Acessibilidade / Reportar erro

Vertical distribution of epiphytic bryophytes in Atlantic Forest fragments in northeastern Brazil

ABSTRACT

The microclimatic gradient established from the forest understory to the canopy provides a range of different conditions for the establishment of bryophytes along the height of a tree. We investigated epiphytic bryophyte communities of four fragments of Atlantic Forest with the aim of describing their vertical zonation and assessing differentiation among the communities of the different fragments. In each fragment, five host trees were selected from which bryophyte samples were collected in four height zones from the base to the canopy. Furthermore, 10 plots were demarcated in each fragment where bryophytes were collected from the understory. In total, 114 bryophyte species were found on the 20 sampled phorophytes, plus an additional 51 species in the understory, for a total of 165 species. Species composition of height zones differed significantly between communities of the trunk base and the canopy. The samples from the understory included 77% of all species. Among all species found, 10 showed a significant preference for a specific height. Around 70% of the bryophyte species grew as mats; this life form occurred in all trees and height zones. The results showed a weak, yet significant, vertical gradient, which differs from what is usually found in the Atlantic Forest.

Keywords
canopy; ecology; liverworts; mosses; tropical rain forest

Introduction

In tropical forests, the epiphytic habit is clearly the most important for bryophytes, where communities colonize trees from the base to the outermost canopy twigs (Richards 1984Richards PW. 1984. The ecology of tropical forest bryophytes. In: Schuster RM. (ed.) New manual of Bryology. Nichinan, Hattori Botanical Laboratory. p. 1233-1270.). This pattern of colonization is facilitated by the elevated moisture in tropical environments, which provides ideal microclimatic conditions for the development of these plants (Frahm 2003Frahm JP. 2003. Manual of Tropical Bryology . Tropical Bryology 23: 1-195.). The variety of microhabitats and the constantly moist microclimates facilitate the cooccurrence of numerous species, which results in high species diversity (Gradstein et al. 2008Gradstein SR, Homeier J, Gansert D. 2008. The Tropical Mountain Forest, patterns and processes in a biodiversity hotspot. Göttingen, Göttingen University Press.).

The microclimatic gradient from the forest understory to the canopy, with significant variation in air temperature, relative humidity, light availability and CO2 concentration, is well documented in the literature (Whitmore 1998Whitmore TC. 1998. An introduction to Tropical Rain Forests. New York, Oxford University Press.; Lowman & Rinker 2004Lowman MD, Rinker HB. 2004. Forest canopies. Oxford, Elsevier Academic Press.). A clear turnover in bryophyte species composition along this vertical gradient on host trees has been documented in tropical rainforest sites (Sporn et al. 2010Sporn SG, Bos MM, Kessler M, Gradstein SR. 2010. Vertical distribution of epiphytic bryophytes in an Indonesian rainforest. Biodiversity and Conservation19: 745-760.). In the Guianas, species turnover along the vertical gradient has been shown to be consistent accross local as well as regional scales, with data from three different localities following the same pattern (Cornelissen & Steege 1989Cornelissen JHC, Steege H. 1989. Distribution and ecology of epiphytic bryophytes and lichens in dry evergreen forest of Guyana. Journal of Tropical Ecology 5: 131-150.; Oliveira et al. 2009Oliveira SM, Steege H, Cornelissen JHC, Gradstein SR. 2009. Niche assembly of epiphytic bryophyte communities in the Guianas: A regional approach. Journal of Biogeography 36: 2076-2084.).

Studies dealing with vertical zonation of bryophytes in the Atlantic Forest have found little differentiation between understory and canopy, since a significant turnover in bryophyte species composition was not present, which can be a consequence of the altered microclimatic conditions in fragmented forests (Germano 2003Germano SR. 2003. Florística e Ecologia das Comunidades de Briófitas em um Remanescente de Floresta Atlântica (Reserva Ecológica do Grajaú, Pernambuco, Brasil). PhD Thesis, Universidade Federal de Pernambuco, Brazil.; Campelo & Pôrto 2007Campelo MJA, Pôrto KC. 2007. Brioflora epífita e epífila da RPPN Frei Caneca, Jaqueira, PE, Brasil. Acta Botanica Brasilica 21: 185-192.; Silva & Pôrto 2013Silva MPP, Pôrto KC. 2013. Bryophyte communities along horizontal and vertical gradients in a human-modified Atlantic forest remnant. Botany91: 155-166.). The closed canopy of undisturbed forests is more favorable to vertical species turnover of epiphytes, because microenvironmental conditions are clearly distinct (Werneck & Espírito-Santo 2002Werneck MS, Epírito-Santo MM. 2002. Species diversity and abundance of vascular epiphytes on Vellozia piresiana in Brazil. Biotropica 34: 51-57. ).

Epiphytic bryophytes with different life forms and with different water conduction and retention structures could be indicators of the height zone where they occurr due to their potential strict reliance on a particular level of water supply, gas exchange and light availability (Proctor 1990Proctor MCF. 1990. The physiological basis of bryophyte production. Botanical Journal of the Linnean Society 104: 61-77.). The strong correlation of life forms with moisture and light conditions has been extensively discussed (Thiers 1988Thiers BM. 1988. Morphological adaptations of the Jungermanniales (Hepaticae) to the tropical rain forest habitat. Journal of the Hattori Botanical Laboratory 64: 5-14.; Sporn et al. 2010Sporn SG, Bos MM, Kessler M, Gradstein SR. 2010. Vertical distribution of epiphytic bryophytes in an Indonesian rainforest. Biodiversity and Conservation19: 745-760.), however, there have been few studies of these patterns in the Atlantic Forest. According to Gimingham & Birse (1957Gimingham CH, Birse EM. 1957. Ecological studies on growth-form in bryophytes: I. Correlations between growth-form and habitat. Journal of ecology 45: 533-545.), cushions, mats and turf can be roughly considered as xeric life forms with increasing drought-tolerance, whereas dendroids, fans, pendants, and wefts are life forms characteristic of wet environments.

In this work, we investigated the epiphytic bryophyte communities of four fragments of Atlantic Forest that belong to the same forest matrix, with the aim of assessing their vertical zonation and the relationship between canopy and understory. We also evaluated species richness and composition among the studied fragments.

Materials and methods

Study area

The study was carried out in the Ecological Reserve Michelin, property of Plantações Michelin da Bahia LTDA, in the municipalities of Igrapiúna and Ituberá (13°48'08''S, 39°10'03''W) in Bahia State, Brazil (Fig. 1). The total area of the reserve is 3,096 ha and includes a range of environments that are part of the Atlantic Forest biome, such as restinga, mangrove and estuary, besides the main tropical rainforest matrix. The region has a humid tropical climate (Af) with an average annual precipitation of 2,051 mm, which is distributed throughout the year. The average temperature varies between 18 °C and 30 °C (CEI/CONDER 1993CEI/CONDER - Centro de Estatística e Informação/Companhia do Desenvolvimento da Região Metropolitana de Salvador. 1993. Informações básicas dos municípios baianos: Região Litoral Sul. Salvador, Secretaria do Planejamento, Ciência e Tecnologia. ). The forest is classified as a Lowland Tropical Rainforest and its altitude varies from sea level to 327 m (Veloso et al. 1991Veloso HP, Rangel Filho ALR, Lima JC. 1991. Classificação da vegetação brasileira, adaptada a um sistema universal. Rio de Janeiro, IBGE.).

Figure 1
Location of the Michelin Ecological Reserve in the state of Bahia, Brazil, with the sampling points indicated.

Sampling design

Bryophyte sampling took place in four forest fragments within this complex: Mata de Pacangê (PAC) - 550 ha; Mata da Vila Cinco (MV) - 180 ha; Mata de Pancada Grande (PG) - 172 ha; and Mata do Luis Inácio (LI) - 140 ha, with a total forest cover of 1,042 ha. In each of the four fragments, five canopy trees were selected, considering size for climbing and bryophyte cover, that were separated from each other by at least 100 m, for a total of 20 trees. Four height zones were established to standardize sampling of vertical zonation, following Pócs (1982)Pócs T. 1982. Tropical forest bryophytes. In: Smith AJE. (ed.) Bryophyte Ecology. London, Chapman and Hall. p. 59-104. . Samples were collected in each height zone of all selected trees, which ammounted to 80 sample units (20 trees x 4 zones). Additionally, 10 plots of 100 m² were established in each forest fragment, where understory bryophytes were sampled from the stems of five young trees of CAP >0.30<1.0 m, up to a maximum of 2 m high, and from leaves of shrubs or young trees (Gradstein et al. 1996Gradstein SR, Hietz P, Lücking R, et al. 1996. How to sample the epiphytic diversity of tropical rain forests? Ecotropica 2: 59-72.).

Plant Material

Samples were identified to species level using identification keys, illustrations and descriptions from specialized literature (Florschütz 1964Florschütz PA. 1964. The Mosses of Suriname. Leiden, E. J. Brill.; Fulford 1963Fulford MH. 1963. Manual of the leafy Hepaticae of Latin America I. Memoirs of the New York Botanical Garden 11: 1-172. , 1966Fulford MH. 1966. Manual of the leafy Hepaticae of Latin America II. Memoirs of the New York Botanical Garden 11: 173-276. , 1968Fulford MH. 1968. Manual of the leafy Hepaticae of Latin America III. Memoirs of the New York Botanical Garden 11: 277-392. , 1976Fulford MH. 1976. Manual of the leafy Hepaticae of Latin America IV. Memoirs of the New York Botanical Garden 11: 393-535. ; Sharp et al. 1994Sharp AJ, Crum H, Eckel PM. 1994. The moss flora of Mexico. Memoirs of the New York Botanical Garden 69: 1-1113.; Buck 1998Buck WR. 1998. Pleurocarpous mosses of the West Indies. Memoirs of the New York Botanical Garden 82: 1-400. ; Dauphin 2003Dauphin G. 2003. Ceratolejeunea. Flora Neotropica Monograph 90: 1-86.; Gradstein & Costa 2003Gradstein SR, Costa DP. 2003. The Hepaticae and Anthocerotae of Brazil. Memoirs of the New York Botanical Garden 87: 1-318.). The classification system used was Crandall-Stotler et al. (2009Crandall-Stotler B, Stotler RE, Long DG. 2009. Phylogeny and classification of the Marchantiophyta. Edinburgh Journal of Botany 66: 155-198.) for liverworts and Goffinet et al. (2009)Goffinet B, Buck WR, Shaw AJ. 2009. Morphology and classification of the Bryophyta. In: Goffinet B, Shaw AJ. Bryophyte Biology 2: 55-138. for mosses. After observation and identification, the samples were deposited in the Herbarium ALCB of the Federal University of Bahia.

To verify changes in life forms and water conduction and retention structures, six life forms were recorded following the classification of Frahm (2003Frahm JP. 2003. Manual of Tropical Bryology . Tropical Bryology 23: 1-195.): cushion, dendroid, mat, pendent, turf and weft. Presence/absence of seven morphological adaptations related to water conduction/retention (Glime 2007Glime JM. 2007. Economic and ethnic uses of bryophytes. In: Flora of North America Editorial Committee. (Eds.). Flora of North America and North of Mexico. Oxford, Oxford University Press. p. 14-41.) were also recorded: long acuminate or piliferous apex, lobules, concave leaves, hyalocysts, tomentose gametophyte, allar cells and papillae.

Data analysis

Species accumulation curves were built using bulk samples of all height zones of each tree, with host tree as surrogate unit of area. We used Analysis of Variance (ANOVA) to test for differences in species richness between height zones in the data subset from the canopy trees (N=80), as well as to test for differences in species richness between fragments in the complete data set (N=120). Furthermore, we performed a Kolmogorov-Smirnov Analysis of Similarities (ANOSIM) with all sample pairs of the complete dataset to detect significant differences in species composition between fragments.

In order to identify intrinsic gradients in community composition, we carried out Detrended Correspondence Analysis (DCA), an indirect ordination method (solely based on species composition of communities) which assumes that species show an unimodal response along the gradient (McCune & Grace 2002Mccune B, Grace JB. 2002. Analysis of ecological communities. Gleneden Beach, MjM Software Design.). The ordination was carried out for the dataset obtained from canopy host trees, first with all fragments and then by fragment. After visual inspection including sample labels, such as height zone, elevation and spatial location (fragment), correlation between the axis and height zone was tested for significance through permutation procedure, since grouping of samples were only detected for height zone. All analyses were performed in R 2.14.1 (R Core Team 2012R Core Team. 2012. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. ISBN 3-900051-07-0, ISBN 3-900051-07-0, http://www.R-project.org/ . 19 Nov. 2012.
http://www.R-project.org/...
).

Indicator Species Analysis (ISA, McCune & Grace 2002Mccune B, Grace JB. 2002. Analysis of ecological communities. Gleneden Beach, MjM Software Design.) was carried out to explore species preference for a specific height zone on host trees and for a given fragment. Indicator species of zones 1 and 2 were classified as understory specialists; indicators species of zones 3 and 4 were classified as canopy specialists. In order to score the occurrence of the life forms and morphological features per height zone, the percentage of occurrence per zone relative to the percentage of incidence of the life form/morphological feature in the species list was used.

Results

A total of 114 bryophyte species were found on the 20 sampled phorophytes, of which 76 belonged to Marchantiophyta (liverworts), distributed among six families and 24 genera, and 38 to Bryophyta (mosses), distributed among 11 families and 21 genera. Including samples from small trees of the understory, the total number of species recorded increases to 165, with 62 moss species belong to 16 families and 32 genera and 103 liverwort species belong to eight families and 30 genera (Tab. S1 in supplementary material). Among the liverworts, Lejeuneaceae was the most represented family with 71 species, while Aneuraceae was the least represented with only one species. Among the mosses, Calymperaceae was the most represented family with 16 especies, while Bryaceae, Fissidentaceae, Hypnaceae and Stereophyllaceae had only one species each. Species accumulation curves (species - host tree) show that a sample of ten trees is sufficient to capture 75% of the number of species that are captured when the number of sampled trees is doubled (i.e., the total of 20 trees; Fig. 2).

Figure 2
Randomized species phorophyte curve for the 114 species found in 20 trees sampled, showing that only half of the effort is needed to sample 75% of the species found in 20 trees.

Species richness did not vary significantly either among fragments, with 89 species on average, or among height zones (P > 0.05). Average number of species per host tree was also similar among fragments (Fig. 3). The greatest similarities in species composition among height zones were found between adjacent zones, and decreased with increasing distance between zones (Tab. 1).

Figure 3
Average number and standard deviation of epiphytic bryophyte species per host tree in each of the four forest fragments.

Table 1
Sørensen index and number of species shared for each pair of height zones, including samples from the base of understory trees (UN).

The DCA analysis of the 114 bryophyte species produced three primary axes with eigenvalues of 0.73, 0.63 and 0.44. Axis I explained 41% of the variation and axis II 14.38% (Fig. 4). On the first axis, sample units from heigh zones 1 and 4 were clearly separated, while zones 2 and 3 exhibited overlap. Pearson correlation of the scores of the first axis for the DCA and height zone was significant (P < 0.001) after permutation test. None of the explanatory variables we measured could be correlated with axis II or III. When the same analysis was applied for fragments separetely, only Pacangê showed a slightly, yet significant, correlation (P < 0.01) between axis I scores and height zones, where axis I had an eigenvalue of 0.56 and explained 91% of the variation (Fig. 5).

Figure 4
Scatterplot showing first and second axes of DCA of 114 species of epiphytic bryophytes in different height zones in the studied fragments. Height zone to which the plot belongs are as follows: triangles = zone 1; plus sign = zone 2; squares = zone 3; circles = zone 4. Eigenvalues were 0.73 (Axis 1), 0.63 (Axis 2). Top right corner: correlation between height zones and DCA values (Axis 1; P = 0.0000008).

Figure 5
Scatterplot showing first and second axes of DCA of 54 species of epiphytic bryophytes in different height zones in Pacangê fragment. Height zone to which the plot belongs are as follows: triangles = zone 1; plus sign = zone 2; squares = zone 3; circles = zone 4. Eigenvalues were 0.74 (Axis 1) and 0.47 (Axis 2). Top right corner: correlation between height zones and the DCA values (Axis 1) in Pacangê fragment (P = 0.003).

Among the 114 species recorded, 10 showed significative preferences for a given height zone (P < 0.05), according to the Indicator Species Analysis (ISA). Most of these species, here classified as specialists, were found at the extremes of the gradient: Calymperes lonchophyllum, Plagiochila aerea, Prionolejeunea scaberula and Taxithelium planum were indicator species of zone 1 and Cheilolejeunea trifaria, Cheilolejeunea xanthocarpa, Frullania caulisequa and Groutiella apiculata were indicator species of zone 4. Ceratolejeunea rubiginosa and Ceratolejeunea fallax were indicators of zones 2 and 3, respectively (Fig. 6).

Figure 6
The relative abundance of individuals of each of the specialist species per height zone in Michelin Ecological Reserve.

About 70% of the bryophyte species grew as mats (Fig. 7). Species with this life form occurred on all trees and in all height zones. Each of the other life forms represented less than 10% of the species. The number of dendrois, pendents and crusts did not vary significantly among zones, while cushions were restricted to zones 2 and 3, and wefts were significantly more abundant in zone 1.

Figure 7
Distribution of life forms along the vertical gradient.

Among the morphological features that can be related to water retention and/or water conduction, lobules and papillae did not vary significatively among height zone. Long acuminate leaf apex, concave leaf shape and well developed alar cells were present more in zone 1, while hyalocysts and tomentose were present more on the higher strata- zones 3 and 4 (Fig. 8).

Figure 8
Distribution of morphological adaptations for water retention and/or conduction along the vertical gradient.

Discussion

The distribution of bryophytes among height zones and understory exhibited a well-known pattern for tropical rain forests; liverworts were more frequent than mosses in all height zones, both in number of species and in abundance, and mosses were mainly concentrated in the understory (Gradstein & Culmsee 2010Gradstein SR, Culmsee H. 2010. Bryophyte diversity on tree trunks in montane forests of Central Sulawesi, Indonesia. Tropical Bryology 31: 95-105.). The dominance of liverworts was due to the high number of species of Lejeuneaceae, which represented 44% of the total number of species. Lejeuneaceae can represent up to 70% of the bryoflora of a tropical rainforest (Cornelissen & Gradstein 1990Cornelissen JHC, Gradstein SR. 1990. On the occurrence of Bryophytes and macrolichens in different lowland rain fo rest types at Mabura Hill, Guyana. Tropical Bryology 3: 29-35.; Gradstein et al. 1990Gradstein SR, Montfoort D, Cornelissen JHC. 1990. Species richness and phytogeography of the bryophyte flora of the Guianas, with special reference to the lowland forest. Tropical Bryology 2: 117-126.), and, together with Plagiochillaceae, comprise the most important pantropical families of liverworts in terms of number of species (Gradstein et al. 2001Gradstein SR, Griffin D, Morales MI, Nadkarni NM. 2001. Diversity and habitat differentiation of mosses and liverworts in the cloud forest of Monteverde, Costa Rica. Caldasia 23: 203-212.; Holz et al. 2002Holz I, Gradstein SR, Heinrichs J, Kappelle M. 2002. Bryophyte diversity, microhabitat differentiation and distribution of life forms in Costa Rican upper montane Quercus forest. Bryologist 105: 334-348.).

The samples from the understory included 77% of all species of bryophytes sampled. This result is in agreement with other studies carried out in the Atlantic Forest, where the turnover of species along the vertical gradient is absent or not pronounced (Germano 2003Germano SR. 2003. Florística e Ecologia das Comunidades de Briófitas em um Remanescente de Floresta Atlântica (Reserva Ecológica do Grajaú, Pernambuco, Brasil). PhD Thesis, Universidade Federal de Pernambuco, Brazil.; Campelo & Pôrto 2007Campelo MJA, Pôrto KC. 2007. Brioflora epífita e epífila da RPPN Frei Caneca, Jaqueira, PE, Brasil. Acta Botanica Brasilica 21: 185-192.; Alvarenga et al. 2009Alvarenga LDP, Pôrto KC, Silva MPP. 2009. Relations between regional-local habitat loss and metapopulation properties of epiphyllous bryophytes in the Brazilian Atlantic Forest. Biotropica 41: 682-691.). Although species zonation along the vertical gradient was not very clear, the changes observed from zone 1 to zone 4 were statistically significant. This differs from what has been found in other studies in the Atlantic Forest in the Northeast Region of Brazil, where no significant differences in species composition along the vertical gradient were found. Generally, the lack of zonation in tropical forests is thought to be related to homogeinization of environmental conditions due to forest fragmentation, fragment isolation and forest exploitation, which changes forest structure by creating openings and therefore decreasing moisture and shade (Alvarenga et al. 2009Alvarenga LDP, Pôrto KC, Silva MPP. 2009. Relations between regional-local habitat loss and metapopulation properties of epiphyllous bryophytes in the Brazilian Atlantic Forest. Biotropica 41: 682-691.). Such alteration allows canopy species to colonize lower strata of the forest and hampers the occurrence of several understory species. The mechanisms that influence species establishment in specific microhabitats are, mostly, physiological (Glime 2007Glime JM. 2007. Economic and ethnic uses of bryophytes. In: Flora of North America Editorial Committee. (Eds.). Flora of North America and North of Mexico. Oxford, Oxford University Press. p. 14-41.). Shade especialists are innibited by long periods of exposure to light and are frequently limited to the understory of a forest, sheltered by the canopy cover of individual trees. Therefore, shade specialists tend experience a drastic reduction in abundance with forest fragmentation (Alvarenga & Pôrto 2007Alvarenga LDP, Pôrto KC. 2007. Patch size and isolation effects on epiphytic and epiphyllous bryophytes in the fragmented Brazilian Atlantic forest. Biological Conservation 134: 415-427.) or disapear altogether after deforestation (Costa 1999Costa DP. 1999. Epiphytic bryophyte diversity in primary and secondary lowland rainforests in southeastern Brazil. Bryologist 102: 320-326.; Acebey et al. 2003Acebey C, Gradstein SR, Krömer T. 2003. Species richness and habitat diversification of bryophytes in submontane rain forest and fallows in Bolivia. Journal of Tropical Ecology 18: 1-16.). Indeed, species classified as shade specialists in this study (zones 1 and 2 indicators) are commonly found in the undersory of other tropical forests and hardly ever in the canopy (Montfoort & Ek 1990Montfoort D, Ek RC. 1990. Vertical distribution and ecology of epiphytic Bryophytes and lichens in a lowland rain forest in French Guiana. MSc Thsis, University of Utrecht, The Netherlands.; Acebey et al. 2003Acebey C, Gradstein SR, Krömer T. 2003. Species richness and habitat diversification of bryophytes in submontane rain forest and fallows in Bolivia. Journal of Tropical Ecology 18: 1-16.; Campelo & Pôrto 2007Campelo MJA, Pôrto KC. 2007. Brioflora epífita e epífila da RPPN Frei Caneca, Jaqueira, PE, Brasil. Acta Botanica Brasilica 21: 185-192.). On the other hand, some species require high levels of solar radiation, such as some highly pigmented species in the families Frullaniaceae and Orthotrichaceae (Glime 2007Glime JM. 2007. Economic and ethnic uses of bryophytes. In: Flora of North America Editorial Committee. (Eds.). Flora of North America and North of Mexico. Oxford, Oxford University Press. p. 14-41.). Species of Frullania are particularly tolerant to dissecation (Romero et al. 2006Romero C, Putz FE, Kitajima K. 2006. Ecophysiology in relation to exposure of pendant epiphytic bryophytes in the canopy of a tropical montane oak forest. Biotropica 38: 35-41.) and are comonly found in the higher zones of trees, and especially in the outer canopy, as we have found in our study with Frullania caulisequa and Groutiella apiculata, being among the few species classified as zone 4 indicators (sun specialists).

The percentage of species classified as specialists according to ISA (8%, 10 spp.) is much lower than what was found in a similar study of epiphytic bryophytes in the Guianas, where 96 species, corresponding to 66% of the species found, were classified as specialists (Oliveira et al. 2009Oliveira SM, Steege H, Cornelissen JHC, Gradstein SR. 2009. Niche assembly of epiphytic bryophyte communities in the Guianas: A regional approach. Journal of Biogeography 36: 2076-2084.). The low percentage of specialists in the present study might be indicative of a high degree of disturbance to the reserve. Because bryophytes are sensitive to changes in the microclimate, habitat fragmentation can negatively affect species richness and composition and increase the representation of generalist species and diminish both sun and shade specialists (Alvarenga & Pôrto 2007Alvarenga LDP, Pôrto KC. 2007. Patch size and isolation effects on epiphytic and epiphyllous bryophytes in the fragmented Brazilian Atlantic forest. Biological Conservation 134: 415-427.). According to Werneck & Espírito-Santo (2002Werneck MS, Epírito-Santo MM. 2002. Species diversity and abundance of vascular epiphytes on Vellozia piresiana in Brazil. Biotropica 34: 51-57. ), the closed canopy of undisturbed forests is more favorable for the establishment of a vertical species turnover of epiphytes, because microenvironmental conditions are more distinct.

The predominant life form found in our study (mats), is in agreement with what has been documented in other tropical forests (Cornelissen & Steege 1989Cornelissen JHC, Steege H. 1989. Distribution and ecology of epiphytic bryophytes and lichens in dry evergreen forest of Guyana. Journal of Tropical Ecology 5: 131-150.; Montfoort & Ek 1990Montfoort D, Ek RC. 1990. Vertical distribution and ecology of epiphytic Bryophytes and lichens in a lowland rain forest in French Guiana. MSc Thsis, University of Utrecht, The Netherlands.; Costa 1999Costa DP. 1999. Epiphytic bryophyte diversity in primary and secondary lowland rainforests in southeastern Brazil. Bryologist 102: 320-326.; Pardow et al. 2012Pardow A, Gehrig-Downie C, Gradstein SR, Lakatos M. 2012. Functional diversity of epiphytes in two tropical lowland rainforests, French Guiana: using bryophyte life-forms to detect areas of high biodiversity. Biodiversity and Conservation 21: 3637-3655. ). It is widely accepted that life forms reflect habitat characteristics, and are closely related to moisture availability (Pócs 1982Pócs T. 1982. Tropical forest bryophytes. In: Smith AJE. (ed.) Bryophyte Ecology. London, Chapman and Hall. p. 59-104. ; Thiers 1988Thiers BM. 1988. Morphological adaptations of the Jungermanniales (Hepaticae) to the tropical rain forest habitat. Journal of the Hattori Botanical Laboratory 64: 5-14.; Proctor 1990Proctor MCF. 1990. The physiological basis of bryophyte production. Botanical Journal of the Linnean Society 104: 61-77.). Life forms such as mats or wefts have a high capacity for water retention, and grow strongly appressed to the surface of the substrate (Glime 2007Glime JM. 2007. Economic and ethnic uses of bryophytes. In: Flora of North America Editorial Committee. (Eds.). Flora of North America and North of Mexico. Oxford, Oxford University Press. p. 14-41.). They are abundant growing along trunks in tropical secondary forests, which have high levels of humidity and light (Sillet et al. 1995Sillet C, Gradstein SR, Griffin D. 1995. Bryophyte diversity of Ficus tree crowns from cloud forest and pasture in Costa Rica. Bryologist 25: 1-260.), as was recorded in the present study.

Life form classifications are based on morphological characters, and so those structures that seem to be adaptations to water conduction and retention can be used as indicators of environmental conditions (Frahm 2003Frahm JP. 2003. Manual of Tropical Bryology . Tropical Bryology 23: 1-195.). The higher frequency of occurrence of some of these morphological adaptations in certain height zones at Michelin ecological reserve reflects the microenvironmental gradient along tree trunks. The predominance of leaves with a long acuminate apex in lower zones, as well as concave leaves and well developed allar cells, was expected since these are typical of understory families like Calymperaceae, Pylaisiadelphaceae and Sematophyllaceae. Among the other morphological characteristics investigated, lobules are present in almost all Lejeuneaceae, which were distributed throughout the entire height gradient with no tendency for preference to a specific height zone. Likewise, papillae were abundant in all zones and revealed no pattern of altitudinal distribution, as was also found in a study of altitudinal zonation in Borneo (Frey et al. 1990Frey W, Gossow R, Kürschner H. 1990. Verteilungsmuster von Lebensformen, wasserleitenden und wasserspeichernden Strukturen in epiphytischen Moosgesellschaften am Mt. Kinabalu (Nord Borneo). Nova Hedwigia 51: 87-119). Until now, lobules and papillae had been successfully used as indicators of environmental conditions in altitudinal zonation, such as in a study in Central Africa (Kürschner & Seifert 1995Kürschner H, Seifert UH. 1995. Wissenschaftliche Ergebnisse der BRYOTROP Expedition nach Zaire und Rwanda 6. Lebensformen und adaptationen zur wasserleitung und wasserspeicherung in epiphytischen moosgesellschaften im östlichen Kongobecken. Tropical Bryology 11: 87-118.). Adaptation to water retention and conduction along the vertical gradient in tropical forests needs further study.

We verified that there is substitution of species of epiphytic bryophytes along the height zones of host trees, however, this zonation is relatively weak in comparison with the results usually found in tropical forest environments. The morphological adaptations for water conduction or retention, as well as life forms, are strongly associated with microhabitat. A better understanding of the spatial distribution of these organisms in Atlantic Forest fragments is essential, since bryophytes have important roles in the proper functioning of ecosystems and can serve as bioindicators for monitoring changes in biodiversity, be they natural or caused by human actions.

Acknowledgements

The authors thank Dr. P.E.A.S. Câmara, of the University of Brasília (UnB), Brazil, for confirmation of the Taxithelium species identifications and Dr. H. ter Steege for comments on data analysis.

References

  • Acebey C, Gradstein SR, Krömer T. 2003. Species richness and habitat diversification of bryophytes in submontane rain forest and fallows in Bolivia. Journal of Tropical Ecology 18: 1-16.
  • Alvarenga LDP, Pôrto KC. 2007. Patch size and isolation effects on epiphytic and epiphyllous bryophytes in the fragmented Brazilian Atlantic forest. Biological Conservation 134: 415-427.
  • Alvarenga LDP, Pôrto KC, Silva MPP. 2009. Relations between regional-local habitat loss and metapopulation properties of epiphyllous bryophytes in the Brazilian Atlantic Forest. Biotropica 41: 682-691.
  • Buck WR. 1998. Pleurocarpous mosses of the West Indies. Memoirs of the New York Botanical Garden 82: 1-400.
  • Campelo MJA, Pôrto KC. 2007. Brioflora epífita e epífila da RPPN Frei Caneca, Jaqueira, PE, Brasil. Acta Botanica Brasilica 21: 185-192.
  • CEI/CONDER - Centro de Estatística e Informação/Companhia do Desenvolvimento da Região Metropolitana de Salvador. 1993. Informações básicas dos municípios baianos: Região Litoral Sul. Salvador, Secretaria do Planejamento, Ciência e Tecnologia.
  • Cornelissen JHC, Gradstein SR. 1990. On the occurrence of Bryophytes and macrolichens in different lowland rain fo rest types at Mabura Hill, Guyana. Tropical Bryology 3: 29-35.
  • Cornelissen JHC, Steege H. 1989. Distribution and ecology of epiphytic bryophytes and lichens in dry evergreen forest of Guyana. Journal of Tropical Ecology 5: 131-150.
  • Costa DP. 1999. Epiphytic bryophyte diversity in primary and secondary lowland rainforests in southeastern Brazil. Bryologist 102: 320-326.
  • Crandall-Stotler B, Stotler RE, Long DG. 2009. Phylogeny and classification of the Marchantiophyta. Edinburgh Journal of Botany 66: 155-198.
  • Dauphin G. 2003. Ceratolejeunea. Flora Neotropica Monograph 90: 1-86.
  • Florschütz PA. 1964. The Mosses of Suriname. Leiden, E. J. Brill.
  • Frahm JP. 2003. Manual of Tropical Bryology . Tropical Bryology 23: 1-195.
  • Frey W, Gossow R, Kürschner H. 1990. Verteilungsmuster von Lebensformen, wasserleitenden und wasserspeichernden Strukturen in epiphytischen Moosgesellschaften am Mt. Kinabalu (Nord Borneo). Nova Hedwigia 51: 87-119
  • Fulford MH. 1963. Manual of the leafy Hepaticae of Latin America I. Memoirs of the New York Botanical Garden 11: 1-172.
  • Fulford MH. 1966. Manual of the leafy Hepaticae of Latin America II. Memoirs of the New York Botanical Garden 11: 173-276.
  • Fulford MH. 1968. Manual of the leafy Hepaticae of Latin America III. Memoirs of the New York Botanical Garden 11: 277-392.
  • Fulford MH. 1976. Manual of the leafy Hepaticae of Latin America IV. Memoirs of the New York Botanical Garden 11: 393-535.
  • Germano SR. 2003. Florística e Ecologia das Comunidades de Briófitas em um Remanescente de Floresta Atlântica (Reserva Ecológica do Grajaú, Pernambuco, Brasil). PhD Thesis, Universidade Federal de Pernambuco, Brazil.
  • Gimingham CH, Birse EM. 1957. Ecological studies on growth-form in bryophytes: I. Correlations between growth-form and habitat. Journal of ecology 45: 533-545.
  • Glime JM. 2007. Economic and ethnic uses of bryophytes. In: Flora of North America Editorial Committee. (Eds.). Flora of North America and North of Mexico. Oxford, Oxford University Press. p. 14-41.
  • Goffinet B, Buck WR, Shaw AJ. 2009. Morphology and classification of the Bryophyta. In: Goffinet B, Shaw AJ. Bryophyte Biology 2: 55-138.
  • Gradstein SR, Costa DP. 2003. The Hepaticae and Anthocerotae of Brazil. Memoirs of the New York Botanical Garden 87: 1-318.
  • Gradstein SR, Culmsee H. 2010. Bryophyte diversity on tree trunks in montane forests of Central Sulawesi, Indonesia. Tropical Bryology 31: 95-105.
  • Gradstein SR, Griffin D, Morales MI, Nadkarni NM. 2001. Diversity and habitat differentiation of mosses and liverworts in the cloud forest of Monteverde, Costa Rica. Caldasia 23: 203-212.
  • Gradstein SR, Hietz P, Lücking R, et al. 1996. How to sample the epiphytic diversity of tropical rain forests? Ecotropica 2: 59-72.
  • Gradstein SR, Homeier J, Gansert D. 2008. The Tropical Mountain Forest, patterns and processes in a biodiversity hotspot. Göttingen, Göttingen University Press.
  • Gradstein SR, Montfoort D, Cornelissen JHC. 1990. Species richness and phytogeography of the bryophyte flora of the Guianas, with special reference to the lowland forest. Tropical Bryology 2: 117-126.
  • Holz I, Gradstein SR, Heinrichs J, Kappelle M. 2002. Bryophyte diversity, microhabitat differentiation and distribution of life forms in Costa Rican upper montane Quercus forest. Bryologist 105: 334-348.
  • Kürschner H, Seifert UH. 1995. Wissenschaftliche Ergebnisse der BRYOTROP Expedition nach Zaire und Rwanda 6. Lebensformen und adaptationen zur wasserleitung und wasserspeicherung in epiphytischen moosgesellschaften im östlichen Kongobecken. Tropical Bryology 11: 87-118.
  • Lowman MD, Rinker HB. 2004. Forest canopies. Oxford, Elsevier Academic Press.
  • Mccune B, Grace JB. 2002. Analysis of ecological communities. Gleneden Beach, MjM Software Design.
  • Montfoort D, Ek RC. 1990. Vertical distribution and ecology of epiphytic Bryophytes and lichens in a lowland rain forest in French Guiana. MSc Thsis, University of Utrecht, The Netherlands.
  • Oliveira SM, Steege H, Cornelissen JHC, Gradstein SR. 2009. Niche assembly of epiphytic bryophyte communities in the Guianas: A regional approach. Journal of Biogeography 36: 2076-2084.
  • Pardow A, Gehrig-Downie C, Gradstein SR, Lakatos M. 2012. Functional diversity of epiphytes in two tropical lowland rainforests, French Guiana: using bryophyte life-forms to detect areas of high biodiversity. Biodiversity and Conservation 21: 3637-3655.
  • Pócs T. 1982. Tropical forest bryophytes. In: Smith AJE. (ed.) Bryophyte Ecology. London, Chapman and Hall. p. 59-104.
  • Proctor MCF. 1990. The physiological basis of bryophyte production. Botanical Journal of the Linnean Society 104: 61-77.
  • R Core Team. 2012. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. ISBN 3-900051-07-0, ISBN 3-900051-07-0, http://www.R-project.org/ 19 Nov. 2012.
    » http://www.R-project.org/
  • Richards PW. 1984. The ecology of tropical forest bryophytes. In: Schuster RM. (ed.) New manual of Bryology. Nichinan, Hattori Botanical Laboratory. p. 1233-1270.
  • Romero C, Putz FE, Kitajima K. 2006. Ecophysiology in relation to exposure of pendant epiphytic bryophytes in the canopy of a tropical montane oak forest. Biotropica 38: 35-41.
  • Sharp AJ, Crum H, Eckel PM. 1994. The moss flora of Mexico. Memoirs of the New York Botanical Garden 69: 1-1113.
  • Sillet C, Gradstein SR, Griffin D. 1995. Bryophyte diversity of Ficus tree crowns from cloud forest and pasture in Costa Rica. Bryologist 25: 1-260.
  • Silva MPP, Pôrto KC. 2013. Bryophyte communities along horizontal and vertical gradients in a human-modified Atlantic forest remnant. Botany91: 155-166.
  • Sporn SG, Bos MM, Kessler M, Gradstein SR. 2010. Vertical distribution of epiphytic bryophytes in an Indonesian rainforest. Biodiversity and Conservation19: 745-760.
  • Thiers BM. 1988. Morphological adaptations of the Jungermanniales (Hepaticae) to the tropical rain forest habitat. Journal of the Hattori Botanical Laboratory 64: 5-14.
  • Veloso HP, Rangel Filho ALR, Lima JC. 1991. Classificação da vegetação brasileira, adaptada a um sistema universal. Rio de Janeiro, IBGE.
  • Werneck MS, Epírito-Santo MM. 2002. Species diversity and abundance of vascular epiphytes on Vellozia piresiana in Brazil. Biotropica 34: 51-57.
  • Whitmore TC. 1998. An introduction to Tropical Rain Forests. New York, Oxford University Press.

Publication Dates

  • Publication in this collection
    10 Nov 2016
  • Date of issue
    Oct-Dec 2016

History

  • Received
    18 Aug 2016
  • Accepted
    21 Oct 2016
Sociedade Botânica do Brasil SCLN 307 - Bloco B - Sala 218 - Ed. Constrol Center Asa Norte CEP: 70746-520 Brasília/DF. - Alta Floresta - MT - Brazil
E-mail: acta@botanica.org.br