Acessibilidade / Reportar erro

Effect of Leaf Quality on Herbivory of Three Atlantic Forest Species

ABSTRACT

The present study evaluated the relationship between herbivory and leaf quality of three abundant tree species (Cupania oblongifolia Mart., Siparuna guianensis Aubl. and Xylopia sericea St. Hill.) in the understory of a eucalyptus plantation and an adjacent Atlantic Rainforest fragment. Herbivory differed among species, since X. sericea was less attacked, while C. oblongifolia and S. guianensis were more. The species showed similar herbivory patterns between the investigated areas (eucalyptus plantation versus forest), although X. sericea showed higher herbivory rates in the forest fragment. Fiber values (NDF, CEL and LIG) varied significantly among the species, being lower in S. guianensis. X. sericea, the species with the highest C/N ratio, higher concentrations of total phenols, smaller leaf area and higher trichome density, showed the lowest leaf herbivory rates, corroborating the hypothesis that herbivory is lower in species with lower nutritional quality and a greater set of leaf defense features.

Keywords:
defenses; insect-plant interactions; leaf nutrients

1. INTRODUCTION AND OBJECTIVES

Herbivores and plants are connected by one of the most important ecological relationships in nature, herbivory (Strong et al., 1984Strong DR, Lawton JH, Southwood R. Insects on plants: community patterns and mechanisms. Massachusetts: Harvard University Press; 1984.; Wilson, 2001Wilson EO. The diversity of life. Toronto: Penguin; 2001.). Existing herbivores, which include arthropods, mollusks, nematodes and vertebrates, consume about 15% of the annual plant biomass production, both in temperate and tropical forests. This makes herbivory a fundamental process for the energy flow between food webs (Agrawal, 2011Agrawal AA. Current trends in the evolutionary ecology of plant defence. Functional Ecology 2011; 25(2): 421-433. 10.1111/j.1365-2435.2010.01796.x
https://doi.org/10.1111/j.1365-2435.2010...
; Cyr & Pace, 1993Cyr H, Pace ML. Magnitude and patterns of herbivory in aquatic and terrestrial ecosystems. Nature 1993; 361(6408): 148-150. 10.1038/361148a0
https://doi.org/10.1038/361148a0...
). In addition, herbivore pressure has allowed the development of different types of plant defenses, which have influenced their palatability to insects (Marquis, 2012Marquis RJ. Uma abordagem geral das defesas das plantas contra ação dos herbívoros. In: Del-Claro K, Torezan-Silingardi V, editores. Ecologia das interações plantas-animais: uma abordagem ecológica-evolutiva. Rio de Janeiro: Technical Books; 2012. p. 55-66.).

Although herbivory may positively affect plants, most often it causes negative effects, slowing their development and their ability to reproduce and survive in natural communities (Boege & Marquis, 2005Boege K, Marquis RJ. Facing herbivory as you grow up: the ontogeny of resistence in plants. Trends in Ecology & Evolution 2005; 20(8): 441-448. 10.1016/j.tree.2005.05.001
https://doi.org/10.1016/j.tree.2005.05.0...
; Coley & Barone, 1996Coley PD, Barone JA. Herbivory and plant defenses in tropical forests. Annual Review of Ecology and Systematics 1996; 27(1): 305-335. 10.1146/annurev.ecolsys.27.1.305
https://doi.org/10.1146/annurev.ecolsys....
). Different plant characteristics can both directly or indirectly influence leaf damage rates by insects, such as leaf age and size, light, seasonality, spatial distribution, abundance, and certain particular characteristics of each species (Ballaré, 2014Ballaré CL. Light regulation of plant defense. Annual Review of Plant Biology 2014; 65: 335-363. 10.1146/annurev-arplant-050213-040145
https://doi.org/10.1146/annurev-arplant-...
; Engelkes et al., 2016Engelkes T, Meisner A, Morriën E, Kostenko O, Van der Putten WH, Macel M. Herbivory and dominance shifts among exotic and congeneric native plant species during plant community establishment. Oecologia 2016; 180(2): 507-517. 10.1007/s00442-015-3472-6
https://doi.org/10.1007/s00442-015-3472-...
; Gonçalves-Alvim et al., 2010Gonçalves-Alvim SJ, Lana TC, Ranieri BD, Silveira FAO, Ribeiro V, Fernandes GW. Growth, defense and herbivory on young leaves of Qualea parviflora (Vochysiaceae) in three different Cerrado habitats. Neotropical Biology and Conservation 2010; 5(2): 86-92. 10.4013/nbc.2010.52.03
https://doi.org/10.4013/nbc.2010.52.03...
; Karolewski et al., 2013Karolewski P, Giertych MJ, Żmuda M, Jagodziński AM, Oleksyn J. Season and light affect constitutive defenses of understory shrub species against folivorous insects. Acta Oecologica 2013; 53: 19-32. 10.1016/j.actao.2013.08.004
https://doi.org/10.1016/j.actao.2013.08....
; Lowman, 1985Lowman MD. Temporal and spatial variability in insect grazing of the canopies of five Australian rainforest tree species. Australian Journal of Ecology 1985; 10(1): 7-24. 10.1111/j.1442-9993.1985.tb00859.x
https://doi.org/10.1111/j.1442-9993.1985...
; Silva & Neves, 2014Silva JO, Neves FS. Insect herbivores associated with an evergreen tree Goniorrhachis marginata Taub. (Leguminosae: Caesalpinioideae) in a tropical dry forest. Brazilian Journal of Biology 2014; 74(3): 623-631. 10.1590/bjb.2014.0093
https://doi.org/10.1590/bjb.2014.0093...
).

Plants with lower nutritional value, for example, low in nutrient concentrations and high in C (carbon) concentrations, are less palatable to herbivorous insects (Coley & Barone, 1996Coley PD, Barone JA. Herbivory and plant defenses in tropical forests. Annual Review of Ecology and Systematics 1996; 27(1): 305-335. 10.1146/annurev.ecolsys.27.1.305
https://doi.org/10.1146/annurev.ecolsys....
; Kurokawa & Nakashizuka, 2008Kurokawa H, Nakashizuka T. Leaf herbivory and decomposability in a Malaysian tropical rain forest. Ecology 2008; 89(9): 2645-2656. 10.1890/07-1352.1
https://doi.org/10.1890/07-1352.1...
; Mundim et al., 2009Mundim FM, Costa AN, Vasconcelos HL. Leaf nutrient content and host plant selection by leaf-cutter ants, Atta laevigata, in a Neotropical savanna. Entomologia Experimentalis et Applicata 2009; 130(1): 47-54. 10.1111/j.1570-7458.2008.00789.x
https://doi.org/10.1111/j.1570-7458.2008...
). Plants also contain a variety of special chemical compounds and morphological traits that can influence their attractiveness to herbivores (Agrawal & Fishbein, 2006Agrawal AA, Fishbein, M. Plant defense syndromes. Ecology 2006; 87(7): S132-S149. 10.1890/0012-9658(2006)87[132:PDS]2.0.CO;2
https://doi.org/10.1890/0012-9658(2006)8...
; Marquis, 2012Marquis RJ. Uma abordagem geral das defesas das plantas contra ação dos herbívoros. In: Del-Claro K, Torezan-Silingardi V, editores. Ecologia das interações plantas-animais: uma abordagem ecológica-evolutiva. Rio de Janeiro: Technical Books; 2012. p. 55-66.; Trigo et al., 2012Trigo JR, Pareja M, Massuda KF. O papel das substâncias químicas nas interações entre plantas e insetos herbívoros. In: Del-Claro K, Torezan-Silingardi H, editores. Ecologia das interações plantas-animais: uma abordagem ecológico-evolutiva. Rio de Janeiro: Technical Books ; 2012. p. 69-88.) such as cellulose and hemicellulose, and special compounds, like tannins (Kitajima et al., 2012Kitajima K, Llorens A-M, Stefanescu C, Timchenko MV, Lucas PW, Wright SJ. How cellulose-based leaf toughness and lamina density contribute to long leaf lifespans of shade-tolerant species. New Phytologist 2012; 195(3): 640-652. 10.1111/j.1469-8137.2012.04203.x
https://doi.org/10.1111/j.1469-8137.2012...
; Linton & Greenaway, 2007Linton SM, Greenaway P. A review of feeding and nutrition of herbivorous land crabs: adaptations to low quality plant diets. Journal of Comparative Physiology 2007; 177(3): 269-286. 10.1007/s00360-006-0138-z
https://doi.org/10.1007/s00360-006-0138-...
).

Certain morphological traits may also affect the palatability of plants to herbivores, such as thorns (Cooper & Ginnet, 1998Cooper SM, Ginnett TF. Spines protect plants against browsing by small climbing mammals. Oecologia 1998; 113(2): 219-221. 10.1007/s004420050371
https://doi.org/10.1007/s004420050371...
), leaf toughness (Kitajima et al., 2012Kitajima K, Llorens A-M, Stefanescu C, Timchenko MV, Lucas PW, Wright SJ. How cellulose-based leaf toughness and lamina density contribute to long leaf lifespans of shade-tolerant species. New Phytologist 2012; 195(3): 640-652. 10.1111/j.1469-8137.2012.04203.x
https://doi.org/10.1111/j.1469-8137.2012...
; Lucas et al., 2000Lucas PW, Turner IM, Dominy NJ, Yamashita N. Mechanical defences to herbivory. Annals of Botany 2000; 86(5): 913-920. 10.1006/anbo.2000.1261
https://doi.org/10.1006/anbo.2000.1261...
) and glandular and non-glandular trichomes (Barônio, 2012Barônio GJ. Pilosidade foliar reduz herbivoria em folhas jovens e maduras de Qualea multiflora Mart. em cerrado stricto sensu. Neotropical Biology and Conservation 2012; 7(2): 122-128. 10.4013/nbc.2012.72.06
https://doi.org/10.4013/nbc.2012.72.06...
; Glas et al., 2012Glas JJ, Schimmel BCJ, Alba JM, Escobar-Bravo R, Schuurink RC, Kant MR. Plant glandular trichomes as targets for breeding or engineering of resistance to herbivores. International Journal of Molecular Sciences 2012; 13(12): 17077-17103. 10.3390/ijms131217077
https://doi.org/10.3390/ijms131217077...
), among others. Glandular trichomes generally produce sticky substances and/or toxic exudates, thus becoming a chemical barrier, as well as mechanical, against herbivorous insects. The major classes of metabolites produced in trichomes include terpenoids, flavonoids and phenylpropenes, methyl ketones and sugar acyl defensive proteins (Glas et al., 2012Glas JJ, Schimmel BCJ, Alba JM, Escobar-Bravo R, Schuurink RC, Kant MR. Plant glandular trichomes as targets for breeding or engineering of resistance to herbivores. International Journal of Molecular Sciences 2012; 13(12): 17077-17103. 10.3390/ijms131217077
https://doi.org/10.3390/ijms131217077...
).

In this context, the objectives of the present study were to: (1) evaluate herbivory percentages of chewing insects and leaf traits (nutrients, fibers and total phenols) of Cupania oblongifolia Mart., Siparuna guianensis Aubl. and Xylopia sericea St. Hill. individuals in the regeneration of the understory of an abandoned eucalyptus plantation and an Atlantic forest fragment; and (2) test the hypothesis that leaf herbivory is lower in species with lower nutritional quality and higher chemical and morphological defenses.

2. MATERIALS AND METHODS

2.1. Study site and species

The present study was conducted at the União Biological Reserve, located in the northern region of Rio de Janeiro (22°27’30” S and 42°02’15” W). Two sites were investigated: a former eucalyptus stand 39a (Corymbia citriodora (Hook) L.A, Jonhoson; K.D. Hill) with approximately 11.5 ha and which is 44 years old, abandoned for 16 years, suffering no silvicultural treatments (Ibama, 2008Instituto Brasileiro do Meio Ambiente e dos Recursos Naturais e Renováveis - Ibama. Plano de Manejo Reserva Biológica União: encarte 3: análise da unidade de conservação. Rio de Janeiro: Ibama; 2008.); and a fragment of Lowland Atlantic Forest, classified as a Tropical Rainforest (Carvalho et al., 2008Carvalho FA, Nascimento MT, Oliveira Filho AT. Composição, riqueza e heterogeneidade da flora arbórea da bacia do rio São João, RJ, Brasil. Acta Botanica Brasilica 2008; 22(4): 929-940. 10.1590/S0102-33062008000400004
https://doi.org/10.1590/S0102-3306200800...
; Ibama, 2008Instituto Brasileiro do Meio Ambiente e dos Recursos Naturais e Renováveis - Ibama. Plano de Manejo Reserva Biológica União: encarte 3: análise da unidade de conservação. Rio de Janeiro: Ibama; 2008.). The forest fragment of this biological reserve is one of the most preserved fragments of the São João River Basin (Carvalho et al., 2008Carvalho FA, Nascimento MT, Oliveira Filho AT. Composição, riqueza e heterogeneidade da flora arbórea da bacia do rio São João, RJ, Brasil. Acta Botanica Brasilica 2008; 22(4): 929-940. 10.1590/S0102-33062008000400004
https://doi.org/10.1590/S0102-3306200800...
; Carvalho et al., 2016Carvalho FA, Braga JMA, Nascimento MT. Tree structure and diversity of lowland Atlantic forest fragments: comparison of disturbed and undisturbed remnants. Journal of Forestry Research 2016; 27(3): 605-609. 10.1007/s11676-015-0165-2
https://doi.org/10.1007/s11676-015-0165-...
) and is part of the Serra do Mar Ecological Corridor, thus considered an area of great ecological and conservationist relevance, since it shelters the largest continuous stretch of Atlantic Forest (Ayers & Marigo, 2005Ayers JM, Marigo LC. Os corredores ecológicos das florestas tropicais do Brasil. Belém: Sociedade Civil Mamirauá; 2005.).

The climate of the study site is characterized as tropical humid, with annual average temperature of 24 °C, and annual rainfall around 2,200 mm/year. Much of the rainfall is concentrated between October and April (Nascimento et al., 2011Nascimento AA, Vieira AF, Silva AP, Villela DM, Nascimento MT. Herbivoria foliar de Xylopia sericea St. Hil. (Annonaceae) em sub-bosque de plantio de Eucalipto e de Mata Atlântica. Scientia Forestalis 2011; 39(89): 77-86.).

The selected species, Cupania oblongifolia Mart. (Sapindaceae), Siparuna guianensis Aubl. (Siparunaceae) and Xylopia sericea St. Hil (Anonnaceae), are considered the main and most abundant tree species in the understory of eucalyptus plantation areas (Evaristo et al., 2011Evaristo VT, Braga JMA, Nascimento MT. Atlantic Forest regeneration in abandoned plantations of eucalypt (Corymbia citriodora (Hook.) K. D. Hill and L. A. S. Johnson) in Rio de Janeiro, Brazil. Interciencia 2011; 36(6): 431-436.) and Atlantic Forest fragments in the state of Rio de Janeiro (Carvalho et al., 2006Carvalho FA, Nascimento MT, Braga JMA. Composição e riqueza florística do componente arbóreo da Floresta Atlântica submontana na região de Imbaú, município de Silva Jardim, RJ. Acta Botanica Brasilica 2006; 20(3): 727-740. 10.1590/S0102-33062006000300022
https://doi.org/10.1590/S0102-3306200600...
, 2007Carvalho FA, Nascimento MT, Braga JMA. Estrutura e composição florística do estrato arbóreo de um remanescente de Mata Atlântica submontana no município de Rio Bonito, RJ, Brasil (Mata Rio Vermelho). Revista Árvore 2007; 31(4): 717-730. 10.1590/S0100-67622007000400017
https://doi.org/10.1590/S0100-6762200700...
; Carvalho et al., 2008Carvalho FA, Nascimento MT, Oliveira Filho AT. Composição, riqueza e heterogeneidade da flora arbórea da bacia do rio São João, RJ, Brasil. Acta Botanica Brasilica 2008; 22(4): 929-940. 10.1590/S0102-33062008000400004
https://doi.org/10.1590/S0102-3306200800...
). However, they are still not widely used in ecological studies, especially those focused on interactions between plants and animals (Costa et al., 2007Costa UAS, Oliveira M, Tabarelli M, Leal IR. Dispersão de sementes por formigas em remanescentes de Floresta Atlântica nordestina. Revista Brasileira de Biociências 2007; 5(1): 231-233.; Nascimento et al., 2011Nascimento AA, Vieira AF, Silva AP, Villela DM, Nascimento MT. Herbivoria foliar de Xylopia sericea St. Hil. (Annonaceae) em sub-bosque de plantio de Eucalipto e de Mata Atlântica. Scientia Forestalis 2011; 39(89): 77-86.).

2.2. Leaf samplings

A total of 15 young individuals from each species, with heights ranging from 1.50 m to 3.0 m, were randomly collected at each of the studied areas. Twelve mature leaves were removed from three branches per plant. Mature leaves were considered as being dark green and arranged below the fourth leaf, counting from the apex to the base branch (Nascimento & Proctor, 2001Nascimento MT, Proctor J. Leaf herbivory on three species in a monodominant and two other Terra firme forests on Maracá Island, Brazil. Acta Amazonica 2001; 31(1): 27-38. 10.1590/1809-43922001311038
https://doi.org/10.1590/1809-43922001311...
; Nascimento et al., 2011Nascimento AA, Vieira AF, Silva AP, Villela DM, Nascimento MT. Herbivoria foliar de Xylopia sericea St. Hil. (Annonaceae) em sub-bosque de plantio de Eucalipto e de Mata Atlântica. Scientia Forestalis 2011; 39(89): 77-86.). In the case of compound leaves in C. oblongifolia, the leaflets were considered leaves. The leaves were first sampled in April 2009 (end of the rainy season), followed by August 2009 (end of the dry season). Only samples from the first sampling were used for chemical composition analyses and investigations of morphological/structural defenses. Bulked samples of undamaged leaves from five individuals among the 15 trees (n = 3 per area) from each species were used to obtain the mass (g) required for the chemical and morphological/structural analyses.

2.3. Leaf area (LA) and herbivory determinations

The sampled leaves were photographed (resolution: 1,200 × 1,600) using a digital camera (Samsung Digimax D53) on a white board with the aid of a tripod at a standard distance of 30 cm, according to the criteria used by Paul et al. (2012Paul GS, Montagnini F, Berlyn GP, Craven DJ, van Breugel M, Hall JS. Foliar herbivory and leaf traits of five native tree species in a young plantation of Central Panama. New Forests 2012; 43(1): 69-87. 10.1007/s11056-011-9267-7
https://doi.org/10.1007/s11056-011-9267-...
). The ImageJ software package was used for the leaf area and herbivory analyses (Rasband, 2003Rasband, W. ImageJ software 1.42. Bethesda: National Institutes of Health; 2003.), since this method is considered very precise (Paul et al., 2012Paul GS, Montagnini F, Berlyn GP, Craven DJ, van Breugel M, Hall JS. Foliar herbivory and leaf traits of five native tree species in a young plantation of Central Panama. New Forests 2012; 43(1): 69-87. 10.1007/s11056-011-9267-7
https://doi.org/10.1007/s11056-011-9267-...
). Herbivory damage was defined as the percentage damage at one moment in time. The percentage of herbivory for each leaf was determined by dividing the area consumed by the total leaf area (cm2) (Aide, 1993Aide TM. Patterns of leaf development and herbivory in a tropical understory community. Ecology 1993; 74(2): 455-466. 10.2307/1939307
https://doi.org/10.2307/1939307...
; Paul et al., 2012Paul GS, Montagnini F, Berlyn GP, Craven DJ, van Breugel M, Hall JS. Foliar herbivory and leaf traits of five native tree species in a young plantation of Central Panama. New Forests 2012; 43(1): 69-87. 10.1007/s11056-011-9267-7
https://doi.org/10.1007/s11056-011-9267-...
). Holes and missing leaf parts were considered to be due to herbivory. In cases of damage by herbivores at the end of the leaf tissues, leaf contours were redrawn using ImageJ drawing tools.

2.4. Chemical analyses

After being photographed, the leaves collected in April 2009 were washed with both tap water and deionized water and left in a circulation oven at 60 °C for 72 hours. After drying, they were ground in a knife mill. The total carbon (C) and nitrogen (N) determinations were performed on a CHNS/O auto-analyzer (Perkin Elmer 2,400) using 2 mg to 4 mg aliquots of leaf sub-samples.

The determinations of neutral detergent fibers (NDF), cellulose (CEL) and lignin (LIG) were performed at the Animal Science and Nutrition Laboratory (LZNA), in CCTA/UENF, according to the method reported by van Soest et al. (1991van Soest PJ, Robertson JB, Lewis BA. Symposium: carbohydrate methodology, metabolism, and nutritional implications in dairy cattle: methods for dietary fiber, neutral detergent fiber and nonstarch polysaccharides in relation to animal nutrition. Journal of Dairy Science 1991; 74(10): 3583-3597.). The dry and pulverized material (0.5 g) was used for the sequential analyses.

LIG content was determined by the 72% sulphuric acid method, followed by incineration of the residue at 550 °C for 3 h. Cellulose content was determined by calculating the difference between ADF (acid-detergent fibers) and LIG. NDF include cellulose, hemicellulose and lignin as main components, in addition to heat-damaged proteins, cell wall proteins and minerals (ashes). In contrast, ADF consist primarily of cellulose and lignin (lignocellulose) and heat-damaged proteins, with small amounts of cell wall proteins and insoluble minerals (ashes).

For the determination of total phenols (TP), the hydroalcoholic leaf extracts (EtOH) were prepared using 0.2 g of the ground material, according to the method reported by Phillips & Henshaw (1977Phillips R, Henshaw GG. The regulation of synthesis of phenolics in stationary phase cell cultures of Acer pseudoplatanus L. Journal of Experimental Botany 1977; 28(4): 785-794. 10.1093/jxb/28.4.785
https://doi.org/10.1093/jxb/28.4.785...
), modified. Total phenol concentrations were obtained through a standard curve of gallic acid at concentrations ranging from 50 mg/mL to 175 mg/mL, with interpolation of the sample absorbance (760 nm) from the calibration curve. All leaf analyses were conducted in replicate in cases of value disagreements.

2.5. Leaf trichomes

The trichome densities on the adaxial (AD) and abaxial (AB) leaf surfaces from five leafs per species were determined using a stereomicroscope (40×) in four fields taken at random (Gilardón et al., 2001Gilardón E, Pocovi M, Hernández C, Collavino G, Olsen A. Papel da 2-tridecanona e dos tricomas glandulares tipo VI na resistência do tomateiro a Tuta absoluta. Pesquisa Agropecuária Brasileira 2001; 36(7): 929-933. 10.1590/S0100-204X2001000700001
https://doi.org/10.1590/S0100-204X200100...
). Visual estimates of the leaf areas occupied by the trichomes were considered, categorized into four classes of trichome leaf coverage percentage: glabra (0%), rare (1-20%), abundant (21-50%), and dense (> 50%).

2.6. Data analyses

The variables of the residues analysis that showed normality and homoscedasticity issues were transformed by arcsines (herbivory) and square root (LA and phenols) for data standardization (Zar, 1984Zar JH. Biostatistical analysis. New Jersey: Prentice Hall International; 1984.). Herbivory values, even after this transformation, did not meet normality and homoscedasticity criteria. Thus, for intraspecific leaf herbivory comparisons between the two study sites, the Mann-Whitney U test was applied and interspecific comparisons of this variable in each environment were performed by a Kruskal-Wallis test, followed by Dunn’s comparison test. A factorial Anova and post-hoc Tukey test were used for variable comparisons (NDF, CEL, LIG, TP, LA, C, N and C/N), where study site (forest and eucalyptus) and species (C. oblongifolia, S. guianensis and X. sericea) were tested simultaneously. All variables were subjected to a Principal Component Analysis (PCA). Leaf areas between species for 15 individuals were compared, using the leaves from the first sampling. All data analyses were conducted using the Statistica software package (version 7.0).

3. RESULTS

3.1. Herbivory and leaf area

Herbivory damage varied from 1% to 13% among species, with significant differences within the study sites observed for leaves from both sampling (Kruskal-Wallis, p ≤ 0.01) (Figure 1). X. sericea showed the lowest percentage of herbivory, significantly differing from C. oblongifolia from the forest (Kruskal-Wallis test, p = 0.005) and the eucalyptus stand (Kruskal-Wallis, p ≤ 0.0001); differing, however, from S. guianensis in the first sampling only at the eucalyptus stand (Kruskal-Wallis test, p = 0.003) (Figure 1). Intraspecific loss of leaf area due to herbivory was similar between sites, considering leaves from both the first and the second samplings. Only X. sericea differed significantly (Mann-Whitney test, p = 0.01) between the sites, with higher percentages of leaf herbivory in the forest (4.6 % ± 5.1 %) compared to the eucalyptus stand (1.23 % ± 1.13 %), although for the first sampling only (Figure 1).

Figure 1
Leaf herbivory values (%) for Cupania oblongifolia (Cup), Siparuna guianensis (Sip) and Xylopia sericea (Xyl) sampled from the forest (for) and the eucalyptus stand (euc) for the first (A) and second (B) leaf samplings.The boxplots represent the median values, 1st and 3rd quartiles, extreme values and outlier. Different letters represent significant differences between species within each site (Kruskal-Wallis, p ≤ 0.05) and different numbers represent significant differences within each species between sites (Mann-Whitney test, p ≤ 0.05, n = 15).

Interspecific differences were observed regarding LA, with C. oblongifolia presenting higher mean leaf areas, followed by S. guianensis and X. sericea (factorial Anova, p ≤ 0.05) (Figure 2). Within the species, only X. sericea showed differences in mean leaf area when comparing the sites, with the highest value observed in individuals from the eucalyptus stand (Mann-Whitney test, p ≤ 0.001) (Figure 2).

Figure 2
Total leaf area (means and standard deviation) for Cupania oblongifolia (Cup), Siparuna guianensis (Sip) and Xylopia sericea (Xyl) sampled from the forest (for) and the eucalyptus stand (euc) regarding leaves from the first sampling.Different letters represent significant differences between species within each site and different numbers represent significant differences within each species between sites (factorial Anova, p ≤ 0.05, n = 15).

3.2. Foliar nutrient concentrations

Interspecific comparisons revealed that C. oblongifolia presented the highest C values, differing significantly from S. guianensis (factorial Anova, p = 0.00002) and X. sericea in the forest fragment (factorial Anova, p = 0.0005), with no significant difference between S. guianensis and X. sericea in this same area (p > 0.05). In the eucalyptus stand, the only significant difference was observed for C values between X. sericea and C. oblongifolia (factorial Anova, p = 0.003), with C. oblongifolia presenting the highest values (Figure 3). X. sericea showed higher C/N ratios, significantly differing from S. guianensis (factorial Anova, p = 0.0001) and C. oblongifolia (factorial Anova, p = 0.0005), with no significant difference for C/N values between S. guianensis and C. oblongifolia (p > 0.05) in the forest area. No significant differences in the C/N ratios between species were observed for the eucalyptus stand (p > 0.05) (Figure 3). The intraspecific analysis of C and N foliar concentrations between the forest and eucalyptus sites indicated that only C concentrations for S. guianensis were significantly higher in the eucalyptus stand individuals (factorial Anova, p = 0.011), while X. sericea showed higher C/N ratios in forest individuals (factorial Anova, p = 0.008). Regarding N, no significant differences were observed within species between the forest and eucalyptus areas, or between the investigated species within areas (factorial Anova, p > 0.05) (Figure 3).

Figure 3
Carbon (C) and nitrogen (N) percentages and C/N ratios (means ± standard deviation) for Cupania oblongifolia (Cup), Siparuna guianensis (Sip) and Xylopia sericea (Xyl) sampled from the forest (for) and the eucalyptus sites (euc). Different letters represent significant differences between species within site and different numbers represent significant differences within species between sites (factorial Anova, p ≤ 0.05, n = 3).

3.3. Structural and chemical defenses

The fiber analyses indicated no patterns between the three investigated species for NDF, CEL or LIG. Regarding interspecific variations of NDF values in the forest fragment, S. guianensis showed significantly lower values than C. oblongifolia and X. sericea (factorial Anova, p = 0,0001). The same pattern was observed for these species in the eucalyptus stand (factorial Anova, p ≤ 0.05); however, no significant difference was observed for intraspecific NDF concentrations (factorial Anova, p > 0.05) between the sites (Figure 4).

Figure 4
Fiber concentrations (%) and total phenol (µg/mL) (means ± standard deviation) for Cupania oblongifolia (Cup), Siparuna guianensis (Sip) and Xylopia sericea (Xyl) sampled from the forest (for) and the eucalyptus stand (euc).Fiber components: neutral detergent fiber (NDF), cellulose (CEL) and lignin (LIG). TP: total phenol. Different letters represent significant differences between species within each site and different numbers represent significant differences within species between sites (factorial Anova, p < 0.05, n = 3).

Regarding cellulose concentrations, interspecific variations - both in the forest fragment and the eucalyptus stand - were observed, with X. sericea always showing the highest mean cellulose values (factorial Anova, p ≤ 0.05). However, no intraspecific difference was observed when comparing sites (factorial Anova, p > 0.05) (Figure 4).

Interspecies comparisons revealed that S. guianensis presented significantly lower LIG values (p < 0.05) at both sites. On the other hand, S. guianensis and X. sericea showed significant variations in LIG concentrations between sites, with S. guianensis showing higher values in the eucalyptus stand (factorial Anova, p = 0.004), whereas X. sericea showed higher LIG values at the forest site (factorial Anova, p = 0.03) (Figure 4).

Interspecies variations in total phenol were observed, with X. sericea presenting higher total phenol concentrations at both studied sites (p < 0.05) (Figure 4). These concentrations also differed between environments (forest and eucalyptus) for X. sericea (factorial Anova, p = 0.006) and S. guianensis (factorial Anova, p = 0.05), being higher in forest individuals (Figure 4).

The three species also presented leaf trichomes, although differing in location (abaxial or adaxial) and/or density. C. oblongifolia and S. guianensis presented scarce (< 20% of leaf area covered with trichomes) and small trichomes on both sides, while X. sericea presented trichomes only on the abaxial surface, but more densely packed (> 70% of leaf area covered with trichomes) than the other species.

3.4. Relationship between leaf traits of the three investigated species

The PCA ordination diagram (Figure 5a and 5b), regarding the species data, indicates that the first axis represented 58.9 % of the variance, followed by the second axis at 34.9 %. Both accounted for 93.8 % of the total data variance. The eigenvalues produced by the PCA (axis 1 = 4.71, axis 2 = 2.79) were higher than the ones for the third and fourth axis (all less than 0.3), which were, thus, considered unlikely to have any meaning. This indicates that the species are distributed along gradients. The three species formed two groups, indicating interspecific discrimination, with X. sericea positioned on the negative end of axis 1 and C. oblongifolia and S. guianensis on the positive end of it (Figure 5a). However, no intraspecific difference was observed between the study sites. In addition, C. oblongifolia was positioned at the negative end of axis 2, while the other two species were positioned on the positive end of the same axis. Nitrogen (r = 0.93) showed a positive and significant correlation with axis 1, while TP (r = −0.97), CEL (r = −0.91) and C/N (r = −0.87) were negatively correlated (Figure 5b; Table 1). For axis 2, LIG (r = −0.91) and C (r = −0.87) showed the highest significant correlations (Figure 5b; Table 1).

Figure 5
Biplot of the principal component analysis (PCA) for Siparuna guianensis (Sip_gui), Cupania oblongifolia (Cup_obl) and Xylopia sericea (Xyl_ser) sampled from the forest fragment (1) and eucalyptus stand (2). Projection of species (a) and variables (b) in relation to axes 1 and 2. NDF: neutral detergent fiber; CEL: cellulose; LIG: lignin; C: carbon; N: nitrogen; TP: total phenols; LA: leaf area.

Table 1
Correlation coefficients between the investigated variables and eigenvectors of the principal component analysis (PCA) regarding the species sampled from the forest fragment and eucalyptus stand.

4. DISCUSSION

The mean values of leaf herbivory found for the three investigated species (1 %-13 %) are near or within the range found for tropical rainforest species (0.09 % to 30 %) using the point method (Cárdenas et al., 2014Cárdenas RE, Valencia R, Kraft NJ, Argoti A, Dangles O. Plant traits predict inter- and intraspecific variation in susceptibility to herbivory in a hyperdiverse Neotropical rain forest tree community. Journal of Ecology 2014; 102(4): 939-952. 10.1111/1365-2745.12255
https://doi.org/10.1111/1365-2745.12255...
; Peñuelas et al., 2013Peñuelas J, Sardans J, Llusia J, Silva J, Owen SM, Bala-Ola B et al. Foliar chemistry and standing folivory of early and late-successional species in a Bornean rainforest. Plant Ecology & Diversity 2013; 6(2): 245-256. 10.1080/17550874.2013.768713
https://doi.org/10.1080/17550874.2013.76...
).

Non-glandular trichomes generally show negative relationships with herbivory (Barônio, 2012Barônio GJ. Pilosidade foliar reduz herbivoria em folhas jovens e maduras de Qualea multiflora Mart. em cerrado stricto sensu. Neotropical Biology and Conservation 2012; 7(2): 122-128. 10.4013/nbc.2012.72.06
https://doi.org/10.4013/nbc.2012.72.06...
; Rodrigues et al., 2012Rodrigues DM, Souza CR, Aguiar RWS, Melo AV, Silva JC, Ootani MA et al. Tricomas conferem resistência contra herbivoria de Cerotoma arcuata em cultivares de soja. Agroecossistemas 2012; 4(2): 33-39. 10.18542/ragros.v5i2.1211
https://doi.org/10.18542/ragros.v5i2.121...
), since trichomes obstruct herbivorous arthropod movements on leaf surfaces or prevent these predators from reaching the leaf surface with their mouthparts (Fürstenberg-Hägg et al., 2013Fürstenberg-Hägg J, Zagrobelny M, Bak S. Plant defense against insect herbivores. International Journal of Molecular Sciences 2013; 14(5): 10242-10297. 10.3390/ijms140510242
https://doi.org/10.3390/ijms140510242...
; Glas et al., 2012Glas JJ, Schimmel BCJ, Alba JM, Escobar-Bravo R, Schuurink RC, Kant MR. Plant glandular trichomes as targets for breeding or engineering of resistance to herbivores. International Journal of Molecular Sciences 2012; 13(12): 17077-17103. 10.3390/ijms131217077
https://doi.org/10.3390/ijms131217077...
; Marquis, 2012Marquis RJ. Uma abordagem geral das defesas das plantas contra ação dos herbívoros. In: Del-Claro K, Torezan-Silingardi V, editores. Ecologia das interações plantas-animais: uma abordagem ecológica-evolutiva. Rio de Janeiro: Technical Books; 2012. p. 55-66.). It is worth noting that X. sericea leaves are densely covered by non-glandular trichomes on the abaxial surface of the leaves, which are much larger and denser than those in C. oblongifolia and S. guianensis. This may have contributed to the lower herbivory values observed for X. sericea, suggesting that the mechanical strength of trichomes in this species is probably more efficient against herbivory. In contrast to the results observed for X. sericea, the highest leaf herbivory percentages found for C. oblongifolia and S. guianensis (> 8%) appear to be related to the low leaf trichome density in these species, facilitating access to the leaves and their intake by herbivorous insects. Furthermore, C. oblongifolia has larger and wider leaves, followed by S. guianensis, whereas X. sericea shows the lowest leaf area. Insects belonging to the order Lepidoptera, such as some species of butterflies, can be influenced by leaf size, as they prefer wider leaves for oviposition (Godfray, 1986Godfray HCJ. Clutch size in a leaf-mining fly (Pegomya nigritarsis: Anthomyidae). Ecological Entomology 1986; 11(1): 75-81. 10.1111/j.1365-2311.1986.tb00281.x
https://doi.org/10.1111/j.1365-2311.1986...
). In addition, some studies have found higher insect species richness in plants with wider leaves compared to narrower ones (Kennedy & Southwood, 1984Kennedy CEJ, Southwood TRE. The number of species of insects associated with British trees: a re-analysis. Journal of Animal Ecology 1984; 53(2): 455-478. 10.2307/4528
https://doi.org/10.2307/4528...
; Moran & Southwood, 1982Moran VC, Southwood TRE. The guild composition of arthropod communities on trees. Journal of Animal Ecology 1982; 51(1): 289-306. 10.2307/4325
https://doi.org/10.2307/4325...
), suggesting that larger leaves may attract more insects and thus, increase herbivory percentages (Garibaldi et al., 2011Garibaldi LA, Kitzberger T, Ruggiero A. Latitudinal decrease in folivory within Nothofagus pumilio forests: dual effect of climate on insect density and leaf traits? Global Ecology and Biogeography 2011; 20(4): 609-619. 10.1111/j.1466-8238.2010.00623.x
https://doi.org/10.1111/j.1466-8238.2010...
; Gonçalves-Alvim et al., 2010Gonçalves-Alvim SJ, Lana TC, Ranieri BD, Silveira FAO, Ribeiro V, Fernandes GW. Growth, defense and herbivory on young leaves of Qualea parviflora (Vochysiaceae) in three different Cerrado habitats. Neotropical Biology and Conservation 2010; 5(2): 86-92. 10.4013/nbc.2010.52.03
https://doi.org/10.4013/nbc.2010.52.03...
).

The total phenolic concentrations found in X. sericea were about two to three times higher than those found for C. oblongifolia and S. guianensis, respectively, at both sites. Phenolic compounds are highly diverse in the plant kingdom, including about 10,000 different compounds. They are known to act in plant protection against herbivory, repelling and intoxicating several types of insects (Ballaré, 2014Ballaré CL. Light regulation of plant defense. Annual Review of Plant Biology 2014; 65: 335-363. 10.1146/annurev-arplant-050213-040145
https://doi.org/10.1146/annurev-arplant-...
; Fürstenberg-Hägg et al., 2013Fürstenberg-Hägg J, Zagrobelny M, Bak S. Plant defense against insect herbivores. International Journal of Molecular Sciences 2013; 14(5): 10242-10297. 10.3390/ijms140510242
https://doi.org/10.3390/ijms140510242...
; Karolewski et al., 2013Karolewski P, Giertych MJ, Żmuda M, Jagodziński AM, Oleksyn J. Season and light affect constitutive defenses of understory shrub species against folivorous insects. Acta Oecologica 2013; 53: 19-32. 10.1016/j.actao.2013.08.004
https://doi.org/10.1016/j.actao.2013.08....
; Mazid et al., 2011Mazid M, Khan T, Mohammad F. Role of secondary metabolites in defense mechanisms of plants. Biology and Medicine 2011; 3(2): 232-249.).

However, the lowest percentage of leaf herbivory found in X. sericea seems to be related not only to the high phenol concentrations in this species leaves, but also the highest leaf trichome density, smaller leaf area and lower nutritional quality (> C/N ratio and fibers). Such foliar traits appear to jointly contribute to increased X. sericea leaf resistance to herbivory, since they reduce leaf access or digestibility (Fürstenberg-Hägg et al., 2013Fürstenberg-Hägg J, Zagrobelny M, Bak S. Plant defense against insect herbivores. International Journal of Molecular Sciences 2013; 14(5): 10242-10297. 10.3390/ijms140510242
https://doi.org/10.3390/ijms140510242...
; Marquis, 2012Marquis RJ. Uma abordagem geral das defesas das plantas contra ação dos herbívoros. In: Del-Claro K, Torezan-Silingardi V, editores. Ecologia das interações plantas-animais: uma abordagem ecológica-evolutiva. Rio de Janeiro: Technical Books; 2012. p. 55-66.; Poorter et al., 2004Poorter L, van de Plassche M, Willems S, Boot RGA. Leaf traits and herbivory rates of tropical tree species differing in successional status. Plant Biology 2004; 6(6): 746-754. 10.1055/s-2004-821269
https://doi.org/10.1055/s-2004-821269...
), making this species less susceptible to herbivory.

Herbivory values did not differ between the forest fragment (richer in tree species) and the eucalyptus stand (poorer in tree species), for two (C. oblongifolia and S. guianensis) of the three species, and were higher for X. sericea in the forest fragment, as observed by Nascimento et al. (2011Nascimento AA, Vieira AF, Silva AP, Villela DM, Nascimento MT. Herbivoria foliar de Xylopia sericea St. Hil. (Annonaceae) em sub-bosque de plantio de Eucalipto e de Mata Atlântica. Scientia Forestalis 2011; 39(89): 77-86.). Thus, the results observed herein did not corroborate the results of Brown & Ewel (1987Brown BJ, Ewel JJ. Herbivory in complex and simple tropical successional ecosystems. Ecology 1987; 68(1): 108-116. 10.2307/1938810
https://doi.org/10.2307/1938810...
) and Andow (1991Andow DA. Vegetational diversity and arthropod population response. Annual Reviews of Entomology 1991; 36: 561-586. 10.1146/annurev.en.36.010191.003021
https://doi.org/10.1146/annurev.en.36.01...
), that observed higher herbivory in plant communities with lower species richness.

The principal component analysis did not indicate species separation between the forest and eucalyptus stand with regard to leaf quality (nutrients, total phenols and fibers), suggesting that these leaf traits are not considerably altered in these two areas, which may have contributed to the herbivory patterns observed, mainly for C. oblongifolia and S. guianensis.

5. CONCLUSIONS

The species that suffered the lowest percentage of leaf herbivory (X. sericea) showed the lowest nutritional quality and a better set of defensive leaf traits, corroborating the tested hypothesis.

ACKNOWLEDGEMENTS

The authors would like to thank technicians HS Carvalho and GR of the Purification for aid in the fieldwork and the staff at the União Biological Reserve/ICMBIO for the logistic support. Thanks also given to CNPq for the Scientific Initiation scholarship awarded by Jefferson Cabral Luiz and the productivity research grant conferred to Marcelo Trindade Nascimento, to Coordination for the Improvement of Higher Education Personnel (Capes - Finance Code 001) for granting the master’s scholarship to Aline Alves do Nascimento and Research Support Foundation for the State of Rio de Janeiro (Faperj) for financial support. In addition, the authors also thank Drs. Alberto Magno Fernandes and Ricardo Augusto Mendonça Vieira and technician Claudio Teixeira Lombardi for the support during the chemical analyses.

REFERENCES

  • Agrawal AA. Current trends in the evolutionary ecology of plant defence. Functional Ecology 2011; 25(2): 421-433. 10.1111/j.1365-2435.2010.01796.x
    » https://doi.org/10.1111/j.1365-2435.2010.01796.x
  • Agrawal AA, Fishbein, M. Plant defense syndromes. Ecology 2006; 87(7): S132-S149. 10.1890/0012-9658(2006)87[132:PDS]2.0.CO;2
    » https://doi.org/10.1890/0012-9658(2006)87[132:PDS]2.0.CO;2
  • Aide TM. Patterns of leaf development and herbivory in a tropical understory community. Ecology 1993; 74(2): 455-466. 10.2307/1939307
    » https://doi.org/10.2307/1939307
  • Andow DA. Vegetational diversity and arthropod population response. Annual Reviews of Entomology 1991; 36: 561-586. 10.1146/annurev.en.36.010191.003021
    » https://doi.org/10.1146/annurev.en.36.010191.003021
  • Ayers JM, Marigo LC. Os corredores ecológicos das florestas tropicais do Brasil. Belém: Sociedade Civil Mamirauá; 2005.
  • Ballaré CL. Light regulation of plant defense. Annual Review of Plant Biology 2014; 65: 335-363. 10.1146/annurev-arplant-050213-040145
    » https://doi.org/10.1146/annurev-arplant-050213-040145
  • Barônio GJ. Pilosidade foliar reduz herbivoria em folhas jovens e maduras de Qualea multiflora Mart. em cerrado stricto sensu. Neotropical Biology and Conservation 2012; 7(2): 122-128. 10.4013/nbc.2012.72.06
    » https://doi.org/10.4013/nbc.2012.72.06
  • Boege K, Marquis RJ. Facing herbivory as you grow up: the ontogeny of resistence in plants. Trends in Ecology & Evolution 2005; 20(8): 441-448. 10.1016/j.tree.2005.05.001
    » https://doi.org/10.1016/j.tree.2005.05.001
  • Brown BJ, Ewel JJ. Herbivory in complex and simple tropical successional ecosystems. Ecology 1987; 68(1): 108-116. 10.2307/1938810
    » https://doi.org/10.2307/1938810
  • Cárdenas RE, Valencia R, Kraft NJ, Argoti A, Dangles O. Plant traits predict inter- and intraspecific variation in susceptibility to herbivory in a hyperdiverse Neotropical rain forest tree community. Journal of Ecology 2014; 102(4): 939-952. 10.1111/1365-2745.12255
    » https://doi.org/10.1111/1365-2745.12255
  • Carvalho FA, Braga JMA, Nascimento MT. Tree structure and diversity of lowland Atlantic forest fragments: comparison of disturbed and undisturbed remnants. Journal of Forestry Research 2016; 27(3): 605-609. 10.1007/s11676-015-0165-2
    » https://doi.org/10.1007/s11676-015-0165-2
  • Carvalho FA, Nascimento MT, Braga JMA. Composição e riqueza florística do componente arbóreo da Floresta Atlântica submontana na região de Imbaú, município de Silva Jardim, RJ. Acta Botanica Brasilica 2006; 20(3): 727-740. 10.1590/S0102-33062006000300022
    » https://doi.org/10.1590/S0102-33062006000300022
  • Carvalho FA, Nascimento MT, Braga JMA. Estrutura e composição florística do estrato arbóreo de um remanescente de Mata Atlântica submontana no município de Rio Bonito, RJ, Brasil (Mata Rio Vermelho). Revista Árvore 2007; 31(4): 717-730. 10.1590/S0100-67622007000400017
    » https://doi.org/10.1590/S0100-67622007000400017
  • Carvalho FA, Nascimento MT, Oliveira Filho AT. Composição, riqueza e heterogeneidade da flora arbórea da bacia do rio São João, RJ, Brasil. Acta Botanica Brasilica 2008; 22(4): 929-940. 10.1590/S0102-33062008000400004
    » https://doi.org/10.1590/S0102-33062008000400004
  • Coley PD, Barone JA. Herbivory and plant defenses in tropical forests. Annual Review of Ecology and Systematics 1996; 27(1): 305-335. 10.1146/annurev.ecolsys.27.1.305
    » https://doi.org/10.1146/annurev.ecolsys.27.1.305
  • Cooper SM, Ginnett TF. Spines protect plants against browsing by small climbing mammals. Oecologia 1998; 113(2): 219-221. 10.1007/s004420050371
    » https://doi.org/10.1007/s004420050371
  • Costa UAS, Oliveira M, Tabarelli M, Leal IR. Dispersão de sementes por formigas em remanescentes de Floresta Atlântica nordestina. Revista Brasileira de Biociências 2007; 5(1): 231-233.
  • Cyr H, Pace ML. Magnitude and patterns of herbivory in aquatic and terrestrial ecosystems. Nature 1993; 361(6408): 148-150. 10.1038/361148a0
    » https://doi.org/10.1038/361148a0
  • Engelkes T, Meisner A, Morriën E, Kostenko O, Van der Putten WH, Macel M. Herbivory and dominance shifts among exotic and congeneric native plant species during plant community establishment. Oecologia 2016; 180(2): 507-517. 10.1007/s00442-015-3472-6
    » https://doi.org/10.1007/s00442-015-3472-6
  • Evaristo VT, Braga JMA, Nascimento MT. Atlantic Forest regeneration in abandoned plantations of eucalypt (Corymbia citriodora (Hook.) K. D. Hill and L. A. S. Johnson) in Rio de Janeiro, Brazil. Interciencia 2011; 36(6): 431-436.
  • Fürstenberg-Hägg J, Zagrobelny M, Bak S. Plant defense against insect herbivores. International Journal of Molecular Sciences 2013; 14(5): 10242-10297. 10.3390/ijms140510242
    » https://doi.org/10.3390/ijms140510242
  • Garibaldi LA, Kitzberger T, Ruggiero A. Latitudinal decrease in folivory within Nothofagus pumilio forests: dual effect of climate on insect density and leaf traits? Global Ecology and Biogeography 2011; 20(4): 609-619. 10.1111/j.1466-8238.2010.00623.x
    » https://doi.org/10.1111/j.1466-8238.2010.00623.x
  • Gilardón E, Pocovi M, Hernández C, Collavino G, Olsen A. Papel da 2-tridecanona e dos tricomas glandulares tipo VI na resistência do tomateiro a Tuta absoluta. Pesquisa Agropecuária Brasileira 2001; 36(7): 929-933. 10.1590/S0100-204X2001000700001
    » https://doi.org/10.1590/S0100-204X2001000700001
  • Glas JJ, Schimmel BCJ, Alba JM, Escobar-Bravo R, Schuurink RC, Kant MR. Plant glandular trichomes as targets for breeding or engineering of resistance to herbivores. International Journal of Molecular Sciences 2012; 13(12): 17077-17103. 10.3390/ijms131217077
    » https://doi.org/10.3390/ijms131217077
  • Godfray HCJ. Clutch size in a leaf-mining fly (Pegomya nigritarsis: Anthomyidae). Ecological Entomology 1986; 11(1): 75-81. 10.1111/j.1365-2311.1986.tb00281.x
    » https://doi.org/10.1111/j.1365-2311.1986.tb00281.x
  • Gonçalves-Alvim SJ, Lana TC, Ranieri BD, Silveira FAO, Ribeiro V, Fernandes GW. Growth, defense and herbivory on young leaves of Qualea parviflora (Vochysiaceae) in three different Cerrado habitats. Neotropical Biology and Conservation 2010; 5(2): 86-92. 10.4013/nbc.2010.52.03
    » https://doi.org/10.4013/nbc.2010.52.03
  • Instituto Brasileiro do Meio Ambiente e dos Recursos Naturais e Renováveis - Ibama. Plano de Manejo Reserva Biológica União: encarte 3: análise da unidade de conservação. Rio de Janeiro: Ibama; 2008.
  • Karolewski P, Giertych MJ, Żmuda M, Jagodziński AM, Oleksyn J. Season and light affect constitutive defenses of understory shrub species against folivorous insects. Acta Oecologica 2013; 53: 19-32. 10.1016/j.actao.2013.08.004
    » https://doi.org/10.1016/j.actao.2013.08.004
  • Kennedy CEJ, Southwood TRE. The number of species of insects associated with British trees: a re-analysis. Journal of Animal Ecology 1984; 53(2): 455-478. 10.2307/4528
    » https://doi.org/10.2307/4528
  • Kitajima K, Llorens A-M, Stefanescu C, Timchenko MV, Lucas PW, Wright SJ. How cellulose-based leaf toughness and lamina density contribute to long leaf lifespans of shade-tolerant species. New Phytologist 2012; 195(3): 640-652. 10.1111/j.1469-8137.2012.04203.x
    » https://doi.org/10.1111/j.1469-8137.2012.04203.x
  • Kurokawa H, Nakashizuka T. Leaf herbivory and decomposability in a Malaysian tropical rain forest. Ecology 2008; 89(9): 2645-2656. 10.1890/07-1352.1
    » https://doi.org/10.1890/07-1352.1
  • Linton SM, Greenaway P. A review of feeding and nutrition of herbivorous land crabs: adaptations to low quality plant diets. Journal of Comparative Physiology 2007; 177(3): 269-286. 10.1007/s00360-006-0138-z
    » https://doi.org/10.1007/s00360-006-0138-z
  • Lowman MD. Temporal and spatial variability in insect grazing of the canopies of five Australian rainforest tree species. Australian Journal of Ecology 1985; 10(1): 7-24. 10.1111/j.1442-9993.1985.tb00859.x
    » https://doi.org/10.1111/j.1442-9993.1985.tb00859.x
  • Lucas PW, Turner IM, Dominy NJ, Yamashita N. Mechanical defences to herbivory. Annals of Botany 2000; 86(5): 913-920. 10.1006/anbo.2000.1261
    » https://doi.org/10.1006/anbo.2000.1261
  • Marquis RJ. Uma abordagem geral das defesas das plantas contra ação dos herbívoros. In: Del-Claro K, Torezan-Silingardi V, editores. Ecologia das interações plantas-animais: uma abordagem ecológica-evolutiva. Rio de Janeiro: Technical Books; 2012. p. 55-66.
  • Mazid M, Khan T, Mohammad F. Role of secondary metabolites in defense mechanisms of plants. Biology and Medicine 2011; 3(2): 232-249.
  • Moran VC, Southwood TRE. The guild composition of arthropod communities on trees. Journal of Animal Ecology 1982; 51(1): 289-306. 10.2307/4325
    » https://doi.org/10.2307/4325
  • Mundim FM, Costa AN, Vasconcelos HL. Leaf nutrient content and host plant selection by leaf-cutter ants, Atta laevigata, in a Neotropical savanna. Entomologia Experimentalis et Applicata 2009; 130(1): 47-54. 10.1111/j.1570-7458.2008.00789.x
    » https://doi.org/10.1111/j.1570-7458.2008.00789.x
  • Nascimento AA, Vieira AF, Silva AP, Villela DM, Nascimento MT. Herbivoria foliar de Xylopia sericea St. Hil. (Annonaceae) em sub-bosque de plantio de Eucalipto e de Mata Atlântica. Scientia Forestalis 2011; 39(89): 77-86.
  • Nascimento MT, Proctor J. Leaf herbivory on three species in a monodominant and two other Terra firme forests on Maracá Island, Brazil. Acta Amazonica 2001; 31(1): 27-38. 10.1590/1809-43922001311038
    » https://doi.org/10.1590/1809-43922001311038
  • Paul GS, Montagnini F, Berlyn GP, Craven DJ, van Breugel M, Hall JS. Foliar herbivory and leaf traits of five native tree species in a young plantation of Central Panama. New Forests 2012; 43(1): 69-87. 10.1007/s11056-011-9267-7
    » https://doi.org/10.1007/s11056-011-9267-7
  • Peñuelas J, Sardans J, Llusia J, Silva J, Owen SM, Bala-Ola B et al. Foliar chemistry and standing folivory of early and late-successional species in a Bornean rainforest. Plant Ecology & Diversity 2013; 6(2): 245-256. 10.1080/17550874.2013.768713
    » https://doi.org/10.1080/17550874.2013.768713
  • Phillips R, Henshaw GG. The regulation of synthesis of phenolics in stationary phase cell cultures of Acer pseudoplatanus L. Journal of Experimental Botany 1977; 28(4): 785-794. 10.1093/jxb/28.4.785
    » https://doi.org/10.1093/jxb/28.4.785
  • Poorter L, van de Plassche M, Willems S, Boot RGA. Leaf traits and herbivory rates of tropical tree species differing in successional status. Plant Biology 2004; 6(6): 746-754. 10.1055/s-2004-821269
    » https://doi.org/10.1055/s-2004-821269
  • Rasband, W. ImageJ software 1.42. Bethesda: National Institutes of Health; 2003.
  • Rodrigues DM, Souza CR, Aguiar RWS, Melo AV, Silva JC, Ootani MA et al. Tricomas conferem resistência contra herbivoria de Cerotoma arcuata em cultivares de soja. Agroecossistemas 2012; 4(2): 33-39. 10.18542/ragros.v5i2.1211
    » https://doi.org/10.18542/ragros.v5i2.1211
  • Silva JO, Neves FS. Insect herbivores associated with an evergreen tree Goniorrhachis marginata Taub. (Leguminosae: Caesalpinioideae) in a tropical dry forest. Brazilian Journal of Biology 2014; 74(3): 623-631. 10.1590/bjb.2014.0093
    » https://doi.org/10.1590/bjb.2014.0093
  • Strong DR, Lawton JH, Southwood R. Insects on plants: community patterns and mechanisms. Massachusetts: Harvard University Press; 1984.
  • Trigo JR, Pareja M, Massuda KF. O papel das substâncias químicas nas interações entre plantas e insetos herbívoros. In: Del-Claro K, Torezan-Silingardi H, editores. Ecologia das interações plantas-animais: uma abordagem ecológico-evolutiva. Rio de Janeiro: Technical Books ; 2012. p. 69-88.
  • van Soest PJ, Robertson JB, Lewis BA. Symposium: carbohydrate methodology, metabolism, and nutritional implications in dairy cattle: methods for dietary fiber, neutral detergent fiber and nonstarch polysaccharides in relation to animal nutrition. Journal of Dairy Science 1991; 74(10): 3583-3597.
  • Wilson EO. The diversity of life. Toronto: Penguin; 2001.
  • Zar JH. Biostatistical analysis. New Jersey: Prentice Hall International; 1984.

Publication Dates

  • Publication in this collection
    23 Sept 2019
  • Date of issue
    2019

History

  • Received
    27 July 2017
  • Accepted
    30 June 2018
Instituto de Florestas da Universidade Federal Rural do Rio de Janeiro Rodovia BR 465 Km 7, CEP 23897-000, Tel.: (21) 2682 0558 | (21) 3787-4033 - Seropédica - RJ - Brazil
E-mail: floram@ufrrj.br