensing. Glob. Ecol. Biogeogr. 2019; 28: 548-556. 31217748. 6559161 Bétard F. Patch-scale relationships between geodiversity and biodiversity in hard rock quarries: Case study from a disused quartzite quarry in NW France. Geoheritage. 2013; 5: 59-71 Tukiainen H. Spatial relationship between biodiversity and geodiversity across a gradient of land-use intensity in high-latitude landscapes. Landsc. Ecol. 2017; 32: 1049-1063 Anderson MG. Case studies of conservation plans that incorporate geodiversity. Conserv. Biol. 2015; 29: 680-691. 1:STN:280:DC%2BC2MjoslKnsw%3D%3D. 25924074 Ren Y, Lü Y, Hu J, Yin L. Geodiversity underpins biodiversity but the relations can be complex: Implications from two biodiversity proxies. Glob. Ecol. Conserv. 2021; 31: e01830 Homeier J, Breckle S-W, Günter S, Rollenbeck RT, Leuschner C. Tree diversity, forest structure and productivity along altitudinal and topographical gradients in a species-rich Ecuadorian montane rain forest: Ecuadorian Montane forest diversity and structure. Biotropica. 2010; 42: 140-148 Krashevska V, Bonkowski M, Maraun M, Scheu S. Testate amoebae (protista) of an elevational gradient in the tropical mountain rain forest of Ecuador. Pedobiologia. 2007; 51: 319-331 Zhalnina K. Soil pH determines microbial diversity and composition in the park grass experiment. Microb. Ecol. 2015; 69: 395-406. 1:CAS:528:DC%2BC2cXhvFKlt7jK. 25395291 Fierer N, Craine JM, McLauchlan K, Schimel JP. Litter quality and the temperature sensiticity of decomposition. Ecology. 2005; 86: 320-326 Gibb H. Climate mediates the effects of disturbance on ant assemblage structure. Proc. R. Soc. B Biol. Sci. 2015; 282: 20150418 Sanders NJ, Lessard J-P, Fitzpatrick MC, Dunn RR. Temperature, but not productivity or geometry, predicts elevational diversity gradients in ants across spatial grains. Glob. Ecol. Biogeogr. 2007; 16: 640-649 Paaijmans KP. Temperature variation makes ectotherms more sensitive to climate change. Glob. Change Biol. 2013; 19: 2373-2380. 2013GCBio.19.2373P McCain CM. Global analysis of bird elevational diversity. Glob. Ecol. Biogeogr. 2009; 18: 346-360 Tews J. Animal species diversity driven by habitat heterogeneity/diversity: The importance of keystone structures: Animal species diversity driven by habitat heterogeneity. J. Biogeogr. 2004; 31: 79-92 Rahbek C. Humboldt's enigma: What causes global patterns of mountain biodiversity?. Science. 2019; 365: 1108-1113. 2019Sci.365.1108R. 1:CAS:528:DC%2BC1MXhslGksbfE. 31515383 Hofhansl F. Climatic and edaphic controls over tropical forest diversity and vegetation carbon storage. Sci. Rep. 2020; 10: 5066. 2020NatSR.10.5066H. 1:CAS:528:DC%2BB3cXlvFCqtLs%3D. 32193471. 7081197 Peters MK. Predictors of elevational biodiversity gradients change from single taxa to the multi-taxa community level. Nat. Commun. 2016; 7: 13736. 2016NatCo.713736P. 1:CAS:528:DC%2BC28XitFGisLjP. 28004657. 5192166 Gagic V. Functional identity and diversity of animals predict ecosystem functioning better than species-based indices. Proc. R. Soc. B Biol. Sci. 2015; 282: 20142620 Kraft NJB, Godoy O, Levine JM. Plant functional traits and the multidimensional nature of species coexistence. Proc. Natl. Acad. Sci. 2015; 112: 797-802. 2015PNAS.112.797K. 1:CAS:528:DC%2BC2MXjtVSlsw%3D%3D. 25561561. 4311865 Cadotte MW. Functional traits explain ecosystem function through opposing mechanisms. Ecol. Lett. 2017; 20: 989-996. 28639274 Hillebrand H. Biodiversity change is uncoupled from species richness trends: Consequences for conservation and monitoring. J. Appl. Ecol. 2018; 55: 169-184 Whittaker RH. Evolution and measurement of species diversity. Taxon. 1972; 21: 213-251 Socolar JB, Gilroy JJ, Kunin WE, Edwards DP. How should beta-diversity inform biodiversity conservation?. Trends Ecol. Evol. 2016; 31: 67-80. 26701706 Legendre P, De Cáceres M. Beta diversity as the variance of community data: Dissimilarity coefficients and partitioning. Ecol. Lett. 2013; 16: 951-963. 23809147 Lichstein JW. Multiple regression on distance matrices: A multivariate spatial analysis tool. Plant Ecol. 2007; 188: 117-131 Tuomisto H, Ruokolainen K. Analyzing or explaining beta diversity? Understanding the targets of different methods of analysis. Ecology. 2006; 87: 2697-2708. 17168014 Peres-Neto PR, Jackson DA. How well do multivariate data sets match? The advantages of a Procrustean superimposition approach over the Mantel test. Oecologia. 2001; 129: 169-178. 2001Oecol.129.169P. 28547594 Peres-Neto PR, Legendre P, Dray S, Borcard D. Variation partitioning of species data matrices: Estimation and comparison of fractions. Ecology. 2006; 87: 2614-2625. 17089669 Hillebrand H, Matthiessen B. Biodiversity in a complex world: Consolidation and progress in functional biodiversity research: Consolidation and progress in BDEF research. Ecol. Lett. 2009; 12: 1405-1419. 19849711 Myers N, Mittermeier RA, Mittermeier CG, da Fonseca GAB, Kent J. Biodiversity hotspots for conservation priorities. Nature. 2000; 403: 853-858. 2000Natur.403.853M. 1:CAS:528:DC%2BD3cXhs1Olsr4%3D. 10706275 Bendix J. A research framework for projecting ecosystem change in highly diverse tropical mountain ecosystems. Oecologia. 2021; 195: 589-600. 2021Oecol.195.589B. 33515062. 7940296 Beck E, Bendix J, Kottke I, Makeschin F, Mosandl R. Gradients in a Tropical Mountain Ecosystem of Ecuador. 2008; Springer Landscape Restoration, Sustainable Use and Cross-Scale Monitoring of Biodiversity and Ecosystem Functions - A Science-directed Approach for South Ecuador (PAK823–825 Platform for Biodiversity and Ecosystem Monitoring and Research in South Ecuador, 2017). Beck E. Ecosystem Services, Biodiversity and Environmental Change in a Tropical Mountain Ecosystem of South Ecuador. 2013; Springer Homeier J, Leuschner C. Factors controlling the productivity of tropical Andean forests: Climate and soil are more important than tree diversity. Biogeosciences. 2021; 18: 1525-1541. 2021BGeo.18.1525H. 1:CAS:528:DC%2BB3MXhvFCisr3N Krashevska V, Sandmann D, Maraun M, Scheu S. Consequences of exclusion of precipitation on microorganisms and microbial consumers in montane tropical rainforests. Oecologia. 2012; 170: 1067-1076. 2012Oecol.170.1067K. 22614263. 3496542 Krashevska V, Sandmann D, Maraun M, Scheu S. Moderate changes in nutrient input alter tropical microbial and protist communities and belowground linkages. ISME J. 2014; 8: 1126-1134. 1:CAS:528:DC%2BC2cXmvV2htb4%3D. 24285360 Tiede Y. Ants as indicators of environmental change and ecosystem processes. Ecol. Indic. 2017; 83: 527-537 Santillán V. Spatio-temporal variation in bird assemblages is associated with fluctuations in temperature and precipitation along a tropical elevational gradient. PLoS ONE. 2018; 13: e0196179. 29746478. 5945003 Hsieh TC, Ma KH, Chao A. iNEXT: An R package for rarefaction and extrapolation of species diversity (H ill numbers). Methods Ecol. Evol. 2016; 7: 1451-1456 Hsieh, T. C, Ma, K. H. & Chao, A. iNEXT: Interpolation and Extrapolation for Species Diversity. R package version 2.0.20, http://chao.stat.nthu.edu.tw/wordpress/software%5fdownload/ (2020). Chao A. Rarefaction and extrapolation with Hill numbers: A framework for sampling and estimation in species diversity studies. Ecol. Monogr. 2014; 84: 45-67 Wallis CIB. Modeling tropical montane forest biomass, productivity and canopy traits with multispectral remote sensing data. Remote Sens. Environ. 2019; 225: 77-92. 2019RSEnv.225.77W Keuskamp JA, Dingemans BJJ, Lehtinen T, Sarneel JM, Hefting MM. Tea Bag Index: A novel approach to collect uniform decomposition data across ecosystems. Methods Ecol. Evol. 2013; 4: 1070-1075 Quitián M. Elevation-dependent effects of forest fragmentation on plant-bird interaction networks in the tropical Andes. Ecography. 2018; 41: 1497-1506 Fries A. Thermal structure of a megadiverse Andean mountain ecosystem in southern Ecuador and its regionalization. Erdkunde. 2009; 63: 321-335 Fries A, Rollenbeck R, Nauß T, Peters T, Bendix J. Near surface air humidity in a megadiverse Andean mountain ecosystem of southern Ecuador and its regionalization. Agric. For. Meteorol. 2012; 152: 17-30. 2012AgFM.152.17F Zvoleff, A. glcm: calculate textures from grey-level co-occurrence matrices (GLCMs). R package version 1.6.1 (2016). Wallis CIB. Remote sensing improves prediction of tropical montane species diversity but performance differs among taxa. Ecol. Indic. 2017; 83: 538-549 Wolf K, Veldkamp E, Homeier J, Martinson GO. Nitrogen availability links forest productivity, soil nitrous oxide and nitric oxide fluxes of a tropical montane forest in southern Ecuador: N2 O + NO flux of tropical montane forests. Glob. Biogeochem. Cycles. 2011. 10.1029/2010GB003876 Fisher WD. On Grouping for Maximum Homogeneity. J. Am. Stat. Assoc. 1958; 53: 789-798. 101577. 0084.35904 Bivand, R. classInt: Choose Univariate Class Intervals (2020). Oksanen, J. et al. vegan: Community Ecology Package (2020). vegan: Community Ecology Package. https://CRAN.R-project.org/package=vegan. Wood SN. Generalized Additive Models: An Introduction with R. 2017; Chapman and Hall/CRC. 1368.62004 Barbosa AM, Real R, Munoz AR, Brown JA. New measures for assessing model equilibrium and prediction mismatch in species distribution models. Divers. Distrib. 2013; 19: 1333-1338 Lotz T, Nieschulze J, Bendix J, Dobbermann M, König-Ries B. Diverse or uniform? Intercomparison of two major German project databases for interdisciplinary collaborative functional biodiversity research. Ecol. Inform. 2012; 8: 10-19 Göttlicher D. Land-cover classification in the Andes of southern Ecuador using Landsat ETM+ data as a basis for SVAT modelling. Int. J. Remote Sens. 2009; 30: 1867-1886 Deng Y, Wilson JP, DEM Bauer BO. resolution dependencies of terrain attributes across a landscape. Int. J. Geogr. Inf. Sci. 2007; 21: 187-213 Weiss, M. & Baret, F. S2ToolBox Level 2 products: LAI, FAPAR, FCOVER Version 1.1. in S2 Toolbox Level 2 Product algorithms v1.1 53. Unger M, Homeier J, Leuschner C. Relationships among leaf area index, below-canopy light availability and tree diversity along a transect from tropical lowland to montane forests in NE Ecuador. Trop. Ecol. 2013; 54: 33-45 Krashevska V, Maraun M, Scheu S. Micro- and macroscale changes in density and diversity of Testate amoebae of tropical montane rain forests of southern Ecuador. Acta Protozool. 2010; 49: 17-28

By Christine I. B. Wallis; Yvonne C. Tiede; Erwin Beck; Katrin Böhning-Gaese; Roland Brandl; David A. Donoso; Carlos I. Espinosa; Andreas Fries; Jürgen Homeier; Diego Inclan; Christoph Leuschner; Mark Maraun; Katrin Mikolajewski; Eike Lena Neuschulz; Stefan Scheu; Matthias Schleuning; Juan P. Suárez; Boris A. Tinoco; Nina Farwig and Jörg Bendix

Reported by Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author; Author

Titel:
Biodiversity and ecosystem functions depend on environmental conditions and resources rather than the geodiversity of a tropical biodiversity hotspot.
Autor/in / Beteiligte Person: Wallis, CIB ; Tiede, YC ; Beck, E ; Böhning-Gaese, K ; Brandl, R ; Donoso, DA ; Espinosa, CI ; Fries, A ; Homeier, J ; Inclan, D ; Leuschner, C ; Maraun, M ; Mikolajewski, K ; Neuschulz, EL ; Scheu, S ; Schleuning, M ; Suárez, JP ; Tinoco, BA ; Farwig, N ; Bendix, J
Link:
Zeitschrift: Scientific reports, Jg. 11 (2021-12-31), Heft 1, S. 24530
Veröffentlichung: London : Nature Publishing Group, copyright 2011-, 2021
Medientyp: academicJournal
ISSN: 2045-2322 (electronic)
DOI: 10.1038/s41598-021-03488-1
Schlagwort:
  • Climate
  • Forests
  • Models, Theoretical
  • Soil
  • Biodiversity
  • Ecosystem
  • Environment
  • Tropical Climate
Sonstiges:
  • Nachgewiesen in: MEDLINE
  • Sprachen: English
  • Publication Type: Journal Article; Research Support, Non-U.S. Gov't
  • Language: English
  • [Sci Rep] 2021 Dec 31; Vol. 11 (1), pp. 24530. <i>Date of Electronic Publication: </i>2021 Dec 31.
  • MeSH Terms: Biodiversity* ; Ecosystem* ; Environment* ; Tropical Climate* ; Climate ; Forests ; Models, Theoretical ; Soil
  • References: PLoS One. 2010 Jul 14;5(7):e11554. (PMID: 20644646) ; Sci Rep. 2020 Mar 19;10(1):5066. (PMID: 32193471) ; Oecologia. 2021 Mar;195(3):589-600. (PMID: 33515062) ; PLoS One. 2018 May 10;13(5):e0196179. (PMID: 29746478) ; Science. 2019 Sep 13;365(6458):1108-1113. (PMID: 31515383) ; Oecologia. 2001 Oct;129(2):169-178. (PMID: 28547594) ; Proc Natl Acad Sci U S A. 2015 Jan 20;112(3):797-802. (PMID: 25561561) ; Conserv Biol. 2015 Jun;29(3):668-79. (PMID: 25923191) ; Ecology. 2006 Nov;87(11):2697-708. (PMID: 17168014) ; Proc Natl Acad Sci U S A. 2019 Aug 13;116(33):16155-16158. (PMID: 31409742) ; Nature. 2000 May 11;405(6783):212-9. (PMID: 10821281) ; Conserv Biol. 2015 Jun;29(3):618-29. (PMID: 25922899) ; Nat Ecol Evol. 2020 Jan;4(1):2-3. (PMID: 31745262) ; Ecology. 2006 Oct;87(10):2614-25. (PMID: 17089669) ; Trends Ecol Evol. 2016 Jan;31(1):67-80. (PMID: 26701706) ; Proc Biol Sci. 2015 Jun 07;282(1808):20150418. (PMID: 25994675) ; Science. 2019 Dec 13;366(6471):. (PMID: 31831642) ; Ecol Lett. 2017 Aug;20(8):989-996. (PMID: 28639274) ; Glob Ecol Biogeogr. 2019 May;28(5):548-556. (PMID: 31217748) ; Microb Ecol. 2015 Feb;69(2):395-406. (PMID: 25395291) ; Oecologia. 2012 Dec;170(4):1067-76. (PMID: 22614263) ; Environ Manage. 2013 Sep;52(3):541-52. (PMID: 23775492) ; Proc Biol Sci. 2015 Feb 22;282(1801):20142620. (PMID: 25567651) ; Conserv Biol. 2015 Jun;29(3):680-91. (PMID: 25924074) ; Conserv Biol. 2015 Jun;29(3):630-9. (PMID: 25923307) ; Conserv Biol. 2010 Jun;24(3):701-10. (PMID: 20067491) ; Ecol Lett. 2013 Aug;16(8):951-63. (PMID: 23809147) ; Conserv Biol. 2015 Jun;29(3):692-701. (PMID: 25923052) ; Glob Chang Biol. 2013 Aug;19(8):2373-80. (PMID: 23630036) ; Ecol Lett. 2009 Dec;12(12):1405-19. (PMID: 19849711) ; Nat Commun. 2016 Dec 22;7:13736. (PMID: 28004657) ; Nature. 2000 Feb 24;403(6772):853-8. (PMID: 10706275) ; Philos Trans R Soc Lond B Biol Sci. 2011 Aug 27;366(1576):2426-37. (PMID: 21768157) ; ISME J. 2014 May;8(5):1126-34. (PMID: 24285360)
  • Substance Nomenclature: 0 (Soil)
  • Entry Date(s): Date Created: 20220101 Date Completed: 20220131 Latest Revision: 20220131
  • Update Code: 20231215
  • PubMed Central ID: PMC8720099

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -