SciELO - Scientific Electronic Library Online

 
vol.26 número2GENETIC ANALYSES SUGGEST BURROW SHARING BY RÍO NEGRO TUCO-TUCOS (Ctenomys rionegrensis)ACTIVIDAD REPRODUCTIVA Y DINÁMICA POBLACIONAL DE Rhipidomys fulviventer (RODENTIA: CRICETIDAE) EN LA CORDILLERA ORIENTAL COLOMBIANA índice de autoresíndice de assuntospesquisa de artigos
Home Pagelista alfabética de periódicos  

Serviços Personalizados

Journal

Artigo

Indicadores

  • Não possue artigos citadosCitado por SciELO

Links relacionados

Compartilhar


Mastozoología neotropical

versão impressa ISSN 0327-9383versão On-line ISSN 1666-0536

Mastozool. neotrop. vol.26 no.2 Mendoza jun. 2019  Epub 07-Jan-2019

 

ARTÍCULOS

INFLUENCE OF BAT MORPHOLOGY ON STRUCTURAL PROPERTIES OF A PLANT-FRUGIVORE NETWORK

Influencia de la morfología de los murciélagos en las propiedades estructurales de una red de frugívoros y plantas

Tatiana Velásquez-Roa1 

Oscar E Murillo-García3 

1Grupo de Investigación en Ecología Animal, Departamento de Biología, Facultad de Ciencias Naturales y Exactas, Universidad del Valle

3Grupo de Investigación en Ecología Animal, Departamento de Biología, Facultad de Ciencias Naturales y Exactas, Universidad del Valle

Abstract

Identifying factors that influence the importance of species for mutualistic interactions is crucial for designing management and conservation strategies that aim to conserve ecosystem functions. Even though bats are key seed-dispersers, factors that influence their importance for structural properties of seed-dispersal networks are poorly studied. We evaluated the influence of cranial morphology, phylogenetic legacy and abundance of frugivorous bats on the structural properties of an interaction network in a tropical dry forest. Consequently, we captured bats and identified interactions with plants by collecting and identifying seeds from fecal samples. Then, we described network structure by using network-level indexes and assessed the contribution of bats to network structure by using species-level indexes. Analysis of the mutualistic network indicated that: (1) there are few species with high number of interactions, (2) the structure was more compartmentalized and less nested than expected, (3) there are subgroups of bats and plants that are highly connected among them, and (4) network would collapse promptly with the loss of bats with most interactions (generalist). Furthermore, bats that contribute the most to network structure tended to have short rostrum and wide cranium and/or were abundant, with Artibeus lituratus being the most important bat; however, contribution to network structure did not have phylogenetic signal. According to the interaction neutrality hypothesis, species’ abundances drive the chances of individuals encountering and interacting with other species. However, our results suggest that frugivores bats are not functionally equivalent since contribution of species to network’s structure was associated with both cranial morphologically and abundance.

Palabras clave abundancia; dispersión de semillas; estructura de redes; morfología craneal; señal filogenética

Resumen

La identificación de los factores que influencian la importancia de las especies para las interacciones mutualistas es crucial para diseñar estrategias de conservación y manejo destinadas a conservar la funcionalidad de los ecosistemas. A pesar de que los murciélagos son dispersores claves de semillas, los factores que influencian la importancia de estos para las propiedades estructurales de las redes de dispersión de semillas han sido poco estudiados. Evaluamos la influencia de la morfología craneal, el legado filogenético y la abundancia de murciélagos frugívoros sobre las propiedades estructurales de una red de interacción en un bosque seco tropical. Para esto, capturamos murciélagos e identificamos interacciones con plantas a través de las semillas colectadas en sus heces. Luego, describimos la estructura de la red utilizando índices a nivel de red y evaluamos la contribución de murciélagos a la estructura de la red utilizando índices a nivel de especie. El análisis de la estructura de la red mutualista entre murciélagos frugívoros y planas indicó que: (1) hay pocas especies con alto número de interacciones, (2) la estructura fue más compartimentalizada y menos anidada de lo esperado, (3) hay subgrupos de murciélagos y plantas que están altamente conectadas entre ellas y (4) la red podría colapsar rápidamente debido a la pérdida de murciélagos con muchas interacciones (generalistas). Los murciélagos que más contribuyen a la estructura de la red tienden a presentar rostro corto y cráneo ancho y/o a ser abundantes, siendo Artibeus lituratus el murciélago con mayor importancia para la estructura de la red; sin embargo, la contribución a la estructura de la red no presentó señal filogenética. Según la hipótesis de neutralidad en interacciones mutualistas, la abundancia determina las posibilidades de que los individuos se encuentren e interactúen con otras especies. Sin embargo, nuestros resultados sugieren que los murciélagos frugívoros no son funcionalmente equivalentes para la estructura de la red, dado que la contribución de estos a la estructura de la red estuvo asociada tanto con la morfología craneal como con su abundancia.

Palabras clave abundancia; dispersión de semillas; estructura de redes; morfología craneal; señal filogenética

INTRODUCTION

Mutualistic interactions are important to determine demographic dynamics and co-evolution of the species in communities, as a consequence their study is key to understand ecological processes that generate and sustain biodiversity (Bascompte et al. 2007; Bascompte & Jordano 2008). In particular, identifying factors that determine the importance of species to interaction network is fundamental to developing management strategies that lead to conservation of ecological processes. The structure of interaction networks is derived from the accumulation of interactions between species within communities. In particular, the complex interaction between plants and animals across time and space results in distinctive topological properties that differentiate mutualistic networks from other ecological networks (Vázquez et al. 2009a). This structure arises from spatial and temporal structures and/or evolutionary processes leading to non-random patterns of interactions, and describes the extent what species are organized into networks (Olesen et al. 2007). Despite the removal of keystone species could trigger secondary extinctions and cascading effects (Montoya et al. 2006), the structural properties of mutualistic networks can buffer the effect of primary extinctions (Memmott et al. 2004), conferring robustness to mutualistic networks. Therefore, understanding the architecture of the network and identifying species that are important for its structure are fundamental to conserve and manage ecosystems (Solé & Montoya 2001).

Mutualistic networks are more heterogeneous than expected by chance with many species having scarce interactions and a few species having a very large number of interactions (Bascompte & Jordano 2014; Vázquez et al. 2009a; Vázquez & Aizen 2004). Besides, mutualistic networks in general are highly nested (Jordano et al. 2003; Dupont & Olesen 2009; Bascompte et al. 2003; Montesinos-Navarro et al. 2012), which means that there is a core of generalist species interacting among them and specialists tend to interact with the most generalist species.

Finally, plant-animal mutualistic networks are modular (Olesen et al. 2007) since there are groups of species (modules) that tend to interact more between them than with species from other modules (Krause et al. 2003). This structure increases the stability and coexistence of mutualistic communities (Bastolla et al. 2009; Okuyama & Holland 2008) and makes mutualistic network more robust to loss of keystone species (Solé & Bascompte 2006; Solé et al. 2002). Evidence suggests that species traits (Bascompte et al. 2003; Jordano et al. 2003; Vázquez et al. 2009b; Olesen et al. 2011), evolutionary history (Peralta 2016; Vitória et al. 2018) and abundance (Vázquez & Aizen 2004) of species influence the structure of mutualistic interactions.

Morphological and phenological traits are important for interactions in plant-animal mutualism by constraining interactions between species (Olesen et al. 2011; Stang et al. 2009; Bascompte & Jordano 2007). Morphological constraints generate mismatches between morphology of the species involved in the interaction (Olesen et al. 2011; Stang et al. 2009). On the other hand, the presence of a phylogenetic signal in mutualistic networks, would suggest that inherited ancestral phenotypic traits determine, at least in part, ecological interactions (Thompson 2005; Vázquez et al. 2009a). However, other evidence indicates that neutral processes have an important influence on those observed patterns (Krishna et al. 2008; Canard et al. 2012; Vázquez et al. 2007). The ecological equivalence (neutrality) hypothesis proposes that interactions between individuals are random; consequently, all individuals have the same probability of interaction in ecological networks regardless of species differences (Vázquez et al. 2007). Hence, species abundance and random interactions between individuals are important drivers of mutualistic network structure. Consequently, biological constraints, evolutionary history and neutral processes can determine species role in structuring mutualistic networks (Ives & Godfray 2006; Jordano 2010; Bascompte & Jordano 2014); however, the relative contribution of these factors to network structure is poorly known and requires studying highly diverse biological group with important ecological roles for ecosystem functioning.

Seed dispersal and pollination are mutualistic interactions that contribute to maintenance of the structure and stability of tropical communities (Bascompte & Jordano 2014). Interactions between plants and frugivores, in particular, are crucial in tropical forest since from 50% to 90% of the woody plants are dispersed by animals (Fleming 1987). Bats are the most diverse group of seeds dispersal among mammals, playing a key role for conservation of plant diversity in tropical forests (Muscarella & Fleming 2007). Bats mainly disperse plants with opportunistic strategies that are important for colonization of degraded areas, which are characterized by quick growing and fruits with large number of seeds (Charles-Dominique 1986). Therefore, bats disperse seeds of important plants for regeneration of disturbed forests by intense human activities (Memmott et al. 2007; Hegland et al. 2009; Laliberté & Tylianakis 2010). This process is particularly important for the tropical dry forest, an ecosystems heavily affected by loss of coverage due to agricultural and livestock expansion (Castaño-Salazar 2009). Consequently, understanding the role of mutualistic networks for tropical forests conservation requires assessing factors that determine species’ importance for the structure of interaction networks.

Our goal was to evaluate the influence of cranial morphology, phylogenetic history and abundance of frugivorous bats on the structural properties of a mutualistic network between frugivorous bats and plants in a tropical dry forest. We expected that: (a) interaction network would have be heterogeneous, nested and modular; and (b) closely related species and species with similar cranial morphology would have similar importance for network structure and that abundant species would be the most important for network structure.

MATERIALS AND METHODS

Study area

The study was conducted at a fragment of tropical dry forest (Holdridge 1967) corresponding to “Parque Regional Natural El Vínculo” (PRN El Vínculo), Buga, Valle del Cauca, Colombia. The PRN El Vínculo is surrounded by a matrix of sugar cane crops and grasslands, and is covered by forests in different successional stages (intervened forest, secondary forest and scrub land) with different levels of anthropogenic disturbance (Torres 2012). The park ranges from 977 to 1150 m. of elevation, and the climate is characterized by average annual temperature of 24°C and average annual of precipitation 1379 mm (Torres 2012). The area presents two periods of heavy rainfall (March-June and SeptemberDecember) and two dry periods (January-April and July- August), this seasonality is typical of the tropical dry forest (Marulanda et al. 2003).

Field methods

We captured bats monthly, from July 2017 to February 2018, at the different successional stages of PRN El Vínculo during five nights per month. In each sampling period, five 12m x 3m mist nets were set at different zones, during each day, for a total sampling effort of 40 nights (246 hours) during eight months. We randomly selected sites for net placement, before each sampling day, with consecutive nest separated by at least 100 m. To capture bats during the period of greatest activity, the nets were set in the evening (18:00 h) and pulled at midnight (24:00 h); mist nets were checked every 30 minutes. Identification of the bats captured in field, to lowest possible taxonomic category, was based on specialized taxonomic keys (Gardner 2007; Diaz et al. 2016). We held captured bats individually in cloth bags, for one hour, in order to collect their fecal samples (Thomas 1988), which were deposited individually in properly-labeled paper bags. In addition, we walked tracks monthly with the purpose of collecting fruits potentially consumed by bats in order to create a seed reference collection for the study area. Vegetation samples were identified until the least possible taxonomic level with specialized literature (Ríos et al. 2004; Lobova et al. 2009; Linares & Moreno-Mosquera 2010).

Seeds were sorted out from fecal samples of bats, and identified (whenever possible) using taxonomic keys (Ríos et al. 2004; Lobova et al. 2009; Linares & Moreno-Mosquera 2010) and comparing with the seed reference collection for the study area. Since we were interested in the potential function of bats as seed dispersers, we identified the links between bats and plants by seeds from bat feces.

Cranial morphology

To quantify skull morphology, we measured 20 linear variable traits reflecting shape and size that are associated with bite performance and diet in Phyllostomidae bats (Santana et al. 2012; Dumont et al. 2012): Total skull length (TL), Postorbital Width (PW), Zygomatic Breadth (MZB), Posterior Skull width (PSW), Palatal Width at canines (PW), Palatal Width at M1 (PW1), Total Palate Length (TPL), Anterior Skull Length (ASL), Post-Palatal Length (PPL), Maxillary Toothrow Length (MTL), Dentary Depth under m1 (DD), Coronoid Process Height (CPH), Condyle Height (CH), Condyle to Canine bite point (CC), Condyle to m1 point (Cm1), Condyle to m3 point (Cm3), Total Dentary Length (TDL), Condyle-Coronoid Length (CCL), Coronoid-Angular Length (CAL) and Mandibular toothrow length (MAN). For measurements, we used a digital caliper with 0.01-mm accuracy. To describe shape and size variation in cranium, we performed a principal components analysis of functional units (skull and jaw) based on the morphometric variables we measured. Then, we averaged species scores for the first three principal components and used them to represent cranial morphology in subsequent analyses.

Data analysis

We constructed an interaction matrix with bat species in rows and plants species in columns; in every cell we registered the number of interactions (seeds from bat feces) observed between each bat and plant species.

Sampling effort

We assessed sampling representativeness by comparing observed and expected number of species (Bascompte & Jordano 2014). Thus, we calculated the asymptotic richness of species (bats and plants) and interactions, with the nonparametric estimator of abundance Jackknife 1 (Chacoff et al. 2012; Chiu et al. 2014). Posteriorly, we calculated the percentage of representativeness of sampling using the ratio between observed and expected species richness. For analysis, we used vegan package 2.2-3 (Oksanen 2015) in R 3.4.1 (R Core Team 2013).

Network structure

To describe network heterogeneity, we separately calculated the accumulated distributions of interactions for plants and animals, and assessed the fit to three models (Jordano et al. 2003): exponential, power law and truncated power law; using the degreedistr function of bipartite Rpackage (Dorman et al. 2017). To assess for network nestedness, we measured the Nestedness metric based on Overlap and Decreasing Fill index (NODF) that varies from 0 (no nestedness) to 100 (perfect nestedness). In a non-nested network, interactions tend to be symmetric, generating a vulnerable structure to the loss of interactions (Bascompte et al. 2003). The significance of NODF was estimated with a resampling procedure in which we randomly generated 1000 bootstrap samples from the original matrix with Patefield’s (1981) algorithm, as implemented in function r2dtable in R statistical software version 3.4.1 (R Core Team 2013); this algorithm fix marginal totals to distribute the interactions and produce networks in which all species are randomly associated. We tested the estimates of nestedness by calculating 95% confidence intervals for NODF values generated from the random matrixes, and we considered a network as antinested (less nested than expected) or nested (more nested than expected) if the observed NODF was lower or higher than values in confidence interval, respectively.

We measured modularity with the QuaBiMo (Q) algorithm (Dormann & Strauss 2014); which computes modules on pondered bipartite networks, based on a hierarchical representation of the heights of the species’ links and the optimal assignation to the modules (Dormann & Strauss 2014). Q index ranges from 0 (random network without modules) to 1 (maximum modularity) (Dormann & Strauss 2014). To test for significance of modularity, we generate 1000 random matrixes from the original matrix with a bootstrap procedure using Patefield’s (1981) algorithm as before. We tested the estimates of modularity by calculating 95% confidence intervals for Q values generated from the random matrixes, and we considered a network as modular if the observed Q was higher than values in the confidence interval.

We assessed robustness of the network to species extinction with index R (Burgos 2007), which corresponds to the area under the curve of extinctions: relationship between the amount of species that survive in the network against the accumulated number of extinct species of the other group. This index measures how fast one group of the network collapses with the extinction of species from the other group. Values of R close to 1 indicate a curve that decrease slowly with the elimination of species, while values close to 0 indicate a network that collapses quickly when the first species are removed. We generated 1000 simulations of extinctions for each one of the tropic levels of the network (bats and plants) with a bootstrap approach, in which species were removed from the network following three strategies: random elimination, elimination from specialists to generalists and elimination from generalists to specialists. Analyses were performed with the bipartite package (Dorman et al. 2017), implemented in program R 3.4.1 (R Core Team 2013).

Contribution of bats to network structure

We calculated several species-level indexes that estimate the importance of bats to different aspects of network structure: (1) z measure the importance of species to connectivity inside modules (Vidal et al. 2014), (2) c measure the importance of species to connectivity among modules (Vidal et al. 2014), (3) species strength measures the extent to which an assemblage depends on a specific species of the other group (Bascompte et al. 2006), (4) partner diversity is a measure of species’ generality of interactions (Dormann 2011), (5) closeness centrality measures the proximity of a species to all other species in the network (Martín-González et al. 2010; Mello et al. 2015) and, (6) betweenness centrality quantifies species’ importance as a network connector (Martín-González et al. 2010; Mello et al. 2015). These indexes can be correlated since species may have similar importance for distinct aspects of network structure, so we performed a principal component analysis (PCA) on the correlation matrix, and considered the first principal component (PC1) as a measure of species contribution to the structure of the interaction network (Vidal et al. 2014; Sazima et al. 2010). For index estimation, we used bipartite R-package (Dorman et al. 2017) in R 3.4.1 (R Core Team 2013).

We evaluate the importance of abundance and morphology of the bats to the network structure by performing Pearson’s correlations between the index of species importance for the network and: a) bat abundance, and b) the first three principal components of cranial morphology. Finally, to evaluate the contribution of phylogenetic closeness to network structure, we assessed the phylogenetic signal of the index of species contribution to network structure using the lambda method (λ) (Pagel 1999) with Phylosig function of Phytools R-package (Revell 2012). For this, we used the phylogenic tree for Noctilionoidea of Rojas et al. (2016); from which we extracted a tree with the bat species captured during the fieldwork. For analysis, we used software R 3.4.1 (R Core Team 2013).

RESULTS

We captured 11 species of frugivorous bats (111 individuals) during the sampling period; which belong to the Phyllostomidae family and to the genera Artibeus, Carollia, Chiroderma, Dermanura, Sturnira and Uroderma (Table 1. We obtained 45 out of 75 samples of bat feces with seeds; the remaining came from individuals that consumed fleshy fruits with large seeds such as Mango (Mangifera indica, which was one of the most abundant trees around the sampling site. Uroderma species were not considered for analyses since we did not collect seeds in their feces (Table 1). We found 38 plant species (32 genera and 23 families) potentially consumed by frugivorous bats during the monthly censuses in the study area. The most abundant families in those censuses were Moraceae, Piperaceae, Rubiaceae and Solanaceae. However, we only found seeds of nine of those plants in bat feces (Table 2). We identified 55 interactions between nine species of bats and nine species of plants for the network of PRN El Vínculo (Fig. 1). The species of bats considered generalists, those that consumed the most variety of fruits, were A. lituratus (seven interactions) and, C. perspicillata (three interactions). For plants, Ficus sp. (five) and Piper holtonii (four) presented the most interactions (Fig. 1).

Table 1 Number of individuals (ni) of the species of frugivorous bats captured in the PRN El Vínculo, number of feces samples by species (FSS) and number of feces samples with seeds (FSWS). 

Fig. 1 Species with the higher number of interactions (generalists): (A) Artibeus lituratus and (B) Carollia perspicillata for bats, and (C) Ficus sp. and (D) Piper holtonii for plants. 

Sampling effort

The sampling was representative for bats and plants, since we found that the richness corresponded to 75% and to 81% of the expected richness, respectively. On the other hand, we recorded 65% of the expected pairwise interactions between bats and plants.

Table 2 Plant species consumed by frugivorous bats in the PRN El Vínculo. 

Network structure

The general structure of the mutualistic network was heterogeneous since distribution of interactions fitted to the truncated power-law model (truncated power-law AIC=-16. 50; Exponential AIC=-14.37; power law AIC=-4.75); which indicated that most species have a few interactions and few species have a very large number if interactions (Fig. 2). Furthermore, connectance was low since we only registered the 26% of total possible interactions, whereas density indicated that in average each species interacts with 2.5 species from the other group. On the other hand, the network was significantly less nested than expected (antinested) since the observed value of the NODF (0.39) was lower than confidence interval values of the null model (95% Confidence interval=0.41-0.76). Additionally, the mutualistic network presented a modular structure since the observed value (Q=0.48) was not included in the confidence interval of the null model (95% CI=0.19-0.29). This indicated that the network is organized in subgroups of bats and plants (modules) that are more connected among them than with the other species (Fig. 3). Finally, the simulated extinction of specialist to generalist presented high values of robustness for plants and for bats (R=0.90 and 0.83, respectively), with the random elimination of species presented intermediated values (R=0.67 plants and 0.66 bats), and with extinctions from generalists to specialists presented low values (R=0.43 and 0.41); which indicated that the network of interactions is most sensitive to the loss of generalist species. In addition, our results showed that bat assemblage is equally resilient to loss of generalist species as plants, but is less resilient to the loss of specialist plants than bats.

Contribution of bats to network structure

The first principal component (index of importance) had similar and positive factor loads (z=0.42, c=0.36, species strength=0.39, partner diversity=0.45, closeness centrality=0.36, betweenness centrality=0.44), which indicated that species with high values (PC1 scores) contribute the most to network structure. Consequently, A. lituratus was the species thatmost contributed to the structure of the interaction network (species scores: A. lituratus=5.34; A.planirostris=0.32; C. perspicillata=0.26; Dermanura sp.=0.12; C. brevicauda=0.03; S. lilium=-0.67; A. jamaicensis=-1.62; C. salvini=-1.66; D. phaeotis=-2.11). Furthermore, the most abundant species were themost important for network structure since we found a positive correlation between bat abundance and importance index (r=0.95, p<=0.01).

On the other hand, the first three PCs account for 98.05% of the variation in cranium and jaw morphology; with PC1 scores representing skull size, PC2 scores representing relationship between skull width (Posterior Skull Width) and rostrum length (Total Palate Length), and PC3 scores representing relationship between cranial width (Zygomatic and Palatal Amplitude at M1) and length (Post Palatal Length) (Table 3). Consequently, these components represented changes in skull size and shape of bats. The first two morphological components were not associated with the index of importance (PC1 r=0.39, p=0.34; PC2 r=0.05, p=0.91), whereas association with the third morphological principal component was between 1.6 and 12.8 times higher than for first two PCs (r=0.64, p=0.08). This indicated a trend for bat species with a short and wide skull to be the most important for network structure. Finally, the contribution of the species to network structure was not associated with phylogenetic relationships since the index of importance did not have a strong phylogenetic signal (λ=6.61e-05, p=1.00).

Table 3 Results of the Principal Component Analysis based on cranial morphometric traits of frugivorous bats from PRN El Vínculo. TL: Total skull length, PSW: Posterior skull width, TPL: Total palate length, MZB: Zygomatic breadth, PW1: Palatal width at M1, PPL: Post-palatal length. 

Fig. 2 Interaction network between plants and frugivorous bats (black and grey, respectively) in the tropical dry forest at PRN El Vínculo. Grey lines represent interactions between species with width of the line representing interaction strength width of the line; so thicker lines represent a higher strength between species. 

DISCUSSION

The mutualistic network of interactions between frugivorous bats and plants was heterogeneous, antinested, modular, and was sensitive to the loss of generalist species. The importance of species to network structure was determined by cranial morphology and abundance; with Artibeus lituratus being the bat species contributing the most to network structure. However, phylogenetically related species were not equally important. Thus, our results suggest that bats are not functionally equivalent for network structure.

Network structure

Contrary to species inventories, there is a considerable fraction of non-possible interactions, due to biological constraints or non-co-occurrence in space or time that cannot be sampled. Besides, using seeds from feces would underestimate the network between large bats and plants with large fruits since they can disperse large seeds without swallow them. In plant-seed disperser networks, around 15% of all potential links can be forbidden because of morphological constraints (Olesen et al. 2011) and 46-61% due to phenological uncoupling, especially in highly seasonal habitats (Jordano 2016) such as dry forest. In addition, the truncated nature of the distribution suggests that there is a limit to the number of interactions of the most connected species in our study area (Bascompte & Jordano 2014). Consequently, the fact that we observed more than 65% of the interactions seems to be appropriate for confidence in results based on mutualistic networks analyses (Jordano 2016).

High robustness is caused by some structural patterns such as heterogeneity, nestedness and modularity (Krause et al. 2003; Memmott et al. 2004; Teng & Mccann 2004; Burgos 2007; Olesen et al. 2007; Bascompte & Jordano 2014). As expected for mutualistic interactions between frugivorous bats and plants (Mello et al. 2011), the network was heterogeneous and modular, but was less nested than expected based on null models (NODF observed=0.39, 95% Confidence interval of NODF expected=0.410.76). The network was heterogeneous since some species displayed a large number of interactions (generalist), which correspond to the bats A. lituratus and C. perspicillata and to the plants Ficus sp. and P. holtonii, whereas other displayed few interactions (specialists). However, previous works have reported heterogeneous bat fruit networks fitting to an exponential function (Zapata-Mesa et al. 2017) not a truncated power-law, as we found. The truncated nature of the distribution suggests that there is a limit to the number of interactions of the most connected species (Bascompte & Jordano 2014). Heterogeneity implies that each plant in the network is connected with a high number of frugivores, and this redundancy increases network robustness to bats extinction (Memmott et al. 2004).

Fig. 3 Modules resulting from the mutualistic interaction network between frugivorous bats and plants in the tropical dry forest at PRN El Vínculo. Modules were computed by applying QuaBiMo modularity measure (Dormann & Strauss 2014). Interactions between bats and plants are represented by the squares with intensity representing interaction strength; so darker squares represent a higher strength between species.  

Contrary to our expectations, the interaction network studied did not exhibit a nested linkage pattern; in which specialist interact with subsets of the partners of generalists. Nine bat-fruit datasets were analyzed (Mello et al. 2011) and most net works showed higher nestedness (NODF x=0.566, minimum=0.41, maximum=0.75) than our dataset (NODF=0.39). We found an antinested structure since the missing observations of specialist bats feeding on generalists’ preferred plant species could result in a lower nestedness than in the null models; which can result from competition- induced host utilization and compartmentalization (Dorman et al. 2017). This pattern can result if the most generalist bats (A. lituratus and C. perspicillata) can act as dominant species that monopolizes specific plant species (Ficus sp. and Piper holtonii, respectively) forcing specialist bats to feed onto other plant species (Dorman et al. 2017). Consequently, whether species substantially decrease plant overlap in bat assemblages need to be tested.

Furthermore, the modular structure indicates that the network consists of subgroups of plants and bats that interact more strongly among them than with the other groups. This modular structure has been reported previously and corresponds with the associations that have been reported between particular genera of bats and plants (Mello et al. 2011; Zapata Mesa et al. 2017). Bat species of Stenodermatinae subfamily are consumers of Ficus plants (Giannini & Kalko 2004), with exception of Sturnira bats that prefer plants of Solanum (Saldaña-Vázquez et al. 2013; Montoya-Bustamante et al. 2016); while species of subfamily Carollinae are associated with plants of genus Piper (Marinho-Filho 1991; Thies & Kalko 2004; Saldaña-Vázquez et al. 2013; Montoya Bustamante et al. 2016). We observed these associations in the modules corresponding to A. lituratus, A. jamaicensis, C. salvini and C. brevicauda with Ficussp.; and C. perspicillata and C. brevicauda with P. adumcun and C. rhomboideum. Consequently, our results suggest that bat species from the same genus share morphological traits that allow them to exploit similar plant resources (Olesen et al. 2007).

As expected the network can be considered to be robust to species extinction, particularly when such extinctions occur at random or affect the least linked species in the network first (Mello et al. 2011; Bascompte & Jordano 2014). Network robustness to the random loss of interacting species (R=0-66 bats and 0.67 plants) tended to be higher for bats and intermediate for plants with respect to previous studies (bats=0.41-0.69 and plants 0.58-0.84, Mello et al. 2011); suggesting that the network is resilient to the changes in species and interactions, and to the arrival of other species (Díaz-Castelazo et al. 2010). When specialist species were extinct first, bats presented a lower robustness in comparison to the plants, suggesting that bats have higher dependence for specialist plants.

Contribution of bats to network structure

According to our expectations, cranial morphology was important for determining bats’ contribution to network. Cranial morphology has been associated with bite’s force (Santana et al. 2012; Dumont et al. 2012) and strict frugivorous diet in Stenodermatinae bats (Dumont et al. 2012); which, in turn, has been a key factor in the diversification of phyllostomids (Dumont et al. 2012). Morphological traits associated with the importance of the species for the interaction network we studied (zygomatic amplitude, palatal amplitude and the post palatal longitude) has been identified as important for bite force, which is related with the consumption of fruits in Phyllostomidae (Dumont et al. 2012). The importance of A. lituratus for network structure was derived from having a short and wide skull, which is associated with a strong bite (Dumont et al. 2012), to feed on fruit of different characteristics; which make this bat a module connector. Results suggest that cranial morphology of A. lituratus is translated in an advantage for the use of diverse fruit resources in the study area, since these characteristics enable A. lituratus to have a stronger bite than the other abundant species (i.e. C. perspicillata) and to consume not only hard (i.e. Ficus plants) but also soft fruits (i.e. Piper plants).

As we expected, abundance of bats was associated with their contributions to network structure. This has been found in mutualistic networks, where variations in animal abundance can potentially affect the structure and robustness of the mutualistic networks (Rooney et al. 2006; Bascompte & Jordano 2014; Ramos-Robles et al. 2016; Laurindo et al. 2017). Although species from Artibeus genus reported in the study area (such as A. lituratus and A. planirostris) are morphologically similar, they did not have the same level of importance since A. lituratus had a greater abundance than A. planirostris. However, the relative importance of A. lituratus with respect to C. perspicillata was greater than expected based on differences in abundance. Furthermore, in spite of the differences in abundance between C. perspicillata and A. planirostris, these species presented a similar level of importance for the interactions network. Collectively, these results, suggest that morphology actually plays a determining role in the contribution of bats to network structure. Several studies have suggested that species abundance is the leading factor in the structuration of the mutualistic networks (Jordano 1987; Olesen et al. 2002; Vázquez & Aizen 2004; Gonzalez & Loiselle 2016), which has relevant ecological or evolutionary implications for interactions network (Bascompte & Jordano 2007). This agrees with the neutral theory of diversity that proposes that species are functionally similar (Hubbell 2001), that is biological characteristics of individuals do not matter but the factors like their abundance determine interactions in ecological networks (Krishna et al. 2008; Bascompte & Jordano 2014; Ramos-Robles et al. 2016; Laurindo et al. 2017). On the contrary, our results suggest that frugivorous bats are not functionally equivalent since differences in morphology grants them differential capacities to exploit feeding resources, indicating that the importance A. lituratus is not exclusively derived of its abundance; as the neutral hypotheses would predict.

CONCLUSIONS

Assessing the differential contribution of species to the structure of interaction networks is essential to develop conservation and management strategies that lead to the safeguarding of the ecological processes that generate and maintain diversity. The removal of keystone species could trigger secondary extinctions and cascading effects affecting network stability (Montoya et al. 2006). Therefore, identifying those species is important to conserve and manage ecosystems (Solé & Montoya 2001). In our case, A. lituratus was the species that contributed the most to network structuration, so this species is fundamental for the regeneration process of PRN El Vínculo by the virtue of its seed dispersal capability. This species is generalist and common, representing almost 30% of captures in tropical forest (Muylaert 2017); but their functional relevance for ecological processes in tropical forest has been overlooked. Generalist species connect peripheral species together into modules, but also connect modules keeping the cohesiveness of the network (Guimarães et al. 2007). Therefore, they are considered very important for the conservation of mutualistic interaction networks. The importance of A. lituratus was derived from both a cranial morphology that allows access to several types of fruits and a high abundance. We conclude that cranial morphology is important for interactions between frugivores and plants, and that the contribution of morphologically similar species to network structure may depend on their relative abundances. Taken together, our results suggest that there is not functional equivalence among bats of the studied assemblage; on the contrary, species’ contribution to the network was derived from a combination of biological restrictions and neutral processes but not evolutionary history.

Acknowledgments

We want to thank to the Jóvenes Investigadores program of Colciencias for the funding. We thank to officials of Instituto para la investigación y la preservación del patrimonio cultural y natural del Valle del Cauca (INCIVA) for allowing the access to Parque Regional Natural El Vínculo. We thank to A.F. Ríos, Y. Ramos, J.S. Herrera, A.F. Vargas, Y.L. Villegas, L. Castrillón, M.E. González and V. González for their fieldwork assistance. We also thank to Y. Ramos and L. Alvarez for helping with plant identification. Finally, we want to thank research group Animal Ecology of Department of Biology at Universidad del Valle and to Centro de estudios e investigaciones en biodiversidad y biotecnología (CIBUQ) group from Universidad del Quindío for allowing the use of equipment and access to laboratories.

There are not conflicts of interest.

REFERENCIAS

B01 Arenas, D., & A. Giraldo. 2013. Chiroptera of the Regional Natural Park El Vínculo and its buffer zone (Buga, Valle del Cauca, Colombia). Biota Colombiana 14:51-56. [ Links ]

B02 Bascompte, J., & P. Jordano (eds). 2014. Mutualistic networks, second ed. Princeton University Press. New Yersey USA. [ Links ]

B03 Bascompte, J., & P. Jordano. 2008. Mutualistic networks of species. Investigación y ciencia 384:50-59. [ Links ]

B04 Bascompte, J., & P. Jordano. 2007. Plant-animal mutualistic network: The architecture of biodiversity. Annual Reviews 38:567-593. https://doi.org/10.1146/annurev.ecolsys.38.091206.095818Links ]

B05 Bascompte, J., P. Jordano, M. Pascual, & J. Dunne. 2007. The Structure of Plant-Animal Mutualistic Networks. Ecological networks: Linking structure to dynamics in food webs (M. Pascual & J . Dunne, eds.). Oxford University Press, Oxford, UK. https://doi.org/10.23943/princeton/9780691131269.003.0003Links ]

B06 Bascompte, J., P. Jordano, & J. M. Olesen. 2006. Asymmetric coevolutionary networks facilitate biodiversity maintenance. Science 431:3-6. https://doi.org/10.1126/science.1123412Links ]

B07 Bascompte, J., P. Jordano, C. J. Melian, & J. M. Olesen. 2003. The nested assembly of plant-animal mutualistic networks. Proceedings of the National Academy of Sciences 100:9383-9387. https://doi.org/10.1073/pnas.1633576100Links ]

B08 Bastolla, U., M. A. Fortuna, A. Pascual-García, A. Ferrera, B. Luque & J. Bascompte. 2009. The architecture of mutualistic networks minimizes competition and increases biodiversity. Nature 458:1018-1020. https://doi.org/10.1038/nature07950Links ]

B09 Burgos, E. et al. 2007. Why nestedness in mutualistic networks? Journal of Theoretical Biology 249:307-313. [ Links ]

B10 Canard, E., N. Mouqet, L. Marescot, K. J. Gaston, D. Gravel, & D. Mouillot. 2012. Emergence of structural patterns in neutral trophic networks. PLoS ONE 7:e38295 https://doi.org/10.1371/journal.pone.0038295Links ]

B11 Castaño-Salazar, J. H. 2009. Frugivorous bats and chiropterocorian plants: Discovering the structure of their mualist interactions in a semi-caducifolia jungle. Master of Thesis, Universidad de los Andes, Mérida, Venezuela. [ Links ]

B12 Chacoff, N. P., D. P. Vázqez, S. B. Lomáscolo, E. L. Stevani, J. Dorado, & B. Padrón. 2012. Evaluating sampling completeness in a desert plant-pollinator network. Journal of Animal Ecology 81:190-200. https://doi.org/10.1111/j.1365-2656.2011.01883.xLinks ]

B13 Charles-Dominique, P. 1986. Inter-relations between frugivorous vertebrates and pioneer plants: Cecropia, birds and bats in French Guyana. Frugivores and seed dispersal (A. Estrada & T. Fleming, eds.). Dr W. Junk Publishers, Dordrecht, The Netherlands. https://doi.org/10.1007/978-94-009-4812-9 12 [ Links ]

B14 Chiu, C. H., Y. T. Wang, B. A. Walther, & A. Chao. 2014. An improved nonparametric lower bound of species richness via a modified good-turing frequency formula. Biometrics 70:671-682. https://doi.org/10.1111/biom.12200Links ]

B15 Díaz-Castelazo, C., P. R. Guimarães Jr, P. Jordano, J. N. Thompson, R. J. Marquis, & V. Rico-Gray. 2010. Changes of a mutualistic network over time: Reanalysis over a 10-year period. Ecology 91:793-801. https://doi.org/10.1890/08-1883.1Links ]

B16 Diaz, M., S. Solari, L. Aguirre, L. Aguilar, & R. Barqez. 2016. Identification key of the South American bats. Special Publication No. 2, PCMA (Argentine Bat Conservation Program). Tucumán, Argentina. [ Links ]

B17 Dormann, C. F., J. Fründ, & H. M. Schaefer. (2017) Identifying Causes of Patterns in Ecological Networks: Opportunities and Limitations. Annual Review of Ecology, Evolution, and Systematics 48:559-584. https://doi.org/10.1146/annurev-ecolsys-110316-022928Links ]

B18 Dorman, C. F., J. F. Fruend, & B. Grube. 2017. Visualising Bipartite Networks and Calculating Some (Ecological). R package documentation 2:1-158. [ Links ]

B19 Dormann, C. F., & R. Strauss. 2014. A method for detecting modules in quantitative bipartite networks. Methods in Ecology and Evolution 5:90-98. https://doi.org/10.1111/2041-210x.12139Links ]

B20 Dormann, C. F. 2011. How to be a specialist? Quantifying specialisation in pollination networks. Network Biology 1:1-20. [ Links ]

B21 Dumont, E.R., L. Dávalos, A. Goldberg, S. E. Santana, K. Rex, & C. Voigt. 2012. Morphological innovation, diversification and invasion of a new adaptive zone. Proceedings of the Royal Society of Biology 279:1797-1805. https://doi.org/10.1098/rspb.2011.2005Links ]

B22 Dupont, Y.l., & J. M. Olesen. 2009. Ecological modules and roles of species in heathland plant-insect flower visitor networks. Journal of Animal Ecology 78(2):346-53. https://doi.org/10.1111/j.1365-2656.2008.01501.xLinks ]

B23 Fleming, T. H. 1987. Patterns of tropical vertebrate frugivore diversity. Annual Review of Ecology, Evolution, and Systematics 18:91-109. https://doi.org/10.1146/annurev.ecolsys.18.1.91Links ]

B24 Gardner, A (ed). 2007. Mammals of South America. 1st. edition. University of Chicago Pres, Chicago, USA. [ Links ]

B25 Giannini, N. P., & E. K. V. Kalko. 2004. Trophic structure in a large assemblage of phyllostomid bats. Oikos 2:209-220. https://doi.org/10.1111/j.0030-1299.2004.12690.xLinks ]

B26 Gonzalez, O., & B. A. Loiselle. 2016. Species interactions in an Andean bird–flowering plant network: phenology is more important than abundance or morphology. PeerJ 4:2-22. https://doi.org/10.7717/peerj.2789Links ]

B27 Guimarães, P. R. et al. 2007. Build-up mechanisms determining the topology of mutualistic networks. Journal of Theoretical Biology 249:181-189. [ Links ]

B28 Hegland, S. J., A. Nielsen, A. Lázaro, A. L. Bjerknes, & O. Totland. 2009. How does climate warming affect plant-pollinator interactions? Ecology Letters 12:184-195. https://doi.org/10.1111/j.1461-0248.2008.01269.xLinks ]

B29 Holdridge, L. R (ed). 1967. Life zone ecology. 1st edition. Tropical Science Center. San José, Costa Rica. [ Links ]

B30 Hubbell, S. P. 2001. The unified neutral theory of biodiversity and biogeography. Princeton University Press, New Yersey, USA. [ Links ]

B31 Ives, A. R., & H. C. J. Godfray. 2006. Phylogenetic Analysis of Trophic Associations. The American Naturalist 168:E1-E14. https://doi.org/10.1086/505157Links ]

B32 Jordano, P. 2016. Advances and challenges in the study of ecological networks sampling networks of ecological interactions. Functional Ecology 30: 1883-1893. https://doi.org/10.1111/1365-2435.12763Links ]

B33 Jordano, P. 2010. Coevolution in Multispecific Interactions among Free-Living Species. Evolution: Education and Outreach 3:40-46. https://doi.org/10.1007/s12052-009-0197-1Links ]

B34 Jordano, P. 1987. Patterns of mutualistic interactions in pollination and seed dispersal: connectance, dependence asymmetries, and coevolution. The American Naturalist 129:657-677. https://doi.org/10.1086/284665Links ]

B35 Jordano, P., J. Bascompte, & M. J. Olesen. 2003. Ivariant Properties in Coevolutionary Networks of Plant – Animal Interactions. Ecology letters 6:69-81. https://doi.org/10.1046/j.1461-0248.2003.00403.xLinks ]

B36 Krause, A. E., K. J. Frank, D. M. Mason, R. E. Ulanowicz, W. W. Taylor. 2003. Compartments revealed in food web structure. Nature 426:282-285. https://doi.org/10.1038/nature02115Links ]

B37 Krishna, A., P. R. Guimarães Jr, P. Jordano, & P. Bascompte 2008. A neutral-niche theory of nestedness in mutualistic networks. Oikos 117:1609-1618. https://doi.org/10.1111/j.1600-0706.2008.16540.xLinks ]

B38 Laliberté, E., & J. M. Tylianakis. 2010. Deforestation homogenizes tropical parasitoid-host networks. Ecology 91:1740-1747. https://doi.org/10.1890/09-1328.1Links ]

B39 Laurindo, R.S., R. Gregorin, & D. C. Tavares. 2017. Effects of biotic and abiotic factors on the temporal dynamic of bat-fruit interactions. Acta Oecologica 83:38-47. https://doi.org/10.1016/j.actao.2017.06.009Links ]

B40 Linares, É., & E. Moreno-Mosqera. 2010. Morphology of cecropia frutioli (Cecropiaceae) of the colombian pacific and its taxonomic value in the study of diets of bats. Caldasia 32:275-287. [ Links ]

B41 Lobova, T. A., C. K. Geiselman, & S. A. Mori. 2009. Seed dispersal by bats in the neotropics. Volume 101 de Memoirs of the New York Botanical Garden. New York Botanical Garden, New York. https://doi.org/10.5962/bhl.title.59555Links ]

B42 Marinho-Filho, J. S. 1991. The coexistence of two frugivorous bats and the phenology of their food plants in Brazil. Journal of Tropical Ecology, London, 7 (1): 59-67. https://doi.org/10.1017/S0266467400005083Links ]

B43 Martín-González, A. M., B. Dalsgaard, & J. M. Olesen. 2010. Centrality measures and the importance of generalist species in pollination networks. Ecological complexity 7:36-43. https://doi.org/10.1016/j.ecocom.2009.03.008Links ]

B44 Marulanda, L. O. et al. 2003. Structure and composition of vegetation in a tropical dry forest remnant in San Sebastián, Magdalena (Colombia). I. composition of vascular plants. Actualidades Biológicas 25-78: 17-30. [ Links ]

B45 Mello, M. A. R. et al. 2015. Keystone species in seed despersal networks are mainly determined by dietary specialization. Oikos 124:1031-1039. https://doi.org/10.1111/oik.01613Links ]

B46 Mello, M. A. R., F. M. D. Marquitti, P. R. Guimarães Jr, E. K. V. Kalko, P. Jordano, & M. A. Martinez De Aguiar. 2011. The Missing Part of Seed Dispersal Networks: Structure and Robustness of Bat-Fruit Interactions. PLoS ONE 6: e17395. https://doi.org/10.1371/journal.pone.0017395Links ]

B47 Memmott, J., P. G. Craze, N. M. Waser, & M. V. Price. 2007. Global warming and the disruption of plant – pollinator interactions. Ecology letters 10:710-717. https://doi.org/10.1111/j.1461-0248.2007.01061.xLinks ]

B48 Memmott, J., N. M. Waser, & M. V. Price. 2004. Tolerance of pollinator networks to species extinctions. Proceedings of the Royal Society London Biology 271:2605-2611. [ Links ]

B49 Montesinos-Navarro, A., J. G. Segarra-Moragues, A. Valiente Banuet, & M. Verdú. 2012. The network structure of plant–arbuscular mycorrhizal fungi. New Phytologist 194:536-547. https://doi.org/10.1111/j.1469-8137.2011.04045.xLinks ]

B50 Montoya-Bustamante, S., V. Rojas-Díaz, & A. M. Torres- Gonzalez. 2016. Interactions between frugivorous bats (Chiroptera: Phyllostomidae) and Piper tuberculatum (Piperaceae) in a tropical dry forest. Revista de Biología Tropical 64:701-713. https://doi.org/10.15517/rbt.v64i2.20689Links ]

B51 Montoya, J. M.., L. S. Pimm, & V. R. Sole. 2006. Ecological networks and their fragility. Nature Reviews 442:20. [ Links ]

B52 Muscarella, R., & T. H. Fleming. 2007. The role of frugivorous bats in tropical forest succession. Biological Reviews of the Cambridge Philosophical Society 82:573-590. https://doi.org/10.1111/j.1469-185x.2007.00026.xLinks ]

B53 Muylaert, R. D. L. et al. 2017. Atlantic bats: a data set of bat communities from the Atlantic Forests of South America. Ecology 98(12):3227. [ Links ]

B54 Oksanen, J. 2015. Multivariate analysis of ecological communities in R: vegan tutorial. R documentation 43. [ Links ]

B55 Okuyama, T., & J. N. Holland. 2008. Network structural properties mediate the stability of mutualistic communities. Ecology Letters 11:208-216. https://doi.org/10.1111/j.1461-0248.2007.01137.xLinks ]

B56 Olesen, J. M., J. Bascompte, Y. L. Dupont, H. Elberling, C. Rasmussen, & P. Jordano. 2011. Missing and forbidden links in mutualistic networks. Proceedings of the Royal Society B: Biological Sciences 278:725-732. https://doi.org/10.1098/rspb.2010.1371Links ]

B57 Olesen, J. M., J. Bascompte, Y. L. Dupont, & P. Jordano. 2007. The modularity of pollination networks. Proceedings of the National Academy of Sciences 104:19891-19896. https://doi.org/10.1073/pnas.0706375104Links ]

B58 Olesen, J. M., L. I. Eskildsen, & S. Venkatasamy. 2002. Invasion of pollination networks on oceanic islands: Importance of invader complexes and endemic super generalists. Diversity and Distributions 8:181-192. https://doi.org/10.1046/j.1472-4642.2002.00148.xLinks ]

B59 Pagel, M. 1999. Inferring the historical patterns of biological evolution. Nature 401:877-884. https://doi.org/10.1038/44766Links ]

B60 Patefield, W. M. 1981. Algorithm AS 159: An efficient method of generating random R × C tables with givenrow and column totals. Journal of the Royal Statistical Society 30:91-97. https://doi.org/10.2307/2346669Links ]

B61 Peralta, G. 2016. Merging evolutionary history into species interaction networks. Functinal Ecology 30:1917-1925. https://doi.org/10.1111/1365-2435.12669Links ]

B62 R Core Team. 2013. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing. Vienna, Austria. [ Links ]

B63 Ramos-Robles, M., E. Andresen, & C. Díaz-Castelazo. 2016. Temporal changes in the structure of a plant-frugivore network are influenced by bird migration and fruit availability. PeerJ 4:20-48. https://doi.org/10.7717/peerj.2048Links ]

B64 Revell, L. J. 2012. Phytools: An R package for phylogenetic comparative biology (and other things). Methods in Ecology and Evolution 3:217-223. https://doi.org/10.1111/j.2041-210x.2011.00169.xLinks ]

B65 Ríos, M., P. Giraldo, & D. Correa (eds). 2004. Guide to fruits and seeds from the middle basin of the Otún River. 1st edition. Fundación EcoAndina, WCS-Colombia, Santiago de Cali, Colombia. [ Links ]

B66 Rojas, D., O. M. Warsi, & L. M. Davalos. 2016. Bats (Chiroptera: Noctilionoidea) challenge a recent origin of extant neotropical diversity. Systematic Biology 65:432-448. https://doi.org/10.1093/sysbio/syw011Links ]

B67 Rooney, N., K. Mccann, G. Gellner, & J. C. Moore. 2006. Structural asymmetry and the stability of diverse food webs. Nature 442:265-269. https://doi.org/10.1038/nature04887Links ]

B68 Saldaña-Vázquez, R. A., V. J. Sosa, L. I. Iñiguez-Dávalos, & J. E. Schondube. 2013. The role of extrinsic and intrinsic factors in Neotropical fruit bat–plant interactions. Journal of Mammalogy 94:632-639. https://doi.org/10.1644/11-mamm-a-370.1Links ]

B69 Santana, S. E., I. R. Grosse, & E. R. Dumont. 2012. Dietary hardness, loading behavior, and the evolution of skull form in bats. Evolution 66:2587-2598. https://doi.org/10.1111/j.1558-5646.2012.01615.xLinks ]

B70 Sazima, C., P. R. Guimarães Jr, S. F. Dos Reis, & I. Sazima. 2010. What makes a species central in a cleaning mutualism network? Oikos 119:1319-1325. https://doi.org/10.1111/j.1600-0706.2009.18222.xLinks ]

B71 Solé, R. V., & J. Bascompte (eds). 2006. Self-organization in Complex Ecosystems. Princeton University Press, Princeton, NJ. [ Links ]

B72 Solé, R. V., D. Alonso, & A. Mckane. 2002. Self-organized instability in complex ecosystems. Philosophical Transactions of the Royal Society B: Biological Sciences 357:667-681. https://doi.org/10.1098/rstb.2001.0992Links ]

B73 Solé, R. V., & M. Montoya. 2001. Complexity and fragility in ecological networks. Proceedings of the royal society B 268:2039-2045. [ Links ]

B74 Stang, M., P. G. L. Klinkhamer, N. M. Waser, I. Stang, & E. Van Der Meijden. 2009. Size-specific interaction patterns and size matching in a plant –pollinator interaction web. Annals of Botany 103:1459-1469. https://doi.org/10.1093/aob/mcp027Links ]

B75 Teng, J., & K. S. Mccann. 2004. Dynamics of compartmented and reticulate food webs in relation to energetic flow. The American Naturalist 164:85-100. https://doi.org/10.1086/421723Links ]

B76 Thies, W., & E. K. V. Kalko. 2004. Phenology of neotropical pepper plants (Piperaceae) and their association with their main dispersers, two short-tailed fruit bats, Carollia perspicillata and C. castanea (Phyllostomidae). Oikos 104:362-376. https://doi.org/10.1111/j.0030-1299.2004.12747.xLinks ]

B77 Thomas, D. W. 1988. Analysis of diet of plant-visiting bats. Ecological and behavioral methods for the study of bats. (T. H. Kunz, ed) Smithsonian Institution Press, Whasington, USA. [ Links ]

B78 Thompson, J. N. (ed) 2005. The geographic mosaic of coevolution. The University of Chicago Press. Chicago and London. [ Links ]

B79 Torres, A. M. et al. 2012. Successional dynamics of a fragment of tropical dry forest of Valle del Cauca, Colombia. Biota Colombiana 13:66-84. [ Links ]

B80 Vázquez, D. P., & M. A. Aizen. 2004. Asymmetric specialization: A pervasive feature of plant-pollinator interactions. Ecology 85:1251-1257. https://doi.org/10.1890/03-3112Links ]

B81 Vázquez, D. P., C. J. Melián, N. M. Williams, N. Blüthgen, B. R. Krasnov, & R. Poulin. 2007. Species abundance and asymmetric interaction strength in ecological networks. Oikos 116:1120-1127. https://doi.org/10.1111/j.2007.0030-1299.15828.xLinks ]

B82 Vázquez, D. P., N. Blüthgen, L. Cagnolo & N. P. Chacoff. 2009a. Uniting pattern and process in plant–animal mutualistic networks: a review. Annals of Botany 103:1445-1457. https://doi.org/10.1093/aob/mcp057Links ]

B83 Vázquez, D. P., N. P. Chacoff, & L. Cagnolo. 2009b. Evaluating multiple determinants of the structure of plant-animal mutualistic networks. Ecoloy 90:2039-2046. https://doi.org/10.1890/08-1837.1Links ]

B84 Vidal, M. M., E. Hasui, M. A. Pizo, J. Y. Tamashiro, W. R. Silva, & P. R. Guimarães Jr. 2014. Frugivores at higher risk of extinction are the key elements of a mutualistic network. Ecology 95:3440-3447. https://doi.org/10.1890/13-1584.1Links ]

B85 Vitória, R. S., J. Vizentin-Bugoni, & L. D. S. Duarte. 2018. Evolutionary history as a driver of ecological networks: a case study of plant–hummingbird interactions. Oikos 127:561-569. https://doi.org/10.1111/oik.04344Links ]

B86 Zapata-Mesa, N., S. Montoya-Bustamante, & O. E. Murillo- García. 2017. Temporal variation in bat-fruit interactions: Foraging strategies influence network structure over time. Acta oecologica 85: 9-17. https://doi.org/10.1016/j.actao.2017.09.003Links ]

Recibido: 05 de Septiembre de 2018; Aprobado: 07 de Enero de 2019