Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-19T04:12:44.829Z Has data issue: false hasContentIssue false

Palm functional trait responses to local environmental factors in the Colombian Amazon

Published online by Cambridge University Press:  15 November 2021

William Trujillo*
Affiliation:
Grupo Investigaciones territoriales para el uso y conservación de la biodiversidad. Fundación Reserva Natural La Palmita, Centro de Investigación. Cra 4 # 58-59, Bogotá, Colombia
Carlos A. Rivera-Rondón
Affiliation:
UNESIS, Departamento de Biología, Pontificia Universidad Javeriana, 110231, Bogotá, Colombia
Jorge Jácome
Affiliation:
UNESIS, Departamento de Biología, Pontificia Universidad Javeriana, 110231, Bogotá, Colombia
Néstor García
Affiliation:
UNESIS, Departamento de Biología, Pontificia Universidad Javeriana, 110231, Bogotá, Colombia
Wolf L. Eiserhardt
Affiliation:
Department of Biology, Ecoinformatics & Biodiversity, Aarhus University, Ny Munkegade 116, DK-8000Aarhus C., Denmark
Henrik Balslev
Affiliation:
Department of Biology, Ecoinformatics & Biodiversity, Aarhus University, Ny Munkegade 116, DK-8000Aarhus C., Denmark
*
Author for correspondence: William Trujillo, Email: williamtrujilloca@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

Functional traits play a key role in driving plant community effects on ecosystem function. We examined nine functional traits in various palm (Arecaceae) species and their relationships with moisture, tree-fall gaps, slope, and forest type at 29 transects (500×5 m) in the northeastern region of the Colombian Amazon. Redundancy analysis of mean trait values of species within a plot weighted by their abundance and Pearson correlations were used to evaluate the relationships between traits and environmental factors. The community trait composition was correlated with local environmental factors, which explained 23% of the trait variance. We detected functional dominance of the tallest palms in soils with high moisture and in floodplain forests (p ≤0.05). Palms with relatively long leaves were dominant in the flooded forests. Acaulescent and small palms were dominant on high slopes, and in terra firme forests, long-petioled palms were dominant in forest gaps. The number of seeds per fruit was not correlated with any environmental variable. Thus, hydrology is one of the main drivers of the functional composition of neotropical palm communities at the local scale, segregating tall palms with competitive and evasive strategies from small understory palms, which are mainly stress tolerant.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2021. Published by Cambridge University Press

Introduction

Hydrology, topography, and the spatiotemporal dynamics of canopy gaps are among the main abiotic factors that determine the composition, structure, and richness of palm communities at both local and landscape scales (Eiserhardt et al. Reference Eiserhardt, Svenning, Kissling and Balslev2011). These factors strongly influence the functional response of the plants (Gratani Reference Gratani2014, Gross et al. Reference Gross, Robson, Lavorel, Albert, Le Bagousse-Pinguet and Guillemin2008, Ishii & Asano Reference Ishii and Asano2010, Pearcy Reference Pearcy, Pugnaire and Valladares2007, Salgado-Negret et al. Reference Salgado-Negret, Canessa, Valladares, Armesto and Pérez2015).

Hydrology determines the availability or amount of excess water in the soil. Elongation of shoots is a strategy to avoid anoxic stress by keeping the aerial parts of plants above the water level. This has been identified as a functional response of plants to survive in flooded areas (Blom & Voesenek Reference Blom and Voesenek1996, Voesenek & Bailey-Serres Reference Voesenek and Bailey-Serres2015, Vormisto et al. Reference Vormisto, Tuomisto and Oksanen2004). There are few published works on the functional response of palms to flooding (Balslev et al. Reference Balslev, Kahn, Millan, Svenning, Kristiansen, Borchsenius, Pedersen and Eiserhardt2011). Compared with palms growing in well-drained soils, palms that grow in floodplain forests (FPs) have different tolerances and preferences (Clark et al. Reference Clark, Clark, Sandoval and Castro1995); nevertheless, it is not clear how functional traits vary in response to these environmental traits. Palms may also be specific to particular microhabitats within these habitat types, and fine-scale palm distributions may depend on heterogeneity in terms of soil moisture and drainage (Svenning Reference Svenning1999, Reference Svenning2001).

The density and size of tree-fall gaps importantly affect forest dynamics because they generate variation in light availability and temperature (Kern et al. Reference Kern, Montgomery, Reich and Strong2014). The gradient in light availability gives rise to increased richness (Muscolo et al. Reference Muscolo, Bagnato, Sidari and Mercurio2014) and causes physiological and morphological responses in some understory species (Gratani Reference Gratani2014, Ishii & Asano Reference Ishii and Asano2010, Pearcy Reference Pearcy, Pugnaire and Valladares2007) depending on their tolerance to shade and the amount of increase in light after the formation of a gap. For palms, it has been suggested that large arborescent species depend on tree-fall gaps for recruitment (De Granville Reference De Granville1991, Kahn Reference Kahn1986). However, some species, such as Euterpe edulis, do not grow in open areas, such as large gaps; this phenomenon has been attributed to growth, survival and hydraulic traits, because photoinhibition and hydraulic limitations can affect growth rates (Gatti et al. Reference Gatti, Campanello, Villagra, Montti and Goldstein2014).

Topography is another environmental factor with important effects on plant growth and vegetation development (Den Dubbelden & Knops Reference Den Dubbelden and Knops1993). Soil on slopes experiences a relatively high degree of runoff, and such habitats favor species that evade wet and flooded areas (Méndez-Toribio et al. Reference Méndez-Toribio, Ibarra-Manríquez, Navarrete-Segueda and Paz2017). The local distributions of many palm species are significantly influenced by topography (Clark et al. Reference Clark, Clark, Sandoval and Castro1995, Kahn Reference Kahn1986, Svenning Reference Svenning1999). However, palm species distribution patterns vary from one place to another because topography has no direct influence on itself but rather functions through its correlations with other environmental variables, such as drainage, formation of catenas, exposure of originally different sediment layers, and forest architecture (Vormisto et al. Reference Vormisto, Tuomisto and Oksanen2004).

The ecology and evolution of palms is strongly linked to interspecific variation in the growth, reproduction, and morphology of stems, leaves, inflorescences, fruits, and seeds (Kissling et al. Reference Kissling, Balslev, Baker, Dransfield, Göldel, Lim, Onstein and Svenning2019). Moreover, the functional traits of the dominant species in a community determine ecosystem processes (Conti & Díaz Reference Conti and Díaz2013, Finegan et al. Reference Finegan, Peña-Claros, De Oliveira, Ascarrunz, Bret-Harte, Carreño-Rocabado, Casanoves, Díaz, Eguiguren Velepucha and Fernandez2015, Hobbie Reference Hobbie2015). Palms constitute a noticeable structural and functional component in the Amazon Forest (Gentry Reference Gentry1988); compared with other common species, they are well represented among trees and can reach very high population densities (Ter Steege et al. Reference Ter Steege, Pitman, Sabatier, Baraloto, Salomão, Guevara, Phillips, Castilho, Magnusson and Molino2013), making them an ideal group for evaluating ecological responses because they are diverse, they are gregarious, and they are locally abundant (Terborgh & Andresen Reference Terborgh and Andresen1998).

This article investigates the relationships between local environmental variables and functional traits of palms in the Colombian Amazon, with the goal of contributing to the knowledge of the functional response and strategies of neotropical groups of high ecological importance, such as palms. The question asked is as follows: How are the functional composition of the palm community and the functional traits of palm species related to environmental characteristics (soil moisture, slope and gaps) in a locality in the northeastern region of the Colombian Amazon?

Methods

Study area

The study was performed along the Guaviare River in the eastern region of the Colombian Amazon in the departments of Guainía and Vichada (Figure 1), where three forest types are recognized. One type is terra firme forests, which are never flooded and generally have infertile and well-drained soils, with high local diversity. The second type is FPs, which are usually flooded annually. The third type is forests on terraces that are never or only very rarely flooded, and these include several ancient never-flooded terraces derived from rivers where they used to flow (Balslev et al. Reference Balslev, Copete, Pedersen, Bernal, Galeano, Duque, Berrio, Sanchez and Myster2017, Villota Reference Villota2005). The data collected along the Guaviare River form part of a larger dataset covering 546 transects in palm communities covering the western region of the Amazon (Balslev et al. Reference Balslev, Copete, Pedersen, Bernal, Galeano, Duque, Berrio, Sanchez and Myster2017, Reference Balslev, Kristiansen and Muscarella2019).

Figure 1. Study area along the Guaviare River in the Colombian Amazon. The triangles represent the position of 29 transects (500×5 m) in which all adult individuals of 25 species of palms were identified and counted.

Data collection

We established 29 transects of 500×5 m each, which were subdivided into 100 subunits of 5×5 m. The transects were established in the terra firme forests (11), in floodplains (14), and on terraces (4). In each transect, all adult palm individuals were registered, and on the basis of their size, it was determined whether they had reached the reproductive stage. The palm species were documented with voucher specimens under Rodrigo Bernal’s numbers 4442–4541 deposited at the Colombian National Herbarium in Bogotá (COL), with duplicates at the Aarhus University Herbarium (AAU). The specimens can be seen on the Aarhus University Herbarium database (https://www.aubot.dk/search_form.php). Further details of the collection and census methodology can be found in the study of Balslev et al. (Reference Balslev, Navarrete, Paniagua-Zambrana, Pedersen, Eiserhardt and Kristiansen2010, Reference Balslev, Kristiansen and Muscarella2019).

In each transect, the following environmental variables were recorded: (1) Presence/absence of gaps in each subunit of 5 × 5 m, with the subunits with gaps subsequently counted to obtain a value between 0 and 100 per transect; (2) moisture on the soil surface in subunits of 5×5 m within each transect, recorded as 0 or 1 for dry soil or muddy soil and/or stagnant water, respectively; with the subunits with values of 1 subsequently counted to obtain a value between 0 and 100 per transect; and (3) slope measured in degrees (0–90°) in each subunit of 5×5 m, which were averaged to obtain a single value per transect.

Palm functional traits were determined based on the literature (Galeano & Bernal Reference Galeano and Bernal2010, Henderson Reference Henderson2011) and by examining specimens deposited in the Herbario Nacional Colombiano (COL). Nine functional traits were recorded for each palm species (Table 1): lifeform (LF; cespitose = ces, solitary = sol), growth form (GF; acaulescent = aca, erect = ere, climbing = cli), maximum stem height (StH), maximum leaf number (LN), maximum petiole length (PeL), maximum leaf rachis length (RL), maximum fruit diameter (FD), seed number per fruit (SN), and breeding system (BS, dioecious = dio, monoecious = mon) (Supplement 1).

Table 1. Functional traits and abundances of 25 palm species registered in 29 transects, each covering 500×5 m, equalling 0.25 ha each or a total of 7.25 ha for all transects. The species names are according to the World Flora Online (worldfloraonline.org).

Functional traits: LF=lifeform (cespitose=ces, solitary=sol), GF=growth form (acaulescent=aca, erect=ere, climbing=cli), StH= maximum stem height, LN= maximum leaf number, PeL= maximum petiole length, RL= maximum leaf rachis length, FD=maximum fruit diameter, SN=seed number, BS=breeding system (dioecious=dio, monoecious=mon). Forest type: TF=terra firme, FP=floodplain and Ter=terrace.

Data analysis

The following continuous traits were categorized to show their distribution and frequencies among the forest types: maximum stem height (StH; short=0−8 m, medium=9−17 m, tall=18–25 m); maximum leaf number (LN; low=7−21, medium=22−35, high=36−50); maximum petiole length (PeL; short=45−120 cm; medium=121−225 cm; long=226−330 cm); maximum leaf rachis length (RL; =0−263 cm, medium=264−526 cm, long=527−750 cm); maximum fruit diameter (FD; small=7−24 mm, medium=25−42 mm, large=43−60 mm); and seed number per fruit (SN; 1, 2 or 3 seeds). With the field data, three tables were constructed: matrix Q (q x m) = 9 functional traits of 25 palm species (Table 1); matrix R (n x p) = environmental variables of 29 transects; and matrix L (n x q) = palm abundance in the 29 transects (Supplement 2).

The functional trait composition was calculated from the community-weighted means of traits (CWM) using the formula of Garnier et al. (Reference Garnier, Lavorel, Ansquer, Castro, Cruz, Dolezal, Eriksson, Fortunel, Freitas and Golodets2007). The CWM was obtained from the combination of the abundance (L) and trait (Q) matrices. The CWM calculates a trait value for each transect (Kleyer et al. Reference Kleyer, Dray, Bello, Lepš, Pakeman, Strauss, Thuiller and Lavorel2012) from the mean trait value, weighted by the abundance of all the species present in that transect (for continuous traits), or the weighted proportion of species (for categorical traits). The CWM corresponds to the weighted mean of the trait in the community (Díaz et al. Reference Díaz, Lavorel, De Bello, Quétier, Grigulis and Robson2007, Lavorel et al. Reference Lavorel, Grigulis, McIntyre, Williams, Garden, Dorrough, Berman, Quétier, Thébault and Bonis2008).

Redundancy analysis (CWM-RDA) was used to establish the relationships between functional composition (CWM) and environmental factors (Kleyer et al. Reference Kleyer, Dray, Bello, Lepš, Pakeman, Strauss, Thuiller and Lavorel2012). The categorical variable ‘forest type’ was included in the analysis as a passive variable. Additionally, RLQ (Dolédec et al. Reference Dolédec, Chessel, Ter Braak and Champely1996), fourth corner (Dray et al. Reference Dray, Choler, Dolédec, Peres-Neto, Thuiller, Pavoine and Ter Braak2014), and Pearson correlation (Sokal & Rohlf Reference Sokal and Rohlf2000) analyses were used. All analyses were performed using R statistical software (R Core Team 2016) and the Vegan and Ade-4 packages (Dray & Dufour Reference Dray and Dufour2007).

To test if the observed trait–environment relationships are potentially driven by phylogenetic covariation with other traits, we related our data to the most comprehensive palm phylogeny to date (Faurby et al. Reference Faurby, Eiserhardt, Baker and Svenning2016), using the version of the phylogeny that includes full taxonomic constraints. We visually explored the distribution of the traits across the phylogenetic tree, calculated phylogenetic signal, and statistically related mean trait values across communities to phylogenetic eigenvectors. These analyses were performed in R using the phangorn (Schliep Reference Schliep2011), picante (Kembel et al. Reference Kembel, Cowan, Helmus, Cornwell, Morlon, Ackerly, Blomberg and Webb2010), and PCPS (Debastiani & Duarte Reference Debastiani and Duarte2014) packages.

Results

We recorded 3123 mature individuals from 25 palm species in the 29 transects that together encompassed a total of 7.25 ha (Table 1). The most diverse genus was Bactris, with seven species, followed by Attalea, Desmoncus, and Geonoma, each with three species. The most abundant species was Iriartella setigera, with 492 individuals, followed by Geonoma deversa and Astrocaryum gynacanthum, with 453 and 402 individuals, respectively. Astrocaryum jauari, Bactris bidentula, and Astrocaryum acaule were identified and collected in the area, but they were not included in the analysis because they were collected only outside the transects.

Most individuals (62%) were solitary. The cespitose palms were evenly distributed between the floodplains and terra firme forests, while the solitary palms were more frequent in the terra firme forests (Figure 2A). Most individuals (83%) had erect growth forms. Notably, 99% of the acaulescent palms grew in the terra firme and terrace forests (Figure 2B).

Figure 2. Relative frequency distribution of functional traits in a palm community in floodplain (FP), terrace (Ter), and terra firme forest types along the Guaviare River in the northeastern region of the Colombian Amazon.

The highest proportion of individuals had short stems (52%), particularly in the terra firme forests, whereas tall stems were more frequent in the FPs (Figure 2C). Most individuals had few leaves (96%). The individuals with high and medium numbers of leaves were in the FP (Figure 2D). Of all individuals, 75% had short petioles (Figure 2E). Moreover, most individuals had short leaf rachises (73%); medium-rachis individuals and species were predominant in the FPs, and long-rachis individuals were better represented in the terra firme forests (Figure 2F).

Most individuals (70%) had small fruits, and all large-fruited palms grew in the FPs (Figure 2G). Most individuals (86%) had one seed in each fruit. Individuals with three seeds were best represented in the terra firme forests, while all individuals with two seeds were in the FPs (Figure 2H). Most of the individuals palm were monoecious (99.9%); the only two dioecious individuals were in the terrace forests (Figure 2I).

Environmental variables and palm functional traits

Redundancy analysis (CWM-RDA) showed that the two first axes of the environmental variables (gaps, moisture, slope, and FT) explained 22% (P = 0.002) of the variation in the functional composition (Figure 3). Water in the soil surface (moisture) and the FPs was associated with positive values of axis 1, while the slope (slope) and terra firme forests (TF) were associated with the negative values of these axes. The presence of gaps (Gap) and terrace forests (Ter) were associated with positive values of axis 2 (Figure 3).

Figure 3. Ordination obtained by redundancy analysis (RDA) of the functional trait composition (CWM) in the palm community along the Guaviare River in the Colombian Amazon as a function of environmental variables. The white-tipped arrows represent environmental variables: Gap=presence of tree-fall gaps, Slope=inclination, Moist=soil moisture. The gray triangles are the forest types: TF=terra firme forests, Fp=floodplain forests and Ter=terrace forests. The black-tipped arrows represent traits: LF=lifeform (cespitose=ces, solitary=sol), GF=growth form (acaulescent=aca, erect=ere, climbing=cli), StH= maximum stem height, LN= maximum leaf number, RL= maximum leaf rachis length, PeL= maximum petiole length, FD=maximum fruit diameter, SN=seed number, BS=breeding system (dioecious=dio, monoecious=mon).

StH and FD were clearly associated with positive values of axis 1, while the number of seeds per fruit (SN) and the acaulescent growth form (GFaca) were related to negative values in the CWM-RDA. In general, the species with the relatively long petioles are present in greater numbers in sites with high numbers of forest gaps, while tall, large-leaved and climbing species occur mainly in sites with high moisture (Figure 3).

The RLQ analysis showed that species traits, environmental factors, and abundance were correlated (P <0.001). Most of the variation was captured by the first axis of the RLQ (81%, Supplement 3). In the FP transects, the maximum height of the stem (StH) and the soil moisture were associated with positive values of this axis. Species of the Socratea, Oenocarpus, and Euterpe genera (all tall palms) were associated with axis 1, whereas Attalea and Manicaria species were associated with negative values of axis 2. Likewise, tree-fall gaps and terrace forests were associated with the positive values of axis 2, where mainly species of the genera Geonoma and Bactris were abundant (medium to low number of palms) (Supplement 3).

Individual trait-environment relationships

The acaulescent growth form was positively correlated with slope and terra firme forests and negatively correlated with moisture and FPs. The StH was positively correlated with soil moisture and FPs and negatively correlated with slope and terra firme forests. We found functional dominance of tall palms in FPs with high soil moisture and functional dominance of small palms in terra firme and sloping forests (Figure 4).

Figure 4. Scatter plots of palm functional traits and environmental variables that showed significant Spearman correlations in a palm community in the Amazon rainforest of Colombia, as studied in 29 transects of 500×5 m.

Leaf size was positively correlated with soil moisture and FPs and negatively correlated with terrace forests. The FD was strongly and positively correlated with FPs and negatively correlated with terra firme forests (Table 2). The seed number was not strongly correlated with any environmental variable measured.

Table 2. Spearman’s rank correlations between environmental variables and functional composition (CWM) of a palm community in the Colombian Amazon.

N=29 transects. ***, P< 0.001; **, P< 0.05. Environmental variables: moist=soil moisture, gaps=presence of tree-fall gaps, slope=inclination; traits: LF=lifeform (cespitose=ces, solitary=sol), GF=growth form (acaulescent=aca, erect=ere, climbing=cli), StH= maximum stem height, LN= maximum leaf number, PeL= maximum petiole length, RL= maximum leaf rachis length, FD=maximum fruit diameter, SN=seed number, BS=breeding system (dioecious=dio, monoecious=mon).

Phylogenetic patterns

All traits except fruit diameter and acaulescent growth form showed phylogenetic signal significantly larger than random in a significant number of the phylogenies sampled from the posterior distribution of Faurby et al. (Reference Faurby, Eiserhardt, Baker and Svenning2016) (Supplement 4). Cespitose life form, climbing habit, seed number, and leaf dimensions (rachis and PeL) showed particularly strong phylogenetic signal. However, when testing for phylogenetic patterns in the distribution of traits across communities, only cespitose habit and – to a much lesser degree – PeL deviated from a phylogenetically random distribution (Supplement 4).

Discussion

Relationship between functional traits and environmental variables

Our results are in agreement with the plant ecology strategy scheme, which states that leaf morphology, stem height, and seed size (LHS) capture the main axes of plant responses to competition, stress, and disturbance (Westoby Reference Westoby1998). Tall and large-leaved palms were associated with moist and flooded areas, and small and acaulescent palms were associated with terra firme forests and sloping areas (Figure 3). This suggests that hydrology is the main environmental filter of functional composition, as has been documented in other studies of palm functional traits in the Amazonia region (Kristiansen et al. Reference Kristiansen, Svenning, Grández, Salo and Balslev2009, Pereira De Souza Reference Pereira De Souza2019). Hydrology segregates tall canopy palms with a competitive and avoidance strategy (Voesenek et al. Reference Voesenek, Rijnders, Peeters, Van de Steeg and De Kroon2004) from small understory palms, which are mainly stress tolerant (Grime Reference Grime1974).

We found that slope was strongly correlated with acaulescent growth (Table 2). Previous studies of palms have shown that the effects of topography on the distribution and abundance of palms depend on species and varies between sites (Svenning Reference Svenning1999). It was suggested that topography has no direct influence on species but that slope is correlated with other environmental variables such as drainage, which does have an effect (Vormisto et al. Reference Vormisto, Tuomisto and Oksanen2004). This is the most plausible explanation of our results.

The strong correlation found between StH and flooded and moist environments (FP and moist) (Figure 4) suggests that being a tall palm is a functional response to prevent leaves and leaf buds from becoming submerged, making survival and growth in flooded zones possible. Therefore, it is an evasion strategy to avoid anoxic effects caused by water (Garssen et al. Reference Garssen, Baattrup-Pedersen, Voesenek, Verhoeven and Soons2015, Gross et al. Reference Gross, Robson, Lavorel, Albert, Le Bagousse-Pinguet and Guillemin2008, Mommer et al. Reference Mommer, Lenssen, Huber, Visser and De Kroon2006, Voesenek et al. Reference Voesenek, Rijnders, Peeters, Van de Steeg and De Kroon2004). Other studies have demonstrated strong dominance of tall palms in FPs and a low representation of medium and small palms in these environments (Balslev et al. Reference Balslev, Kahn, Millan, Svenning, Kristiansen, Borchsenius, Pedersen and Eiserhardt2011, Kahn & De Castro Reference Kahn and De Castro1985). On a local scale in Peru, the composition of species is determined by moisture at the soil surface (Normand et al. Reference Normand, Vormisto, Svenning, Grandez and Balslev2006); the results of this work show that these variables also determine the functional composition (Figure 4).

A proportional increase in petiole length with respect to leaf size reduces the overlap between leaves of adjacent Geonoma cuneata and Asterogyne martiana palms (Figure 3) (Chazdon Reference Chazdon1985, Reference Chazdon1986). Therefore, we interpret long petioles in the presence of gaps as the ability of the palms to compete for light, since the length of the petiole increases the diameter of the crown, allowing the palm to optimize the uptake of resources and grow rapidly. In this sense, it has been proposed that canopy palms can reach their adult state only in the presence of gaps because their light requirements increase with ontogeny (Kahn Reference Kahn1986, Salm et al. Reference Salm, Jalles-Filho and Schuck-Paim2005), adapting to the opportunity for growth between disturbances through an acquisitive strategy (Reich et al. Reference Reich, Wright, Cavender-Bares, Craine, Oleksyn, Westoby and Walters2003). However, additional studies are needed to corroborate this relationship given the low statistical support we found (Table 1), and the finding that the distribution of petiole length across communities is phylogenetically structured (Supplement 4), indicating that the observed patterns may be driven by phylogenetic co-variation with other traits.

Long-leaved palm species are well represented in flooded habitats. This trait might represent an adaptation to floodplains to capture more light, as the canopy is more open in flooded environments (Kahn & De Castro Reference Kahn and De Castro1985). However, the abundance of palms with long leaves may be due to the abundance of members of the genus Astrocaryum in FPs, and the correlation could be driven by a phylogenetic artifact (Kahn & De Castro Reference Kahn and De Castro1985). On the other hand, leaf size may have an allometric correlation with plant size. Knowing the possible constraining effects of allometry, ontogeny, and phylogeny on leaf size may help us interpret the adaptive significance of this trait (Cornelissen Reference Cornelissen1999).

We found a clear relationship between FD and moist and flooded environments (Table 2). Generally, fruit diameter is related to the capability to colonize new habitats and to recruit in resource-limiting environments (Salgado-Negret & Paz Reference Salgado-Negret, Paz and Salgado-Negret2016). Likewise, it has been documented that seed and fruit size determine the success of the establishment of plants because a relatively large seed mass increases the ability of seedlings to survive drought (Baskin & Baskin Reference Baskin and Baskin2014). In particular, some studies in varzea and igapo forests in the central region of the Brazilian Amazonia have shown that large seeds are adapted to areas with high levels of flooding by promoting rapid seedling establishment (Parolin et al. Reference Parolin, Ferreira and Junk2003). Seed size can affect germination by other means, but evidence that corroborates further relationships between large fruits and FPs is currently limited.

Conclusion

Soil moisture and flood regimes are the main environmental factors that affect the functional trait composition of this Amazonian rainforest palm community. Avoidance strategies allow palms to keep their aerial parts above the water level, preventing anoxic stress under waterlogged conditions. Acaulescence and small palm size are stress tolerance mechanisms on sloping and dry terrain (terra firme) under shady conditions, and palms with these traits do not tolerate flooded areas. Hydrology is one of the main determinants of the functional trait composition of neotropical palms on a local scale, and habitats with prolonged floods have communities of palms in which species are adapted to the stressful ecophysiological conditions that flooding creates.

Acknowledgements

WT thanks the program ‘Socios para la Conservación de la Amazonia Colombiana’ for partially financing his Master at Pontificia Universidad Javeriana. Rodrigo Bernal is thanked for help in organizing the fieldwork and identifying the palms.

Financial support

This work was supported by the Independent Research Fund Denmark (HB grant number 9040-00136B). WLE considers this work a contribution to his VILLUM Young Investigator project ‘Explaining the hyperdiversity of Tropical rainforests using the Tree of Life (TropiToL)’ funded by VILLUM FONDEN (grant 00025354).

Conflict of interest

None

Ethical approval

None

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0266467421000493

References

Balslev, H, Copete, J, Pedersen, D, Bernal, R, Galeano, G, Duque, A, Berrio, J and Sanchez, M (2017) Palm diversity and abundance in the Colombian Amazon. In Myster, R (ed.), Forest Structure, Function and Dynamics in Western Amazonia. New Jersey, NJ: John Wiley & Sons Ltd., pp. 101123.Google Scholar
Balslev, H, Kahn, F, Millan, B, Svenning, JC, Kristiansen, T, Borchsenius, F, Pedersen, D and Eiserhardt, WL (2011) Species diversity and growth forms in tropical American palm communities. The Botanical Review 77, 381425.CrossRefGoogle Scholar
Balslev, H, Kristiansen, SM and Muscarella, R (2019) Palm community transects and soil properties in Western Amazonia. Ecology 100, e02841.CrossRefGoogle ScholarPubMed
Balslev, H, Navarrete, H, Paniagua-Zambrana, N, Pedersen, D, Eiserhardt, W and Kristiansen, T (2010) El uso de transectos para el estudio de comunidades de palmas. Ecología en Bolivia 45, 822.Google Scholar
Baskin, C and Baskin, J (2014) Seeds. Ecology, Biogeography, and Evolution of Dormancy and Germination. Amsterdam: Elsevier.Google Scholar
Blom, C and Voesenek, L (1996) Flooding: the survival strategies of plants. Trends in Ecology & Evolution 11, 290295.CrossRefGoogle Scholar
Chazdon, RL (1985) Leaf display, canopy structure, and light interception of two understory palm species. American Journal of Botany 72, 14931502.CrossRefGoogle Scholar
Chazdon, RL (1986) The costs of leaf support in understory palms: economy versus safety. The American Naturalist 127, 930.CrossRefGoogle Scholar
Clark, DA, Clark, DB, Sandoval, RM and Castro, MVC (1995) Edaphic and human effects on landscape-scale distributions of tropical rain forest palms. Ecology 76, 25812594.CrossRefGoogle Scholar
Conti, G and Díaz, S (2013) Plant functional diversity and carbon storage–an empirical test in semi-arid forest ecosystems. Journal of Ecology 101, 1828.CrossRefGoogle Scholar
Cornelissen, J (1999) A triangular relationship between leaf size and seed size among woody species: allometry, ontogeny, ecology and taxonomy. Oecologia 118, 248255.CrossRefGoogle ScholarPubMed
De Granville, J (1991) Life forms and growth strategies of Guianan palms as related to their ecology. Bulletin de l’Institut Français dʼÉtudes Andines 21, 533548.Google Scholar
Debastiani, VJ and Duarte, LdS (2014) PCPS – an R-package for exploring phylogenetic eigenvectors across metacommunities. Frontiers of Biogeography 6, https://doi.org/10.21425/F5FBG22943.CrossRefGoogle Scholar
Den Dubbelden, KC and Knops, JM (1993) The effect of competition and slope inclination on aboveground biomass allocation of understorey ferns in subtropical forest. Oikos 67, 285290.CrossRefGoogle Scholar
Díaz, S, Lavorel, S, De Bello, F, Quétier, F, Grigulis, K and Robson, TM (2007) Incorporating plant functional diversity effects in ecosystem service assessments. Proceedings of the National Academy of Sciences of the United States of America 104, 2068420689.CrossRefGoogle ScholarPubMed
Dolédec, S, Chessel, D, Ter Braak, CJ and Champely, S (1996) Matching species traits to environmental variables: a new three-table ordination method. Environmental and Ecological Statistics 3, 143166.CrossRefGoogle Scholar
Dray, S and Dufour, AB (2007) The ade4 package: implementing the duality diagram for ecologists. Journal of Statistical Software 22, 120.CrossRefGoogle Scholar
Dray, S, Choler, P, Dolédec, S, Peres-Neto, PR, Thuiller, W, Pavoine, S and Ter Braak, CJ (2014) Combining the fourth-corner and the RLQ methods for assessing trait responses to environmental variation. Ecology 95, 1421.CrossRefGoogle ScholarPubMed
Eiserhardt, WL, Svenning, JC, Kissling, WD and Balslev, H (2011) Geographical ecology of the palms (Arecaceae): determinants of diversity and distributions across spatial scales. Annals of Botany 108, 13911416.CrossRefGoogle Scholar
Faurby, S, Eiserhardt, WL, Baker, WJ and Svenning, J-C (2016) An all-evidence species-level supertree for the palms (Arecaceae). Molecular Phylogenetics and Evolution 100, 5769.CrossRefGoogle Scholar
Finegan, B, Peña-Claros, M, De Oliveira, A, Ascarrunz, N, Bret-Harte, MS, Carreño-Rocabado, G, Casanoves, F, Díaz, S, Eguiguren Velepucha, P and Fernandez, F (2015) Does functional trait diversity predict above-ground biomass and productivity of tropical forests? Testing three alternative hypotheses. Journal of Ecology 103, 191201.CrossRefGoogle Scholar
Galeano, G and Bernal, R (2010) Palmas de Colombia: Guía de Campo. Bogotá, D.C: Universidad Nacional de Colombia, Facultad de Ciencias, Instituto de Ciencias Naturales.Google Scholar
Garnier, E, Lavorel, S, Ansquer, P, Castro, H, Cruz, P, Dolezal, J, Eriksson, O, Fortunel, C, Freitas, H and Golodets, C (2007) Assessing the effects of land-use change on plant traits, communities and ecosystem functioning in grasslands: a standardized methodology and lessons from an application to 11 European sites. Annals of Botany 99, 967985.CrossRefGoogle ScholarPubMed
Garssen, AG, Baattrup-Pedersen, A, Voesenek, LA, Verhoeven, JT and Soons, MB (2015) Riparian plant community responses to increased flooding: a meta-analysis. Global Change Biology 21, 28812890.CrossRefGoogle ScholarPubMed
Gatti, MG, Campanello, PI, Villagra, M, Montti, L and Goldstein, G (2014) Hydraulic architecture and photoinhibition influence spatial distribution of the arborescent palm Euterpe edulis in subtropical forests. Tree Physiology 34, 630639.CrossRefGoogle ScholarPubMed
Gentry, AH (1988) Changes in plant community diversity and floristic composition on environmental and geographical gradients. Annals of the Missouri Botanical Garden, 134. https://doi.org/10.2307/2399464.CrossRefGoogle Scholar
Gratani, L (2014) Plant phenotypic plasticity in response to environmental factors. Advances in Botany 2014, 117.CrossRefGoogle Scholar
Grime, JP (1974) Vegetation classification by reference to strategies. Nature 250, 2631.CrossRefGoogle Scholar
Gross, N, Robson, T, Lavorel, S, Albert, C, Le Bagousse-Pinguet, Y and Guillemin, R (2008) Plant response traits mediate the effects of subalpine grasslands on soil moisture. New Phytologist 180, 652662.CrossRefGoogle ScholarPubMed
Henderson, A (2011) A revision of geonoma (Arecaceae). Phytotaxa 17, 1271.CrossRefGoogle Scholar
Hobbie, SE (2015) Plant species effects on nutrient cycling: revisiting litter feedbacks. Trends in Ecology & Evolution 30, 357363.CrossRefGoogle ScholarPubMed
Ishii, H and Asano, S (2010) The role of crown architecture, leaf phenology and photosynthetic activity in promoting complementary use of light among coexisting species in temperate forests. Ecological Research 25, 715722.CrossRefGoogle Scholar
Kahn, F (1986) Life forms of Amazonian palms in relation to forest structure and dynamics. Biotropica 18, 214218.CrossRefGoogle Scholar
Kahn, F and De Castro, A (1985) The palm community in a forest of Central Amazonia, Brazil. Biotropica 17, 210216.CrossRefGoogle Scholar
Kembel, SW, Cowan, PD, Helmus, MR, Cornwell, WK, Morlon, H, Ackerly, DD, Blomberg, SP and Webb, CO (2010) Picante: R tools for integrating phylogenies and ecology, Bioinformatics 26, 14631464.CrossRefGoogle ScholarPubMed
Kern, CC, Montgomery, RA, Reich, PB and Strong, TF (2014) Harvest-created canopy gaps increase species and functional trait diversity of the forest ground-layer community. Forest Science 60, 335344.CrossRefGoogle Scholar
Kissling, WD, Balslev, H, Baker, WJ, Dransfield, J, Göldel, B, Lim, JY, Onstein, RE and Svenning, JC (2019) PalmTraits 1.0, a species-level functional trait database of palms worldwide. Scientific Data 6, 178191.CrossRefGoogle ScholarPubMed
Kleyer, M, Dray, S, Bello, F, Lepš, J, Pakeman, RJ, Strauss, B, Thuiller, W and Lavorel, S (2012) Assessing species and community functional responses to environmental gradients: which multivariate methods? Journal of Vegetation Science 23, 805821.CrossRefGoogle Scholar
Kristiansen, T, Svenning, JC, Grández, C, Salo, J and Balslev, H (2009) Commonness of Amazonian palm (Arecaceae) species: cross-scale links and potential determinants. Acta Oecologica 35, 554562.CrossRefGoogle Scholar
Lavorel, S, Grigulis, K, McIntyre, S, Williams, NSG, Garden, D, Dorrough, J, Berman, S, Quétier, F, Thébault, A and Bonis, A (2008) Assessing functional diversity in the field–methodology matters! Functional Ecology 22, 134147.Google Scholar
Méndez-Toribio, M, Ibarra-Manríquez, G, Navarrete-Segueda, A and Paz, H (2017) Topographic position, but not slope aspect, drives the dominance of functional strategies of tropical dry forest trees. Environmental Research Letters 12, 085002.CrossRefGoogle Scholar
Mommer, L, Lenssen, JP, Huber, H, Visser, EJ and De Kroon, H (2006) Ecophysiological determinants of plant performance under flooding: a comparative study of seven plant families. Journal of Ecology 94, 11171129.CrossRefGoogle Scholar
Muscolo, A, Bagnato, S, Sidari, M and Mercurio, R (2014) A review of the roles of forest canopy gaps. Journal of Forestry Research 25, 725736.CrossRefGoogle Scholar
Normand, S, Vormisto, J, Svenning, JC, Grandez, C and Balslev, H (2006) Geographical and environmental controls of palm beta diversity in paleo-riverine terrace forests in Amazonian Peru. Plant Ecology 186, 161176.CrossRefGoogle Scholar
Parolin, P, Ferreira, L and Junk, W (2003) Germination characteristics and establishment of trees from Central Amazonian flood plains. Tropical Ecology 44, 155168.Google Scholar
Pearcy, R (2007) Responses of plants to heterogeneous light environments. In Pugnaire, F and Valladares, F (eds.), Functional Plant Ecology. New York, NY: Taylor & Francis Group, pp. 213258.CrossRefGoogle Scholar
Pereira De Souza, H (2019) Estrutura Funcional de uma Comunidade de Palmeiras (Arecaceae) ao Longo de Gradientes Ambientais em uma Floresta da Amazônia Central (Masters’ dessertation). Manaus: Instituto Nacional de Pesquisas da Amazônia.Google Scholar
R Core Team (2016) R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna. https://www.R-project.org/ Google Scholar
Reich, PB, Wright, I, Cavender-Bares, J, Craine, J, Oleksyn, J, Westoby, M and Walters, M (2003) The evolution of plant functional variation: traits, spectra, and strategies. International Journal of Plant Sciences 164, S143S164.CrossRefGoogle Scholar
Salgado-Negret, B and Paz, H (2016) Escalando de los rasgos funcionales a procesos poblacionales, comunitarios y ecosistémicos. In Salgado-Negret, B (ed.), La Ecología Funcional como Aproximación al Estudio, Manejo y Conservación de la Biodiversidad: Protocolos y Aplicaciones. Bogotá, D.C: Instituto de Investigación de Recursos Biológicos Alexander von Humboldt, pp. 1235.Google Scholar
Salgado-Negret, B, Canessa, R, Valladares, F, Armesto, JJ and Pérez, F (2015) Functional traits variation explains the distribution of Aextoxicon punctatum (Aextoxicaceae) in pronounced moisture gradients within fog-dependent forest fragments. Frontiers in Plant Science 6, 511.CrossRefGoogle ScholarPubMed
Salm, R, Jalles-Filho, E and Schuck-Paim, C (2005) A model for the importance of large arborescent palms in the dynamics of seasonally-dry Amazonian forests. Biota Neotropica 5, 151156.CrossRefGoogle Scholar
Schliep, KP (2011) phangorn: phylogenetic analysis in R. Bioinformatics 27, 592593.CrossRefGoogle ScholarPubMed
Sokal, R and Rohlf, J (2000) Biometry. The Principles and Practice of Statistics in Biological Research. New York, NY: W.H. Freeman and Company.Google Scholar
Svenning, JC (1999) Microhabitat specialization in a species-rich palm community in Amazonian Ecuador. Journal of Ecology 87, 5565.CrossRefGoogle Scholar
Svenning, JC (2001) On the role of microenvironmental heterogeneity in the ecology and diversification of neotropical rain-forest palms (Arecaceae). The Botanical Review 67, 153.CrossRefGoogle Scholar
Ter Steege, H, Pitman, NC, Sabatier, D, Baraloto, C, Salomão, RP, Guevara, JE, Phillips, OL, Castilho, CV, Magnusson, WE and Molino, JF (2013) Hyperdominance in the Amazonian tree flora. Science 342, 1243092.CrossRefGoogle ScholarPubMed
Terborgh, J and Andresen, E (1998) The composition of Amazonian forests: patterns at local and regional scales. Journal of Tropical Ecology 14, 645664.CrossRefGoogle Scholar
Villota, H (2005) Geomorfología Aplicada a Levantamientos Edafológicos y Zonificación Física de Tierras. Bogotá: Instituto Geográfico Agustin Codazzi (IGAC).Google Scholar
Voesenek, L, Rijnders, J, Peeters, A, Van de Steeg, H and De Kroon, H (2004) Plant hormones regulate fast shoot elongation under water: from genes to communities. Ecology 85, 1627.CrossRefGoogle Scholar
Voesenek, LA and Bailey-Serres, J (2015) Flood adaptive traits and processes: an overview. New Phytologist 206, 5773.CrossRefGoogle ScholarPubMed
Vormisto, J, Tuomisto, H and Oksanen, J (2004) Palm distribution patterns in Amazonian rainforests: what is the role of topographic variation? Journal of Vegetation Science 15, 485494.CrossRefGoogle Scholar
Westoby, M (1998) A leaf-height-seed (LHS) plant ecology strategy scheme. Plant and Soil 199, 213227.CrossRefGoogle Scholar
Figure 0

Figure 1. Study area along the Guaviare River in the Colombian Amazon. The triangles represent the position of 29 transects (500×5 m) in which all adult individuals of 25 species of palms were identified and counted.

Figure 1

Table 1. Functional traits and abundances of 25 palm species registered in 29 transects, each covering 500×5 m, equalling 0.25 ha each or a total of 7.25 ha for all transects. The species names are according to the World Flora Online (worldfloraonline.org).

Figure 2

Figure 2. Relative frequency distribution of functional traits in a palm community in floodplain (FP), terrace (Ter), and terra firme forest types along the Guaviare River in the northeastern region of the Colombian Amazon.

Figure 3

Figure 3. Ordination obtained by redundancy analysis (RDA) of the functional trait composition (CWM) in the palm community along the Guaviare River in the Colombian Amazon as a function of environmental variables. The white-tipped arrows represent environmental variables: Gap=presence of tree-fall gaps, Slope=inclination, Moist=soil moisture. The gray triangles are the forest types: TF=terra firme forests, Fp=floodplain forests and Ter=terrace forests. The black-tipped arrows represent traits: LF=lifeform (cespitose=ces, solitary=sol), GF=growth form (acaulescent=aca, erect=ere, climbing=cli), StH= maximum stem height, LN= maximum leaf number, RL= maximum leaf rachis length, PeL= maximum petiole length, FD=maximum fruit diameter, SN=seed number, BS=breeding system (dioecious=dio, monoecious=mon).

Figure 4

Figure 4. Scatter plots of palm functional traits and environmental variables that showed significant Spearman correlations in a palm community in the Amazon rainforest of Colombia, as studied in 29 transects of 500×5 m.

Figure 5

Table 2. Spearman’s rank correlations between environmental variables and functional composition (CWM) of a palm community in the Colombian Amazon.

Supplementary material: File

Trujillo et al. supplementary material

Trujillo et al. supplementary material 1

Download Trujillo et al. supplementary material(File)
File 25.2 KB
Supplementary material: File

Trujillo et al. supplementary material

Trujillo et al. supplementary material 2

Download Trujillo et al. supplementary material(File)
File 316.2 KB
Supplementary material: File

Trujillo et al. supplementary material

Trujillo et al. supplementary material 3

Download Trujillo et al. supplementary material(File)
File 41.8 KB
Supplementary material: File

Trujillo et al. supplementary material

Trujillo et al. supplementary material 4

Download Trujillo et al. supplementary material(File)
File 676.2 KB