Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 3207 word(s) 3207 2021-04-25 11:22:34 |
2 format change Meta information modification 3207 2021-04-27 03:47:25 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Sher, F. Biomass and Waste Produced. Encyclopedia. Available online: https://encyclopedia.pub/entry/9016 (accessed on 26 April 2024).
Sher F. Biomass and Waste Produced. Encyclopedia. Available at: https://encyclopedia.pub/entry/9016. Accessed April 26, 2024.
Sher, Farooq. "Biomass and Waste Produced" Encyclopedia, https://encyclopedia.pub/entry/9016 (accessed April 26, 2024).
Sher, F. (2021, April 26). Biomass and Waste Produced. In Encyclopedia. https://encyclopedia.pub/entry/9016
Sher, Farooq. "Biomass and Waste Produced." Encyclopedia. Web. 26 April, 2021.
Biomass and Waste Produced
Edit

In the late twentieth century, the only cost-effective opportunity for waste removal cost at least several thousand dollars, but nowadays, a lot of improvement has occurred. The biomass and waste generation problems attracted concerned authorities to identify and provide environmentally friendly sustainable solutions that possess environmental and economic benefits. The present study emphasises the valorisation of biomass and waste produced by domestic and industrial sectors.

waste to value bio-residues

1. Introduction

In the 21st century, it has become challenging to provide clean, affordable, and reliable energy sources, which are very important from the socio-economic and environmental perspectives. To manage these crucial problems, biomass is the most favourable renewable source at the moment [1]. Biomass has drawn attention in the latest years as the only continuous carbon source available on earth. Therefore, it is regarded as a perfect substitute for fossil fuels. The use of biomass is a modern approach for power generation and is considered an excellent neutral resource where CO2 emissions are reduced. Biomass is an attractive feedstock due to several reasons. Firstly, it reduces greenhouse gas (GHG) emissions and is a new source, which will be established in the future. Lastly, it has the benefit of providing economic benefits to society. However, biomass possesses an adverse effect when utilised in conventional stoves. It emits polycyclic aromatic hydrocarbons (PAHs). PAHs are composed of more than 100 chemicals that include furans (which are toxic to health either inhaled or ingested), volatile organic compounds, and heavy metals [2][3]. Biomass consists of diverse organic materials which contain chemical energy. Green plants generate this chemical energy by photosynthesis, where sunlight serves as an energy source [4]. Figure 1 illustrates the derived carriers of the resultant energy. These carriers utilise biomass and fix CO2 (i.e., photosynthesis) with a concurrent transformation of solar energy into chemical energy.

Sustainability 13 04200 g001 550Figure 1. Model of carbon cycle illustrating how energy carriers are derived from biomass. Biomass carbon is generated via photosynthesis upon fixing atmospheric CO2 with the simultaneous conversion of solar energy into chemical energy stored in biomass. Biomass carbon could be transformed into several energy carriers through either an environmentally amicable route (shown in green) or an environmentally unfriendly route (shown in red). If biomass carbon, harvested crops, or wastes are converted into fuel, the process is renewable with no atmospheric CO2 build-up. Conversely, biomass decomposed over several epochs (geologic carbon) can also be partially recovered and utilised. However, the later process is slow, non-sustainable, and potentially harmful to the natural environment.

Biomass, if not used for alternate energy sources, may lead to severe hazards to the environment in the form of discarded materials like waste, which has become a striking and crucial problem of civilised human societies [5][6]. In the US, people throw away the waste/garbage equivalent to their body weight on a monthly basis where a significant part of discarded material is composted or recycled and/or incinerated or landfilled. Waste production has increased ten-fold because of the affluent and urban population of the world, which will double by 2025. Dealing with waste is one of the most critical issues in developing countries, resulting in increased municipal budgets [7]. Landfills like Laogang (China), Bordo Poniente (Mexico), and Jardim Gramacho (Brazil) are competing for the name of the biggest dump in the world. The urbanisation of the world’s population increases solid-waste generation around the globe. It is hard to predict the expanding socio-economic trends leading towards the 22nd century. We are putting our efforts to avoid the maximum rate of waste production in this century. Our efforts will be in vain until we decrease people’s growth and the rate of consumption material, otherwise, the world will have to bear a growing problem of waste [8].

2. Transformation of Biomass and Waste into Bioproducts

The transformation of biomass into biofuels and chemicals increases worldwide moderate energy sources and reduces global warming [9]. Biomass is a neutral resource of renewable energy and generally burns cleaner than fossil fuels [10][11][12][13]. Different biomass forms are converted into fuels and chemicals, like wood and timber waste, plants in agriculture, industrial waste, sludge, and waste from food processing. In contrast, wood logs can be used directly. Wood waste is easily recycled and used as fertilizer as sawmill residues, slashed from logging and municipal wood waste by most plants and crops [12].

It is possible to use various procedures to transform biomass into energy and other industrially relevant high-interest products (Figure 2) [13][14][15]. Biomass can be converted into fuel, biogas, bioethanol, biodiesel, or changed into syngas [16][17]. Bio-refinery, where several significant processes are integrated and joined with downstream upgrades and cultured separations, will be very attractive to get the most out of feed use and product value [18]. The production of biofuels and chemicals by transforming biomass is becoming popular worldwide to expand energy resources and mitigate global warming. The following characteristics, such as renewable natural sources and carbon-neutral sources, underpin the significance of the biomass. Moreover, resultant fuels acquired from biomass are an important source of sustainable energy because they burn cleanly compared to fossil fuels. Evaluations of the estimated contribution to global energy (i.e., biomass) range from a minimum of 100 EJ/year to a maximum of 400 EJ/year [9]. For example, the total consumption of energy in the USA is around 100 EJ/year and, in 2005, the global commercial consumption of energy was 440 EJ/year. Primary sources of this energy were hydroelectricity, natural gas, oil, coal, and nuclear energy [19]. This means that biomass could contribute between 20% and 90% of world energy demand.

Sustainability 13 04200 g002 550

Figure 2. The biomass conversion process and their resulted products of industrial interest.

3. Sources of Waste/Production of Biomass

3.1. Agricultural Aspects

In agricultural production systems, large quantities of harvested crop residues are generated. The cultivation of wheat, rice, vegetables, fruits, and other crops creates substantial amounts of residues essential for the yearly production of biomass. Therefore, biomass produced from agriculture significantly enhances the generation of sustainable energy used in the industry and the domestic processes [15][20][21][22]. Unused biomass from agricultural sources can be split into two categories: (1) parts of the plant kept in the field, (2) loss of plant components owing to harvesting techniques. However, all types of residues could not be used for biomass in the past. Nowadays, such biomass residues can be used for many purposes like biofuels (biogas/bioethanol) or energy production and forage, which contribute significantly to a country’s economic growth [23][24][25]. In horticulture, fruit plants, particularly drupes, have endocarp tissues. A drupe fruit’s endocarp is the hard and uneatable portion of the fruit, whereas the eatable part is mesocarp. The woody biomass of drupe endocarp is the leading lignin source that usually generates 50 percent of the total [26][27][28]. Lignin has a greater energy density than cellulose as a biofuel. Lignin plays an essential role in energy production by using endocarp of fruits in horticulture [29][30][31]. Lignin degradation and the function of the pre-treatment are graphically represented in Figure 3.

Sustainability 13 04200 g003 550

Figure 3. Schematic of the role of biological pre-treatment in biomass lignin degradation. The initial step of lignin biodegradation is when the oxidative enzymes induce new functional groups into lignin’s macromolecular structure, making lignin vulnerable to other enzymes’ consequent degradation. The natural form of lignin has numerous different functional groups that could be specifically functionalized.

3.2. Industrial Aspects

Microorganisms can be used for biomass production in industry. Particularly yeast which serves as biocatalysts and used in bakeries, lactic acid (used as starter culture in manufacturing of dairy products), breweries, probiotics, aqua and animal feed production [32][33]. On a large scale, fermentation can be used with economical substrates and products for biomass production. These economical substrates include soya bean meal, sugar cane molasses, and various industries’ wastes [34][35][36]. Solid biomass is preferred to use in the industry. Liquid waste and biogas are also used in industrial applications, but particularly solid biomass is important for the past 7 years. In other countries, 80 percent of the global biomass has significant application in industry. Most of the industry’s biomass is obtained from forests in the United States and mostly black liquor [1]. In Sweden, 41 percent of the industries consumed biomass in 2010. The Finnish timber processing sector accounted for 54 percent of total industrial power consumption in 2010. In particular, wood-based biomass is used by the wood processing sector, which accounted for 45 percent of total industrial power use in 2010.

3.3. Domestic Food Waste and Considerations

Economic growth is facing pivotal problems worldwide due to increased waste products that cause energy loss, damage to the environment and human health, and adversely affect quality of life [37][38]. The European Union’s goal to discover an effective source for the “Recycling Society” can be a better resource to prevent waste generation. Biological waste is the most significant portion of municipal solid waste (MSW), primarily food waste [39][40]. As a result, the rise in the population increased the utilisation of food, and the production of waste (i.e., food waste) is also expected to increase substantially. Food waste is primarily controlled at the EU level by the Waste Framework Directive [41]. Therefore, it is evident that efficient and alternative methods of handling biodegradable waste generated in homes need to be found [36][42][43]. Domestic composting and anaerobic digestion are both familiar, distributed, and compact processes. Still, they fail many times and cause more problems than fixing them, such as odours and greenhouse gas (GHG) emissions [44][45]. Approximately 23% of the 4.8 million tons of municipal solid waste produced yearly are reused (mostly waste from packaging), however, the remaining 77% are disposed of without proper treatment. In the shortage of new innovative projects, the proportion of produced waste is expected to increase continuously with severe environmental issues in the foreseeable future.

4. Characteristics and Composition of Biomass

There is a broad range of produced biomass, thereby there are significant variations in the composition of industrial or domestically generated biomass. Cellulose, hemicelluloses, lignin, starch, and proteins are some of the main elements [46][47][48][49]. Cellulose: A polysaccharide where β-glucoside bonds evenly connect D-glucose. Organic compound cellulose is an important component in the primary cell wall of the plants. It supports the structural assembly of the cell wall [50][51]. Cellulose assembles into unbranched and long micro fibrils that provide supports. Several bacterial species secrete cellulose to facilitate biofilm formation [52]. Its molecular formula is (C6H10O5)n. The polymerisation degree stated by n is broad, ranging from thousands to many tens of thousands. A schematic illustration of plant cellulose is given in Figure 4. Hemicellulose: Hemicellulose comprises several heteropolymers. It is a polysaccharide with 5-carbon monosaccharide units including D-xylose, D-arabinose, and 6-carbon monosaccharide units which have D-Galactose, D-glucose, and D-mannose. Hemicellulose possesses a lower molecular weight than cellulose and has a less specialised structure than cellulose [53][54][55]. Hemicellulose is known as a second main component of the biomass from the plant. The industrial use of hemicellulose has expanded with time and the integration of modern techniques into the existing methods. Lignin: A compound whose constituent units, phenyl-propane, and its derivatives are 3-dimensionally linked. Cellulose, hemicelluloses, and lignin are widely discovered in many types of biomass [55]. A network of irregular polymers that consists of cellulose and hemicellulose fibres is recognised as lignin. The said network provides structural support to the plants. The irregular formation and complex structural identity make its exploitation challenging in the industry sector. The complex structure is impervious to most of the chemical and hydrolysis treatment methods. Plant biomass chemical composition differs between species. Generally, crops are comprised of about 25 percent lignin and 75 percent carbohydrates. The percentage of carbohydrates is composed of many sugar molecules in the form of polymers. There are two distinctive categories: cellulose and hemicellulose. The moisture content is considered the most significant property of biomass feedstock. The moisture content negatively affects the fuel’s energy when used for incineration and other processes (e.g., thermo-chemical) [13][14]. Dry biomass has a higher heating value because it utilises little energy to evaporate any moisture. There is a direct association between the energy ratio and humidity, an increase in humidity entails less energy [14][56][57]. Biomass materials contain a certain amount of humidity, from 10% for a dried straw to over 50% for new cut timber [58].Sustainability 13 04200 g004 550Figure 4. Schematic illustration of plant cellulose.Carbon: Biomass carbon content is around 45%, while coal has 60% or higher. A greater carbon content results in a higher heating value [58]. Hydrogen: Biomass hydrogen content is about 6% [59]. A greater content of hydrogen results in a greater value of heating. Nitrogen: Biomass nitrogen content ranges from 0.2 percent to over 1 percent [59]. Fuel-bound nitrogen is responsible for most emissions from the biomass combustion of nitrogen oxide (NOx). Lower nitrogen content should result in reduced emissions of NOx in the fuel (e.g., diesel). Sulphur: Biomass fuels have a sulphur content of less than 0.2 percent, with a few exceptions as elevated as 0.5–0.7 percent. Coals range from 0.5 percent to 0.75 percent [58]. Chloride: Biomass combustion with elevated levels of chloride may result in enhanced fouling of ash. The boiler tubes’ high chloride content results in hydrochloric acid formation, leading to corrosion resulting in pipe failure and water leakage in the boiler.

5. Green Technologies for Biomass and Waste Valorisation

Waste valorisation is a process that converts waste materials into valuable products like chemicals, fuels, and materials. The waste valorisation concept relies on the thought; the waste products contain unused polymeric products that can convert into energy and different forms of chemicals [60][61]. These types of products make the waste residues a valuable source that could not be left unharnessed. This idea can be applied to artificial waste or bio-waste, and it will become the basis of the waste-to-energy approach. Waste valorisation is not magnificent because of the deletion of natural resources, but it is much-needed technology for waste management and renewable energy production and also produces high values of nano-bioplastic products and ethanol, which are cost-effective and sustainable [62]. Figure 5 illustrates a graphical representation of green technologies applied for biomass and waste valorisation.Sustainability 13 04200 g005 550Figure 5. Green technologies applied for biomass and waste valorisation.In this context, waste-to-value (WtV) and, more specifically, waste-to-energy (WtE) have noteworthy potential that should be considered in future considerations prior to the development and implementation of tools like Life Cycle Assessment (LCA), Ecological Indicators (EI), and/or Ecological Footprint Analysis (EFA) [63][64]. Using WtV and/or WtE thematic concepts with a zero-waste approach, fine chemicals and/or valuable energy are produced from waste residues using processes, such as green processing technologies, using organic and inorganic chemicals, genetically engineered organisms, and others. Other technological WtE solutions, such as anaerobic digestion, incineration, gasification, pyrolysis, landfill waste, and agro-industrial waste biomass residues to bioenergy, are considered valuable drivers to optimise the waste supply chain management to strengthens the modern WtE facilities [64]. Some of the technological and in practice strategies are taken as model examples and thus are discussed in the following subsections.

5.1. Fermentation Technology

On a commercial scale, fermentation is used in different countries to make ethanol. The synthesis of ethanol is performed using sugar crops (i.e., sugar beet) and starch crops (i.e., wheat). Starch combines with enzymes to convert sugars, sugars are transformed with yeast into ethanol [65][66]. Distillation of the ethanol is an energy-intensive phase and produces approximately 450 mL of ethanol per ton of dry maize. The strong residues used as a bovine feed form the fermentation method. For boilers or gasification, it can be used as a bagasse fuel. The more complex biomass is converting lingo-cellulosic biomass because long-chain polysaccharide molecules are present and require acid before the sugars are fermented to ethanol [14][67].

5.2. Flow Technology

When considering fresh procedures, the elevated amount of biomass’s physical and chemical complexity is a major problem. One prospective key to overcoming the difficulty of biomass includes converting it into simpler fractions that are easier to handle in downstream procedures. In a single petroleum industry facility, bio-refinery is the source of chemicals, energy, and fuel production [20][68][69]. These significant biomass derivatives (so-called platform molecules or construction blocks) are comparatively easy compounds with various functionalities in their structures that are appropriate for a range of useful chemical compound transformations [70]. Continuous flow processing enables the response conditions to be better controlled. This is beneficial when dealing with extremely reactive feedstock such as biomass-derived feedstock. Flow processing also promotes scaling up, considering that many biomass procedures are still on the laboratory scale, which is a significant point. Therefore, the development of flow technologies in the near future will contribute to the marketing of biomass technologies [71][72]. Since the chemical structure of biomass feedstock is usually very distinct from that of the final products, various processing steps are typically needed in such transformations, negatively influencing the process economy. Using stream processing techniques enables chemical processes to be intensified, thereby contributing considerably to technology simplification [73].

5.3. Gasification

The gasification process is the partial oxidation of organic/natural products producing syngas at a constant temperature (i.e., 500–1800 °C). The gasification method appears as a char that responds to hydrogen and carbon monoxide with water vapour and CO2. In equilibrium reaction, the concentration of steam, CO2, hydrogen, and carbon monoxide become stable very fast at the given temperature in a gasifier [74][75], to produce heat or electricity, which is used as fuel through syngas. The gasification agents, which are the combination of carbon dioxide, oxygen, and steam, are used in a gasifier. The gasification process can be used as a cleaner and is a more logical technology than combustion [76]. Before its commercial organisation, biomass gasification must control some barriers. The main application of gasification is to remove tars, problems related to the production and pre-treatment of biomass feedstock, and the effect of biomass properties must be clearly understood [77]. Supercritical gasification of water in wet biomass is an advanced technology nowadays and seeks the attention of all big countries like the USA, Germany, Netherlands, and Japan [78]. Super Critical Water Gasification (SCWG) has the advantage that this technique does not require any dry method for the wet biomass before subjecting to gasification [79].

5.4. Microbial Digestion

Microbial digestion of organic matter through metabolic pathways leads to CO2 and methane formation [80][81]. Biogas is called a combination of coal dioxide and methane [82][83][84]. Anaerobic digestion gives renewable energy production possibilities and a higher quality of agro-residue-waste treatment. Recently, this technology became an appealing technique for biodegrading strong municipal waste organic fractions in Europe [85]. The method takes place in well-designed ships called the anaerobic bioreactor/anaerobic digester. A biogas plant includes the entire feedstock, digester, biogas holder, and digestive tank system.

5.5. Microwave Technology

Biomass feedstock is traditionally heated with pressure (i.e., ~60–100 bar). The heating activity is carried out between 180 and 200 °C in the closed vessel where oxygen (O2) is not present for at least 12 h [86][87][88]. Through conduction and convection, heat transfer is accomplished through temperature gradients. Potential disadvantages of standard heating techniques include extended periods of residence and surface heating [89]. The microwave hydrothermal carbonization method was suggested to solve these disadvantages. Over the years, microwave technology has been used to substitute standard heating with carbohydrate [90][91] digestion, sterilisation [92], synthesis [93], and recently the dry pyrolysis method of the biomass feedstock to produce char and gas [56][94][95]. Besides the benefit of shorter periods of residence, microwave heating provides fast, selective, and volumetric heating, promoting fresh response paths and appropriate conditions for new products [96][97].In conclusion, the operational considerations are met by the accurate and well-regulated nature of microwave technology and its prospective portable processing capability (owing to the comparatively small size of a reactor) [98]. The suitability of microwave heating to process human bio-wastes arises from their comparatively elevated water molecule, which is easily coupled with electromagnetic fields that cause “microwave dielectric heating” [99][100][101].

References

  1. Junginger, H.; Jonker, J.; Faaij, A. Summary, Synthesis and Conclusions from IEA Bioenergy Task 40 Country Reports on International Bioenergy Trade; Copernicus Institute, Utrecht University: Utrecht, The Netherlands, 2011; pp. 1–25. Available online: (accessed on 31 March 2021).
  2. Demirbas, M.F. Emissions of polychlorinated dibenzo-p-dioxins and dibenzofurans from biomass combustion and solid waste incineration. Energy Sources Part A Recover. Util. Environ. Eff. 2007, 29, 1041–1047.
  3. Monien, B.H.; Herrmann, K.; Florian, S.; Glatt, H. Metabolic activation of furfuryl alcohol: Formation of 2-methylfuranyl DNA adducts in Salmonella typhimurium strains expressing human sulfotransferase 1A1 and in FVB/N mice. Carcinogenesis 2011, 32, 1533–1539.
  4. Hall, D.O.; Rosillo-Calle, F.; Williams, R.H.; Woods, J. Biomass for energy: Supply prospects. In Biomass Energy Supply Prospect; Earthscan: London, UK, 1993; pp. 593–651. ISBN 1853831557.
  5. Brunner, P.H.; Rechberger, H. Practical Handbook of Material Flow Analysis; CRC Press: Boca Raton, FL, USA, 2016; ISBN 9780203507209.
  6. Awasthi, M.K.; Zhao, J.; Soundari, P.G.; Kumar, S.; Chen, H.; Awasthi, S.K.; Duan, Y.; Liu, T.; Pandey, A.; Zhang, Z. Sustainable Management of Solid Waste. In Sustainable Resource Recovery and Zero Waste Approaches; Elsevier: Amsterdam, The Netherlands, 2019; pp. 79–99.
  7. Korhonen, J.; Honkasalo, A.; Seppälä, J. Circular Economy: The Concept and its Limitations. Ecol. Econ. 2018, 143, 37–46.
  8. Agarwal, M.; Jareda, K.; Bajpai, M. A review on solid waste management for smart city. SSRG Int. J. Civil Eng. (SSRG-IJCE) 2016, 3, 109–112.
  9. Berndes, G.; Hoogwijk, M.; Van Den Broek, R. The contribution of biomass in the future global energy supply: A review of 17 studies. Biomass Bioenergy 2003, 25, 1–28.
  10. Kopetz, H. Renewable resources: Build a biomass energy market. Nature 2013, 494, 29–31.
  11. Chum, H.L.; Overend, R.P. Biomass and renewable fuels. Fuel Process. Technol. 2001, 71, 187–195.
  12. Saldarriaga-Hernández, S.; Velasco-Ayala, C.; Flores, P.L.I.; de Jesús Rostro-Alanis, M.; Parra-Saldivar, R.; Iqbal, H.M.; Carrillo-Nieves, D. Biotransformation of lignocellulosic biomass into industrially relevant products with the aid of fungi-derived lignocellulolytic enzymes. Int. J. Biol. Macromol. 2020, 161, 1099–1116.
  13. Razik, A.H.A.; Khor, C.S.; Elkamel, A. A model-based approach for biomass-to-bioproducts supply Chain network planning optimization. Food Bioprod. Proc. 2019, 118, 293–305.
  14. Bilal, M.; Iqbal, H.M. Recent Advancements in the Life Cycle Analysis of Lignocellulosic Biomass. Curr. Sustain. Renew. Energy Rep. 2020, 7, 100–107.
  15. Demirbaş, A. Biomass resource facilities and biomass conversion processing for fuels and chemicals. Energy Convers. Manag. 2001, 42, 1357–1378.
  16. Parikka, M. Global biomass fuel resources. Biomass Bioenergy 2004, 27, 613–620.
  17. Arevalo-Gallegos, A.; Ahmad, Z.; Asgher, M.; Parra-Saldivar, R.; Iqbal, H.M. Lignocellulose: A sustainable material to produce value-added products with a zero waste approach—A review. Int. J. Biol. Macromol. 2017, 99, 308–318.
  18. Cherubini, F. The biorefinery concept: Using biomass instead of oil for producing energy and chemicals. Energy Convers. Manag. 2010, 51, 1412–1421.
  19. John, R.P.; Sukumaran, R.K.; Nampoothiri, K.M.; Pandey, A. Statistical optimization of simultaneous saccharification and l(+)-lactic acid fermentation from cassava bagasse using mixed culture of lactobacilli by response surface methodology. Biochem. Eng. J. 2007, 36, 262–267.
  20. Naik, S.N.; Goud, V.V.; Rout, P.K.; Dalai, A.K. Production of first and second generation biofuels: A comprehensive review. Renew. Sustain. Energy Rev. 2010, 14, 578–597.
  21. Kumar, A.; Kumar, N.; Baredar, P.; Shukla, A. A review on biomass energy resources, potential, conversion and policy in India. Renew. Sustain. Energy Rev. 2015, 45, 530–539.
  22. Balat, M.; Balat, H. Recent trends in global production and utilization of bio-ethanol fuel. Appl. Energy 2009, 86, 2273–2282.
  23. Deublein, D.; Steinhauser, A. Biogas from Waste and Renewable Resources: An Introduction, 2nd, Revised and Expanded Edition; Current Reviews for Academic Libraries; John Wiley & Sons: New Jersey, NJ, USA, 2010; p. 578. ISBN 978-3-527-32798-0.
  24. Mata, T.M.; Martins, A.A.; Caetano, N.S. Microalgae for biodiesel production and other applications: A review. Renew. Sustain. Energy Rev. 2010, 14, 217–232.
  25. Zeng, X.; Ma, Y.; Ma, L. Utilization of straw in biomass energy in China. Renew. Sustain. Energy Rev. 2007, 11, 976–987.
  26. Mendu, V.; Harman-Ware, A.E.; Crocker, M.; Jae, J.; Stork, J.; Morton, S.; Placido, A.; Huber, G.; Debolt, S. Identification and thermochemical analysis of high-lignin feedstocks for biofuel and biochemical production. Biotechnol. Biofuels 2011, 4.
  27. Mendu, V.; Shearin, T.; Campbell, J.E.; Stork, J.; Jae, J.; Crocker, M.; Huber, G.; DeBolt, S. Global bioenergy potential from high-lignin agricultural residue. Proc. Natl. Acad. Sci. USA 2012, 109, 4014–4019.
  28. Welker, C.M.; Balasubramanian, V.K.; Petti, C.; Rai, K.M.; De Bolt, S.; Mendu, V. Engineering plant biomass lignin content and composition for biofuels and bioproducts. Energies 2015, 8, 7654–7676.
  29. Li, W.; Amos, K.; Li, M.; Pu, Y.; DeBolt, S.; Ragauskas, A.J.; Shi, J. Fractionation and characterization of lignin streams from unique high-lignin content endocarp feedstocks. Biotechnol. Biofuels 2018, 11.
  30. Harman-Ware, A.E.; Crocker, M.; Pace, R.B.; Placido, A.; Morton, S.; DeBolt, S. Characterization of Endocarp Biomass and Extracted Lignin Using Pyrolysis and Spectroscopic Methods. Bioenergy Res. 2015, 8, 350–368.
  31. Dardick, C.D.; Callahan, A.M.; Chiozzotto, R.; Schaffer, R.J.; Piagnani, M.C.; Scorza, R. Stone formation in peach fruit exhibits spatial coordination of the lignin and flavonoid pathways and similarity to Arabidopsis dehiscence. BMC Biol. 2010, 8.
  32. Johnson, E.A.; Echavarri-Erasun, C. Chapter 3—Yeast biotechnology. In The Yeasts; Elsevier: San Diego, CA, USA, 2011; pp. 21–44.
  33. Siqueira, P.F.; Karp, S.G.; Carvalho, J.C.; Sturm, W.; Rodríguez-León, J.A.; Tholozan, J.L.; Singhania, R.R.; Pandey, A.; Soccol, C.R. Production of bio-ethanol from soybean molasses by Saccharomyces cerevisiae at laboratory, pilot and industrial scales. Bioresour. Technol. 2008, 99, 8156–8163.
  34. John, R.P.; Nampoothiri, K.M.; Pandey, A. Fermentative production of lactic acid from biomass: An overview on process developments and future perspectives. Appl. Microbiol. Biotechnol. 2007, 74, 524–534.
  35. Makkar, R.S.; Cameotra, S.S. Biosurfactant production by microorganisms on unconventional carbon sources. J. Surfactants Deterg. 1999, 2, 237–241.
  36. Nigam, P.S.; Singh, A. Production of liquid biofuels from renewable resources. Prog. Energy Combust. Sci. 2011, 37, 52–68.
  37. Panwar, N.L.; Kaushik, S.C.; Kothari, S. Role of renewable energy sources in environmental protection: A review. Renew. Sustain. Energy Rev. 2011, 15, 1513–1524.
  38. Bhutto, A.W.; Bazmi, A.A.; Zahedi, G. Greener energy: Issues and challenges for Pakistan-Biomass energy prospective. Renew. Sustain. Energy Rev. 2011, 15, 3207–3219.
  39. Turan, N.G.; Çoruh, S.; Akdemir, A.; Ergun, O.N. Municipal solid waste management strategies in Turkey. Waste Manag. 2009, 29, 465–469.
  40. Ionescu, G.; Rada, E.C.; Ragazzi, M.; Mǎrculescu, C.; Badea, A.; Apostol, T. Integrated municipal solid waste scenario model using advanced pretreatment and waste to energy processes. Energy Convers. Manag. 2013, 76, 1083–1092.
  41. Kupper, T.; Bürge, D.; Bachmann, H.J.; Güsewell, S.; Mayer, J. Heavy metals in source-separated compost and digestates. Waste Manag. 2014, 34, 867–874.
  42. Bridgwater, A.V.; Meier, D.; Radlein, D. An overview of fast pyrolysis of biomass. Org. Geochem. 1999, 30, 1479–1493.
  43. Sonesson, U.; Björklund, A.; Carlsson, M.; Dalemo, M. Environmental and economic analysis of management systems for biodegradable waste. Resour. Conserv. Recycl. 2000, 28, 29–53.
  44. Kiyasudeen, K.; Ibrahim, M.H.; Quaik, S.; Ismail, S.A. An introduction to anaerobic digestion of organic wastes. In Prospects of Organic Waste Management and the Significance of Earthworms; Springer: Cham, Switzerland, 2016; pp. 23–44.
  45. Vasco-Correa, J.; Khanal, S.; Manandhar, A.; Shah, A. Anaerobic digestion for bioenergy production: Global status, environmental and techno-economic implications, and government policies. Bioresour. Technol. 2018, 247, 1015–1026.
  46. Yokoyama, S. The Asian Biomass Handbook A Guide for Biomass Production and Utilization Support Project for Building Asian-Partnership for Environmentally Conscious Agriculture; Entrusted by Ministry of Agriculture, Forestry, and Fisheries; The Japan Institute of Energy: Tokyo, Japan, 2008.
  47. Kircher, M. Sustainability of biofuels and renewable chemicals production from biomass. Curr. Opin. Chem. Biol. 2015, 29, 26–31.
  48. Brar, S.K.; Dhillon, G.S.; Soccol, C.R. Biotransformation of Waste Biomass into High Value Biochemicals; Springer Science & Business Media: New York, NY, USA, 2013; ISBN 9781461480044.
  49. Duku, M.H.; Gu, S.; Hagan, E. Ben A comprehensive review of biomass resources and biofuels potential in Ghana. Renew. Sustain. Energy Rev. 2011, 15, 404–415.
  50. Kögel-Knabner, I. The macromolecular organic composition of plant and microbial residues as inputs to soil organic matter. Soil Biol. Biochem. 2002, 34, 139–162.
  51. Zhang, N.; Li, S.; Xiong, L.; Hong, Y.; Chen, Y. Cellulose-hemicellulose interaction in wood secondary cell-wall. Model. Simul. Mater. Sci. Eng. 2015, 23.
  52. Flemming, H.C.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8, 623–633.
  53. Maki, M.; Leung, K.T.; Qin, W. The prospects of cellulase-producing bacteria for the bioconversion of lignocellulosic biomass The prospects of cellulase-producing bacteria for the bioconversion of lignocellulosic biomass Page 2 sur 8. Int. J. Biol. Sci. 2013, 5, 1–8.
  54. Ekman, A.; Börjesson, P. Environmental assessment of propionic acid produced in an agricultural biomass-based biorefinery system. J. Clean. Prod. 2011, 19, 1257–1265.
  55. Bilal, M.; Asgher, M.; Iqbal, H.M.; Hu, H.; Zhang, X. Biotransformation of lignocellulosic materials into value-added products—A review. Int. J. Biol. Macromol. 2017, 98, 447–458.
  56. Zhang, L.; Xu, C.; Champagne, P. Overview of recent advances in thermo-chemical conversion of biomass. Energy Convers. Manag. 2010, 51, 969–982.
  57. Van der Stelt, M.J.C.; Gerhauser, H.; Kiel, J.H.A.; Ptasinski, K.J. Biomass upgrading by torrefaction for the production of biofuels: A review. Biomass Bioenergy 2011, 35, 3748–3762.
  58. Demirbas, A. Combustion characteristics of different biomass fuels. Prog. Energy Combust. Sci. 2004, 30, 219–230.
  59. Jenkins, B.M.; Baxter, L.L.; Miles, T.R.; Miles, T.R. Combustion properties of biomass. Fuel Process. Technol. 1998, 54, 17–46.
  60. Lin, C.S.K.; Pfaltzgraff, L.A.; Herrero-Davila, L.; Mubofu, E.B.; Abderrahim, S.; Clark, J.H.; Koutinas, A.A.; Kopsahelis, N.; Stamatelatou, K.; Dickson, F.; et al. Food waste as a valuable resource for the production of chemicals, materials and fuels. Current situation and global perspective. Energy Environ. Sci. 2013, 6, 426–464.
  61. Lin, C.S.K.; Koutinas, A.A.; Stamatelatou, K.; Mubofu, E.B.; Matharu, A.S.; Kopsahelis, N.; Pfaltzgraff, L.A.; Clark, J.H.; Papanikolaou, S.; Kwan, T.H.; et al. Current and future trends in food waste valorization for the production of chemicals, materials and fuels: A global perspective. Biofuels Bioprod. Biorefin. 2014, 8, 686–715.
  62. Matharu, A.S.; de Melo, E.M.; Houghton, J.A. Food Supply Chain Waste: A Functional Periodic Table of Bio-Based Resources. In Waste Biorefinery; Elsevier: Amsterdam, The Netherlands, 2018; pp. 219–236.
  63. Vlachokostas, C.; Achillas, C.; Diamantis, V.; Michailidou, A.V.; Baginetas, K.; Aidonis, D. Supporting decision making to achieve circularity via a biodegradable waste-to-bioenergy and compost facility. J. Environ. Manag. 2021, 285, 112215.
  64. Vlachokostas, C.; Michailidou, A.V.; Achillas, C. Multi-Criteria Decision Analysis towards promoting Waste-to-Energy Management Strategies: A critical review. Renew. Sust. Energ. Rev. 2021, 138, 110563.
  65. Gray, K.; Zhao, L.; Emptage, M. Bioethanol. Curr. Opin. Chem. Biol. 2006, 10, 141–146.
  66. You, C.; Chen, H.; Myung, S.; Sathitsuksanoh, N.; Ma, H.; Zhang, X.Z.; Li, J.; Zhang, Y.H.P. Enzymatic transformation of nonfood biomass to starch. Proc. Natl. Acad. Sci. USA 2013, 110, 7182–7187.
  67. Lin, Y.; Tanaka, S. Ethanol fermentation from biomass resources: Current state and prospects. Appl. Microbiol. Biotechnol. 2006, 69, 627–642.
  68. Bozell, J.J.; Petersen, G.R. Technology development for the production of biobased products from biorefinery carbohydrates—The US Department of Energy’s “top 10” revisited. Green Chem. 2010, 12, 539–554.
  69. Tuck, C.O.; Pérez, E.; Horváth, I.T.; Sheldon, R.A.; Poliakoff, M. Valorization of biomass: Deriving more value from waste. Science 2012, 337, 695–699.
  70. Yue, D.; You, F.; Snyder, S.W. Biomass-to-bioenergy and biofuel supply chain optimization: Overview, key issues and challenges. Comput. Chem. Eng. 2014, 66, 36–56.
  71. Richard, T.L. Challenges in scaling up biofuels infrastructure. Science 2010, 329, 793–796.
  72. Agbor, V.B.; Cicek, N.; Sparling, R.; Berlin, A.; Levin, D.B. Biomass pretreatment: Fundamentals toward application. Biotechnol. Adv. 2011, 29, 675–685.
  73. Glasnov, T.N.; Kappe, C.O. The microwave-to-flow paradigm: Translating high-temperature batch microwave chemistry to scalable continuous-flow processes. Chem. A Eur. J. 2011, 17, 11956–11968.
  74. Ruiz, J.A.; Juárez, M.C.; Morales, M.P.; Muñoz, P.; Mendívil, M.A. Biomass gasification for electricity generation: Review of current technology barriers. Renew. Sustain. Energy Rev. 2013, 18, 174–183.
  75. Molino, A.; Chianese, S.; Musmarra, D. Biomass gasification technology: The state of the art overview. J. Energy Chem. 2016, 25, 10–25.
  76. Mondal, P.; Dang, G.S.; Garg, M.O. Syngas production through gasification and cleanup for downstream applications—Recent developments. Fuel Process. Technol. 2011, 92, 1395–1410.
  77. Lapuerta, M.; Hernández, J.J.; Pazo, A.; López, J. Gasification and co-gasification of biomass wastes: Effect of the biomass origin and the gasifier operating conditions. Fuel Process. Technol. 2008, 89, 828–837.
  78. Tock, J.Y.; Lai, C.L.; Lee, K.T.; Tan, K.T.; Bhatia, S. Banana biomass as potential renewable energy resource: A Malaysian case study. Renew. Sustain. Energy Rev. 2010, 14, 798–805.
  79. Gasafi, E.; Reinecke, M.Y.; Kruse, A.; Schebek, L. Economic analysis of sewage sludge gasification in supercritical water for hydrogen production. Biomass Bioenergy 2008, 32, 1085–1096.
  80. Antoni, D.; Zverlov, V.V.; Schwarz, W.H. Biofuels from microbes. Appl. Microbiol. Biotechnol. 2007, 77, 23–35.
  81. Gumisiriza, R.; Mshandete, A.M.; Rubindamayugi, M.S.T.; Kansiime, F.; Kivaisi, A.K. Enhancement of anaerobic digestion of Nile perch fish processing wastewater. Afr. J. Biotechnol. 2009, 8, 328–333.
  82. Cirne, D.G.; Paloumet, X.; Björnsson, L.; Alves, M.M.; Mattiasson, B. Anaerobic digestion of lipid-rich waste-Effects of lipid concentration. Renew. Energy 2007, 32, 965–975.
  83. Das Neves, L.C.M.; Converti, A.; Penna, T.C.V. Biogas production: New trends for alternative energy sources in rural and urban zones. Chem. Eng. Technol. 2009, 32, 1147–1153.
  84. Gupta, P.; Gupta, A. Biogas production from coal via anaerobic fermentation. Fuel 2014, 118, 238–242.
  85. Scarlat, N.; Motola, V.; Dallemand, J.F.; Monforti-Ferrario, F.; Mofor, L. Evaluation of energy potential of Municipal Solid Waste from African urban areas. Renew. Sustain. Energy Rev. 2015, 50, 1269–1286.
  86. Titirici, M.M.; Thomas, A.; Antonietti, M. Back in the black: Hydrothermal carbonization of plant material as an efficient chemical process to treat the CO2 problem? New J. Chem. 2007, 31, 787–789.
  87. Huber, G.W.; Iborra, S.; Corma, A. Synthesis of transportation fuels from biomass: Chemistry, catalysts, and engineering. Chem. Rev. 2006, 106, 4044–4098.
  88. Silveira, M.H.L.; Morais, A.R.C.; Da Costa Lopes, A.M.; Olekszyszen, D.N.; Bogel-Łukasik, R.; Andreaus, J.; Pereira Ramos, L. Current Pretreatment Technologies for the Development of Cellulosic Ethanol and Biorefineries. ChemSusChem 2015, 8, 3366–3390.
  89. Meyer, S.; Glaser, B.; Quicker, P. Technical, economical, and climate-related aspects of biochar production technologies: A literature review. Environ. Sci. Technol. 2011, 45, 9473–9483.
  90. Richel, A.; Paquot, M. Conversion of Carbohydrates Under Microwave Heating. In Carbohydrates—Comprehensive Studies on Glycobiology and Glycotechnology; InTech: Rijeka, Croatia, 2012.
  91. Kappe, C.O. Controlled microwave heating in modern organic synthesis. Angew. Chem. Int. Ed. 2004, 43, 6250–6284.
  92. Guo, L.; Li, X.M.; Bo, X.; Yang, Q.; Zeng, G.M.; Liao, D.X.; Liu, J.J. Impacts of sterilization, microwave and ultrasonication pretreatment on hydrogen producing using waste sludge. Bioresour. Technol. 2008, 99, 3651–3658.
  93. Hu, Y.; Liu, C.; Zhang, Y.; Ren, N.; Tang, Y. Microwave-assisted hydrothermal synthesis of nanozeolites with controllable size. Microporous Mesoporous Mater. 2009, 119, 306–314.
  94. Menéndez, J.A.; Inguanzo, M.; Pis, J.J. Microwave-induced pyrolysis of sewage sludge. Water Res. 2002, 36, 3261–3264.
  95. Digman, B.; Joo, H.S.; Kim, D.S. Recent progress in gasification/ pyrolysis technologies for biomass conversion to energy. Environ. Prog. Sustain. Energy 2009, 28, 47–51.
  96. Yin, C. Microwave-assisted pyrolysis of biomass for liquid biofuels production. Bioresour. Technol. 2012, 120, 273–284.
  97. Luque, R.; Menéndez, J.A.; Arenillas, A.; Cot, J. Microwave-assisted pyrolysis of biomass feedstocks: The way forward? Energy Environ. Sci. 2012, 5, 5481–5488.
  98. Guiotoku, M.; Rambo, C.R.; Hansel, F.A.; Magalhães, W.L.E.; Hotza, D. Microwave-assisted hydrothermal carbonization of lignocellulosic materials. Mater. Lett. 2009, 63, 2707–2709.
  99. Amr Sohby, J.C. Microwave-assisted biorefinery. Chem. Eng. Trans. 2010, 19, 211–212.
  100. Nomanbhay, S.; Salman, B.; Hussain, R.; Ong, M.Y. Microwave pyrolysis of lignocellulosic biomass—A contribution to power Africa. Energy Sustain. Soc. 2017, 7, 1–24.
  101. Narzari, R.; Bordoloi, N.; Sarma, B.; Gogoi, L.; Gogoi, N.; Borkotoki, B.; Kataki, R. Fabrication of biochars obtained from valorization of biowaste and evaluation of its physicochemical properties. Bioresour. Technol. 2017, 242, 324–328.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 1.5K
Revisions: 2 times (View History)
Update Date: 05 Nov 2021
1000/1000