Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-06-24T09:31:47.323Z Has data issue: false hasContentIssue false

Thermogravimetric analysis of agricultural residue carbonized briquettes for domestic and industrial applications

Published online by Cambridge University Press:  27 December 2019

Vianney Andrew Yiga*
Affiliation:
Department of Mechanical Engineering, Makerere University Kampala, Uganda
Michael Lubwama
Affiliation:
Department of Mechanical Engineering, Makerere University Kampala, Uganda Africa Centre of Excellence in Materials, Product Development and Nanotechnology, MAPRONANO, Makerere University Kampala, Uganda
Get access

Abstract

For a material to be used as energy, understanding its combustion behavior is crucial. Biomass is one such material that is increasingly gaining traction. Biomass may be utilized by direct combustion or transformation into fluid or solid biomass-based fuels. In this work, slow pyrolysis of groundnut shells, bagasse, rice husks and coffee husks was done to produce briquettes with cassava starch binder. Thermogravimetric analysis (TGA) was carried out using an Eltra Thermostep thermogravimetric analyzer. The samples were heated from ambient to 920 deg. Celsius. This analysis provided combustion explanations in terms of the weight loss, burning rates, peak temperatures, char residues and mean reactivity. TGA results showed that binder inclusion reduced the amount of fixed carbon present in the developed briquettes, thus slightly reducing their calorific values. Rice husks briquettes yielded the least weight loss (20.9% and 24.7% for 30g and 50g binder incorporations respectively) compared to others, owing to former’s higher ash contents. Increase in binder contents reduced the amount of char residues, caused by reducing ash contents in the developed briquettes. Peak temperatures and char residues generally increased with increasing binder content. This signifies increasing thermal stabilities of the developed briquettes. Highest char residues were obtained by briquettes developed with rice husks at 30g binder while briquettes developed with bagasse briquettes developed at 50g binder had the least char residues. The highest mean reactivities were obtained in briquettes developed from bagasse and coffee husks while briquettes developed from rice husks had lowest mean reactivities. Briquettes developed in this study showed sufficient combustion properties suitable to provide energy for domestic and industrial applications.

Type
Articles
Copyright
Copyright © Materials Research Society 2019

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Warner, K. J., & Jones, G. A. (2017). A population-induced renewable energy timeline in nine world regions. Energy Policy, 101, 65-76.CrossRefGoogle Scholar
Chen, L., Xing, L., & Han, L. (2009). Renewable energy from agro-residues in China: Solid biofuels and biomass briquetting technology. Renewable and Sustainable Energy Reviews, 13(9), 2689-2695.CrossRefGoogle Scholar
Yiga, V. A., Lubwama, M., & Olupot, P. W. (2019). Effect of Alkaline Surface Modification & Carbonization on Biochemical Properties of Rice and Coffee Husks for Use in Briquettes and Fiber-Reinforced Plastics. Journal of Natural Fibers, DOI: 10.1080/15440478.2019.1642824.CrossRefGoogle Scholar
Okello, C., Pindozzi, S., Faugno, S., & Boccia, L. (2013). Bioenergy potential of agricultural and forest residues in Uganda. Biomass and bioenergy, 56, 515-525.CrossRefGoogle Scholar
Giorio, C., Pizzini, S., Marchiori, E., Piazza, R., Grigolato, S., ZanettiM., ... M., ... & Tapparo, A. (2019). Sustainability of using vineyard pruning residues as an energy source: Combustion performances and environmental impact. Fuel, 243, 371-380.CrossRefGoogle Scholar
Lubwama, M., & Yiga, V. A. (2018). Characteristics of briquettes developed from rice and coffee husks for domestic cooking applications in Uganda. Renewable energy, 118, 43-55.CrossRefGoogle Scholar
Lubwama, M., & Yiga, V. A. (2017). Development of groundnut shells and bagasse briquettes as sustainable fuel sources for domestic cooking applications in Uganda. Renewable energy, 111, 532-542.CrossRefGoogle Scholar
Martinez, C. L. M., Sermyagina, E., Carneiro, A. D. C. O., Vakkilainen, E., & Cardoso, M. (2019). Production and characterization of coffee-pine wood residue briquettes as an alternative fuel for local firing systems in Brazil. Biomass and Bioenergy, 123, 70-77.CrossRefGoogle Scholar
Lubwama, M., Yiga, V.A., Muhairwe, F., Kihedu, J. (2019). Physical and combustion properties of agricultural residue bio-char bio-composite briquettes as sustainable domestic energy sources. Renewable energy . DOI: 10.1016/j.renene.2019.10.085.Google Scholar
Okot, D. K., Bilsborrow, P. E., & Phan, A. N. (2019). Briquetting characteristics of bean straw-maize cob blend. Biomass and Bioenergy, 126, 150-158.CrossRefGoogle Scholar
Onifade, T. B., Orisaleye, J. I., Pecenka, R., & Jekayinfa, S. O. (2019). Effect of densification variables on water resistance of corn cob briquettes.Google Scholar
Safdar, H. M., Nasir, A., & Ahmad, R. (2020). Enhancing the quality of maize, wheat, rice and cotton residue briquettes by optimizing the operational parameters.Google Scholar
Soares, L. D. S., Maia, A. A., Moris, V. A., & De Paiva, J. M. (2019). Study of the Effects of the Addition of Coffee Grounds and Sugarcane Fibers on Thermal and Mechanical Properties of Briquettes. Journal of Natural Fibers, 1-9.CrossRefGoogle Scholar
Kpelou, P., Kongnine, D. M., Kombate, S., Mouzou, E., & Napo, K. (2019). Energy Efficiency of Briquettes Derived from Three Agricultural Waste’s Charcoal Using Two Organic Binders. Journal of Sustainable Bioenergy Systems, 9(02), 79.CrossRefGoogle Scholar
Trubetskaya, A., Leahy, J. J., Yazhenskikh, E., Müller, M., Layden, P., JohnsonR., ... R., ... & Monaghan, R. F. (2019). Characterization of woodstove briquettes from torrefied biomass and coal. Energy, 171, 853-865.CrossRefGoogle Scholar
Zhang, L., Xu, C. C., & Champagne, P. (2010). Overview of recent advances in thermo-chemical conversion of biomass. Energy Conversion and Management, 51(5), 969-982.CrossRefGoogle Scholar
Wang, T., Li, Y., Zhi, D., Lin, Y., He, K., Liu, B., & Mao, H. (2019). Assessment of combustion and emission behavior of corn straw biochar briquette fuels under different temperatures. Journal of environmental management, 250, 109399.CrossRefGoogle ScholarPubMed
Song, A., Zha, F., Tang, X., & Chang, Y. (2019). Effect of the additives on combustion characteristics and desulfurization performance of cow dung briquette. Chemical Engineering and Processing-Process Intensification, 143, 107585.CrossRefGoogle Scholar
da Silva, J. E., de Araújo Melo, D. M., de Freitas Melo, M. A., de Aguiar, E. M., Pimenta, A. S., de MedeirosE. P., ... E. P., ... & Braga, R. M. (2019). Energetic characterization and evaluation of briquettes produced from naturally colored cotton waste. Environmental Science and Pollution Research, 26(14), 14259-14265.CrossRefGoogle ScholarPubMed
Trubetskaya, A., Jensen, P. A., Jensen, A. D., Steibel, M., Spliethoff, H., Glarborg, P., & Larsen, F. H. (2016). Comparison of high temperature chars of wheat straw and rice husk with respect to chemistry, morphology and reactivity. Biomass and Bioenergy, 86, 76-87.CrossRefGoogle Scholar
Song, X., Zhang, S., Wu, Y., & Cao, Z. (2019). Investigation on the properties of the bio-briquette fuel prepared from hydrothermal pretreated cotton stalk and wood sawdust. Renewable Energy.Google Scholar
Cong, H., Zhao, L., Mašek, O., Yao, Z., Meng, H., Huo, L., ... & Wu, Y. (2019). Evaluating the performance of honeycomb briquettes produced from semi-coke and corn stover char: Co-combustion, emission characteristics, and a value-chain model for rural China. Journal of Cleaner Production, 118770.Google Scholar
Yiga, V. A., Pagel, S., Lubwama, M., Epple, S., Olupot, P. W., & Bonten, C. (2019). Development of fiber-reinforced polypropylene with NaOH pretreated rice and coffee husks as fillers: mechanical and thermal properties. Journal of Thermoplastic Composite Materials, DOI: 10.1177/0892705718823255.CrossRefGoogle Scholar
Shen, J., Zhu, S., Liu, X., Zhang, H., & Tan, J. (2010). The prediction of elemental composition of biomass based on proximate analysis. Energy Conversion and Management, 51(5), 983-987.CrossRefGoogle Scholar
Amarasekara, A., Tanzim, F.S., Asmatulu, E. (2017). Briquetting and carbonization of naturally grown algae biomass for low-cost fuel and activated carbon production, Fuel, 208, 612-617.CrossRefGoogle Scholar
Yiga, V. A. and Lubwama, M. (2019). Thermal stability of compression molded rice and coffee husk fiber-reinforced polypropylene composites. 14th HEFAT conference, Dublin, Ireland.Google Scholar
Munir, S., Daood, S.S., Nimmo, W., Cunliffe, A.M., Gibbs, B.M. (2009). Thermal analysis and devolatilization kinetics of cotton stalk, sugar cane bagasse and shea meal under nitrogen and air atmospheres. Bioresource Technology, 100(3), 1413-1418.CrossRefGoogle ScholarPubMed
Setter, C., Silva, F. T. M., Assis, M. R., Ataíde, C. H., Trugilho, P. F., & Oliveira, T. J. P. (2020). Slow pyrolysis of coffee husk briquettes: Characterization of the solid and liquid fractions. Fuel, 261, 116420.CrossRefGoogle Scholar
Liu, Z., Quek, A., Hoekman, S. K., & Balasubramanian, R. (2013). Production of solid biochar fuel from waste biomass by hydrothermal carbonization. Fuel, 103, 943-949.CrossRefGoogle Scholar