Open Access
How to translate text using browser tools
8 February 2018 Effects of Secondary Forest Succession on Amphibians and Reptiles: A Review and Meta-analysis
Michelle E. Thompson, Maureen A. Donnelly
Author Affiliations +
Abstract

Over the past century, humans have cleared the Earth's forests at an alarming rate and intensity. The majority of global forest cover is categorized as secondary forest, and it is becoming increasingly important to consider secondary forests in addition to old-growth forest in conservation planning for biota. We reviewed the literature to synthesize information on amphibian and reptile communities during secondary forest succession. We summarized literature on mechanisms of community change during forest succession and conducted a meta-analysis to estimate effect sizes for species richness and abundance in human-modified landscapes (agriculture, pasture, and plantation) and old-growth forests compared to regenerating secondary forests. Studies reported strong support for differences in species composition among human-modified landscapes, secondary forest, and old-growth forest as well as species-specific responses to successional forest change. Secondary forest generally had higher species richness and abundance than human-modified landscapes, but lower species richness and abundance than old-growth forests. This result was more pronounced in amphibians than reptiles, and effect size of abundance was more variable than richness among studies. Secondary forests have better conservation value than altered habitats, but they do not necessarily hold the same conservation value for species as old-growth forest.

THE extensive degradation of natural systems caused by anthropogenic activities is a pressing global conservation concern (Raven and Wilson, 1992; Williams and Nowak, 1993; Sodhi et al., 2008). There is hope that some of the negative impacts caused by forest loss such as reduction of ecosystem services and loss of biodiversity may be offset by the regeneration of altered landscapes to secondary forests (Pearce, 2001). However, the value of secondary forests to fauna is poorly understood (Gardner et al., 2007a). Over 60% of the world's forests are degraded or are recovering from a major disturbance (FAO, 2015; Fig. 1), and in some regions of the world, secondary forest cover is increasing (Aide and Grau, 2004). Thus, understanding the role, structure, and function of secondary forests in supporting biodiversity is critical for wildlife in the future.

Fig. 1.

Map of percent of primary forest (black) and other naturally regenerated or planted forests (white) as defined by FAO (2015) by continent.

i0045-8511-106-1-10-f01.tif

For decades, there has been a consistent trend of loss in global forest cover. However, in many regions of the world, forest loss is partially mitigated by secondary forest regeneration (Keenan et al., 2015). Shifting social, political, and economic trends are driving reduction in forest cover loss and secondary forest gain. As a consequence of reduced deforestation and an increase in forest regeneration, the global rate of forest loss was reduced by over 50% between the periods of 1990–2000 and 2010–2015 (FAO, 2015). Many countries are experiencing trends of rural to urban migration (Grau et al., 2003; Barbieri and Carr, 2005; McDonald, 2008), changes in forest and conservation policy (Southworth and Tucker, 2001; Kull et al., 2007), or are developing ecotourism (Kull et al., 2007), resulting in abandonment of agriculture and pasture land and promoting natural regeneration and formation of protected areas (Aide and Grau, 2004; Aide et al., 2012).

One of the main consequences of deforestation is biodiversity loss (Brook et al., 2003; Gibson et al., 2011). Worldwide declines have been reported for amphibians and reptiles (Houlahan et al., 2000; Wake and Vredenburg, 2008), and habitat destruction is one of the primary contributing factors to declines (Stuart et al., 2004; Reading et al., 2010; Böhm et al., 2013). Approximately one third of amphibian species are listed as threatened on the IUCN red list (IUCN et al., 2008). Although a full assessment for reptiles has not yet been completed, it is estimated that somewhere between 15% and 36% of reptiles qualify as threatened by IUCN standards (Böhm et al., 2013). The ecological requirements and physiological limitations of amphibians and reptiles make these animals sensitive to environmental changes such as altered vegetation structure and microclimates after deforestation.

In many animal taxa, species richness recovers asymptotically as forest matures, and recovery has been found to occur in approximately the same amount of time as recovery of tree species richness (Dunn, 2004). Thus, the ecological values of secondary forest to fauna may largely depend on the trajectory of vegetation regrowth. For plant species, overcoming the challenges of recolonization involves species overcoming biotic (e.g., competition with exotic species) and abiotic legacies (e.g., altered soil nutrient content, altered hydrology) of disturbance that can vary considerably as a result of disturbance type (e.g., large-scale hurricane, agriculture, pasture), disturbance intensity, and surrounding landscape matrix (e.g., proximity to patches of remnant forest; Lucas et al., 2002; Cramer et al., 2008). The factors that may contribute to recovery of amphibians and reptiles during secondary succession include disperal to secondary forest, changes in forest structure, temperature and humidity, competition, and prey, predator, and parasite fluctuations over the course of forest succession, all of which are directly or indirectly affected by the course of regeneration of the vegetation.

Across animal taxa, there is support for lower diversity in secondary forests than in old-growth forests (Gibson et al., 2011). Often, species composition differs between secondary and old-growth forests. Subsets of old-growth specialist species are absent from secondary forests (Luja et al., 2008; Chazdon et al., 2009; Gibson et al., 2011; Hernández-Ordóñez et al., 2015) but begin to recover as the forest ages (Chazdon et al., 2009). In some cases, recovery can result in new forests with new combinations of species in comparison to historical sites (Lugo and Helmer, 2004). While secondary forests have been found to be a valuable habitat for a wide array of species, species' use of secondary forests is extremely variable among species and sites differing in land use history (Janzen, 2002; Bowen et al., 2007).

These highly variable species responses to forest succession are reflected in the literature on recovery of amphibian and reptile communities. For example, some authors have found similar species richness (Herrera-Montes and Brokaw, 2010; Hilje and Aide, 2012; Cortés-Gómez et al., 2013), while others have found higher species richness in old-growth forest compared to secondary forests (Petranka et al., 1993, 1994; Vallan, 2002; Pawar et al., 2004; Scott et al., 2006; Gardner et al., 2007a; Basham et al., 2016). For abundance, studies have reported similar (Corn and Bury, 1991; Gardner et al., 2007a), lower (Lieberman, 1986; Heinen, 1992), and higher total community abundance in old-growth forest compared to secondary forest (Petranka et al., 1993; Crawford and Semlitch, 2008; Luja et al., 2008). Measures of total abundance and species diversity tend to be variable, but there is an emerging consensus on changes in species composition and interspecific differences in abundance (Ernst and Rödel, 2006; Gardner et al., 2007a; Ficetola et al., 2008; Hawkes and Gregory, 2012; Beirne et al., 2013; Guerra and Aráoz, 2015; Hernández-Ordóñez et al., 2015). There is a distinct difference in amphibian and reptile composition between secondary forest and anthropogenic land use (Gardner et al., 2007a; Luja et al., 2008; Gillespie et al., 2012; Bruton et al., 2013; Cortés-Gómez et al., 2013; Guerra and Aráoz, 2015) and between secondary forest and old-growth forest (Luja et al., 2008; Cortés-Gómez et al., 2013; Hernández-Ordóñez et al., 2015).

The dominance of secondary forest cover is ubiquitous across continents (Fig. 1). However, we currently have poor knowledge of patterns of and mechanisms of community assembly in secondary forests. To better understand general trends of amphibian and reptile communities in secondary forests, we conducted a review and meta-analysis of the literature. We summarized published literature on mechanisms that drive amphibian and reptile community change during secondary forest succession and conducted a meta-analysis of published studies on amphibian and reptile community recovery in secondary forests to determine the overall effect sizes of amphibian and reptile richness and total abundance in old-growth forest and human-modified landscapes (agriculture, pasture, and plantation) compared to secondary forests. We hypothesized that amphibian and reptile species richness and abundance would be higher in secondary forest than human-modified landscapes and lower in secondary forests than in old-growth forests.

MATERIALS AND METHODS

Literature search.—We searched the database Thompson ISI Web of Knowledge (all years through March 2017) for keywords “herpetofauna” or “amphibian∗” or “reptile∗”, in combination with “secondary forest” or “secondary succession” or “forest regeneration” or “regenerating forest” or “logging”. In addition, we searched the literature cited sections of relevant papers found through the database search.

Literature summary of mechanisms.—As a consequence of the lack of research that explicitly tests mechanistic drivers to amphibian and reptile community change during forest regeneration, we were unable to conduct a formal meta-analysis. Instead, we summarized abiotic and biotic trends in secondary forest succession that have the potential to act as mechanisms for amphibian and reptile community change during secondary forest succession and discussed results of the few studies that have that have tested support for these mechanisms.

Meta-analysis.—The term ‘secondary forest' encompasses many land use types ranging from forests regenerating from complete clearing of land for another use to moderate human use for selective logging and agriculture. For our meta-analysis, we defined secondary forest as forest that had been completely cut and was undergoing natural regeneration. We compiled data on estimated time to recovery for species richness for studies that conduced research in different age classes of secondary forest (at least two different replicated age classes of secondary forest) and reference sites (old-growth forest). We calculated “recovery time” as the age or age class reported by the literature where species richness in secondary forest was not significantly different from reference sites.

To calculate effect size of community parameters across studies (average species richness and average abundance of total community), we included studies that compared secondary forest with undisturbed reference sites or a human-modified land use (agriculture, pasture, and plantation) and that used standardized sampling techniques, replication, and reported values on species richness and abundance. We combined all human-modified habitats together in one category because we found too few studies to analyze each type of modified habitat separately. We used reported values of average species richness and average total abundance and standard deviation or we calculated values using data extracted from tables and figures. We calculated the effect sizes across studies by using the log-transformed ratio of means (Hedges et al., 1999). Because we were interested in how anthropogenic land use and reference sites compared to secondary forests, we calculated effect size as the natural log of the ratio of average species richness or average total community abundance in a given land use or undisturbed natural habitat to species richness or abundance in secondary forest. Negative values indicate average species richness or abundance was lower in old-growth forest or human-modified habitat than in secondary forest. We conducted analysis using the ‘escalc' function and random-effects models with the restricted maximum likelihood estimator in package ‘metafor' (Viechtbauer, 2010) in R v3.3.1 (R Core Team, 2016).

RESULTS

Literature summary of mechanisms

Dispersal.—Before any other mechanisms driving community assembly in secondary forests can take place, species must first disperse to secondary forest sites. Compared to other taxa such as birds and mammals, amphibians and reptiles are generally more limited in dispersal capability (Hillman et al., 2014), and limited dispersal may limit their ability to colonize secondary forests. Dispersal is largely affected by geographic distance between patches (Brown and Kodric-Brown, 1977; Ficetola and De Bernardi, 2004), type of matrix between patches (Fahrig and Merriam, 1994; Gascon et al., 1999; Nowakowski et al., 2013), and species-specific behavior and physiology (Lees and Peres, 2009). Species that are highly mobile and resilient to matrix conditions will be more successful in colonizing isolated secondary forest patches. For amphibians and reptiles, differences in microclimates, predation rates, and movement through substrate type can affect dispersal through matrix habitat (Nowakowski et al., 2015; Kay et al., 2016). However, studies on amphibians and reptiles rarely explicitly incorporated matrix type or distance of secondary forest to old-growth forest in analyses (but see Hilje and Aide, 2012).

Forest structure.—Compared to old-growth stands, secondary forests have been found to differ in vegetation structure and leaf litter structure (Lebrija-Trejos et al., 2008; Letcher and Chazdon, 2009; Chazdon, 2014) which are thought to be important habitat components that regulate amphibian and reptile community composition and density (Lieberman, 1986; Heinen, 1992; Herrera-Montes and Brokaw, 2010; Whitfield et al., 2014). The structure of forest vegetation provides species with microhabitats for perching, foraging, breeding, and fleeing predators. Additionally, forest structure and leaf litter structure mediate temperature and humidity on the forest floor; the leaf litter layer is an important habitat feature for amphibians and reptiles in forests. As secondary forest ages, and forest structure becomes more similar to that of old-growth forest, secondary forest may provide more suitable habitat for amphibian and reptile species that are dependent on the characteristics of old-growth forest. Early stages of secondary forests (<20 years after disturbance) tend to have low plant diversity (Letcher and Chazdon, 2009) and young trees, of similar age and size (Budowski, 1965), providing a uniform habitat of canopy height and perch diameter, and in some studies of herpetofauna in secondary forests, these vegetation characteristics have been linked to change in amphibian and reptile communities. For example, vegetation structure features such as canopy cover and abundance of woody plants (Cortés-Gómez et al., 2013; Hernández-Ordóñez et al., 2015) have been linked to amphibian and reptile community composition. In young secondary forests, there is an absence of large, mature buttressed trees, and there is less course woody debris on the forest floor (Kissing and Powers, 2010) than in old-growth forests which are microhabitats that some amphibian and reptile species specialize on (e.g., Norops humilis in Central American tropical forest [Fitch, 1973] and Ensatina eschscholtzii in the Pacific Northwest of the United States [Jones and Aubry, 1985; Butts and McComb, 2000]). Additionally, absence of trees in riparian zones following clearing can also increase sedimentation in streams that may affect amphibian stream communities (Corn and Bury, 1989). Depth of leaf litter is known to affect densities of amphibians and reptiles (Whitfield et al., 2014), and therefore fluctuations in leaf litter among successional stages can also influence community composition. For example, Ash (1997) found that plethodontid salamander abundance returned in concurrence with return of the leaf litter layer. However, leaf litter fall and depth has been shown to recover rapidly during secondary forest succession (Oliviera, 2008; Ostertag et al., 2008), so leaf litter depth may have a greater effect on species composition and abundance in very early stages of regeneration than in later stages of succession.

Temperature and humidity.—As secondary forest ages, temperature decreases and humidity increases (Lebrija-Trejos et al., 2011). Response of ectothermic animals, such as amphibians and reptiles, to habitat change is thought to be influenced by changes in temperature (Tuff et al., 2016). Regulation of body temperature is important for amphibians and reptiles because temperature affects growth, reproduction (Hillman et al., 2009), ecological interactions, and disease susceptibility in ectotherms (Woodhams et al., 2003; Pounds et al., 2006). Additionally, for amphibians, humidity influences distribution because the highly permeable skin of amphibians increases their vulnerability of desiccation, particularly for species that oviposit terrestrially (Duellman, 1988; Hillman et al., 2009). The eggs, surrounded by a gelatinous coat, are also vulnerable to desiccation. Many studies that conducted amphibian and reptile surveys over the course of forest succession suggest that temperature and humidity likely play a large role in the described patterns of amphibian and reptiles they observed (Lieberman, 1986; Welsh, 1990; Heinen, 1992; Vallan, 2002; Rios-López and Aide, 2007; Herrera-Montes and Brokaw, 2010; Hernández-Ordóñez et al., 2015). One study found that forest structure explained the variability in microclimatic data and microclimate explained best the variation in herpetofaunal diversity (Herrera-Montes and Brokaw, 2010). Rittenhouse et al. (2008) found reduced juvenile anuran survival in recent clear-cut areas because of desiccation, but brush piles helped mitigate negative effects of logging by providing cool, humid microhabitats for amphibians. Despite the general consensus that microclimate likely plays a large role in community assembly, there is a lack of research that specifically tests for temperature and humidity as mechanisms for species response to forest succession.

Biotic factors.—Biotic factors such as the effect of prey and predator fluctuations, competition, and parasitism are known to affect species distributions at both local and broad spatial extents (Wisz et al., 2013). In studies in secondary forests, much less attention has been paid to biotic factors compared to abiotic factors. Competition between ecologically close species has been found to increase with increasing levels of human disturbance (Luiselli, 2006). However, Ernst and Rödel (2006) tested the importance of competition in community organization in secondary forests and did not find evidence for competition shaping species assemblage of anurans in regenerating forests. Arthropods, common prey for amphibians and reptiles, change in abundance and diversity during secondary forest succession, but communities are similar to those in old-growth forests after about 25–50 years (Floren and Linsenmair, 2001; Osorio-Pérez et al., 2007; Hopp et al., 2010). Changes in prey abundance may not only affect composition and abundance of species but can also affect behavior. For example, Greene et al. (2008) found that terrestrial prey abundance for salamanders was lower in late successional forests than early successional forests, causing salamanders to move farther from streams to forage in late successional sites. Predator assemblages change over the course of secondary forests regeneration (e.g., birds: Borges, 2007; Karthik et al., 2009). Therefore, it is likely that predation rates differ during forest regeneration. However, little is known about amphibian and reptile predation rates during forest succession.

Meta-analysis

A total of 24 studies met our requirements for meta-analysis of species richness and total abundance in land use, secondary forests, and old-growth forests (Supplemental Appendix A; see Data Accessibility). Sixteen studies included amphibians and 14 included reptiles. Studies were conducted across the globe but mostly clustered in North America, South America, and Australia (Fig. 2). There was an even distribution of age classes of secondary forests included in studies, but 17% of studies did not include information on age of secondary forest (Fig. 3). Estimates for time to recovery for species richness in secondary forest varied from 10–16 years to more than 80 years of regeneration (Fig. 4).

Fig. 2.

Map of study sites included in meta-analysis by country. Black dots indicate the study locations. Points jittered in the northwestern United States to show overlapping locations.

i0045-8511-106-1-10-f02.tif

Fig. 3.

The age distribution of forest included in 20 of the published articles included in the meta-analysis. Four studies did not provide information on secondary forest age.

i0045-8511-106-1-10-f03.tif

Fig. 4.

Published estimates of time to recovery (years) of amphibian and reptile species richness. Arrow under Petranka et al. (1994) indicates that more than 80 years were required for species richness to recover.

i0045-8511-106-1-10-f04.tif

The effect size across all studies for average amphibian species richness was significantly higher in undisturbed habitats compared to secondary forests. Sites with other types of land use had significantly lower species richness than secondary forests (Fig. 5). However, there was no significant difference in abundance of amphibians between secondary forest and sites of anthropogenic land use. For reptiles, we did not find statistically significant trends in species richness. Results show only a suggestive trend of a positive effect of old-growth forest and negative effect of modified habitat on species richness compared to secondary forest (Fig. 5). We did not find any trends in the comparison of average abundance among secondary forest and old-growth forest and human-modified land use sites for reptiles; there was substantial variation among studies (Fig. 5).

Fig. 5.

Mean effect sizes (and 95% CIs) for the comparison of amphibian and reptile mean species richness and mean abundance in secondary forest to old-growth forest (closed circles) and human-modified habitat (open circles). Response ratios were calculated as the natural log of the ratio of average species richness or average abundance in a given human-modified land use or old-growth forest habitat to species richness or abundance in secondary forest. Negative values indicate average species richness or abundance was lower in areas of old-growth forest or human-modified habitat than in secondary forest. For amphibian richness: Nold-growth = 10, Nland use = 5, and for reptile richness: Nold-growth = 11, Nland use = 7. For amphibian abundance: Nold-growth = 10, Nland use = 6, and for reptile abundance: Nold-growth = 9, Nland use = 7.

i0045-8511-106-1-10-f05.tif

DISCUSSION

With increasing reliance on secondary forest for conservation planning and maintaining biodiversity, it is imperative that we understand how animal communities assemble over the course of forest regeneration. Here, we report the state of knowledge on amphibian and reptile community response over the course of secondary forest succession and summarize information on potential mechanisms for observed patterns in the literature. We found that, in general, old-growth forest tends to have more species than secondary forest and human-modified habitat less species than secondary forest. Secondary forests have better conservation value than altered habitat, but they do not necessarily hold the same conservation value for species as old-growth forest. However, there was substantial variation among studies, especially for reptiles. Our finding of significant differences in community response to secondary forest succession for amphibians but not reptiles suggests that amphibians and reptiles may be affected differently by environmental factors associated with secondary forest succession and supports why they should be considered separately in studies of communities, ecosystems, and landscapes.

Secondary forests provide suitable habitat for many amphibian and reptile species, but there is substantial variation in time to recovery of the animal community. Several studies reported that amphibian and reptile communities recover relatively rapidly. Others reported a period of at least 80 years to recovery (Fig. 4). However, in these forests that are deemed “recovered,” secondary forests may have similar species richness as old-growth forests but secondary forests may not provide suitable habitat for every species in the regional species pool. Some species appear to be unique only to old-growth forests (Barlow et al., 2007; Luja et al., 2008; Gibson et al., 2011). It is critical to identify the old-growth specialists in order to make appropriate conservation decisions for species most at risk. Additionally, it is unclear if amphibian and reptile populations in secondary forest patches are being maintained by internal recruitment, immigration from nearby mature forest, or a combination of the two processes. Although secondary forests do not provide suitable habitat to maintain populations of some species, they may still have other positive effects in comparison to matrix habitat such as increasing connectivity between older forest patches, providing less resistance to movement than matrix habitat, and acting as good corridors for dispersal (Nowakowski et al., 2013).

In some cases, land-use legacy and current surrounding landscape conditions may cause the trajectory of community assembly to vary from historic old-growth conditions. The variation in recovery trajectory has been recorded in plant communities (Janzen, 2002; Cramer et al., 2008). Time to recovery for a forest can also depend on the life zone. In the tropics, vegetation in dry forest recovers more rapidly than wet forest, and cloud forest recovers the slowest of the three forest types (Janzen, 2002). The variation in vegetation trajectory and recovery time is likely to affect amphibian and reptile communities. For example, a species may be less inclined to disperse through or populate a pasture or early stage secondary forest in lowland wet forest habitat than in lowland dry forest habitat because the microclimate conditions in the recently modified landscapes and old-growth forests sites are substantially more disparate in lowland wet forests than dry forests (i.e., hot, dry microclimates; Janzen, 2002).

Not all species in a community respond the same over the course of secondary forest succession. There was a common trend across studies of species-specific effects. These species-specific effects are likely a contributing factor in why many studies found statistically nonsignificant effects between treatments and reference sites and why we found such variation in effect size across studies, especially for the measure of total abundance. Species that are disturbance specialists can weaken observed effects for whole community analysis (Thompson et al., 2016). Some trends in interspecific differences can be explained by particular ecological traits such as tolerance to harsh microclimates, breeding requirements, and other habitat associations. Species that are more resistant to desiccation (Ash, 1997) and species with high thermal tolerances and metabolic rates (Rios-López and Aide, 2007) can tolerate recently disturbed habitats that have high solar irradiation and warm, dry microclimatic conditions. Arboreal species of amphibians (Rios-López and Aide, 2007) and reptiles (Enge and Marion, 1986) increase in abundance and diversity with the return of woody vegetation. One of the most evident trends in the relationship between species traits and forest succession is effect of breeding habitat of anurans. Species with specific breeding habitats and with terrestrial breeding habits are more confined to old-growth forests (Vallan, 2002; Gardner et al., 2007b) whereas pool-breeding species are often able to exploit matrix habitat (Tocher et al., 2002). Terrestrial breeding anurans will likely be one of the groups most at risk in coming decades because of their adverse response to both habitat change (Nowakowski et al., 2017) and climate change (Donnelly and Crump, 1998). However, the presence of many terrestrial breeding anurans in later stages of secondary forest provides hope that secondary forest sites can eventually provide suitable habitat to maintain diversity of these species.

One of the main findings of our review is that there are enormous gaps in the understanding of amphibian and reptile community assembly over the course of secondary forest succession. In addition to calling attention to the dire need for more research, we suggest several recommendations for future studies. First, researchers should pay careful attention to study design. We found few studies that focused on amphibians and reptiles in secondary forest that had well-constructed experimental design, controls, and replication to adequately test hypotheses (and see review by Gardner et al., 2007b). Second, in future studies it is important to try to understand underlying causes of variation and explicitly test mechanisms driving trends in community change. Past studies characterize community patterns over the course of forest regeneration and suggest hypotheses to explain observed patterns. Future work should focus on the underlying processes generating the patterns and to evaluate the strength of mechanisms relative to one another. Success in planning conservation strategies not only depends on knowledge of patterns but the mechanisms driving the patterns (Cushman, 2006; Gardner et al., 2007a). Lastly, it is important to establish long-term research projects. Although several long-term studies on vegetation regeneration in secondary forest exist (Burslem et al., 2000; Sheil, 2001; Chazdon et al., 2007), we know of no long-term research on amphibians and reptiles along the course of secondary forest succession. Many studies stretch over one or two field seasons (one or two years) and substitute space for time by using chronosequences. While these methods are valid and provide valuable inference, ideally, long-term research programs should be established to tease out ecological trends from stochastic fluctuations, detect gradual changes, and detect small but biologically relevant effect sizes.

Future conservation management planning will have to use approaches that integrate the conservation of remaining patches of old-growth forest and surrounding secondary forests to preserve the greatest biodiversity possible. To integrate predictions of the value of secondary forest for a given conservation area, we first need to know for what species secondary forest can provide habitat that can maintain stable populations, and what the most important habitat features are to species at risk of decline. Secondary forest is the dominant global forest cover, is increasing in some regions, and it has been posed that secondary forest may mitigate for biodiversity loss from deforestation (Wright and Muller-Landau, 2006). However, the potential of secondary forests to serve as safety nets for biodiversity is the subject of formidable debate (Laurence, 2007). We still do not know to what extent secondary forest may mitigate for species loss, especially for relatively understudied taxa like amphibians and reptiles. It is urgent to evaluate the capability of secondary forests to host biodiversity comparable to old-growth forests and understand the mechanisms by which communities assemble in secondary forests, especially for taxa at high risk of extinction such as amphibians and reptiles.

DATA ACCESSIBILITY

Supplemental material is available at  http://www.copeiajournal.org/ch-17-654.

ACKNOWLEDGMENTS

Support for M. Thompson was provided by funds from the Fulbright Student Research Award, the College of Arts, Sciences and Education at FIU, and a FIU DEA Fellowship from the University Graduate School.

LITERATURE CITED

1.

Aide, T. M., M. L. Clark, H. R. Grau, D. López-Carr, M. A. Levy, D. Redo, M. Bonilla-Moheno, G. Riner, M. J. Andrade-Núñez, and M. Muñiz. 2012. Deforestation and reforestation of Latin America and the Caribbean (2001–2010). Biotropica 45:262–71. Google Scholar

2.

Aide, T. M., and H. R. Grau. 2004. Globalization, migration, and Latin American ecosystems. Science 305:1915–1916. Google Scholar

3.

Ash, A. 1997. Disappearance and return of plethodontid salamanders to clearcut plots in the southern Blue Ridge Mountains. Conservation Biology 11:983–989. Google Scholar

4.

Barbieri, A. F., and D. L. Carr. 2005. Gender-specific out-migration, deforestation and urbanization in the Ecuadorian Amazon. Global and Planetary Change 47:99–110. Google Scholar

5.

Barlow, J., T. A. Gardner, I. S. Araujo, T. C. Ávila-Pires, A. B. Bonaldo, J. E. Costa, M. C. Esposito, L. V. Ferreira, J. Hawes, M. I. M. Hernandez, M. S. Hoogmoed, R. N. Leite, N. F. Lo-Man-Hung, J. R. Malcolm, M. B. Martins, L. A. M. Mestre, R. Miranda-Santos, A. L. Nunes-Gutjahr, W. L. Overal, L. Parry, S. L. Peters, M. A. Ribeiro-Junior, M. N. F. da Silva, C. da Silva Motta, and C. A. Peres. 2007. Quantifying the biodiversity value of tropical primary, secondary, and plantation forests. Proceedings of the National Academy of Sciences of the United States of America 104:18555–18560. Google Scholar

6.

Basham, E. W., P. González del Pliego, A. R. Acosta-Galvis, P. Woodcock, C. A. Medina Uribe, T. Haugaasen, J. J. Gilroy, and D. P. Edwards. 2016. Quantifying carbon and amphibian co-benefits from secondary forest regeneration in the Tropical Andes. Animal Conservation 19:548–560. Google Scholar

7.

Beirne, C., O. Burdekin, and A. Whitworth. 2013. Herpetofaunal response to anthropogenic habitat change within a small forest reserve in Eastern Ecuador23:209–219. Google Scholar

8.

Böhm, M., B. Collen, J. E. M. Baillie… G. Zug . 2013. The conservation status of the world's reptiles. Biological Conservation 157:372–385. Google Scholar

9.

Borges, S. H. 2007. Assemblages in secondary forests developing after slash-and-burn agriculture in the Brazilian Amazon. Journal of Tropical Ecology 23:469–477. Google Scholar

10.

Bowen, M. E., C. A. McAlpine, A. P. N. House, and G. C. Smith. 2007. Regrowth forests on abandoned agricultural land: a review of their habitat values for recovering forest fauna. Biological Conservation 140:273–296. Google Scholar

11.

Brook, B. W., N. S. Sodhi, and P. K. Nq. 2003. Catastrophic extinctions follow deforestation in Singapore. Nature 24:420–426. Google Scholar

12.

Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445–449. Google Scholar

13.

Bruton, M. J., C. A. McAlpine, and M. Maron. 2013. Regrowth woodlands are valuable habitat for reptile communities. Biological Conservation 165:95–103. Google Scholar

14.

Budowski, G. 1965. Distribution of tropical American rain forest species in the light of successional processes. Turrialba 15:40–42. Google Scholar

15.

Burslem, D. F. R. P., T. C. Whitmore, and G. C. Brown. 2000. Short-term effects of cyclone impact and long-term recovery of tropical rain forest on Kolombangara, Solomon Islands. Journal of Ecology 88:1063–1078. Google Scholar

16.

Butts, S. R., and W. C. McComb. 2000. Associations of forest-floor vertebrates with coarse woody debris in managed forests of western Oregon. Journal of Wildlife Management 64:95–104. Google Scholar

17.

Chazdon, R. L. 2014. Second Growth: The Promise of Tropical Forest Regeneration in an Age of Deforestation. University of Chicago Press, Chicago. Google Scholar

18.

Chazdon, R. L., S. G. Letcher, M. van Breugel, M. Martínez-Ramos, F. Bongers, and B. Finegan. 2007. Rates of changes in tree communities of secondary neotropical forests following major disturbances. Philosophical Transactions of the Royal Society B 362:273–289. Google Scholar

19.

Chazdon, R. L., C. A. Peres, D. Dent, D. Sheil, A. E. Lugo, D. Lamb, N. E. Stork, and S. E. Miller. 2009. The potential for species conservation in tropical secondary forests. Conservation Biology 23:1406–1417. Google Scholar

20.

Corn, P. S., and R. B. Bury. 1989. Logging in western Oregon: reponses of headwater habitats and stream amphibians. Forest Ecology and Management 29:39–57. Google Scholar

21.

Corn, P. S., and R. B. Bury. 1991. Terrestrial amphibian communities in the Oregon Coast Range, p. 241–254. In: Wildlife and Vegetation of Unmanaged Douglas-Fir Forests. L. F. Ruggiero, K. B. Aubry, A. B. Carey, and M. H. Huff (eds.). General Technical Report PNW-GTR-285, USDA, Forest Service, Pacific Northwest Research Station, Portland, Oregon. Google Scholar

22.

Cortés-Gómez, A. M., F. Castro-Herrera, and J. N. Urbina-Cardona. 2013. Small changes in vegetation structure create great changes in amphibian ensembles in the Colombian Pacific rainforest. Tropical Conservation Science 6:749–769. Google Scholar

23.

Cramer, V. A., R. J. Hobbs, and R. J. Standish. 2008. What's new about old fields? Land abandonment and ecosystem assembly. Trends in Ecology and Evolution 23:104–112. Google Scholar

24.

Crawford, J. A., and R. D. Semlitsch. 2008. Post-disturbance effects of even-aged timber harvest on stream salamanders in southern Appalachian forests. Animal Conservation 11:369–376. Google Scholar

25.

Cushman, S. A. 2006. Effects of habitat loss and fragmentation on amphibians: a review and prospectus. Biological Conservation 128:231–240. Google Scholar

26.

Donnelly, M. A., and M. L. Crump. 1998. Potential effects of climate change on two neotropical amphibian assemblages. Climatic Change 39:541–561. Google Scholar

27.

Duellman, W. E. 1988. Patterns of species diversity in anuran amphibians in the American tropics. Annals of the Missouri Botanical Garden 75:79–104. Google Scholar

28.

Dunn, R. R. 2004. Recovery of faunal communities during tropical forest regeneration. Conservation Biology 18:302–309. Google Scholar

29.

Enge, K. M., and W. R. Marion. 1986. Effects of clearcutting and site preparation on herpetofauna of a north Florida flatwoods. Forest Ecology and Management 14:177–192. Google Scholar

30.

Ernst, R., and M. O. Rödel. 2006. Community assembly and structure of tropical leaf litter anurans. Ecotropica 12:113–129. Google Scholar

31.

FAO (Food and Agriculture Organization). 2015. Global Forest Resources Assessment 2015: How are the world's forests changing? Second edition. FAO, Rome. Google Scholar

32.

Fahrig, L., and G. Merriam. 1994. Conservation of fragmented populations. Conservation Biology 8:50–59. Google Scholar

33.

Ficetola, G. F., and F. De Bernardi. 2004. Amphibians in a human-dominated landscape: the community structure is related to habitat features and isolation. Biological Conservation 119:219–230. Google Scholar

34.

Ficetola, G. F., D. Furlani, G. Colombo, and F. De Bernardi. 2008. Assessing the value of secondary forest for amphibians: Eleutherodactylus frogs in a gradient of forest alteration. Biodiversity and Conservation 17:2185–2195. Google Scholar

35.

Fitch, H. S. 1973. A field study of Costa Rican lizards. The University of Kansas Science Bulletin 50:39–126. Google Scholar

36.

Floren, A., and K. E. Linsenmair. 2001. The influence of anthropogenic disturbances on the structure of arboreal arthropod communities. Plant Ecology 153:153–167. Google Scholar

37.

Gardner, T. A., J. Barlow, and C. A. Peres. 2007b. Paradox, presumption and pitfalls in conservation biology: the importance of habitat change for amphibians and reptiles. Biological Conservation 138:166–179. Google Scholar

38.

Gardner, T. A., M. A. Ribeiro-Júnior, J. Barlow, T. C. S. Ávila-Pires, M. S. Hoogmoed, and C. A. Peres. 2007a. The value of primary, secondary, and plantation forests for a neotropical herpetofauna. Conservation Biology 21:775–787. Google Scholar

39.

Gascon, C., T. E. Lovejoy, R. O. Bierregaard, J. R. Malcolm, P. C. Stouffer, H. L. Vasconcelos, W. F. Laurance, B. Zimmerman, M. Tocher, and S. Borges. 1999. Matrix habitat and species richness in tropical forest remnants. Biological Conservation 91:223–229. Google Scholar

40.

Gibson, L., T. M. Lee, L. P. Koh, B. W. Brook, T. A. Gardner, J. Barlow, C. A. Peres, C. J. A. Bradshaw, W. F. Laurance, T. E. Lovejoy, and N. S. Sodhi. 2011. Primary forests are irreplaceable for sustaining tropical biodiversity. Nature 478:378–383. Google Scholar

41.

Gillespie, G. R., E. Ahmad, B. Elahan, A. Evans, M. Ancrenaz, B. Goossens, and M. P. Scroggie. 2012. Conservation of amphibians in Borneo: relative value of secondary tropical forest and non-forest habitats. Biological Conservation 152:136–144. Google Scholar

42.

Grau, H. R., T. M. Aide, J. K. Zimmerman, J. R. Thomlinson, E. Helmer, and X. M. Zou. 2003. The ecological consequences of socioeconomic and land-use changs in postagriculture Puerto Rico. BioScience 53:1159–1168. Google Scholar

43.

Greene, B. T., W. H. Lowe, and G. E. Likens. 2008. Forest succession and prey availability influence the strength and scale of terrestrial-aquatic linkages in a headwater salamander system. Freshwater Biology 53:2234–2243. Google Scholar

44.

Guerra, C., and E. Aráoz. 2015. Amphibian diversity increases in an heterogeneous agricultural landscape. Acta Oecologica 69:78–86. Google Scholar

45.

Hawkes, V. C., and P. T. Gregory. 2012. Temporal changes in the relative abundance of amphibians relative to riparian buffer width in western Washington, USA. Forest Ecology and Management 274:67–80. Google Scholar

46.

Hedges, L. V., J. Gurevitch, and P. S. Curtis. 1999. The meta-analysis of response ratios in experimental ecology. Ecology 80:1150–1156. Google Scholar

47.

Heinen, J. T. 1992. Comparisons of the leaf litter herpetofauna in abandoned cacao plantations and primary rain forest in Costa Rica: some implications for faunal restoration. Biotropica 24:431–439. Google Scholar

48.

Hernández-Ordóñez, O., N. Urbina-Cardona, and M. Martínez-Ramos. 2015. Recovery of amphibian and reptile assemblages during old-field succession of tropical rain forests. Biotropica 47:377–388. Google Scholar

49.

Herrera-Montes, A., and N. Brokaw. 2010. Conservation value of tropical secondary forest: a herpetofaunal perspective. Biological Conservation 143:1414–1422. Google Scholar

50.

Hilje, B., and T. M. Aide. 2012. Recovery of amphibian species richness and composition in a chronosequence of secondary forests, northeastern Costa Rica. Biological Conservation 146:170–176. Google Scholar

51.

Hillman, S. S., R. C. Drews, M. S. Hedrick, and T. V. Hancock. 2014. Physiological vagility and its relationship to dispersal and neutral genetic heterogeneity in vertebrates. The Journal of Experimental Biology 217:3356–3364. Google Scholar

52.

Hillman, S. S., P. C. Withers, R. C. Drewes, and S. D. Hillyard. 2009. Ecological and Environmental Physiology of Amphibians. Oxford University Press, Oxford. Google Scholar

53.

Hopp, P. W., R. Ottermanns, E. Caron, S Meyer, and M. Rob-Nickoll. 2010. Recovery of litter inhabiting beettle assemblages during forest regeneration in the Atlantic forest of Southern Brazil. Insect Conservation and Diversity 3:103–113. Google Scholar

54.

Houlahan, J. E., C. S. Findlay, B. R. Schmidt, A. H. Meyer, and S. L. Kuzmin. 2000. Quantitative evidence for global amphibian population declines. Nature 404:752–755. Google Scholar

55.

IUCN, Conservation International, and NatureServe. 2008. An Analysis of Amphibians on the 2008 IUCN Red List.  http://www.iucnredlist.org/initiatives/amphibians Google Scholar

56.

Janzen, D. H. 2002. Tropical dry forest: area de Conservación Guanacaste, northwestern Costa Rica, p. 559–583. In: Handbook of Ecological Restoration. Volume 2. Restoration in Practice. M. R. Perrowand A. J. Davy (eds.). Cambridge University Press, Cambridge, UK. Google Scholar

57.

Jones, L. L. C., and K. B. Aubry. 1985. Natural history notes: Ensatina eschscholtzii oregonensis (Oregon ensatina). Reproduction. Herpetological Review 16:26. Google Scholar

58.

Karthik, T., G. G. Veeraswami, and P. K. Samal. 2009. Forest recovery following shifting cultivation: an overview of existing research. Tropical Conservation Science 2:374–387. Google Scholar

59.

Kay, G. M., D. A. Driscoll, D. B. Lindenmayer, S. A. Pulsford, and A. Mortelliti. 2016. Pasture height and crop direction influence reptile movement in an agricultural matrix. Agriculture, Ecosystems and Environment 235:164–171. Google Scholar

60.

Keenan, R. J., G. A. Reams, F. Achard, J. V. de Freitas, A. Grainger, and E. Lindquist. 2015. Dynamics of global forest area: results from the FAO global forest resources assessment 2015. Forest Ecology and Management 352:9–20. Google Scholar

61.

Kissing, L. B., and J. S. Powers. 2010. Coarse woody debris stocks are a function of forest type and stand age in Costa Rica tropical dry forest: long-lasting legacies of previous land use. Journal of Tropical Ecology 26:467–471. Google Scholar

62.

Kull, C. A., C. K. Ibrahim, and T. C. Meredith. 2007. Tropical forest transitions ad globalization: neo-liberalism, migration, tourism, and international conservation agendas. Society and Natural Resources 20:723–737. Google Scholar

63.

Laurence, W. F. 2007. Have we overstated the tropical biodiversity crisis?Trends in Ecology and Evolution 22:65–70. Google Scholar

64.

Lebrija-Trejos, E., F. Bongers, E. A. Pérez-García, and J. A. Meave. 2008. Successional change and resilience of a very dry tropical deciduous forest following shifting agriculture. Biotropica 40:422–431. Google Scholar

65.

Lebrija-Trejos, E., E. A. Pérez-García, J. A. Meave, L. Poorter, and F. Bongers. 2011. Environmental changes during secondary succession in a tropical dry forest in Mexico. Journal of Tropical Ecology 27:477–489. Google Scholar

66.

Lees, A. C., and C. Peres. 2009. Gap-crossing movements predict species occupancy in Amazonian forest fragments. Oikos 118:280–290. Google Scholar

67.

Letcher, S. G., and R. L. Chazdon. 2009. Rapid recovery of biomass, species richness, and species composition in a forest chronosequence in northeastern Costa Rica. Biotropica 41:608–617. Google Scholar

68.

Lieberman, S. S. 1986. The ecology of the leaf litter herpetofauna of a neotropical rain forest: La Selva, Costa Rica. Acta Zoológica Mexicana 15:1–72. Google Scholar

69.

Lucas, R. M., M. Honzak, I. D. Amaral, P. J. Curran, and G. M. Foody. 2002. Forest regeneration on abandonded clearances in central Amazonia. International Journal of Remote Sensing 23:965–988. Google Scholar

70.

Lugo, A. E., and E. Helmer. 2004. Emerging forests on abandoned land: Puerto Rico's new forests. Forest Ecology and Management 190:145–161. Google Scholar

71.

Luiselli, L. 2006. Food niche overlap between sympatric potential competitors increases with habitat alteration at different trophic levels in rain-forest reptiles (omnivorous tortoises and carnivorous vipers). Journal of Tropical Ecology 22:695–704. Google Scholar

72.

Luja, V. H., S. Herrando-Pérez, D. González-Solis, and L. Luiselli. 2008. Secondary rain forests are not havens for reptile species in tropical Mexico. Biotropica 40:747–757. Google Scholar

73.

McAlpine, C. A., M. E. Bowen, G. C. Smith, G. Gramotnev, A. G. Smith, A. Lo Cascio, W. Goulding, and M. Maron. 2015. Reptile abundance, but not species richness, increases with regrowth age and spatial extent in fragmented agricultural landscapes of eastern Australia. Biological Conservation 184:174–181. Google Scholar

74.

McDonald, R. I. 2008. Global urbanization: Can ecologists identify a sustainable way forward?Frontiers in Ecology and the Environment 6:99–104. Google Scholar

75.

Nowakowski, A. J., B. Otero Jiménez, M. Allen, A. Diaz-Escobar, and M. A. Donnelly. 2013. Landscape resistance to movement of the poison frog, Oophaga pumilio, in the lowlands od northeaster Costa Rica. Animal Conservation 16:188–197. Google Scholar

76.

Nowakowski, A. J., M. Veiman-Echeverria, D. J. Kurz, and M. A. Donnelly. 2015. Evaluating connectivity for tropical amphibians using empirically, derived resistance surfaces. Ecological Applications 25:928–942. Google Scholar

77.

Nowakowski, A. J., M. E. Thompson, M. A. Donnelly, and B. D. Todd. 2017. Amphibian sensitivity to habitat modification is associated with population trends and species traits. Global Ecology and Biogeography 26:700–712. Google Scholar

78.

Oliviera, R. R. 2008. When the shifting agriculture is gone: functionality of Atlantic Coastal Forest in abandoned farming sites. Boletim do Museu Paraense Emílio Goeldi. Ciências Humanas 3:213–226. Google Scholar

79.

Osorio-Pérez, K., M. F. Barberena-Arias, and T. M. Aide. 2007. Changes in ant species richness and composition during secondary succession in Puerto Rico. Caribbean Journal of Science 43:244–253. Google Scholar

80.

Ostertag, R., E. Marín-Spiotta, W. L. Silver, and J. Schulten. 2008. Litterfall and decomposition in relation to soil carbon pools along a secondary forest chronosequence in Puerto Rico. Ecosystems 11:701–714. Google Scholar

81.

Pawar, S. S., G. S. Rawat, and B. C. Choudhury. 2004. Recovery of frog and lizard communities following primary habitat alteration in Mizoram, northeast India. BMC Ecology 4:1–18. Google Scholar

82.

Pearce, D. 2001. The economic value of forest ecosystems. Ecosystem Health 7:284–296. Google Scholar

83.

Petranka, J. W., M. P. Brannon, M. E. Hopey, and C. K. Smith. 1994. Effects of timber harvesting on low elevation populations of southern Appalachian salamanders. Forest Ecology and Management 67:135–147. Google Scholar

84.

Petranka, J. W., M. E. Eldridge, and K. E. Haify. 1993. Effects of timber harvesting on Southern Appalachian salamanders. Conservation Biology 7:363–370. Google Scholar

85.

Pounds, J. A., M. R. Bustamante, L. A. Coloma, J. A. Consuegra, M. P. L. Fogden, P. N. Foster, E. La Marca, K. L. Masters, A. Merino-Viteri, R. Puschendorf, S. R. Ron, G. A. Sánchez-Azofeifa, C. J. Still, and B. E. Young. 2006. Widespread amphibian extinctions from epidemic disease driven by global warming. Nature 439:161–167. Google Scholar

86.

R Core Team. 2016. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.  http://www.R-project.org/ Google Scholar

87.

Raven, P. H., and E. O. Wilson. 1992. A fifty-year plan for biodiversity surveys. Science 258:1099–1100. Google Scholar

88.

Reading, C. J., L. M. Luiselli, G. C. Akani, X. Bonnet, G. Amori, J. M. Ballouard, E. Filippi, G. Naulleau, D. Pearson, and L. Rugiero. 2010. Are snake populations in widespread decline?Biology Letters 6:777–780. Google Scholar

89.

Rios-López, N., and T. M. Aide. 2007. Herpetofaunal dynamics during secondary succession. Herpetologica 63:35–50. Google Scholar

90.

Rittenhouse, T. A. G., E. B. Harper, L. R. Rehard, and R. D. Semlitsch. 2008. The role of microhabitats in the desiccation and survival of anurans in recently harvested oak–hickory forest. Copeia 2008:807–814. Google Scholar

91.

Scott, D. M., D. Brown, S. Mahood, B. Denton, A. Silburn, and F. Rakotondraparany. 2006. The impacts of forest clearance on lizard, small mammal and bird communities in the arid spiny forest, southern Madagascar. Biological Conservation 127:72–87. Google Scholar

92.

Sheil, D. 2001. Long-term observations of rain forest succession, tree diversity and responses to disturbance. Plant Ecology 155:183–199. Google Scholar

93.

Sodhi, N. S., D. Bickford, A. C. Diesmos, T. M. Lee, L. P. Koh, B. W. Brook, C. H. Sekercioglu, and C. J. Bradshaw. 2008. Measuring the meltdown: drivers of global amphibian extinction and decline. PLOS ONE 3:1–8. Google Scholar

94.

Southworth, J., and C. Tucker. 2001. The influence of accessibility, local institutions, and socioeconomic factors on forest cover change in the mountains of western Honduras. Mountain Research and Development 21:276–283. Google Scholar

95.

Stuart, S. N., J. S. Chanson, N. A Cox, B. E. Young, A. S. L. Rodrigues, D. L. Fischman, and R. W. Waller. 2004. Status and trends of amphibian declines and extinctions worldwide. Science 306:1783–1786. Google Scholar

96.

Thompson, M. E., A. J. Nowakowski, and M. A. Donnelly. 2016. The importance of defining focal assemblages when evaluating amphibian and reptile response to land use. Conservation Biology 30:249–258. Google Scholar

97.

Tocher, M., C. Gascon, and J. R. Meyer. 2002. Community composition and breeding success of Amazonian frogs in continuous forest and matrix aquatic sites, p. 235–247. In: Lessons from Amazonia: The Ecology and Conservation of a Fragmented Forest. R. O. Bierregaard, C. Gascon, T. E. Lovejoy, and R. Mesquita (eds.). Yale University Press, New Haven, Connecticut. Google Scholar

98.

Tuff, K. T., T. Tuff, and K. F. Davies. 2016. A framework for integrating thermal biology into fragmentation research. Ecology Letters 19:361–374. Google Scholar

99.

Vallan, D. 2002. Effects of anthropogenic environmental changes on amphibian diversity in the rain forests of eastern Madagascar. Journal of Tropical Ecology 18:725–742. Google Scholar

100.

Viechtbauer, W. 2010. Conducting meta-analyses in R with the metafor package. Journal of Statistical Software 36:1–48.  http://www.jstatsoft.org/v36/i03/ Google Scholar

101.

Wake, D. B., and V. T. Vredenburg. 2008. Are we in the midst of the sixth mass extinction? A view from the world of amphibians. Proceedings of the National Academy of Sciences of the United States of America 105:11466–11473. Google Scholar

102.

Welsh, H. H., Jr. 1990. Relictual amphibians and old-growth forests. Conservation Biology 4:309–319. Google Scholar

103.

Whitfield, S. M., K. Reider, S. Greenspan, and M. A. Donnelly. 2014. Litter dynamics regulate population densities in a declinging terrestrial herpetofauna. Copeia 2014:454–461. Google Scholar

104.

Williams, J. D., and R. M. Nowak. 1993. Vanishing species in our own backyard: extinct fish and wildlife of the United States and Canada, p. 115–148. In:The Last Extinction. L. Kaufmanand K. Mallory (eds.). MIT Press, Cambridge, Massachusetts. Google Scholar

105.

Wisz, M. S., J. Pottier, W. D. Kissling, L. Pellissier, J. Lenoir, C. F. Damgaard, C. F. Dormann, M. C. Forchhammer, J. Grytnes, A. Guisan, R. K. Heikkinen, T. T. Høye, I. Kühn, M. Luoto, L. Maiorano, M. Nilsson, S. Normand, E. Öckinger, N. M. Schmidt, M. Termansen, A. Tiimmermann, D. A. Wardle, P. Aastrup, and J. Svenning. 2013. The role of biotic interactions in shaping distributions and realised assemblages of species: implications for species distribution modelling. Biological Reviews of the Cambridge Philosophical Society 88:15–30. Google Scholar

106.

Woodhams, D. C., R. A. Alford, and G. Marantelli. 2003. Emerging disease of amphibians cured by elevated body temperature. Diseases of Aquatic Organisms 55:65–67. Google Scholar

107.

Wright, S. J., and H. C. Muller-Landau. 2006. The future of tropical forest species. Biotropica 38:287–301. Google Scholar
© 2018 by the American Society of Ichthyologists and Herpetologists
Michelle E. Thompson and Maureen A. Donnelly "Effects of Secondary Forest Succession on Amphibians and Reptiles: A Review and Meta-analysis," Copeia 106(1), 10-19, (8 February 2018). https://doi.org/10.1643/CH-17-654
Received: 6 July 2017; Accepted: 1 October 2017; Published: 8 February 2018
Back to Top