Acessibilidade / Reportar erro

HUMIFIED FRACTION OF ORGANIC MATTER DUE TO PLANT MIXTURE CULTIVATION

FRAÇÕES HUMIFICADAS DA MATÉRIA ORGÂNICA EM FUNÇÃO DO CULTIVO DE COQUETÉIS VEGETAIS

Abstracts

ABSTRACT - The aim of this study was to assess the organic matter changes in quantity and quality, particularly of the humic fraction in the surface layer (0-20 cm), of a Typic Plinthustalf soil under different management of plant mixtures used as green manure for mango (Mangifera indica L.) crops. The plant mixtures, which were seeded between rows of mango trees, were formed by two groups of leguminous and non -leguminous plants. Prior to sowing, seeds were combined in different proportions and compositions constituting the following treatments: 100% non-leguminous species (NL); 100% leguminous species (L); 75% L and 25% NL; 50% L and 50% NL; 25% L and 75% NL; and 100% spontaneous vegetation, considered a control. The plant mixtures that grew between rows of mango trees caused changes in the chemical composition of the soil organic matter, especially for the treatments 50% L and 50% NL and 25% L and 75% NL, which increased the content of humic substances in the soil organic matter. However, the treatment 25% L and 75% NL was best at minimising loss of total organic carbon from the soil. The humic acids studied have mostly aliphatic characteristics, showing large amounts of carboxylic and nitrogen groups and indicating that most of the organic carbon was formed by humic substances, with fulvic acid dominating among the alkali soluble fractions.

Keywords:
Humic substances; Infrared spectroscopy; Mangifera indica; Soil quality; Soil management


RESUMO - O objetivo desse trabalho foi avaliar as mudanças na quantidade e qualidade da matéria orgânica, particularmente da fração húmica, na camada superficial (0-20 cm) de um Argissolo Vermelho-Amarelo sob diferentes coquetéis vegetais usados como adubação verde da cultura da mangueira (Mangifera indica L.). Os coquetéis vegetais, semeados nas entrelinhas das mangueiras, foram formados por dois grupos de espécies leguminosas e não-leguminosas. As sementes, antes da semeadura, foram combinadas em diferentes composições e proporções que constituíram os tratamentos: 100% de espécies não-leguminosas (NL); 100% de espécies leguminosas (L); 75% L e 25% NL; 50% L e 50% NL; 25% L e 75% NL; e 100% vegetação espontânea, considerado como testemunha. Constatou-se que a aplicação de coquetéis vegetais na entrelinha da cultura da mangueira promoveu alterações na composição química da matéria orgânica do solo, com destaque para os tratamentos 50% L e 50% NL e 25% L e 75% NL que aumentaram o conteúdo de substâncias húmicas na matéria orgânica do solo. Não obstante, o tratamento 25% L e 75% NL foi o indicado para minimizar as perdas de carbono orgânico total do solo. Os ácidos húmicos estudados são de característica mais alifática apresentando grande quantidade de grupos carboxílicos e nitrogenados e sendo a maior parte do carbono orgânico do solo constituído por substâncias húmicas, tendo a fração ácidos fúlvicos como dominante entre as frações alcalino solúveis.

Palavras-chaves:
Espectroscopia de infravermelho; Manejo do solo; Mangifera indica; Qualidade do solo; Substâncias húmicas


INTRODUCTION

The total organic carbon content (TOC) is usually low in tropical soils of semiarid regions, often not exceeding 10 g kg-1 soil mass (SALCEDO; SAMPAIO, 2008SALCEDO, I. H.; SAMPAIO, E. V. S. B. Matéria orgânica do solo no Bioma Caatinga. In: SANTOS, G.; CAMARGO, F. (Eds.). Fundamentos da matéria orgânica do solo: ecossistemas tropicais e subtropicais. 2. ed. Porto Alegre: Metrópole, 2008. p. 419-441.). In addition, irrigated crops in the Brazilian semiarid region can accelerate the decomposition of organic matter (OM) and may change the quality and quantity of the humified fractions of soil OM (CUNHA et al., 2009aCUNHA, T. J. F. et al. Ácidos húmicos em solo fertirrigado no vale do São Francisco. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 6, p. 1583-1592, 2009a.). Carbon (C) is the main component of OM, affecting important physical, chemical and biological soil processes (BRANCALIÃO; MORAES, 2008BRANCALIÃO, S. R.; MORAES, M. H. Alterações de alguns atributos físicos e das frações húmicas de um Nitossolo Vermelho na sucessão milheto-soja em sistema plantio direto. Revista Brasileira de Ciência do Solo, Viçosa, v. 32, n. 1, p. 393-404, 2008.). Briefly, OM can be defined as a heterogeneous mixture of organic molecules arising from the transformation, in varying degrees, of biopolymers (carbohydrates, lignins, proteins and fatty acids) from plants, animals and microorganisms (SIMPSON, 2002SIMPSON, A. J. Determining the molecular weight, aggregation, structures and interactions of natural organic matter using diffusion ordered spectroscopy. Magnetic Resonance in Chemistry, Chichester, v. 40, n. 13, p. S72-S82, 2002.). The humic fractions form an essential part of OM and are operationally defined on the basis of solubility in aqueous alkalis or acids. Since the majority of OM is formed by the humified fraction, its association with soil fertility (KONONOVA, 1982KONONOVA, M. M. Matéria orgânica del suelo: su natureza, propriedades y métodos de investigacion. Barcelona: Oikos-Tau, 1982. 365 p.) and its use as an indicator of soil quality (ORLOV, 1998ORLOV, D. S. Organic substances of russian soils. European Journal of Soil Science , Oxford, v, 31, n. 9, p. 946-953, 1998.) are common. The humification process or humic substance (HS) formation could occur in nature by the progressive accumulation of more hydrophobic and recalcitrant molecules in the soil (PICCOLO, 2012PICCOLO, A. Carbon sequestration in agricultural soils. Heidleberg, Germany: Springer Verlag, 2012. 306p.). Therefore, the relationships of humic acid (HA)/fulvic acid (FA) and humin (H)/TOC can express the relative degree of humification of the OM of soil and its lability.

The elemental analysis of humified soil matter can be obtained from the quantification of C and nitrogen (N) contained in different humic fractions (HA, FA and H), achieved using spectroscopic methods, especially with employment of 13C nuclear magnetic resonance (13C NMR). The application of spectroscopy in the infrared region using diffuse reflectance infrared Fourier transform (DRIFT) on soil samples is comparatively cheaper than 13C NMR, easy to prepare and fast performing (a few minutes per sample) (LIEFELD, 2006LIEFELD, J. Application of diffuse reflectance FT- IR spectroscopy and partial least-squares regression to predict NMR properties of soil organic matter. European Journal of Soil Science, Oxford, v. 57, n. 6, p. 846-857, 2006.; FRANCIOSO et al., 2009FRANCIOSO, O. et al. Structural differences of chernozem soil humic acids SEC-PAGE fractions revealed by thermal (TG-DTA) and spectroscopic (DRIFT) analyses. Geoderma, Amsterdam, v. 152, n. 3/4, p. 264-268, 2009.; CARLETTI et al., 2010CARLETTI, P. et al. Structural characterization of humic-like substances with conventional and surface-enhanced spectroscopic techniques Journal of Molecular Structure, Amsterdam, v. 982, n. 1/3, p. 169-175, 2010.; FERRARI et al., 2011FERRARI, E. et al. DRIFIT and HR MAS NMR characterization of humic substances from a soil treated with different organic and mineral fertilizers. Journal of Molecular Structure , Amsterdam, v. 998, n. 1/3, p. 216-224, 2011.; SIMONETTI et al., 2012SIMONETTI, G. et al. Characterization of humic carbon in soil aggregates in a long-term experiment with manure and mineral fertilization. Soil Science Society of America Journal , Madison, v. 76, n. 3, p. 880-890, 2012.), including the assessment of resulting changes from management.

The use of plant mixtures in irrigated systems can lead to changes in the properties of HSs in the soil as possible changes in their spectroscopic characteristics. Fourier transform infrared spectroscopy is a useful method for the characterization of OM (OLK et al., 2000OLK, D. C. et al. Decrease in humification of organic matter with intensified lowland rice cropping a wet chemical and spectroscopic investigation. Soil Science Society of America Journal, Madison, v. 64, n. 4, p. 1337-1347, 2000.; VAZ Jr. et al., 2015) and HSs (MAO et al., 2008MAO, J. et al. Influence of animal manure application on the chemical tructures of soil organic matter as investigated by advanced solid-state NMR and FT-IR spectroscopy. Geoderma , Amsterdam, v. 146, n.1/2, p. 353-362, 2008.; CUNHA et al., 2009aCUNHA, T. J. F. et al. Ácidos húmicos em solo fertirrigado no vale do São Francisco. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 6, p. 1583-1592, 2009a., b; FRANCIOSO et al., 2009FRANCIOSO, O. et al. Structural differences of chernozem soil humic acids SEC-PAGE fractions revealed by thermal (TG-DTA) and spectroscopic (DRIFT) analyses. Geoderma, Amsterdam, v. 152, n. 3/4, p. 264-268, 2009.). This technique provides detailed information about the oxygen-containing groups present in complex molecules, since polar groups usually show intense infrared signals. An important information that can be obtained from the infrared spectrum is hydrophobicity index, since expresses the relationship between unpolar (-CH3) and polar groups (-OH, C-O) (FREIXO et al., 2002FREIXO, A. A. Propriedades espectrais da matéria orgânica leve-livre e leve intra-agregado de dois Latossolos sob plantio direto e preparo convencional. Revista Brasileira de Ciência do Solo , Viçosa, v. 26, n. 2, p. 445-453, 2002.).

In the São Francisco valley, green manure consisting of a combination of different species of leguminous and non-leguminous plants, commonly called 'plant mixtures', has aroused farmers interested in irrigated agriculture. At the same time, few changes are known in the quantity and quality of soil OM resulting from continued use of this practice. Due to their larger C/N ratio and therefore slower rate of decomposition, it is expected that plant mixtures composed of higher proportions of non-leguminous species (grasses and oilseeds) promote increased C content in the soil and greater stability in the soil system, thus favouring the process of humification. Mango trees are a crop of great economic importance to the region; mango crops have a long cultivation cycle, with the presence of severe humified OM for increasing nutrient availability, not due to the decomposition and slower mineralization of nutrients. In addition, soil OM fractions are sensitive to land use change and are related to several soil properties. There is no information regarding changes in the quantity and quality of HSs caused by use of plant mixture applications, thus justifying its study. The objective of this work was to evaluate the changes in quantity and quality of OM of HSs of a Typic Plinthustalf soil under different plant mixtures used as green manure between rows of mango trees.

MATERIAL AND METHODS

The experiment was conducted at the Bebedouro Experimental Field, from the Embrapa semiarid, Petrolina-PE (09º 09' S, 42º 22' W) on a Typic Plinthustalf, medium-texture soil (USDA, 2010). The climate is of BSwh' type according to the Köppen classification, with average annual rainfall of 570 mm and monthly average temperatures ranging from 24.2-28.1 ºC. Physical and chemical analysis results of soil characterisation following the Classen (1997CLASSEN, M. E. et al. Manual de métodos de análise de solos. 2. ed. Rio de Janeiro, RJ: EMBRAPA-CNPS, 1997. 212 p.) methods showed that the layer of 0-20 cm had the following characteristics: 900, 10 and 90 g kg-1 of sand, silt and clay, respectively; low TOC content (4.1 g kg-1); moderate acidity (pH = 6.0); low electrical conductivity (0.10 dS m-1); high levels of phosphorus (25 mg dm-3) and potassium (0.30 cmolc dm-3); low levels of calcium, magnesium, sodium and aluminium (1.0, 0.5, 0.01 and 0.05 cmolc dm-3, respectively); low acidity potential (0.66 cmolc dm-3); and consequently, low base sums and cation exchange capacity (1.81 and 2.47 cmolc dm-3, respectively). The mango seedlings with 80-90 cm were implemented in 2005 in the spacing of 6 x 8 m. Fertiliser was applied to each planting hole of mango seedlings of the Tomy cultivar: 59 g P2O5, 20 g K2O, 59 g Ca and 40 g S, using simple superphosphate and potassium chloride as nutrients sources, were applied.

The treatments seeded between rows were formed by two species groups, in different proportion: i) leguminous - Calopogonium mucunoides, Crotalaria juncea, Crotalaria spectabilis, Canavalia ensiformis, Cajanus cajan L. and Dolichos lablab L.; ii) non-leguminous plants - Sesamum indicum L., Chrysanthemum peruviamum, Ricinus communis L., Pennisetum americanum L. and Sorghum vulgare Pers.

The seeds species were combined and mixed in different proportions and compositions prior to sowing that constituted the following treatments used: T1 - 100% of non-leguminous species (NL); T2 - 100% of leguminous species (L); T3 - 75% L and 25% NL; T4 - 50% L and 50% NL; T5 - 25% L and 75% NL; and T6 - 100% spontaneous vegetation (SV), which was considered a control treatment, composed of Commelina benghalensis L., Macroptilium atropurpureum Urb., Desmodium tortuosum (Sw.) DC. and Acanthospermum hispidum DC. The experimental design was a randomised block with four replications.

The species were seeded during the rainy season between the mango rows for 3 years beginning 2006 at a distance of 2.0 m from the plant stems in furrows at a spacing of 0.5 m; each parcel corresponds to an area of 160 m2 (16 x 10 m) containing two mango plants. The mango tree sowing was carried out between rows within each parcel. To ensure uniform seed distribution in rows, seeds smaller in size/weight (10-70 g 1,000 seeds-1) were initially distributed, followed by intermediate- sized ones (200-300 g 1,000 seeds-1) and, subsequently, larger (850-1,450 g 1,000 seeds-1) ones, thus avoiding any segregation effect. The seed quantities used in each treatment are presented in Table 1. An experimental design of entirely randomised blocks with four replications was used.

For 3 years, the plant mixtures were cultivated during the rainy season; seeds that make up each plant mixture were sown between rows of mango trees. At the time of flowering, the green manure plants were mowed close to the ground (1 cm), and samples containing parts of plants (stalks and leaves) were collected. After plant mixtures were mowed, phytomass was deposited in the canopy projection of the mango trees. In Table 2, the amount of phytomass produced and nutrients in the different treatments, as Silva (1999SILVA, F. C. da. Manual de análises químicas de solos, plantas e fertilizantes. Brasília, DF: Embrapa Comunicação para Transferência de Tecnologia, 1999. 37p.) analysed, can be observed.

Four soil samples were collected, each consisting of a composed sample, in the crop line of mango trees in 2009, after 3 years of plant mixture cultivation at a depth of 0-20 cm.

Humic substance quantification

The quantitative chemical fractionation of HSs of the soil was performed according to methods of Benites et al. (2003BENITES, V. M. et al. Extração e fracionamento quantitativo de substâncias húmicas do solo: um procedimento simplificado de baixo custo. Rio de Janeiro: Embrapa Solos, 2003. 7p. (Embrapa Solos. Comunicado Técnico, 16).) to obtain H, HA and FA. The quantitative determination of organic C in each fraction was performed by wet combustion with potassium dichromate (YEOMANS; BREMNER, 1988YEOMANS, J. C.; BREMNER, J. M. A rapid and precise method for routine determination of organic carbon in soil. Communications in Soil Science and Plant Analysis, New York, v. 19, n. 13, p. 1467-1476, 1988.) and the values expressed as percentages of TOC. The relations CHA/CFA, CH/(CFA + CHA), CH + CFA + CHA/TOC and CH + CFA + CHA were used to interpret the obtained data, according to methods of Labrador Moreno (1996) and Benites (2002).

Extraction and humic acid purification

The OM extraction and fractionation to obtain HA and its purification were performed according to the International Humic Substances Society method, whose protocol is described by Swift (1996SWIFT, R. S. Organic matter characterization. In: SPARKS, D. L et al. (Eds.). Methods of soil analysis: chemical methods. Madison: SSSA , 1996. v. 3, p. 1011-1020. (SSSA. Book Series, 5).), with the following modifications: pre-treatment of soil samples was made with hydrochloric acid (HCl) to a solution pH between 1 and 2 for reducing ash content (metal removal adsorbed), and suspended solids removal was performed using K+ (NOVOTNY, 2002NOVOTNY, E. H. Estudos espectroscópicos e cromatográficos de substâncias húmicas de solos sob diferentes sistemas de preparo. 2002. 215 p. Tese (Doutorado em Ciências: Área de Concentração em Físico Química) - Universidade de São Paulo, São Carlos, 2002.), as briefly described next. After soil pre-treatment with 0.1 M HCl, sample amounts with around 1,000 mg of HA (as estimated by quantitative fractionation) and up to 200 mL of 0.1 M sodium hydroxide under inert atmospheric dinitrogen (N2) were added. After shaking for 24 h, the material was centrifuged at 10,000 g for 30 min. The supernatant (alkaline extract) was collected and the pH immediately adjusted to 2.0 by dripping of an aqueous solution of 20% HCl. The residue was subjected to extraction again, and the supernatant was joined to the previous setting, with the pH immediately brought to 2.0. The acidified extract was reserved for the HA precipitation. After 18 h, the excess supernatant (FA) was siphoned and discarded. The remaining material was centrifuged at 5,000 g for 10 min, after which the supernatant (FA) was removed. For removal of soluble solids in suspension, HA was re-dissolved in a potassium hydroxide solution to 0.1 M, after which potassium chloride (KCl) was added to achieve a concentration of 0.3 M K+ and then centrifuged at 4,000 g for 15 min. After this step, the HA was purified with a 0.5% hydrogen fluoride + HCl (SCHNITZER, 1982SCHNITZER, M. Organic matter characterization. In: PAGE, A. L. et al. (Ed.). Methods of soil analysis: chemical and microbiological properties. 2. ed. Madison: SSSA, 1982. v. 2, cap. 30, p. 581-594. (SSSA. Agronomy, 9).) solution for 24 h and centrifuged at 5,000 g. The process was repeated two additional times. The purified samples were washed with deionised water and transferred into cellophane bags of approximately 100 mL. Dialysis samples were evaluated in deionised water in a 20-L container (12 bags per battery). The water was replenished twice daily until there was no further increase in microsiemens to measure the electrical conductivity of the dialysis water after 1 h of this exchange. The samples were then frozen, lyophilised and placed in a desiccator.

To measure DRIFT, the HA was mixed in potassium bromide (1:100, w/w) and compressed into tablet form. Spectra were obtained from 100 scans in the range of 400-4,000 cm-1 with a spectral resolution of 2 cm-1 using Shimadzu Prestige 21 equipment. Absorption bands in the regions of 2,928 cm-1 and 1,050-1,080 cm-1 denominate in this the work hydrophobicity index (HI), since they express the relationship between nonpolar methyl (-CH3) and polar (-OH, C-O) groups, according Freixo et al. (2002FREIXO, A. A. Propriedades espectrais da matéria orgânica leve-livre e leve intra-agregado de dois Latossolos sob plantio direto e preparo convencional. Revista Brasileira de Ciência do Solo , Viçosa, v. 26, n. 2, p. 445-453, 2002.). The average results were compared using the Tukey test at the 5% level of probability.

RESULTS AND DISCUSSION

Humic substance quantification

The plant mixtures used between rows in the mango cultivation induced a reduction in TOC in all treatments compared to the control (SV), with the exception of the 75% L and 25% NL treatment (Table 3). This decrease was in the order of 7-28% of the composition according to the plant mixtures. The greatest decreases were observed in treatments formed only by non-leguminous (100% NL) or only leguminous (100% L) plants, which promoted an average loss 27% of TOC. When the proportion of non-leguminous plants in mixtures containing leguminous ones increased (50% L and 50% NL and 75% L and 25% NL), a lower drop was observed, being 15% (Table 3). A decrease in OM content from soil cultivation, especially in the early years, has been reported in the literature (CUNHA et al., 2009aCUNHA, T. J. F. et al. Ácidos húmicos em solo fertirrigado no vale do São Francisco. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 6, p. 1583-1592, 2009a.; CANELLAS et al., 2010CANELLAS, L. P. et al. Chemical composition and bioactivity properties of size-fractions separated from a vermicompost humic acids. Chemosphere, Oxford, v. 78, n. 4, p. 457-466, 2010.). Moreover, the use of irrigation associated with high temperatures and the sandy texture of Typic Plinthustalf soil also may have contributed to the reduced C content. According to Guimarães et al. (2013GUIMARÃES, R. M. L. et. al. Relating visual evaluation of soil structure to other physical properties in soils of contrasting texture and management. Soil and Tillage Research, Amsterdam, v. 127, n. 1, p. 92-99, 2013.) and Oliveira et al. (2016OLIVEIRA, S. P. et al. Conversion of forest into irrigated pasture. I. Changes in the chemical and biological properties of the soil. Catena, Destedt, v. 137, n. 1, p. 508-516, 2016.), irrigation can affect the quality and quantity of soil OM, favouring microbial activity in the soil. Also, the C/N ratio (Table 2) of the deposited material may have contributed to the reduction of TOC. Organic materials added to soils with a low C/N ratio tend to mineralise the N and immobilise the C in native soil, and their addition may result in depletion of OM content. The tillage exposes the previously protected OM from labile OM and therefore promotes microbial activity (STEVENSON, 1994STEVENSON, F. J. Humus chemistry: genesis, composition, reactions. 2. ed. New York: John Willey, 1994. 496p.). This decrease in TOC, however, can be reduced depending on the vegetation cover quality between rows used in cultivation. Thus, the treatment 75% L and 25% NL led to minor losses of TOC (Table 3).

The type of vegetative cover between rows also modified the humified soil OM composition (Table 3). The treatments 50% L and 50% NL as well as 25% L and 75% NL promoted increases in the order of 56% in the FA fraction and 75% and 60% in the HA fraction, respectively, in relation to the control (SV). These increases were significantly greater than those observed in the other treatments (100% NL to 75% L and 25% NL) (Table 3), which varied in the order of 2-6% HA, or were similar to those found in soil samples from the control treatment (SV).

The most C portions of the HSs of the soil were found in the humic fraction (26-56%), which is intimately associated with the mineral phase and therefore the insoluble alkaline extraction solution (Table 3). These results are in agreement with those of Canellas et al. (2010CANELLAS, L. P. et al. Chemical composition and bioactivity properties of size-fractions separated from a vermicompost humic acids. Chemosphere, Oxford, v. 78, n. 4, p. 457-466, 2010.), indicating that the humic fraction typically predominates in tropical soils. The greatest amount of H in this study compared with other alkaline soluble fractions favours the structural stability of sandy textured soils, due to its greater recalcitrance compared to labile fractions, such as fulvic and humic acids. Moreover, most authors refer to this fraction as being chemically stable and subject to little variation in management (STEVENSON, 1994STEVENSON, F. J. Humus chemistry: genesis, composition, reactions. 2. ed. New York: John Willey, 1994. 496p.). However, in this study, there was variation in the humic fraction, probably due to the high C cycling in this environment, maximised by wetting and drying cycles caused by high air temperature and humidity resulting from irrigation. In addition, a 'priming effect' could be happening with the humic fraction C decomposition, which was previously present in the soil, due to the addition of plant mixtures; the addition could stimulate the activity of decomposing microorganisms, which, in addition to using the OM present in plant mixtures, also might have degraded the native soil OM (KUZYAKOV et al., 2000KUZYAKOV, Y. et al. Review of mechanisms and quantification of priming effects. Soil Biology and Biochemistry, Oxford, v. 32, n. 11/12, p. 1485-1498, 2000.). This process is called the 'positive priming' effect (SILVA, 2005SILVA, F. A. M. Qualidade de compostos orgânicos produzidos com resíduos do processamento de plantas medicinais. 2005. 102f. Tese (Doutorado em Agronomia: Área de Concentração em Energia da Agricultura) - Universidade Estadual Paulista Júlio de Mesquita Filho, Botucatu, 2005.), which is characterized by intense microbial activity derived from plant mixtures added to an autochthonous soil (TRUU et al., 2009TRUU, M. et al. Microbial biomass, activity and community composition in constructed wetlands. Science of the Total Environment, Amsterdam, v. 407, n. 13, p. 3958-3971. 2009.). Rather, there was large variation in the C content in the H fraction, and this variation is associated with different compositions of plant mixtures used in the vegetative cover between rows. Cultivation with a non-leguminous plant mixture (100% NL) resulted in a 33% decrease in C in the H fraction, while the plant mixtures formed of only leguminous plants (100% L) increased C by 29% in this fraction (Table 3). The 25% non-leguminous plant mixture introduction (75% L and 25% NL) decreased the C content in H (23%), whereas the plant mixtures consisting of 50% L and 50% NL had little influence on the C content in H. With the 25% L and 75% NL plant mixture, there was an increase of approximately 44% in CH compared to the control (SV). These results demonstrate that leguminous species and non- leguminous species combinations may promote an increase in C content in HSs compared to the spontaneous vegetation (SV).

The C content contained in the three humic fractions (CH + CHA + CFA) represents more than half of the TOC, ranging from 52% in T1 to 91% in T5 (Table 4). The changes observed in the C content in different humic fractions are reflected in the relationships that express the soil OM quality (Table 4).

The CHA/CFA ratio, which is an indicator of the degree of soluble OM condensation, shows the limited OM evolution, probably due to recent contributions from vegetative biomass by plant mixtures (LABRADOR MORENO, 1996LABRADOR MORENO, J. La matéria organica en los agrosistemas. Madrid: Ministerio de Agricultura, Pesca y Alimentación, 1996. 176 p.). On the other hand, the low content of exchangeable bases also has facilitated a reduction in the intensity of the humification processes (condensation and biosynthesis) and, consequently, can be taken to obtain a low CHA/CFA ratio (STEVENSON, 1994STEVENSON, F. J. Humus chemistry: genesis, composition, reactions. 2. ed. New York: John Willey, 1994. 496p.). Therefore, the results obtained indicate that the OM studied is characteristic of FA and a low humification degree. According to Orlov (1985ORLOV, R. Humic acids of soils. Washington, DC: USDA, 1985. 378 p.), soil fertility is directly related to soil biological productivity, and larger and more diverse biological activity promotes the humification process and organic compound stabilisation in the CHA fraction. In this case, the greater FA presence can be associated with the soil quality and sandy loam (900 g kg-1 of sand), which favours low nutrient levels that do not favour biological activity and evolution of soil OM processes.

The ratio CH/(CHA + CFA) is an indirect indicator of OM structural stability (CUNHA et al., 2009bCUNHA, T. J. F. et al. Soil organic matter and fertility of anthropogenic dark earths (terra preta de índio) in the Brazilian Amazon basin. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 1, p. 85-93, 2009b.), and it was verified that the 100% NL as well as 50% L and 50% NL treatments were those that exhibited the lowest values, suggesting that these plant mixtures provide greater HSs soluble base.

The ratio CH + CHA + CFA/TOC reports on the soil OM humification degree, and significant differences were observed among treatments, with the 25% L and 75% NL treatment showing the largest values. This treatment promoted OM formation of a higher humification degree compared to the other treatments, suggesting that the input of organic wastes with a low C/N ratio (Table 2) favours less-developed HS formation. Values ranged from 63.8-118.0, slightly out of the range typically found by Cunha et al. (2009aCUNHA, T. J. F. et al. Ácidos húmicos em solo fertirrigado no vale do São Francisco. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 6, p. 1583-1592, 2009a.), whose values ranged from 65-92. It is likely that the leguminous species consortium and non-leguminous species decreased the TOC content (priming effect) and increased the relative proportion of HSs due to the loss of labile C.

Humification can be characterised as a complex process based on biosynthesis and re- synthesis of organic compounds that are incorporated into the soil (ZECH et al., 1997ZECH, W. et al. Factors controlling humification and mineralization of soil organic matter in the tropics. Geoderma , Amsterdam, v. 79, n. 1/4, p. 117-161, 1997.). These transformations include an oxidation reactions series, dehydration, hydrolysis, decarboxylation and condensation, which lead, as a rule, to a relative increase in N-containing HSs. The extent of these changes is influenced by intrinsic conditions, such as soil reaction and exchangeable base content; by organic waste quality and quantity; and by soil management (CANELLAS et al., 2007CANELLAS, L. P. et al. Estoque e qualidade da matéria orgânica de um solo cultivado com cana-de- açúcar por longo tempo. Revista Brasileira de Ciência do Solo , Viçosa, v. 31, n. 2, p. 331-340, 2007.).

Infrared spectroscopy of humic acids

The spectra obtained and the absorption band located between 900 and 1,400 cm-1 were typical of HAs of soils (STEVENSON, 1994STEVENSON, F. J. Humus chemistry: genesis, composition, reactions. 2. ed. New York: John Willey, 1994. 496p.; SENESI et al., 2003SENESI, N. et al. Humic acids in the first generation of eurosoils. Geoderma , Amsterdam, v. 116, n. 3/4, p. 325-344, 2003.; CUNHA et al., 2009bCUNHA, T. J. F. et al. Soil organic matter and fertility of anthropogenic dark earths (terra preta de índio) in the Brazilian Amazon basin. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 1, p. 85-93, 2009b.; VAZ Jr. et al., 2015). While not expected, insignificant changes were observed among the spectra of the HAs studied (Figure 1), which may be related to the short time period (3 years). The band around 1,080 cm-1 for the C-O stretching of polysaccharides was well defined for the HAs of all treatments. In the region below 1,800 cm-1, called the 'fingerprint', small differences were observed among the spectra. The similar characteristics but with different relative intensities were: (a) an intense and large absorption band centred between 3,300 and 3,470 cm-1, attributed to the stretching of phenolic - OH groups (contributed by OH and an aliphatic amine); (b) one absorption band between 2,928 and 2,934 cm-1 and one less intense between 2,853 and 2,857 cm-1, attributed to the symmetric and asymmetric stretching of aliphatic CH, respectively; (c) one absorption band between 1,246 and 1,250 cm -1, attributed to the symmetric stretching of a carboxylate group; and (d) a strong absorption band in the region of 1,049-1,080 cm-1, attributed to C-O stretching of polysaccharides and silicate impurities. In general, the data obtained by infrared spectroscopy (DRIFT) suggest an aliphatic sample characteristic, a large amount of carboxylic groups and a large number of N groups in all the studied HAs - chain polypeptides, for example (SENESI et al., 2003).

The hydrophobicity indices (Table 5) varied from 0.28-0.41. Analysis of these data reveals that the HA control treatment (SV) is distinct from those HA treatments that involved mixtures (100% NL to 25% L and 75% NL), because HA control treatment is more hydrophobic. The more hydrophobicity observed for the HA of the control is probably related to more OM stability and selective preservation of the more recalcitrant C during the decomposition process of plant material (ZECH et al., 1997ZECH, W. et al. Factors controlling humification and mineralization of soil organic matter in the tropics. Geoderma , Amsterdam, v. 79, n. 1/4, p. 117-161, 1997.). As with other treatments, there was greater plant material amounts with a low C/N ratio; this may have contributed to a lower degree of humification of these HA treatments compared to the control.

CONCLUSIONS

The plant mixtures cultivated between rows of the mango trees caused changes in the chemical composition of the soil OM. The 25% L and 75% NL treatment was the best for minimising losses of soil TOC. Most of the organic C was formed by HSs, with FA being the dominant fraction among the alkali soluble fractions. HAs featuring large amounts of carboxyl and N groups reflect low levels of evolution and recalcitrance of this humic fraction.

REFERENCES

  • BENITES, V. M. Caracterização de solos e de substâncias húmicas em áreas de vegetação rupestre de altitude. 2002. 71f. Tese (Doutorado em Solos e Nutrição de Plantas: Área de Concentração Manejo de Solo e Meio Ambiente) - Universidade Federal de Viçosa, Viçosa, 2002.
  • BENITES, V. M. et al. Extração e fracionamento quantitativo de substâncias húmicas do solo: um procedimento simplificado de baixo custo. Rio de Janeiro: Embrapa Solos, 2003. 7p. (Embrapa Solos. Comunicado Técnico, 16).
  • BRANCALIÃO, S. R.; MORAES, M. H. Alterações de alguns atributos físicos e das frações húmicas de um Nitossolo Vermelho na sucessão milheto-soja em sistema plantio direto. Revista Brasileira de Ciência do Solo, Viçosa, v. 32, n. 1, p. 393-404, 2008.
  • CANELLAS, L. P. et al. Chemical composition and bioactivity properties of size-fractions separated from a vermicompost humic acids. Chemosphere, Oxford, v. 78, n. 4, p. 457-466, 2010.
  • CANELLAS, L. P. et al. Estoque e qualidade da matéria orgânica de um solo cultivado com cana-de- açúcar por longo tempo. Revista Brasileira de Ciência do Solo , Viçosa, v. 31, n. 2, p. 331-340, 2007.
  • CARLETTI, P. et al. Structural characterization of humic-like substances with conventional and surface-enhanced spectroscopic techniques Journal of Molecular Structure, Amsterdam, v. 982, n. 1/3, p. 169-175, 2010.
  • CLASSEN, M. E. et al. Manual de métodos de análise de solos. 2. ed. Rio de Janeiro, RJ: EMBRAPA-CNPS, 1997. 212 p.
  • CUNHA, T. J. F. et al. Ácidos húmicos em solo fertirrigado no vale do São Francisco. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 6, p. 1583-1592, 2009a.
  • CUNHA, T. J. F. et al. Soil organic matter and fertility of anthropogenic dark earths (terra preta de índio) in the Brazilian Amazon basin. Revista Brasileira de Ciência do Solo , Viçosa, v. 33, n. 1, p. 85-93, 2009b.
  • USDA. Department of Agriculture. Keys to soil taxonomy. 11a. ed. Washington, DC, 2010. 338 p.
  • FERRARI, E. et al. DRIFIT and HR MAS NMR characterization of humic substances from a soil treated with different organic and mineral fertilizers. Journal of Molecular Structure , Amsterdam, v. 998, n. 1/3, p. 216-224, 2011.
  • FRANCIOSO, O. et al. Structural differences of chernozem soil humic acids SEC-PAGE fractions revealed by thermal (TG-DTA) and spectroscopic (DRIFT) analyses. Geoderma, Amsterdam, v. 152, n. 3/4, p. 264-268, 2009.
  • FREIXO, A. A. Propriedades espectrais da matéria orgânica leve-livre e leve intra-agregado de dois Latossolos sob plantio direto e preparo convencional. Revista Brasileira de Ciência do Solo , Viçosa, v. 26, n. 2, p. 445-453, 2002.
  • GUIMARÃES, R. M. L. et. al. Relating visual evaluation of soil structure to other physical properties in soils of contrasting texture and management. Soil and Tillage Research, Amsterdam, v. 127, n. 1, p. 92-99, 2013.
  • KONONOVA, M. M. Matéria orgânica del suelo: su natureza, propriedades y métodos de investigacion. Barcelona: Oikos-Tau, 1982. 365 p.
  • KUZYAKOV, Y. et al. Review of mechanisms and quantification of priming effects. Soil Biology and Biochemistry, Oxford, v. 32, n. 11/12, p. 1485-1498, 2000.
  • LABRADOR MORENO, J. La matéria organica en los agrosistemas. Madrid: Ministerio de Agricultura, Pesca y Alimentación, 1996. 176 p.
  • LIEFELD, J. Application of diffuse reflectance FT- IR spectroscopy and partial least-squares regression to predict NMR properties of soil organic matter. European Journal of Soil Science, Oxford, v. 57, n. 6, p. 846-857, 2006.
  • MAO, J. et al. Influence of animal manure application on the chemical tructures of soil organic matter as investigated by advanced solid-state NMR and FT-IR spectroscopy. Geoderma , Amsterdam, v. 146, n.1/2, p. 353-362, 2008.
  • NOVOTNY, E. H. Estudos espectroscópicos e cromatográficos de substâncias húmicas de solos sob diferentes sistemas de preparo. 2002. 215 p. Tese (Doutorado em Ciências: Área de Concentração em Físico Química) - Universidade de São Paulo, São Carlos, 2002.
  • OLIVEIRA, S. P. et al. Conversion of forest into irrigated pasture. I. Changes in the chemical and biological properties of the soil. Catena, Destedt, v. 137, n. 1, p. 508-516, 2016.
  • OLK, D. C. et al. Decrease in humification of organic matter with intensified lowland rice cropping a wet chemical and spectroscopic investigation. Soil Science Society of America Journal, Madison, v. 64, n. 4, p. 1337-1347, 2000.
  • ORLOV, D. S. Organic substances of russian soils. European Journal of Soil Science , Oxford, v, 31, n. 9, p. 946-953, 1998.
  • ORLOV, R. Humic acids of soils. Washington, DC: USDA, 1985. 378 p.
  • PICCOLO, A. Carbon sequestration in agricultural soils. Heidleberg, Germany: Springer Verlag, 2012. 306p.
  • SALCEDO, I. H.; SAMPAIO, E. V. S. B. Matéria orgânica do solo no Bioma Caatinga. In: SANTOS, G.; CAMARGO, F. (Eds.). Fundamentos da matéria orgânica do solo: ecossistemas tropicais e subtropicais. 2. ed. Porto Alegre: Metrópole, 2008. p. 419-441.
  • SCHNITZER, M. Organic matter characterization. In: PAGE, A. L. et al. (Ed.). Methods of soil analysis: chemical and microbiological properties. 2. ed. Madison: SSSA, 1982. v. 2, cap. 30, p. 581-594. (SSSA. Agronomy, 9).
  • SENESI, N. et al. Humic acids in the first generation of eurosoils. Geoderma , Amsterdam, v. 116, n. 3/4, p. 325-344, 2003.
  • SILVA, F. A. M. Qualidade de compostos orgânicos produzidos com resíduos do processamento de plantas medicinais. 2005. 102f. Tese (Doutorado em Agronomia: Área de Concentração em Energia da Agricultura) - Universidade Estadual Paulista Júlio de Mesquita Filho, Botucatu, 2005.
  • SILVA, F. C. da. Manual de análises químicas de solos, plantas e fertilizantes. Brasília, DF: Embrapa Comunicação para Transferência de Tecnologia, 1999. 37p.
  • SIMONETTI, G. et al. Characterization of humic carbon in soil aggregates in a long-term experiment with manure and mineral fertilization. Soil Science Society of America Journal , Madison, v. 76, n. 3, p. 880-890, 2012.
  • SIMPSON, A. J. Determining the molecular weight, aggregation, structures and interactions of natural organic matter using diffusion ordered spectroscopy. Magnetic Resonance in Chemistry, Chichester, v. 40, n. 13, p. S72-S82, 2002.
  • STEVENSON, F. J. Humus chemistry: genesis, composition, reactions. 2. ed. New York: John Willey, 1994. 496p.
  • SWIFT, R. S. Organic matter characterization. In: SPARKS, D. L et al. (Eds.). Methods of soil analysis: chemical methods. Madison: SSSA , 1996. v. 3, p. 1011-1020. (SSSA. Book Series, 5).
  • TRUU, M. et al. Microbial biomass, activity and community composition in constructed wetlands. Science of the Total Environment, Amsterdam, v. 407, n. 13, p. 3958-3971. 2009.
  • VAZ JUNIOR, S. et al. Study of molecular interactions between humic acid from Brazilian soil and the antibiotic oxytetracycline. Environmental Technology & Innovation, Amsterdam, v. 4, p. 260-267, 2015.
  • YEOMANS, J. C.; BREMNER, J. M. A rapid and precise method for routine determination of organic carbon in soil. Communications in Soil Science and Plant Analysis, New York, v. 19, n. 13, p. 1467-1476, 1988.
  • ZECH, W. et al. Factors controlling humification and mineralization of soil organic matter in the tropics. Geoderma , Amsterdam, v. 79, n. 1/4, p. 117-161, 1997.

Publication Dates

  • Publication in this collection
    Jul-Sep 2016

History

  • Received
    14 Aug 2014
  • Accepted
    18 Apr 2016
Universidade Federal Rural do Semi-Árido Avenida Francisco Mota, número 572, Bairro Presidente Costa e Silva, Cep: 5962-5900, Telefone: 55 (84) 3317-8297 - Mossoró - RN - Brazil
E-mail: caatinga@ufersa.edu.br