Acessibilidade / Reportar erro

Isolation and characterization of micro cellulose obtained from waste mango

Abstract

Cellulose fibers are used in polymeric matrices due to their thermal and mechanical capabilities. These biopolymers can be isolated from different natural sources. In this study, micro cellulose was obtained from mango fibrous endocarp (Mangifera caesia Jack ex Wall) waste. Isolation was performed using sulfuric acid and sodium hydroxide as removal agents of lignin and hemicellulose. A comparative analysis between native fiber (NF) and micro cellulose (MC) was performed, using FTIR, DSC and SEM techniques to assess their composition, thermal, structural and crystallinity capacities. The structures of the fibers were not damaged due to the chemical treatments received and their sizes ranged between 40 and 400 μm in length. The extraction of cellulose from mango waste represents an important start in obtaining biopolymers that can be used in the food, pharmaceutical, and other industries.

Keywords:
biopolymers; mango waste; micro cellulose

1. Introduction

In recent years, different ways have been sought to control and reduce pollution caused by petroleum-derived materials[11 Derraik, J. G. (2002). The pollution of the marine environment by plastic debris: a review. Marine Pollution Bulletin, 44(9), 842-852. http://dx.doi.org/10.1016/S0025-326X(02)00220-5. PMid:12405208.
http://dx.doi.org/10.1016/S0025-326X(02)...

2 O’Brine, T., & Thompson, R. C. (2010). Degradation of plastic carrier bags in the marine environment. Marine Pollution Bulletin, 60(1), 2279-2283. https://doi.org/10.1016/j.marpolbul.2010.08.005.
https://doi.org/10.1016/j.marpolbul.2010...
-33 Webb, H. K., Arnott, J., Crawford, R. J., & Ivanova, E. P. (2012). Plastic degradation and its environmental implications with special reference to poly (ethylene terephthalate). Polymers, 5(1), 1-18. https://doi.org/10.3390/polym5010001.
https://doi.org/10.3390/polym5010001...
], and one of the most used alternatives in the world is recycling[44 Das, O., Sarmah, A. K., & Bhattacharyya, D. (2015). A sustainable and resilient approach through biochar addition in wood polymer composites. The Science of the Total Environment, 512-513, 326-336. http://dx.doi.org/10.1016/j.scitotenv.2015.01.063. PMid:25634737.
http://dx.doi.org/10.1016/j.scitotenv.20...
]. However, and even though this conventional method reduces the ecological impact of plastic has not been possible to reduce the large volumes of those kinds of materials that are produced each year[55 Geyer, R., Jambeck, J. R., & Law, K. L. (2017). Production, use, and fate of all plastics ever made. Science advances, 3(7), e1700782. https://doi.org/10.1126/sciadv.1700782.
https://doi.org/10.1126/sciadv.1700782...
]. This is mainly due to the fact that a great many products are manufactured in way that make the plastic content difficult to separate and, therefore very little plastic is currently recycled. An estimated 8.3 billion metric tons of plastic are produced worldwide, and only 9% goes through recycling process[44 Das, O., Sarmah, A. K., & Bhattacharyya, D. (2015). A sustainable and resilient approach through biochar addition in wood polymer composites. The Science of the Total Environment, 512-513, 326-336. http://dx.doi.org/10.1016/j.scitotenv.2015.01.063. PMid:25634737.
http://dx.doi.org/10.1016/j.scitotenv.20...
]. Another alternative for reducing environmental impact of plastics that has been proposed is the use of polymeric materials combined with pro-oxidizing substances[22 O’Brine, T., & Thompson, R. C. (2010). Degradation of plastic carrier bags in the marine environment. Marine Pollution Bulletin, 60(1), 2279-2283. https://doi.org/10.1016/j.marpolbul.2010.08.005.
https://doi.org/10.1016/j.marpolbul.2010...
,66 Ojeda, T. F., Dalmolin, E., Forte, M. M., Jacques, R. J., Bento, F. M. & Camargo, F. A. (2009). Abiotic and biotic degradation of oxo-biodegradable polyethylenes. Polymer degradation and stability, 94(6), 965-970. https://doi.org/10.1016/j.polymdegradstab.2009.03.011.
https://doi.org/10.1016/j.polymdegradsta...
]. However, the wastes the materials generate after they are degraded consist of heavy metals and petroleum products[77 Wiles, D. M., & Scott, G. (2006). Polyolefins with controlled environmental degradability. Polymer degradation and stability, 91(7), 1581-1592. https://doi.org/10.1016/j.polymdegradstab.2005.09.010.
https://doi.org/10.1016/j.polymdegradsta...

8 Kyrikou, I., & Briassoulis, D. (2007). Biodegradation of agricultural plastic films: a critical review. Journal of Polymers and the Environment, 15(2), 125-150. https://doi.org/10.1007/s10924-007-0053-8.
https://doi.org/10.1007/s10924-007-0053-...
-99 Ammala, A., Bateman, S., Dean, K., Petinakis, E., Sangwan, P., Wong, S., & Leong, K. H. (2011). An overview of degradable and biodegradable polyolefins. Progress in Polymer Science, 36(8), 1015-1049. https://doi.org/10.1016/j.progpolymsci.2010.12.002.
https://doi.org/10.1016/j.progpolymsci.2...
], which tend to accumulate in the environment.

To reduce the impact caused by petroleum products, there is need to look for materials that have a polymer matrix formed by biopolymers (natural polymers)[1010 Flieger, M., Kantorova, M., Prell, A., Řezanka, T., & Votruba, J. (2003). Biodegradable plastics from renewable sources. Folia Microbiologica, 48(1), 27-44. http://dx.doi.org/10.1007/BF02931273. PMid:12744074.
http://dx.doi.org/10.1007/BF02931273...
]. Since the polymer matrix is biodegradable, the composite which is its integral part is also expected to be biodegradable and environmentally friendly[1111 Sagnelli, D., Hooshmand, K., Kemmer, G. C., Kirkensgaard, J. J. K., Mortensen, K., Giosafatto, C. V. L., Holse, M., Hebelstrup, K. H., Bao, J., Stelte, W., Bjerre, A. B., & Blennow, A. (2017). Cross-linked amylose bio-plastic: a transgenic-based compostable plastic alternative. International Journal of Molecular Sciences, 18(10), 2075-2086. http://dx.doi.org/10.3390/ijms18102075. PMid:28973963.
http://dx.doi.org/10.3390/ijms18102075...
]. It has an advantage over synthetic polymers because it can be totally biodegraded in products such as CO2, water and also organic fertilizer[1212 Bastioli, C. (2001). Global status of the production of biobased packaging materials. Starch, 53(8), 351-355. https://doi.org/10.1002/1521-379X(200108)53:8<351::AID-STAR351>3.0.CO;2-R.
https://doi.org/10.1002/1521-379X(200108...
]. Bioplastics and biomaterials are composed of biopolymers, biomass derivatives or blends of natural and synthetic polymers, which will help mitigate environmental issues in the long term.

Agriculture in Mexico is an important productive sector. The role played by the agricultural sector in the other sectors of Mexico such as economic, social and environmental determines its greatest impact on the country's development. However, post-harvest residues (bagasse, husks and seeds) represent an environmental problem due to the followings: lack of distribution channels, pests, environmental policies and inadequate management of the large volumes of residues that are generated. Therefore, agricultural residues must be considered as a source sustainable and renewable of products such as cellulose fiber used as a reinforcement for bioplastics and biomaterials, due to their high availability and potential[1313 Burgos, N., Valdés, A., & Jiménez, A. (2016). Valorization of agricultural wastes for the production of protein-based biopolymers. Journal of Renewable Materials, 4(3), 165-177. http://dx.doi.org/10.7569/JRM.2016.634108.
http://dx.doi.org/10.7569/JRM.2016.63410...
,1414 Vestena, M., Gross, I. P., Muller, C. M. O., & Pires, A. T. N. (2016). Isolation of whiskers from natural sources and their dispersed in a non-aqueous medium. Polímeros: Ciência e Tecnologia, 26(4), 327-335. http://dx.doi.org/10.1590/0104-1428.2367.
http://dx.doi.org/10.1590/0104-1428.2367...
].

Waste mango (seeds, shells and cotyledons) currently have little impact on industries. However, these wastes have been studied to obtain active and functional compounds from mango peel[1515 Serna-Cock, L., García-Gonzales, E., & Torres-León, C. (2016). Agro-industrial potential of the mango peel based on its nutritional and functional properties. Food Reviews International, 32(4), 364-376. http://dx.doi.org/10.1080/87559129.2015.1094815.
http://dx.doi.org/10.1080/87559129.2015....
], antioxidants from cotyledon[1616 Torres-León, C., Rojas, R., Serna-Cock, L., Belmares-Cerda, R., & Aguilar, C. N. (2017). Extraction of antioxidants from mango seed kernel: optimization assisted by microwave. Food and Bioproducts Processing, 105, 188-196. http://dx.doi.org/10.1016/j.fbp.2017.07.005.
http://dx.doi.org/10.1016/j.fbp.2017.07....
], to use fibrous endocarp in the paper industry[1717 Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L. A., Rosa, M. F., de sá M. Sousa, M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. Macromolecular Symposia, 344(1), 39-54. http://dx.doi.org/10.1002/masy.201300217.
http://dx.doi.org/10.1002/masy.201300217...
], and pulp to extract reducing sugars, organic acids and obtain alcohols[1818 Giraldo, L. M., Correa, H. M., Gutiérrez, J. B., & Castano, C. (2007). Aprovechamiento del residuo agroindustrial del mango común (Mangifera indica L.) en la obtención de azúcares fermentables. Ingeniería y Ciencia-ing, 3(6), 41-62.].

Unlike other agricultural residues, mango (Mangifera caesia Jack ex Wall) is composed of more than one biopolymer, with starch and cellulose being the highest proportion[1919 Lorenzo‐Santiago, M. A., Juárez-López, A. L., Rosas‐Acevedo, J. L., Rendón‐Villalobos, J. R., Toribio‐Jiménez, J., & García Hernández, E. (2018). Management and final disposal of mango waste in the State of Guerrero, Mexico: a brief review. The Journal of Agricultural Science, 10(12), 34-41. http://dx.doi.org/10.5539/jas.v10n12p34.
http://dx.doi.org/10.5539/jas.v10n12p34...
]. In addition to being the most abundant natural polymer in the world, cellulose is renewable, biodegradable and is a classic example of a matrix reinforcement biopolymer[2020 Béguin, P., & Aubert, J.-P. (1994). The biological degradation of cellulose. FEMS Microbiology Reviews, 13(1), 25-58. http://dx.doi.org/10.1111/j.1574-6976.1994.tb00033.x. PMid:8117466.
http://dx.doi.org/10.1111/j.1574-6976.19...
,2121 Azizi Samir, M. A. S., Alloin, F., & Dufresne, A. (2005). Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field. Biomacromolecules, 6(2), 612-26. http://dx.doi.org/10.1021/bm0493685. PMid:15762621.
http://dx.doi.org/10.1021/bm0493685...
] and is expected to be able to reduce and replace many non-renewable materials[2222 Wegner, T. H., & Jones, P. E. (2006). Advancing cellulose-based nanotechnology. Cellulose (London, England), 13(2), 115-118. http://dx.doi.org/10.1007/s10570-006-9056-1.
http://dx.doi.org/10.1007/s10570-006-905...
]. The fibrils have two regions known as crystalline (ordered regions) and amorphous (disorder regions). These regions can be hydrolysed with acid, which makes the crystalline region less reactive. This results in obtaining microfibrils, nanofibrils and cellulose nanocrystals[2323 Beck-Candanedo, S., Roman, M., & Gray, D. G. (2005). Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules, 6(2), 1048-1054. http://dx.doi.org/10.1021/bm049300p. PMid:15762677.
http://dx.doi.org/10.1021/bm049300p...
].

Cellulose fibers have an inherent structural hierarchy that originates from their different biological sources. Obtaining cellulose from different lignocellulosic fibers has attracted the interest of the scientific and industrial community in recent years[2424 Cerqueira, J. C., Penha, J. D. S., Oliveira, R. S., Guarieiro, L. L. N., Melo, P. D. S., Viana, J. D., & Machado, B. A. S. (2017). Production of biodegradable starch nanocomposites using cellulose nanocrystals extracted from coconut fibers. Polímeros: Ciência e Tecnologia, 27(4), 320-329. http://dx.doi.org/10.1590/0104-1428.05316.
http://dx.doi.org/10.1590/0104-1428.0531...
,2525 Khenblouche, A., Bechki, D., Gouamid, M., Charradi, K., Segni, L., Hadjadj, M., & Boughali, S. (2019). Extraction and characterization of cellulose microfibers from Retama raetam stems. Polímeros: Ciência e Tecnologia, 29(1), e2019011. http://dx.doi.org/10.1590/0104-1428.05218.
http://dx.doi.org/10.1590/0104-1428.0521...
]. The native fibers are longitudinal, rigid and difficult to modify due to the components that form them. These compounds work as a natural glue that gives firmness and shape. The fibers are hydrophobic. This is why, it is necessary to perform a chemical and thermal treatment, that can help to remove glue from the fibers without damaging the structure of the micro-cellulose fibers. The particle size reduction in the cellulose fibers through thermal treatment, generates major areas of contact with chemical substances such as acids and alkalis. Likewise, with the ultrasonic treatments crystallinity increase and morphology of cellulose fibers can be improved; in the same way individual fibrils could be obtained from waste mango and which serve in developing a bio-based material.

In the present study, cellulose fibers were obtained from waste mango fibrous endocarp. The shape, size and structure of the fibers were characterized as well as the decrease in lignin and hemicellulose subjected to a modified acid-alkaline treatment was analysed by wet chemistry methods. Taking advantage of waste residues facilitates the use of cellulose as a biopolymer and reduces the impact generated by inappropriate disposal of agricultural waste and are able to generate profits.

2. Materials and Methods

2.1 Cellulose insulation

Fibrous endocarp of Ataulfo mango fruits (Mangifera caesia Jack ex Wall) was obtained in the local market in El Arenal, Guerrero, Mexico, and used as the raw material for this study. Reagents used to obtain and extract the native cellulose fibre (NF) and chemically modified cellulose fibre (designated micro-cellulose, MC) were: Sodium hydroxide (CAS 1310-73-2), sulfuric acid (CAS 7664-93-9), hydrogen peroxide (CAS 7722-84-1) and distilled water (CAS 7732-18-5).

2.2 Pre-treatment of mango fibre

The mango endocarp was cut into squares of approximately 1 cm2. 50 g was weighed and hydrated using distilled water. It was stirred at room temperature for two hours and subsequently dried in a continuous flow oven at 40 ° C for 24 hours[2626 Salgado-Delgado, R., Coria-Cortés, L., García-Hernández, E., Galarza, Z. V., Rubio-Rosas, E., & Crispín-Espino, I. (2010). Elaboración de materiales reforzados con carácter biodegradable a partir de polietileno de baja densidad y bagazo de caña modificado. Revista Iberoamericana de Polímeros, 11(7), 520-531. Retrieved in 2019, October 28, from https://dialnet.unirioja.es/servlet/articulo?codigo=3694228
https://dialnet.unirioja.es/servlet/arti...
].

2.3 Alkaline treatment

The dried fibrous endocarp was treated in a solution of 2% NaOH (w/v) in a 1:20 ratio (fibre: solution). The solution was continuously stirred at 80 ° C for two hours. After the alkaline treatment was done, the fibres were washed with distilled water and dried at 40 ° C for 24 h. The dried fibres were grinded using a E3303.00 mini cutting mill (Eberbach Corp.), equipped with a stainless steel blades to separate the bonded fibres formed after the drying process. After the milling was completed, the fibres were stored in an airtight bag[2727 Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L., & Rosa, M. F., Souza, M. S. M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. https://doi.org/10.1002/masy.201300217.
https://doi.org/10.1002/masy.201300217...
].

2.4 Whitening

The fibres were bleached in a 1:20 ratio (fibre: solution) in a solution composed of H2O2 (v/v) and 4% NaOH (w/v), and stirred for two hours at 50 ° C. After the treatment, the fibres were washed with distilled water and dried at 40 ° C for 24 h[2828 Orts, W. J., Shey, J., Imam, S. H., Glenn, G. M., Guttman, M. E., & Revol, J. F. (2005). Application of cellulose microfibrils in polymer nanocomposites. Journal of Polymers and the Environment, 13(4), 301-306. http://dx.doi.org/10.1007/s10924-005-5514-3.
http://dx.doi.org/10.1007/s10924-005-551...
].

2.5 Acid Hydrolysis

The fibres were subjected to 52% sulfuric acid (w/w) in a 1:20 ratio (fibre: solution), and stirred for two hours at 45 ° C. The reaction was stopped by thermal shock using a vessel with distilled water at a temperature of 10 ° C[2727 Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L., & Rosa, M. F., Souza, M. S. M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. https://doi.org/10.1002/masy.201300217.
https://doi.org/10.1002/masy.201300217...
].

2.6 Sonication

Sonication process was applied after acid hydrolysis in order to disintegrate the fibres into microfibres in water dispersion to reduce the viscosity and allow efficient propagation of the vibration through the dispersion. Szymańska-Chargot et al.[2929 Szymańska-Chargot, M., Cieśla, J., Chylińska, M., Gdula, K., Pieczywek, P. M., Koziol, A., Cieślak, K. J., & Zdunek, A. (2018). Effect of ultrasonication on physicochemical properties of apple based nanocellulose-calcium carbonate composites. Cellulose (London, England), 25(8), 4603-4621. http://dx.doi.org/10.1007/s10570-018-1900-6.
http://dx.doi.org/10.1007/s10570-018-190...
] and Kasuga et al.[3030 Kasuga, T., Isobe, N., Yagyu, H., Koga, H., & Nogi, M. (2018). Clearly transparent nanopaper from highly concentrated cellulose nanofiber dispersion using dilution and sonication. Nanomaterials (Basel, Switzerland), 8(2), 104. http://dx.doi.org/10.3390/nano8020104. PMid:29439544.
http://dx.doi.org/10.3390/nano8020104...
] techniques were modified and the dispersion was then subjected to sonication for 5 min at a power of 99 W using a Branson 5510MT ultrasonic cleaner. Subsequently, the samples were centrifuged at 6300 x g at 4 °C for 15 min and dried at 40 °C for 24 h[2727 Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L., & Rosa, M. F., Souza, M. S. M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. https://doi.org/10.1002/masy.201300217.
https://doi.org/10.1002/masy.201300217...
].

2.7 Chemical analyses

The lignin content was determined according to Klason method[3131 Technical Association of the Pulp and Paper Industry – TAPPI. (2007). Acid insoluble lignin in wood and pulp. T 222 om-02. Fibrous materials and pulp testing. Atlanta: TAPPI.]. The dried fibres was extracted by using Soxhlet method[3232 Technical Association of the Pulp and Paper Industry – TAPPI. (2007). Solvent extractives of wood and pulp. T 204 cm-97. Fibrous materials and pulp testing. Atlanta: TAPPI.] with ethanol-benzene solvent for 5 h. The residue extract was oven dried at 103 o C for 1 h, and later treated with 72% H2SO4 for 2 h while stirring at 37 o C. The material was then diluted to 3% H2SO4 and then refluxed at 80 o C for 4 h; the sample was oven dried at 105 o C for 1 h and cooled in a desiccator to obtain constant weight. The lignin content was calculate using the following equation[3131 Technical Association of the Pulp and Paper Industry – TAPPI. (2007). Acid insoluble lignin in wood and pulp. T 222 om-02. Fibrous materials and pulp testing. Atlanta: TAPPI.]:

L i g n i n % = A 100 W (1)

Where A is weight of lignin and W is oven-dry weight of test specimen.

Holocellulose was measured by means of the Wise chlorite method[3333 Haykiri-Acma, H., Yaman, S., Alkan, M., & Kucukbayrak, S. (2014). Mineralogical characterization of chemically isolated ingredients from biomass. Energy Conversion and Management, 77, 221-226. http://dx.doi.org/10.1016/j.enconman.2013.09.024.
http://dx.doi.org/10.1016/j.enconman.201...
] through dried fibers treatment with an acidified sodium chlorite water solution in acidic medium at 70-80 ºC for 1 h. The process was repeated until the product became white. Determination of cellulose to dry fiber samples was using the Seifert method[3434 Waliszewska, B., Mleczek, M., Zborowska, M., Goliński, P., Rutkowski, P., & Szentner, K. (2019). Changes in the chemical composition and the structure of cellulose and lignin in elm wood exposed to various forms of arsenic. Cellulose (London, England), 26(10), 6303-6315. http://dx.doi.org/10.1007/s10570-019-02511-z.
http://dx.doi.org/10.1007/s10570-019-025...
] wherein the cellulose is separated from the plant material in the process of its processing with aqueous solutions of acetylacetone, dioxane, and hydrochloric acid. Hemicellulose content was calculated theoretically from the difference in the contents of holocellulose and cellulose.

2.8 Scanning Electron Microscopy (SEM)

Cellulose samples were sprinkled on a double adhesion carbon conductive tape, which was previously fixed on aluminium support of the scanning electron microscope Carl Zeiss EVO LS 10. Subsequently, they were observed at a voltage of 2.5 kV, with a resolution of 3-10 nm; the micrographs were taken at 100, 500 and 1500 magnifications[3535 Rendón-Villalobos, R., García-Hernández, E., Güizado-Rodríguez, M., Salgado Delgado, R., & Rangel-Vázquez, N. A. (2010). Obtención y caracterización de almidón de plátano (Musa paradisiaca L.) acetilado a diferentes grados de sustitución. Afinidad, 67(548), 294-300. Retrieved in 2019, October 28, from https://www.raco.cat/index.php/afinidad /article/view/269205/356773
https://www.raco.cat/index.php/afinidad ...
].

2.9 Fourier Transform Infrared Spectroscopy (FTIR)

Fourier transform infrared spectroscopy (FTIR) studies were carried out using a Perkin-Elmer-Spectrum 100/100 N FT-IR infrared spectrometer. The infrared region was within a range of 4000-650 cm-1 in the transmittance mode, with a resolution of 16 cm-1 and 8 scans[3636 Xu, Y. X., Miladinov, V., & Hanna, M. A. (2004). Synthesis and characterization of starch acetates with high substitution. Cereal Chemistry, 81(6), 735-740. http://dx.doi.org/10.1094/CCHEM.2004.81.6.735.
http://dx.doi.org/10.1094/CCHEM.2004.81....
].

2.10 Differential Scanning Calorimetry (DSC)

A differential scanning calorimeter model TA 2010 was used. The equipment was calibrated with indium (melting point of 156.4 ° C and enthalpy of 6.8 cal/g). Between 5 and 10 mg of the sample was weighed, using a micro balance with an accuracy of ± 0.01 mg, a heating program ranging from 20 to 400 ° C and a heating rate of 10 ° C/min[3030 Kasuga, T., Isobe, N., Yagyu, H., Koga, H., & Nogi, M. (2018). Clearly transparent nanopaper from highly concentrated cellulose nanofiber dispersion using dilution and sonication. Nanomaterials (Basel, Switzerland), 8(2), 104. http://dx.doi.org/10.3390/nano8020104. PMid:29439544.
http://dx.doi.org/10.3390/nano8020104...
]. This generated an inert atmosphere in the cell by circulating 50 mL/min of nitrogen gas (99.9% purity).

2.11 Crystallinity

The degree of crystallinity was determined based on the DSC enthalpy of fusion data using the following equation[3737 Poley, L. H., Siqueira, A. P. L., Silva, M. G., Vargas, H., & Sanchez, R. (2004). Photothermal characterization of low density polyethylene food packages. Polímeros: Ciência e Tecnologia, 14(1), 8-12. http://dx.doi.org/10.1590/S0104-14282004000100007.
http://dx.doi.org/10.1590/S0104-14282004...
]:

X c = Δ H f Δ H o (2)

Where Xc is the crystalline fraction, ΔHf represents the enthalpy of fusion measured by DSC and ΔHo is the enthalpy of fusion for 100% crystalline polymer.

3. Results and Discussions

3.1 Chemical Analyses

Despite the economic importance of agricultural residues there are no studies on the chemical composition of the mango endocarp. In general, the studies of mango chemistry are limited to chemical characteristics based mainly on mango agroindustrial waste, fruit by-products and silage[3838 Guzmán, O., Lemus, C., Martínez, S., Bonilla, J., Plasencia, A., & Ly, J. (2012). Chemical characteristics of silages of mango (Mangifera indica L.) by-products for animal feeding. Canadian Journal of Agricultural Science, 46(4), 369-374.

39 de Carvalho Couto, C. C., Fo., da Silva, J. C., Fo., Prata Neiva, A., Jr.,Magalhães de Souza, R., Rodrigues Nunes, J. A., & Viana Coelho, J. (2010). Fibrous fractions of mango residue silage with additives. Ciência e Agrotecnologia, 34(3), 751-757. http://dx.doi.org/10.1590/S1413-70542010000300031.
http://dx.doi.org/10.1590/S1413-70542010...

40 Serna Cock, L., & Torres León, C. (2014). Potencial agroindustrial de cáscaras de mango de las variedades Keitt, y Tommy Atkins (Mangifera indica). Acta Agronomica, 64(2), 110-115. http://dx.doi.org/10.15446/acag.v64n2.43579.
http://dx.doi.org/10.15446/acag.v64n2.43...
-4141 Balza, M., Garrido, E., García, M., Martínez, J., & García, A. (2017). Chemical characterization of the cellular wall of mango bocado pulp. Revista Agrollanía, 14, 7-13.].

The results of lignin and calculated hemicellulose of native (NF) and alkaline modified (MC) fibers decreased, from 10.32 wt % to 0.64 wt % for NF and 29.75 wt % to 1.11 wt %, respectively. Regarding the lignin content, the data from this investigation (10.32 wt %) is close to those found for Guzmán et al.[3838 Guzmán, O., Lemus, C., Martínez, S., Bonilla, J., Plasencia, A., & Ly, J. (2012). Chemical characteristics of silages of mango (Mangifera indica L.) by-products for animal feeding. Canadian Journal of Agricultural Science, 46(4), 369-374.], different values are obtained when the food-processing industry method is employed (10.32 wt % vs. 9.0 wt %). This method is similar to Klason lignin extraction process, except a final step of 1 h of calcination at 550 ° C. Nevertheless, a large difference is observed between our lignin results: 5.76 wt % and 6.97 wt % of lignin in mango peel and mango by-products, respectively[3939 de Carvalho Couto, C. C., Fo., da Silva, J. C., Fo., Prata Neiva, A., Jr.,Magalhães de Souza, R., Rodrigues Nunes, J. A., & Viana Coelho, J. (2010). Fibrous fractions of mango residue silage with additives. Ciência e Agrotecnologia, 34(3), 751-757. http://dx.doi.org/10.1590/S1413-70542010000300031.
http://dx.doi.org/10.1590/S1413-70542010...
,4040 Serna Cock, L., & Torres León, C. (2014). Potencial agroindustrial de cáscaras de mango de las variedades Keitt, y Tommy Atkins (Mangifera indica). Acta Agronomica, 64(2), 110-115. http://dx.doi.org/10.15446/acag.v64n2.43579.
http://dx.doi.org/10.15446/acag.v64n2.43...
]. Concerning the hemicellulose content from this study (29.75 wt %) is lower than the values (31.1 wt % and 31.75 wt %) found for mango peel and mango by-products[3838 Guzmán, O., Lemus, C., Martínez, S., Bonilla, J., Plasencia, A., & Ly, J. (2012). Chemical characteristics of silages of mango (Mangifera indica L.) by-products for animal feeding. Canadian Journal of Agricultural Science, 46(4), 369-374.,4141 Balza, M., Garrido, E., García, M., Martínez, J., & García, A. (2017). Chemical characterization of the cellular wall of mango bocado pulp. Revista Agrollanía, 14, 7-13.]. This can be explained by the fact that it seems that in the food-processing industry method, the neutral detergent solution used for determining the Neutral Detergent Fibre (NDF = Hemicellulose + Cellulose + Lignin) content does not remove completely hemicelluloses and is, therefore, this method is not indicated for this work.

3.2 Structural characterization

The native fibres presented very dense outer layers covering the cellulose sacs. The layers are considered to be composed of lignin (Figure 1a). Lignin wraps the cellulose in the inner part covering it with its stiffness and hardness property. Small globular particles and pores along the fibres and impurities can be seen on the surface.

Figure 1
Micrograph of (a) native mango fiber, (b) alkaline modified fiber (c) commercial cellulose (Civeq™).

Figure 1b shows an accumulation of long fibres, rolled and small traces of waste. This is attributed to the modification of the alkaline treatment using NaOH. Cellulose fibres can be seen more clearly, but there is still a need to remove part of the lignin and hemicellulose that holds them together. Cellulose fibres are more available due to the removal of lignin, and their characteristics (e.g. shape, surface) are similar to those of the commercially available cellulose (Civeq™); the SEM images (Figure 1b and c) evidence that separate individual fibres have also similar morphologies with smooth surfaces.

The fibres went through a whitening process with the help of hydrogen peroxide. The peroxide helped in eliminating some waxes, tannins, lignin and alkaline treatment residues.

In Figure 2a, completely separated fibres can be seen; there are no traces of lignin, as well as porous components. In this procedure, the decrease in thickness, shape, size and discoloration of the fibres is more evident. After bleaching, the fibres had sizes ranging from 40 to 400 μm in length with a uniform surface and rectangular shape (Figure 2b). Owi et al.[4242 Owi, W. T., Lin, O. H., Sam, S. T., Chia, C. H., Zakaria, S., Mohaiyiddin, M. S., Villagracia, A. R., Santos, G. N., & Md Akil, H. (2016). Comparative study of microcelluloses isolated from two different biomasses with commercial cellulose. BioResources, 11(2), 3453-3465. http://dx.doi.org/10.15376/biores.11.2.3453-3465.
http://dx.doi.org/10.15376/biores.11.2.3...
] characterized micro celluloses chemically isolated from fruit bunches and sugarcane bagasse as well as commercial cellulose (Sigma™) with characteristics similar to those obtained in this study after extraction, particularly in fibre shape and surface. In a similar manner, Nagalakshmaiah et al.[4343 Nagalakshmaiah, M., El Kissi, N., Mortha, G., & Dufresne, A. (2016). Structural investigation of cellulose nanocrystals extracted from chili leftover and their reinforcement in cariflex-IR rubber latex. Carbohydrate Polymers, 136, 945-954. http://dx.doi.org/10.1016/j.carbpol.2015.09.096. PMid:26572433.
http://dx.doi.org/10.1016/j.carbpol.2015...
] report the shape and diameter of chili fibres by an acid hydrolysis treatment. These results are consistent with the ones presented in this study. This is consistent with the report of Atiqah et al.[4444 Atiqah, M. S. N., Gopakumar, D. A., F A T, O., Pottathara, Y. B., Rizal, S., Aprilia, N. A. S., Hermawan, D., Paridah, M. T. T., Thomas, S., & H P S, A. K. (2019). Extraction of cellulose nanofibers via eco-friendly supercritical carbon dioxide treatment followed by mild acid hydrolysis and the fabrication of cellulose nanopapers. Polymers, 11(11), 1813. PMid:31694184.] that amorphous portions like hemicellulose and lignin were removed after applying alkali and bleaching treatment on raw kenaf fibre.

Figure 2
Micrographs of (a) fiber after whitening (b) and thermal sonication.

3.3 FTIR analysis

Figure 3 shows the native and micro cellulose fibre spectra obtained by chemical treatment. A peak at 894 cm-1 can be observed in the fibres with alkaline treatment. This signal is attributed to the anomeric carbon present in cellulose. This signal is usually in the polysaccharide absorbing region of 950-700 cm-1[4545 Sun, J. X., Xu, F., Sun, X. F., Xiao, B., & Sun, R. C. (2005). Physico-chemical and thermal characterization of cellulose from barley straw. Polymer Degradation & Stability, 88(3), 521-531. http://dx.doi.org/10.1016/j.polymdegradstab.2004.12.013.
http://dx.doi.org/10.1016/j.polymdegrads...
], which also reveals the component structures of ß-glucans. In the native fibre, this signal is less intense because the lignin coating prevents specific cellulose signals that are present.

Figure 3
Infrared spectra for native fibre (NF) and micro cellulose (MC).

A narrow absorption band in the region between 1170 and 920 cm-1 ascribed to the polysaccharide bands, shows the stretching vibrations of C – O, C – C and C – OH bonds which are in cellulose, lignin and hemicellulose[4646 Yang, H., Yan, R., Chen, H., Lee, D. H., & Zheng, C. (2007). Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel, 86(12-13), 1781-1788. http://dx.doi.org/10.1016/j.fuel.2006.12.013.
http://dx.doi.org/10.1016/j.fuel.2006.12...
,4747 Naranjo, D. C., Alamilla-Beltran, L., Gutierrez-Lopez, G. F., Terres-Rojas, E., Solorza-Feria, J., Romero-Vargas, S., Yee-Madeira, H., Flores-Morales, A., & Mora-Escobedo, R. (2016). Isolation and characterization of cellulose obtained from Agave salmiana fibers using two acid-alkali extraction methods. Revista Mexicana de Ciencias Agrícolas, 7(1), 31-43. http://dx.doi.org/10.29312/remexca.v7i1.368.
http://dx.doi.org/10.29312/remexca.v7i1....
]. A strong signal was present in the treated fibre (MC). The increase is based on the breaking of bonds during the acid-base treatment.

The native fibre showed a peak at 1230 cm-1. This is attributed to the stretching of C – O – C bonds of the alkyl-aryl ether, a compound present in the structure of lignin, and which disappears after chemical treatments. This means that the lignin was partially removed in MC as further confirmed by lignin results obtained in this work, 0.64 wt %, determined by Klason method.

In the native fibre, a characteristic peak of the C–C unsaturated bond of lignin was observed at 1623 cm-1, as well as the water band at 1636 cm-1 decreasing its intensity slightly in MC. Likewise, a peak of 1740 cm-1 is observed, which is associated with the stretching of C = O groups linked to aliphatic carboxylic acid and ketone. It could also be related to the presence of hemicellulose[4848 Sgriccia, N., Hawley, M. C., & Misra, M. (2008). Characterization of natural fiber surfaces and natural fiber composites. Composites. Part A, Applied Science and Manufacturing, 39(10), 1632-1637. http://dx.doi.org/10.1016/j.compositesa.2008.07.007.
http://dx.doi.org/10.1016/j.compositesa....
]. Both peaks decrease when the sample is subjected to acid treatment, which allows the removal of lignin and hemicellulose, from 10.32 to 0.64 wt % and from 29.75 to 1.11 wt %, respectively.

Between the range of 2940 and 2800 cm-1, there are two peaks in the native cellulose, which are assigned to the asymmetrical and symmetrical stretching vibrations of alkyl, aliphatic and aromatic rings (H – C – H) groups present in hemicellulose and lignin. These bands were ascribed to the aliphatic materials such as cutin, waxes and cutan[4949 Heredia-Guerrero, J. A., Benítez, J. J., Domínguez, E., Bayer, I. S., Cingolani, R., Athanassiou, A., & Heredia, A. (2014). Infrared and Raman spectroscopic features of plant cuticles: a review. Frontiers in Plant Science, 5, 305. http://dx.doi.org/10.3389/fpls.2014.00305. PMid:25009549.
http://dx.doi.org/10.3389/fpls.2014.0030...
]. The signal decreases when the sample is subjected to an acid-base treatment, due to the removal of these components[3838 Guzmán, O., Lemus, C., Martínez, S., Bonilla, J., Plasencia, A., & Ly, J. (2012). Chemical characteristics of silages of mango (Mangifera indica L.) by-products for animal feeding. Canadian Journal of Agricultural Science, 46(4), 369-374.]. 1740 cm-1 band associated with the stretching of C=O in ester groups disappeared after basic hydrolysis, confirming the partial removal of hemicellulose, which presented variation in hemicellulose content (e.g. 1.11 wt % compared to 29.75 wt % for NF sample).

3.4 Thermal characterization

As for the DSC results, both samples studied in nitrogen atmosphere show a similar trend and, all the reactions during the decomposition of the cellulose were endothermic process in inert atmosphere (Figure 4). The native endocarp fibre presented three steps of thermal degradation in relation to cellulose, lignin and hemicellulose. This is in line with a research done by Stevulova et al.[5050 Stevulova, N., Estokova, A., Cigasova, J., Schwarzova, I., Kacik, F., & Geffert, A. (2017). Thermal degradation of natural and treated hemp hurds under air and nitrogen atmosphere. Journal of Thermal Analysis and Calorimetry, 128(3), 1649-1660. http://dx.doi.org/10.1007/s10973-016-6044-z.
http://dx.doi.org/10.1007/s10973-016-604...
].

Figure 4
DCS curves of native (NF) and modified (MC) fibre.

NF shows an endotherm at 113.64 °C and a plateau at 145.3 °C; while that of MC is at 135.11 °C. This is mainly due to the elimination of moisture by the heat range applied in the DSC, as well as reducing sugars and other volatile components joined to the fibres, that are thermally decomposed. Espinosa-Andrews and Urias-Silvas[5151 Espinosa-Andrews, H., & Urias-Silvas, J. E. (2012). Thermal properties of agave fructans (Agave tequilana Weber var. Azul). Carbohydrate Polymers, 87(4), 2671-2676. http://dx.doi.org/10.1016/j.carbpol.2011.11.053.
http://dx.doi.org/10.1016/j.carbpol.2011...
] mention that there is a humidity loss, specifically in fructans, in temperatures under 130 °C. In addition to this, a small endotherm at 232.12 °C can be observed in NF probably due to the degradation of hemicellulose, reducing sugars and those hemicelluloses that are still joined to the cellulose-lignin complex[4848 Sgriccia, N., Hawley, M. C., & Misra, M. (2008). Characterization of natural fiber surfaces and natural fiber composites. Composites. Part A, Applied Science and Manufacturing, 39(10), 1632-1637. http://dx.doi.org/10.1016/j.compositesa.2008.07.007.
http://dx.doi.org/10.1016/j.compositesa....
]. NF and MC presented endotherm peaks at 325.52 °C and 335.25 °C, respectively. The endothermic peak observed in a 220-370 °C temperature range corresponds to that used for the simultaneous decomposition of hemicelluloses, cellulose and lignin. Based on the report of Miyahara et al.[5252 Miyahara, R. Y., Melquiades, F. L., Ligowski, E., Santos, A., Fávaro, S. L., & Antunes, O. R., Jr. (2018). Preparation and characterization of composites from plastic waste and sugar cane fiber. Polímeros: Ciência e Tecnologia, 28(2), 147-154. http://dx.doi.org/10.1590/0104-1428.12216.
http://dx.doi.org/10.1590/0104-1428.1221...
], this endothermic peak represents the degradation of hemicellulose. As reported by Xiao et al.[5353 Xiao, B., Sun, X., & Sun, R. (2001). Chemical, structural, and thermal characterizations of alkali-soluble lignins and hemicelluloses, and cellulose from maize stems, rye straw, and rice straw. Polymer Degradation & Stability, 74(2), 307-319. http://dx.doi.org/10.1016/S0141-3910(01)00163-X.
http://dx.doi.org/10.1016/S0141-3910(01)...
], cellulose is degraded within the range of 250 and 345 °C.

The enthalpies in the initial endotherm peak of the NF sample (121.92 J/g) are lower compared to those of the MC sample (183.27 J/g). This could be due to a greater presence of traces of waste and volatile components that were impregnated through acid-alkali hydrolysis. Therefore, a greater amount of heat energy is needed for degradation to occur. Conversely, with respect to final endotherm peak for both samples, the enthalpy of MC (172.92 J/g) is lower compared to that of the native fibre (178.32 J/g), which could be due to the elimination of larger sugar chains through acid-alkali hydrolysis. Besides, under nitrogen atmosphere, cellulose is more resistant to thermal treatment, probably due to its crystalline structure[5050 Stevulova, N., Estokova, A., Cigasova, J., Schwarzova, I., Kacik, F., & Geffert, A. (2017). Thermal degradation of natural and treated hemp hurds under air and nitrogen atmosphere. Journal of Thermal Analysis and Calorimetry, 128(3), 1649-1660. http://dx.doi.org/10.1007/s10973-016-6044-z.
http://dx.doi.org/10.1007/s10973-016-604...
]. The range at which lignin is decomposed is between 250 and 600 ° C. The analysis was performed up to 400 ° C, which indicates the degradation of 40% of the total lignin in the native sample. In the composition of native cellulose, it can be seen that the compound that degrades first is hemicellulose, followed by cellulose and lignin[5353 Xiao, B., Sun, X., & Sun, R. (2001). Chemical, structural, and thermal characterizations of alkali-soluble lignins and hemicelluloses, and cellulose from maize stems, rye straw, and rice straw. Polymer Degradation & Stability, 74(2), 307-319. http://dx.doi.org/10.1016/S0141-3910(01)00163-X.
http://dx.doi.org/10.1016/S0141-3910(01)...
].

3.5 Crystallinity

Applying the Equation 1 and with the use of the DSC enthalpy of fusion and taking 230J/g as the value for the fully crystalline Polycaprolactam (PA6)[5454 Wunderlich, B. (1990). Thermal analysis. New York: Academic Press. http://dx.doi.org/10.1016/B978-0-12-765605-2.50006-6.
http://dx.doi.org/10.1016/B978-0-12-7656...
], a lower value of crystallinity for native fibre (NF) was found compared to microcellulose (MC) (Table 1).

Table 1
Percentage of crystallinity of native fibre (NF) and microcellulose (MC).

As reported by Li et al.[5555 Li, Q., Zhou, J., & Zhang, L. (2009). Structure and properties of the nanocomposite films of chitosan reinforced with cellulose whiskers. Journal of Polymer Science. Part B, Polymer Physics, 47(11), 1069-1077. http://dx.doi.org/10.1002/polb.21711.
http://dx.doi.org/10.1002/polb.21711...
], the percentage of crystallinity in native fibre will always be lower, due to the presence of hemicellulose and lignin, amorphous compounds that encapsulate the fibrils. It is believed that the crystallinity index in micro cellulose increases due to the elimination of the amorphous regions of cellulose, in which hydrolysis drives the breakdown of glycosidic bonds, releasing individual crystals[5656 Spagnol, C., Rodrigues, F. H. A., Pereira, A. G. B., Fajardo, A. R., Rubira, A. F., & Muniz, E. C. (2012). Superabsorbent hydrogel composite made of cellulose nanofibrils and chitosan–graft–poly (acrylic acid). Carbohydrate Polymers, 87(3), 2038-2045. http://dx.doi.org/10.1016/j.carbpol.2011.10.017.
http://dx.doi.org/10.1016/j.carbpol.2011...
]. The crystallinity of mango micro cellulose reported in this study is similar to that reported by Bolio-López et al.[5757 Bolio-López, G. I., Valadez-González, A., Veleva, L., & Andreeva, A. (2011). Cellulose whiskers from agro–industrial banana wastes: isolation and characterization. Revista Mexicana de Ingeniería Química, 10(2), 291-299.] in banana peel micro cellulose (72% crystallinity) but greater than that obtained by Naranjo et al.[4747 Naranjo, D. C., Alamilla-Beltran, L., Gutierrez-Lopez, G. F., Terres-Rojas, E., Solorza-Feria, J., Romero-Vargas, S., Yee-Madeira, H., Flores-Morales, A., & Mora-Escobedo, R. (2016). Isolation and characterization of cellulose obtained from Agave salmiana fibers using two acid-alkali extraction methods. Revista Mexicana de Ciencias Agrícolas, 7(1), 31-43. http://dx.doi.org/10.29312/remexca.v7i1.368.
http://dx.doi.org/10.29312/remexca.v7i1....
] in agave fibre (50.07%). The greater the crystallinity of a fibre, the better the thermal stability and properties it can provide in a polyblend to obtain biodegradable materials[5858 Klemm, D., Heublein, B., Fink, H. P., & Bohn, A. (2005). Cellulose: fascinating biopolymer and sustainable raw material. Angewandte Chemie International Edition, 44(22), 3358-3393. http://dx.doi.org/10.1002/anie.200460587. PMid:15861454.
http://dx.doi.org/10.1002/anie.200460587...
].

4. Conclusions

Cellulose microfibers were isolated from waste mango fibrous endocarp (Mangifera caesia Jack ex Wall), after which they were subjected to an acid-alkaline treatment and an ultrasonic treatment that helped homogenize their sizes. The acid-alkaline treatment helped decrease the size of the fibres and was also involved in the removal of lignin and cellulose. Ultrasonication process is a very effective way to depolymerize cellulose and that led to reduce the formation of agglomeration and increase the dispersion of cellulose disintegrating into individual microfibers. We hope that with this simple chemical-ultrasonic treatment applied to cellulose will further stimulate interest in isolating microcellulose fibers from natural sources. The use of agricultural waste as a natural source of biopolymers has become an excellent option for innovation and the development of new products with potentially biodegradable properties by having a polymeric matrix formed by biopolymers and, their utilization can minimize negative environmental effects.

  • How to cite: Lorenzo-Santiago, M. A., & Rendón-Villalobos, R. (2020). Isolation and characterization of micro cellulose obtained from waste mango. Polímeros: Ciência e Tecnologia, 30(3), e2020036. https://doi.org/10.1590/0104-1428.09119

5. References

  • 1
    Derraik, J. G. (2002). The pollution of the marine environment by plastic debris: a review. Marine Pollution Bulletin, 44(9), 842-852. http://dx.doi.org/10.1016/S0025-326X(02)00220-5 PMid:12405208.
    » http://dx.doi.org/10.1016/S0025-326X(02)00220-5
  • 2
    O’Brine, T., & Thompson, R. C. (2010). Degradation of plastic carrier bags in the marine environment. Marine Pollution Bulletin, 60(1), 2279-2283. https://doi.org/10.1016/j.marpolbul.2010.08.005
    » https://doi.org/10.1016/j.marpolbul.2010.08.005
  • 3
    Webb, H. K., Arnott, J., Crawford, R. J., & Ivanova, E. P. (2012). Plastic degradation and its environmental implications with special reference to poly (ethylene terephthalate). Polymers, 5(1), 1-18. https://doi.org/10.3390/polym5010001
    » https://doi.org/10.3390/polym5010001
  • 4
    Das, O., Sarmah, A. K., & Bhattacharyya, D. (2015). A sustainable and resilient approach through biochar addition in wood polymer composites. The Science of the Total Environment, 512-513, 326-336. http://dx.doi.org/10.1016/j.scitotenv.2015.01.063 PMid:25634737.
    » http://dx.doi.org/10.1016/j.scitotenv.2015.01.063
  • 5
    Geyer, R., Jambeck, J. R., & Law, K. L. (2017). Production, use, and fate of all plastics ever made. Science advances, 3(7), e1700782. https://doi.org/10.1126/sciadv.1700782
    » https://doi.org/10.1126/sciadv.1700782
  • 6
    Ojeda, T. F., Dalmolin, E., Forte, M. M., Jacques, R. J., Bento, F. M. & Camargo, F. A. (2009). Abiotic and biotic degradation of oxo-biodegradable polyethylenes. Polymer degradation and stability, 94(6), 965-970. https://doi.org/10.1016/j.polymdegradstab.2009.03.011
    » https://doi.org/10.1016/j.polymdegradstab.2009.03.011
  • 7
    Wiles, D. M., & Scott, G. (2006). Polyolefins with controlled environmental degradability. Polymer degradation and stability, 91(7), 1581-1592. https://doi.org/10.1016/j.polymdegradstab.2005.09.010
    » https://doi.org/10.1016/j.polymdegradstab.2005.09.010
  • 8
    Kyrikou, I., & Briassoulis, D. (2007). Biodegradation of agricultural plastic films: a critical review. Journal of Polymers and the Environment, 15(2), 125-150. https://doi.org/10.1007/s10924-007-0053-8
    » https://doi.org/10.1007/s10924-007-0053-8
  • 9
    Ammala, A., Bateman, S., Dean, K., Petinakis, E., Sangwan, P., Wong, S., & Leong, K. H. (2011). An overview of degradable and biodegradable polyolefins. Progress in Polymer Science, 36(8), 1015-1049. https://doi.org/10.1016/j.progpolymsci.2010.12.002
    » https://doi.org/10.1016/j.progpolymsci.2010.12.002
  • 10
    Flieger, M., Kantorova, M., Prell, A., Řezanka, T., & Votruba, J. (2003). Biodegradable plastics from renewable sources. Folia Microbiologica, 48(1), 27-44. http://dx.doi.org/10.1007/BF02931273 PMid:12744074.
    » http://dx.doi.org/10.1007/BF02931273
  • 11
    Sagnelli, D., Hooshmand, K., Kemmer, G. C., Kirkensgaard, J. J. K., Mortensen, K., Giosafatto, C. V. L., Holse, M., Hebelstrup, K. H., Bao, J., Stelte, W., Bjerre, A. B., & Blennow, A. (2017). Cross-linked amylose bio-plastic: a transgenic-based compostable plastic alternative. International Journal of Molecular Sciences, 18(10), 2075-2086. http://dx.doi.org/10.3390/ijms18102075 PMid:28973963.
    » http://dx.doi.org/10.3390/ijms18102075
  • 12
    Bastioli, C. (2001). Global status of the production of biobased packaging materials. Starch, 53(8), 351-355. https://doi.org/10.1002/1521-379X(200108)53:8<351::AID-STAR351>3.0.CO;2-R
    » https://doi.org/10.1002/1521-379X(200108)53:8<351::AID-STAR351>3.0.CO;2-R
  • 13
    Burgos, N., Valdés, A., & Jiménez, A. (2016). Valorization of agricultural wastes for the production of protein-based biopolymers. Journal of Renewable Materials, 4(3), 165-177. http://dx.doi.org/10.7569/JRM.2016.634108
    » http://dx.doi.org/10.7569/JRM.2016.634108
  • 14
    Vestena, M., Gross, I. P., Muller, C. M. O., & Pires, A. T. N. (2016). Isolation of whiskers from natural sources and their dispersed in a non-aqueous medium. Polímeros: Ciência e Tecnologia, 26(4), 327-335. http://dx.doi.org/10.1590/0104-1428.2367
    » http://dx.doi.org/10.1590/0104-1428.2367
  • 15
    Serna-Cock, L., García-Gonzales, E., & Torres-León, C. (2016). Agro-industrial potential of the mango peel based on its nutritional and functional properties. Food Reviews International, 32(4), 364-376. http://dx.doi.org/10.1080/87559129.2015.1094815
    » http://dx.doi.org/10.1080/87559129.2015.1094815
  • 16
    Torres-León, C., Rojas, R., Serna-Cock, L., Belmares-Cerda, R., & Aguilar, C. N. (2017). Extraction of antioxidants from mango seed kernel: optimization assisted by microwave. Food and Bioproducts Processing, 105, 188-196. http://dx.doi.org/10.1016/j.fbp.2017.07.005
    » http://dx.doi.org/10.1016/j.fbp.2017.07.005
  • 17
    Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L. A., Rosa, M. F., de sá M. Sousa, M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. Macromolecular Symposia, 344(1), 39-54. http://dx.doi.org/10.1002/masy.201300217
    » http://dx.doi.org/10.1002/masy.201300217
  • 18
    Giraldo, L. M., Correa, H. M., Gutiérrez, J. B., & Castano, C. (2007). Aprovechamiento del residuo agroindustrial del mango común (Mangifera indica L.) en la obtención de azúcares fermentables. Ingeniería y Ciencia-ing, 3(6), 41-62.
  • 19
    Lorenzo‐Santiago, M. A., Juárez-López, A. L., Rosas‐Acevedo, J. L., Rendón‐Villalobos, J. R., Toribio‐Jiménez, J., & García Hernández, E. (2018). Management and final disposal of mango waste in the State of Guerrero, Mexico: a brief review. The Journal of Agricultural Science, 10(12), 34-41. http://dx.doi.org/10.5539/jas.v10n12p34
    » http://dx.doi.org/10.5539/jas.v10n12p34
  • 20
    Béguin, P., & Aubert, J.-P. (1994). The biological degradation of cellulose. FEMS Microbiology Reviews, 13(1), 25-58. http://dx.doi.org/10.1111/j.1574-6976.1994.tb00033.x PMid:8117466.
    » http://dx.doi.org/10.1111/j.1574-6976.1994.tb00033.x
  • 21
    Azizi Samir, M. A. S., Alloin, F., & Dufresne, A. (2005). Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field. Biomacromolecules, 6(2), 612-26. http://dx.doi.org/10.1021/bm0493685 PMid:15762621.
    » http://dx.doi.org/10.1021/bm0493685
  • 22
    Wegner, T. H., & Jones, P. E. (2006). Advancing cellulose-based nanotechnology. Cellulose (London, England), 13(2), 115-118. http://dx.doi.org/10.1007/s10570-006-9056-1
    » http://dx.doi.org/10.1007/s10570-006-9056-1
  • 23
    Beck-Candanedo, S., Roman, M., & Gray, D. G. (2005). Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules, 6(2), 1048-1054. http://dx.doi.org/10.1021/bm049300p PMid:15762677.
    » http://dx.doi.org/10.1021/bm049300p
  • 24
    Cerqueira, J. C., Penha, J. D. S., Oliveira, R. S., Guarieiro, L. L. N., Melo, P. D. S., Viana, J. D., & Machado, B. A. S. (2017). Production of biodegradable starch nanocomposites using cellulose nanocrystals extracted from coconut fibers. Polímeros: Ciência e Tecnologia, 27(4), 320-329. http://dx.doi.org/10.1590/0104-1428.05316
    » http://dx.doi.org/10.1590/0104-1428.05316
  • 25
    Khenblouche, A., Bechki, D., Gouamid, M., Charradi, K., Segni, L., Hadjadj, M., & Boughali, S. (2019). Extraction and characterization of cellulose microfibers from Retama raetam stems. Polímeros: Ciência e Tecnologia, 29(1), e2019011. http://dx.doi.org/10.1590/0104-1428.05218
    » http://dx.doi.org/10.1590/0104-1428.05218
  • 26
    Salgado-Delgado, R., Coria-Cortés, L., García-Hernández, E., Galarza, Z. V., Rubio-Rosas, E., & Crispín-Espino, I. (2010). Elaboración de materiales reforzados con carácter biodegradable a partir de polietileno de baja densidad y bagazo de caña modificado. Revista Iberoamericana de Polímeros, 11(7), 520-531. Retrieved in 2019, October 28, from https://dialnet.unirioja.es/servlet/articulo?codigo=3694228
    » https://dialnet.unirioja.es/servlet/articulo?codigo=3694228
  • 27
    Cordeiro, E. M. S., Nunes, Y. L., Mattos, A. L., & Rosa, M. F., Souza, M. S. M., Fo., & Ito, E. N. (2014). Polymer biocomposites and nanobiocomposites obtained from mango seeds. Macromolecular Symposia, 344(1), 39-54. https://doi.org/10.1002/masy.201300217
    » https://doi.org/10.1002/masy.201300217
  • 28
    Orts, W. J., Shey, J., Imam, S. H., Glenn, G. M., Guttman, M. E., & Revol, J. F. (2005). Application of cellulose microfibrils in polymer nanocomposites. Journal of Polymers and the Environment, 13(4), 301-306. http://dx.doi.org/10.1007/s10924-005-5514-3
    » http://dx.doi.org/10.1007/s10924-005-5514-3
  • 29
    Szymańska-Chargot, M., Cieśla, J., Chylińska, M., Gdula, K., Pieczywek, P. M., Koziol, A., Cieślak, K. J., & Zdunek, A. (2018). Effect of ultrasonication on physicochemical properties of apple based nanocellulose-calcium carbonate composites. Cellulose (London, England), 25(8), 4603-4621. http://dx.doi.org/10.1007/s10570-018-1900-6
    » http://dx.doi.org/10.1007/s10570-018-1900-6
  • 30
    Kasuga, T., Isobe, N., Yagyu, H., Koga, H., & Nogi, M. (2018). Clearly transparent nanopaper from highly concentrated cellulose nanofiber dispersion using dilution and sonication. Nanomaterials (Basel, Switzerland), 8(2), 104. http://dx.doi.org/10.3390/nano8020104 PMid:29439544.
    » http://dx.doi.org/10.3390/nano8020104
  • 31
    Technical Association of the Pulp and Paper Industry – TAPPI. (2007). Acid insoluble lignin in wood and pulp. T 222 om-02. Fibrous materials and pulp testing Atlanta: TAPPI.
  • 32
    Technical Association of the Pulp and Paper Industry – TAPPI. (2007). Solvent extractives of wood and pulp. T 204 cm-97. Fibrous materials and pulp testing Atlanta: TAPPI.
  • 33
    Haykiri-Acma, H., Yaman, S., Alkan, M., & Kucukbayrak, S. (2014). Mineralogical characterization of chemically isolated ingredients from biomass. Energy Conversion and Management, 77, 221-226. http://dx.doi.org/10.1016/j.enconman.2013.09.024
    » http://dx.doi.org/10.1016/j.enconman.2013.09.024
  • 34
    Waliszewska, B., Mleczek, M., Zborowska, M., Goliński, P., Rutkowski, P., & Szentner, K. (2019). Changes in the chemical composition and the structure of cellulose and lignin in elm wood exposed to various forms of arsenic. Cellulose (London, England), 26(10), 6303-6315. http://dx.doi.org/10.1007/s10570-019-02511-z
    » http://dx.doi.org/10.1007/s10570-019-02511-z
  • 35
    Rendón-Villalobos, R., García-Hernández, E., Güizado-Rodríguez, M., Salgado Delgado, R., & Rangel-Vázquez, N. A. (2010). Obtención y caracterización de almidón de plátano (Musa paradisiaca L.) acetilado a diferentes grados de sustitución. Afinidad, 67(548), 294-300. Retrieved in 2019, October 28, from https://www.raco.cat/index.php/afinidad /article/view/269205/356773
    » https://www.raco.cat/index.php/afinidad
  • 36
    Xu, Y. X., Miladinov, V., & Hanna, M. A. (2004). Synthesis and characterization of starch acetates with high substitution. Cereal Chemistry, 81(6), 735-740. http://dx.doi.org/10.1094/CCHEM.2004.81.6.735
    » http://dx.doi.org/10.1094/CCHEM.2004.81.6.735
  • 37
    Poley, L. H., Siqueira, A. P. L., Silva, M. G., Vargas, H., & Sanchez, R. (2004). Photothermal characterization of low density polyethylene food packages. Polímeros: Ciência e Tecnologia, 14(1), 8-12. http://dx.doi.org/10.1590/S0104-14282004000100007
    » http://dx.doi.org/10.1590/S0104-14282004000100007
  • 38
    Guzmán, O., Lemus, C., Martínez, S., Bonilla, J., Plasencia, A., & Ly, J. (2012). Chemical characteristics of silages of mango (Mangifera indica L.) by-products for animal feeding. Canadian Journal of Agricultural Science, 46(4), 369-374.
  • 39
    de Carvalho Couto, C. C., Fo., da Silva, J. C., Fo., Prata Neiva, A., Jr.,Magalhães de Souza, R., Rodrigues Nunes, J. A., & Viana Coelho, J. (2010). Fibrous fractions of mango residue silage with additives. Ciência e Agrotecnologia, 34(3), 751-757. http://dx.doi.org/10.1590/S1413-70542010000300031
    » http://dx.doi.org/10.1590/S1413-70542010000300031
  • 40
    Serna Cock, L., & Torres León, C. (2014). Potencial agroindustrial de cáscaras de mango de las variedades Keitt, y Tommy Atkins (Mangifera indica). Acta Agronomica, 64(2), 110-115. http://dx.doi.org/10.15446/acag.v64n2.43579
    » http://dx.doi.org/10.15446/acag.v64n2.43579
  • 41
    Balza, M., Garrido, E., García, M., Martínez, J., & García, A. (2017). Chemical characterization of the cellular wall of mango bocado pulp. Revista Agrollanía, 14, 7-13.
  • 42
    Owi, W. T., Lin, O. H., Sam, S. T., Chia, C. H., Zakaria, S., Mohaiyiddin, M. S., Villagracia, A. R., Santos, G. N., & Md Akil, H. (2016). Comparative study of microcelluloses isolated from two different biomasses with commercial cellulose. BioResources, 11(2), 3453-3465. http://dx.doi.org/10.15376/biores.11.2.3453-3465
    » http://dx.doi.org/10.15376/biores.11.2.3453-3465
  • 43
    Nagalakshmaiah, M., El Kissi, N., Mortha, G., & Dufresne, A. (2016). Structural investigation of cellulose nanocrystals extracted from chili leftover and their reinforcement in cariflex-IR rubber latex. Carbohydrate Polymers, 136, 945-954. http://dx.doi.org/10.1016/j.carbpol.2015.09.096 PMid:26572433.
    » http://dx.doi.org/10.1016/j.carbpol.2015.09.096
  • 44
    Atiqah, M. S. N., Gopakumar, D. A., F A T, O., Pottathara, Y. B., Rizal, S., Aprilia, N. A. S., Hermawan, D., Paridah, M. T. T., Thomas, S., & H P S, A. K. (2019). Extraction of cellulose nanofibers via eco-friendly supercritical carbon dioxide treatment followed by mild acid hydrolysis and the fabrication of cellulose nanopapers. Polymers, 11(11), 1813. PMid:31694184.
  • 45
    Sun, J. X., Xu, F., Sun, X. F., Xiao, B., & Sun, R. C. (2005). Physico-chemical and thermal characterization of cellulose from barley straw. Polymer Degradation & Stability, 88(3), 521-531. http://dx.doi.org/10.1016/j.polymdegradstab.2004.12.013
    » http://dx.doi.org/10.1016/j.polymdegradstab.2004.12.013
  • 46
    Yang, H., Yan, R., Chen, H., Lee, D. H., & Zheng, C. (2007). Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel, 86(12-13), 1781-1788. http://dx.doi.org/10.1016/j.fuel.2006.12.013
    » http://dx.doi.org/10.1016/j.fuel.2006.12.013
  • 47
    Naranjo, D. C., Alamilla-Beltran, L., Gutierrez-Lopez, G. F., Terres-Rojas, E., Solorza-Feria, J., Romero-Vargas, S., Yee-Madeira, H., Flores-Morales, A., & Mora-Escobedo, R. (2016). Isolation and characterization of cellulose obtained from Agave salmiana fibers using two acid-alkali extraction methods. Revista Mexicana de Ciencias Agrícolas, 7(1), 31-43. http://dx.doi.org/10.29312/remexca.v7i1.368
    » http://dx.doi.org/10.29312/remexca.v7i1.368
  • 48
    Sgriccia, N., Hawley, M. C., & Misra, M. (2008). Characterization of natural fiber surfaces and natural fiber composites. Composites. Part A, Applied Science and Manufacturing, 39(10), 1632-1637. http://dx.doi.org/10.1016/j.compositesa.2008.07.007
    » http://dx.doi.org/10.1016/j.compositesa.2008.07.007
  • 49
    Heredia-Guerrero, J. A., Benítez, J. J., Domínguez, E., Bayer, I. S., Cingolani, R., Athanassiou, A., & Heredia, A. (2014). Infrared and Raman spectroscopic features of plant cuticles: a review. Frontiers in Plant Science, 5, 305. http://dx.doi.org/10.3389/fpls.2014.00305 PMid:25009549.
    » http://dx.doi.org/10.3389/fpls.2014.00305
  • 50
    Stevulova, N., Estokova, A., Cigasova, J., Schwarzova, I., Kacik, F., & Geffert, A. (2017). Thermal degradation of natural and treated hemp hurds under air and nitrogen atmosphere. Journal of Thermal Analysis and Calorimetry, 128(3), 1649-1660. http://dx.doi.org/10.1007/s10973-016-6044-z
    » http://dx.doi.org/10.1007/s10973-016-6044-z
  • 51
    Espinosa-Andrews, H., & Urias-Silvas, J. E. (2012). Thermal properties of agave fructans (Agave tequilana Weber var. Azul). Carbohydrate Polymers, 87(4), 2671-2676. http://dx.doi.org/10.1016/j.carbpol.2011.11.053
    » http://dx.doi.org/10.1016/j.carbpol.2011.11.053
  • 52
    Miyahara, R. Y., Melquiades, F. L., Ligowski, E., Santos, A., Fávaro, S. L., & Antunes, O. R., Jr. (2018). Preparation and characterization of composites from plastic waste and sugar cane fiber. Polímeros: Ciência e Tecnologia, 28(2), 147-154. http://dx.doi.org/10.1590/0104-1428.12216
    » http://dx.doi.org/10.1590/0104-1428.12216
  • 53
    Xiao, B., Sun, X., & Sun, R. (2001). Chemical, structural, and thermal characterizations of alkali-soluble lignins and hemicelluloses, and cellulose from maize stems, rye straw, and rice straw. Polymer Degradation & Stability, 74(2), 307-319. http://dx.doi.org/10.1016/S0141-3910(01)00163-X
    » http://dx.doi.org/10.1016/S0141-3910(01)00163-X
  • 54
    Wunderlich, B. (1990). Thermal analysis New York: Academic Press. http://dx.doi.org/10.1016/B978-0-12-765605-2.50006-6
    » http://dx.doi.org/10.1016/B978-0-12-765605-2.50006-6
  • 55
    Li, Q., Zhou, J., & Zhang, L. (2009). Structure and properties of the nanocomposite films of chitosan reinforced with cellulose whiskers. Journal of Polymer Science. Part B, Polymer Physics, 47(11), 1069-1077. http://dx.doi.org/10.1002/polb.21711
    » http://dx.doi.org/10.1002/polb.21711
  • 56
    Spagnol, C., Rodrigues, F. H. A., Pereira, A. G. B., Fajardo, A. R., Rubira, A. F., & Muniz, E. C. (2012). Superabsorbent hydrogel composite made of cellulose nanofibrils and chitosan–graft–poly (acrylic acid). Carbohydrate Polymers, 87(3), 2038-2045. http://dx.doi.org/10.1016/j.carbpol.2011.10.017
    » http://dx.doi.org/10.1016/j.carbpol.2011.10.017
  • 57
    Bolio-López, G. I., Valadez-González, A., Veleva, L., & Andreeva, A. (2011). Cellulose whiskers from agro–industrial banana wastes: isolation and characterization. Revista Mexicana de Ingeniería Química, 10(2), 291-299.
  • 58
    Klemm, D., Heublein, B., Fink, H. P., & Bohn, A. (2005). Cellulose: fascinating biopolymer and sustainable raw material. Angewandte Chemie International Edition, 44(22), 3358-3393. http://dx.doi.org/10.1002/anie.200460587 PMid:15861454.
    » http://dx.doi.org/10.1002/anie.200460587

Publication Dates

  • Publication in this collection
    09 Dec 2020
  • Date of issue
    2020

History

  • Received
    28 Oct 2019
  • Reviewed
    28 Sept 2020
  • Accepted
    17 Nov 2020
Associação Brasileira de Polímeros Rua São Paulo, 994, Caixa postal 490, São Carlos-SP, Tel./Fax: +55 16 3374-3949 - São Carlos - SP - Brazil
E-mail: revista@abpol.org.br