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 -- 4006 2024-01-25 15:49:47 |
2 layout & references Meta information modification 4006 2024-01-26 02:13:05 |

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.
Yaverino-Gutiérrez, M.A.; Wong, A.Y.C.; Ibarra-Muñoz, L.A.; Chávez, A.C.F.; Sosa-Martínez, J.D.; Tagle-Pedroza, A.S.; Hernández-Beltran, J.U.; Sánchez-Muñoz, S.; Santos, J.C.D.; Da Silva, S.S.; et al. Bioethanol Processing. Encyclopedia. Available online: https://encyclopedia.pub/entry/54357 (accessed on 03 July 2024).
Yaverino-Gutiérrez MA, Wong AYC, Ibarra-Muñoz LA, Chávez ACF, Sosa-Martínez JD, Tagle-Pedroza AS, et al. Bioethanol Processing. Encyclopedia. Available at: https://encyclopedia.pub/entry/54357. Accessed July 03, 2024.
Yaverino-Gutiérrez, Mario Alberto, Alán Yazid Chávez-Hita Wong, Lizbeth Alejandra Ibarra-Muñoz, Ana Cristina Figueroa Chávez, Jazel Doménica Sosa-Martínez, Ana Sofia Tagle-Pedroza, Javier Ulises Hernández-Beltran, Salvador Sánchez-Muñoz, Julio César Dos Santos, Silvio Silvério Da Silva, et al. "Bioethanol Processing" Encyclopedia, https://encyclopedia.pub/entry/54357 (accessed July 03, 2024).
Yaverino-Gutiérrez, M.A., Wong, A.Y.C., Ibarra-Muñoz, L.A., Chávez, A.C.F., Sosa-Martínez, J.D., Tagle-Pedroza, A.S., Hernández-Beltran, J.U., Sánchez-Muñoz, S., Santos, J.C.D., Da Silva, S.S., & Balagurusamy, N. (2024, January 25). Bioethanol Processing. In Encyclopedia. https://encyclopedia.pub/entry/54357
Yaverino-Gutiérrez, Mario Alberto, et al. "Bioethanol Processing." Encyclopedia. Web. 25 January, 2024.
Bioethanol Processing
Edit

The liquid biofuel bioethanol is extensively produced worldwide through the fermentation of sugars obtained from various raw materials, including lignocellulosic biomass, an abundant renewable resource. Due to its recalcitrant nature, lignocellulosic materials typically is pretreated using mechanical, chemical, physicochemical, or biological methods to enhance sugar recovery. The pretreated lignocellulosic biomass then undergoes a fermentation process, either sequentially or simultaneously with saccharification, resulting in the biofuel called second-generation ethanol. The ethanol yield is influenced by various fermentation strategies and conditions, such as inoculum concentration, medium agitation, temperature, and pH.

lignocellulosic biomass bioethanol metabolic stress strain engineering life cycle assessment

1. Introduction

The global population has grown massively over the last century, increasing from 2.7 billion people in 1955 to over 7.8 billion in 2022, resulting in a greater demand for energy resources that has been fulfilled by fossil fuels (FFs) [1]. FFs are nonrenewable energy sources that pose a threat to the environment as well as the global economy, being the primary contributors of greenhouse gas (GHG) emissions. The use of FFs in the transportation sector is responsible for 24% of global CO2 emissions [2][3]. To combat climate change, several countries are shifting towards renewable fuels such as bioethanol, biodiesel, and bio-hydrogen as alternatives to conventional fuels like oil, coal, and natural gas [4][5]. Biofuels are increasingly seen as a way to improve energy security, supply, GHG mitigation, and transport sector development [6]. According to the Renewable Fuels Association (RFA), bioethanol is the most widely produced liquid biofuel in the world, with 142.6 billion liters generated from vegetable biomass and algae, making it a crucial player in the transport sector [7]. Bioethanol is an alcohol that is produced by the microbial fermentation of carbohydrates. During this process, microorganisms break down organic molecules into simpler products, including ethanol, carbon dioxide, and other by-products. Currently, the focus of bioethanol production centers on its application within the transportation sector. Governmental regulations have played a pivotal role in incentivizing the ethanol market to reduce dependency on fossil fuels [8]. Notably, the USA and Brazil have emerged as major contributors, generating 61.54 and 29.85 billion liters of ethanol, respectively, predominantly derived from food crops such as corn and sugarcane. However, the adverse environmental impacts associated with waste disposal have spurred intensive research into technologies aimed at integrating waste materials into the biofuel conversion process, with a primary emphasis on bioethanol [9]. In the case of Mexico, as reported by Carillo-Nieves et al. [10], the country leads global production in 71 agro-food products and holds the top position in a total of 181. Consequently, diverse scenarios are under evaluation to enhance the efficient production of second-generation (2G) ethanol, employing consolidated bioprocessing methods. Brazil has been at the forefront of integrated biorefinery research, focusing on both first-generation (1G) and second-generation (2G) sugarcane processes. This involves the predominant production of bioethanol and advancements in upgrading for jet fuel applications [11]. The production of bioethanol not only plays a crucial role in reducing carbon emissions but also contributes significantly to fuel security. In general, the bioethanol production process is categorized into four generations based on the raw materials used and the associated processing technologies [12][13][14].
First-generation bioethanol is produced using crops intended for food consumption as raw materials. These crops can be divided into two groups: the first includes sugarcane, sugar beet, sweet sorghum, and fruit, as well as sugar refinery residues. The second group includes starch sources such as cereals, tubers, and roots, legumes, and green and unripe fruits, with maize, sorghum grains, wheat, cassava, potatoes, and barley being commonly used [14][15]. In 2019, Brazil and the USA were the largest bioethanol producers, using corn and sugarcane as their main raw materials, and together accounting for 84% of total world production. In the USA, 94% of bioethanol is produced from corn starch, whereas in Brazil, 99% is obtained from sugarcane [16]. The use of first-generation feedstocks for bioethanol production raises a contentious issue as these crops are primarily intended for food or feed purposes, and their use in biofuels production can cause significant market price increases due to the use of chemical fertilizers to improve yields.
Second-generation bioethanol comes from lignocellulosic biomass (LCB). Due to its high annual production of 170 billion metric tons [15], lignocellulose is a readily available and easily accessible bioresource, which makes it an attractive option for bioethanol production [17]. LCB can be derived from various sources, such as agricultural residues (straws, stover, leaves, bagasse, etc.), forest residues (sawdust, pruning’s, chips, etc.), wood (soft and hardwoods, such as pine, spruce, and aspen), energy crops (alternative feedstocks relieving the food vs. fuel competition) such as Switchgrass or Miscanthus (i.e., Miscanthus giganteus, Miscanthus sacchariflorus, Miscanthus sinensis), and municipal solid waste [18][19]. Additionally, the use of lignocellulosic wastes for bioethanol production significantly reduces environmental contamination [20][21]. Bioethanol production from lignocellulosic biomass has the advantage of using non-edible feedstock, which does not compete with the food industry [13][22]. However, a drawback associated with the use of lignocellulosic biomass (LCB) lies in the elevated production costs, as the substrates require a pretreatment with or without an enzymatic hydrolysis process before fermentation can occur. Additionally, biofuels derived from LCB require careful consideration due to the environmental impact, energy consumption, and maintenance involved throughout the entire process. All these aspects are evaluated through a life cycle assessment (LCA), which will be discussed, although focusing mainly on the second generation of bioethanol production [23].
Third-generation bioethanol comes from algal biomass. Algal biomass is sustainable and abundant since both macroalgae and microalgae can be used [24][25]. They are photosynthetic organisms that utilize carbon dioxide (CO2) and are made up of carbohydrates, proteins, and lipids, and the content of each component varies between species [24]. Macroalgae, which are typically sourced from salty marine waters, are of three types: brown (Phaeophyceae), red (Rhodophyceae), and green (Chlorophyceae) [26][27].
Freshwater macroalgae such as Oedogonium, Rhizoclonium, Ulothrix, and Microspora also offer significant potential for biofuel production [28]. Compared to traditional raw materials that require longer production times, macroalgae can produce significantly higher amounts of bioethanol, with a yield of 0.43 g/g of substrate, making their use more profitable and viable [29]. It has been reported that the use of macroalgae biomass for bioethanol production is more cost-effective and practical than microalgal biomass [27][28][29]. However, microalgae are ideal for wastewater treatment since they are a good source of nitrogen, phosphorus, and other nutrients necessary for their growth [30][31].
Obtaining microalgae biomass and treating wastewater helps to reduce CO2 emissions through sequestration and decrease the cost of bioethanol production [27]. CO2 sequestration and yield depend on the growth rate of the microalgae, which in turn is influenced by the species, nutrient ratio, light intensity, pH, and temperature [31]. Certain oleaginous microalgae, such as Scenedesmus sp., Chlorella sp., Chlamydomonas sp., and Desmodesmus sp., can store more than 70% of lipids and increase their carbohydrate content up to 50% by weight of dry biomass under favorable growth conditions. As a result of their high sugar content, these microalgae are a suitable feedstock for bioethanol production [28][30]. In addition, algal biomass needs to be pretreated to make it suitable for the chemical conversion of intracellular compounds into value-added products [25][32]. Since the pretreatments used are diverse and specific, they will be described in the pretreatment section of this chapter due to their complexity. The utilization of algal biomass can serve as an excellent alternative to replace nonrenewable sources and promote environmentally friendly practices through sustainable bioethanol production.
The fourth generation of biofuels relies on genetic modification of the feedstocks (crops)/microorganisms to enhance the uptake and storage of high concentrations of carbon dioxide (CO2) in feedstocks and as well to improve the metabolic efficiency of microorganisms used in the fermentation process [33][34][35]. By creating a synthetic carbon sink, this approach helps to minimize CO2 emissions into the environment, resulting in a 50% reduction in GHG emissions and contributing to mitigating global warming [34][36]. Genetic engineering can also improve the growth rate of organisms, increase efficiency by boosting bioethanol yield, and lower production costs [33][37]. However, there are potential drawbacks associated with genetically modified organisms, including their potential impact on the environment due to the intentional or unintentional release of transgenic agents, which can alter natural habitats [38]. Despite this, the use of genetically modified organisms in biofuel production has the potential to significantly contribute to sustainable energy production and environmental preservation [38]. The fourth generation of biofuels has been viewed as a promising solution to the limitations of previous generations, which involved the use of crops intended for human consumption and/ or expensive and time-consuming pretreatment processes [36]. However, it is still in the development phase, due to certain limitations that need to be addressed through extensive research, before it can be implemented on an industrial scale. Failure to do so could result in reduced profitability of bioethanol production [35][36]. Despite these challenges, the potential benefits of fourth-generation biofuels make it a worthwhile area of exploration and investment for sustainable energy production in the future. Figure 1 summarizes the difference between generations for bioethanol production.
Figure 1. Graphical representation of the different generations in bioethanol production.

2. Breakdown of Lignocellulosic Substrates to Fermentable Sugars

2.1. Lignocellulosic Biomass Composition

The large-scale availability of LCB has led to interest in its use as feedstock for ethanol production. Typically, LCB consists of three main components: cellulose (40–60%), hemicellulose (25–40%), and lignin (15–25%) [39]. The relative proportions of these components can vary depending on the source of the biomass. The composition of LCB of different feedstocks is presented in Table 1.
Table 1. Lignocellulosic biomass composition.
Cellulose is the most abundant lignocellulosic polymer on Earth, found in plant cell walls consisting of about 10,000–15,000 glucose units linked by β-(1,4) glycosidic bonds [53][54]. The organized microfibrils in cellulose establish multiple hydrogen bonds and van der Waals forces between hydroxyl groups, creating a rigid, resistant, and strong structure. The chains tend to form a crystalline structure, with various polymorphs, including the abundant polymorph I that is divided into cellulose Iα (algae and bacteria) and Iβ (main form in plants) [55]. Cellulose is present in various materials, including cotton, wood, bast fibers, grasses, algae, and bacteria [56][57].
Hemicellulose is a complex polymer composed of different sugars like pentoses (D-xylose, D-arabinose), hexoses (D-glucose, D-galactose, and D-mannose), and sugar acids (D-galacturonic and D-glucuronic acids) [14]. These sugars are connected by β-1,4-glycosidic bonds and, in some cases, β-1,3-glycosidic bonds. It can be degraded more easily by enzymes than cellulose [58]. Different hemicellulose structures can be found in softwood and hardwood. Within hardwoods, xylans and glucomannans are prevalent, while softwoods encompass galactoglucomannans, glucomannans, xylans, and arabinogalactans. Galactoglucomannans are compounds comprising glucose and mannose, interconnected by β-1,4 bonds, while galactose units are linked through α-1,6 bonds. Glucomannans are characterized by β-1,4-linked D-glucose and D-mannose. In softwood xylans, the predominant structure involves xylose residues linked by β-1,4 bonds, accompanied by α-1,3-linked arabinose and glucuronic acids in the side chains. In contrast, hardwoods display side branches of 4-o-methyl glucuronic acid connected via α-1,2 bonds. Arabinogalactans exhibit poly-β-(3-1,3)-galactose featuring side chains of 1,6 arabinose and galactose [59]. Several studies have demonstrated that the removal of hemicellulose through pretreatment processes can enhance cellulose conversion and facilitate enzyme access to cellulose [57].
In the realm of bioethanol production, the primary focus for conversion into ethanol during the fermentation process lies predominantly on sugars, such as cellulose and hemicellulose. However, the presence of lignin presents a formidable challenge. Acting as a recalcitrant constituent, lignin serves as a substantial physical barrier that curtails both enzymatic hydrolysis and microbial degradation [60]. Lignin is a complex heteropolymer formed by phenylpropanoid monomers (monolignols) such as p-coumaryl alcohol (hydroxyphenyl), coniferyl alcohol (guaiacyl), and sinapyl alcohol (syringyl) [61]. It is found in grasses and straws containing the three lignifying monomers, in softwoods composed of coniferyl alcohol, and in hardwoods, which have coniferyl alcohol and sinapyl alcohol in their composition [62]. Monolignols are joined by different bonds, the most common being β-O-4, β-5, 5-5, 4-O-5, β-β, and β-1. The main linkage that is present is β-O-4, which is weak compared to others and plays an important role in limiting the access of enzymes to cellulose [62][63]. Lignin is considered a physical barrier that limits enzymatic hydrolysis for ethanol due to the unproductive binding of cellulose to lignin [64]. Lignin is widely used when it is fractionated from cellulose and hemicellulose, and as a result of fermentation, it produces biopolymers and phenolic compounds [62]. The phenolic compounds, such as vanillin, syringaldehyde, and 4-hydroxybenzaldehyde, inhibit the fermentation of lignocellulosic hydrolysates, mostly despite low concentrations [65]. Moreover, lignin works as an adhesive because these phenolic compounds hold the fibers and keep cellulose and hemicellulose together [57]. Other inhibitors such as organic acids (acetic acid, formic acid, and levulinic acid), hydroxymethylfurfural (HMF), and furfurals exist, which affect the fermentation process, causing a negative impact on microbial viability [66].

2.2. Pretreatment Methods

Pretreatment processes aim to break down the complex structure of raw materials, facilitating access for hydrolytic enzymes to primarily target the carbohydrates fraction from second- and third-generation sources, converting them into fermentable sugars [4][67]. The effectiveness of pretreatment methods depends on the number of fermentable sugars released. Different technologies such as mechanical, physical, chemical, and biological methods have been successfully used to break down LCB [68][69]. A summary of the advancements on different pretreatment methods used for both lignocellulosic and algal biomass as raw materials and their resulting ethanol yield is presented in Table 2.
Table 2. Summary on the efficiency of different pretreatments on bioethanol production from different second-generation feedstocks.
It can be observed from Table 2 that most of the reported physicochemical pretreatment methods are effective in increasing LCB’s degradability for enzymatic hydrolysis and bioconversion to bioethanol in sugar recovery. The formation of inhibitory compounds throughout the process has many limitations for smooth adaptation at a larger scale, primarily due to the adverse environmental effects [72][74][82]. Among these inhibitory compounds are furan aldehydes, organic acids, pseudo-lignins, small lignin units, extractives, and phenolic compounds [88]. Additionally, the lingering chemical reagents further aggravate inhibition of both enzymatic and microbial activity. Yet another significant challenge arises from the substantial volume of water required for rinsing the residues after the pretreatment process. This water, which contains chemical residues, is frequently discarded [89]. In this regard, biological methods are more sustainable as they break down the complex polysaccharides into simple fermentable sugars without much environmental impact [86][90][91][92].

3. Sugar Recovery from Pretreated Biomass through Hydrolysis

After the removal of lignin through pretreatment methods, cellulose remains on the medium, and cellulose is broken down by a chemical or enzymatic hydrolysis to obtain the sugar monomers [87][90][93]. In chemical hydrolysis, acids such as HCl and H2SO4 are used to break the β-1,4 linkages in cellulose, releasing sugar monomers or oligomers. In chemical hydrolysis compounds such as furfural, hydroxymethylfurfural (HMF), phenol, vanillin, vanillic acid, and other phenolic compounds can be released, which decreases ethanol yield [94]. In enzymatic hydrolysis cellulases, β-1,4-endoglucanase, exoglycanase, and β-glucosidase are used to hydrolyze cellulose into glucose monomers [44]. β-1,4-endoglucanase initiates the saccharification process by randomly cleaving the β-1,4 glycosidic linkages, producing cellodextrins with both reducing and non-reducing ends. Exoglucanase acts on the non-reducing ends of the molecule, releasing cellobiose. Finally, β-glucosidase breaks down small chains of glucose, cleaving the β-1,4 linkages in cellobiose and releasing sugar monomers [95]. Cellulases have been reported to exhibit their highest enzymatic activity in the range of 45 °C to 50 °C, a pH level of 4.5 to 5.5, and at atmospheric pressure. However, the specific conditions mentioned can exhibit variations contingent upon the enzymatic source, which includes microorganisms, as well as the substrate [96].
Other factors affecting cellulose hydrolysis are crystallinity of cellulose and its particle size [97]. Crystallinity is a property of cellulose structure where its fibers are linked by non-covalent hydrogen bonds making it more resistant to hydrolysis than the amorphous parts. Alkali pretreatments have a higher degree of conserving the crystalline structure of cellulose compared with other pretreatments previously mentioned [98]. Particle size plays an important role in enzymatic hydrolysis since it affects the reaction rate between cellulose and enzymes. Small particle size of biomass is preferred for a higher conversion rate; however, they are harder to process while bigger LCB particles tend to not hydrolyze completely and have a slower conversion rate [98]. A study conducted by Kapoor et al. [99] compared glucose recovery from glucan with 10, 15, and 20 mm particle size resulted in a 65.6, 80, and 60% sugar recovery using 5 FPU/mL of hydrolytic enzymes complex.
Enzymatic hydrolysis is widely studied for the purpose of increasing the fermentable sugars released from LCB. Its efficiency is intricately linked to the initial substrate concentrations. Maintaining an optimal balance is imperative for achieving a higher yield of fermentable sugars, thereby maximizing overall biofuel production. Controlling and optimizing these concentrations is key to enhancing the cost-effectiveness and sustainability of the entire process. Gao et al. [100] and Mondebach [101] emphasized that most optimization efforts for cellulose and hemicellulose degradation occur at low solid loading (<5% w/v). However, this low concentration often results in a diminished yield of fermentable sugars and lower end-product concentrations.
Furthermore, additional substrate and enzyme loading significantly influence enzymatic hydrolysis efficiency. Carefully adjusting these parameters impacts reaction kinetics, enabling the extraction of a higher quantity of fermentable sugars from lignocellulosic material [100]. Achieving the right balance in substrate and enzyme loading is vital for the economic feasibility of biofuel production, as reported by Amândio et al. [102], who employed a fed-batch operation starting with 11% (w/v) substrate and achieved an impressive concentration of 161 g/L of sugars with an enzymatic hydrolysis conversion efficiency of 76%.
Different modifications to enzymatic hydrolysis have been studied to increase sugar yield. For example, a study by Lee et al. [86] employed an enzymatic complex comprising cellulase, β-glucosidase, and amyloglucosidase, which yielded 28 g of sugar per liter, accompanied by an efficient ethanol conversion rate of 40% from the produced fermentable sugars. Furthermore, there are several studies that have combined pretreatment with enzymatic hydrolysis. For instance, Constantino et al. [103] adopted a chemo-enzymatic hydrolysis approach, utilizing α-amylase and a diluted acid treatment for biomass autoclaving with 4% H2SO4 v/v at 121 °C for 30 min. They employed microalgae as feedstock and obtained 34 ± 1 g of reducing sugars per 100 g of dry biomass. Wang et al. [104] elevated the maximum efficiency to reach 91.7% through enzymatic hydrolysis after the combination of alkaline and ozone pretreatment. Meanwhile, Ostadjoo et al. [105] employed Xylanase from Thermomyces lanuginosus for the hydrolysis of xylans from hemicellulose and recorded a yield surpassing 70%.

4. Metabolic Pathways, Settings, and Factors of Fermentation Process

Ethanol synthesis is carried out by microorganisms through the fermentation of reducing sugars. Glucose monomers are oxidized in the glycolysis process, forming two pyruvic acid molecules and a pair of NADH molecules from two NAD+ molecules. Pyruvic acid is then transformed into acetaldehyde by the enzyme pyruvate decarboxylase, while releasing carbon dioxide. One molecule of acetaldehyde is converted into a molecule of ethanol by the enzyme alcohol dehydrogenase, using the previous pair of NADH and regenerating the NAD+ molecules [106].
Five-carbon molecules such as xylose and mannose can remain in the medium when using substrates containing hexoses and pentoses. Saccharomyces cerevisiae cannot ferment pentose by itself; thus, pentose-fermenting microorganisms (PFMs) can be added to produce ethanol from five-carbon sugars. Zymomonas mobilis, Pichia stipitis, and Candida shehatae are the most common PFMs used in pentose fermentation. Pentose fermentation is carried out through the pentose phosphoketolase (PPK) pathway and is commonly seen in lactic-acid-producing bacteria. The PPK pathway begins with the same five steps as the pentose phosphate pathway (PPP), in which glucose is transformed into ribulose-5-phosphate (R5P) and xylulose-5-phosphate (X5P). Exogenous five-carbon sugars enter the microorganism and are phosphorylated by the 6-phosphogluconate dehydrogenase enzyme to be converted into R5P or X5P. R5P is used by microorganisms to produce ribose for nucleic acids or coenzymes, whereas X5P enters the PPK pathway and is catabolized into acetyl phosphate by pentose phosphoketolase. The acetyl phosphate then acts with phosphotransacetylase and aldehyde dehydrogenase to produce acetaldehyde, which is converted to ethanol by alcohol dehydrogenase [107]. A scheme on the fermentation of hexoses and pentoses into bioethanol is given in Figure 2.
Figure 2. A schematic representation of hexoses and pentoses fermentation for bioethanol production. (1) Pretreatment of lignocellulosic biomass. (2) Enzymatic hydrolysis of pretreated biomass to release hexose and pentose. (3) Addition of five- and six-carbon sugars to the bioreactor for ethanol fermentation. (4) Ethanol fermentation pathway from glucose inside the reactor. (5) Ethanol fermentation pathway from pentoses inside the reactor.

References

  1. Sadigov, R. Rapid Growth of the World Population and Its Socioeconomic Results. Sci. World J. 2022, 2022, 8110229.
  2. Ritchie, Hannah, and Max Roser. “Cars, Planes, Trains: Where Do CO2 Emissions from Transport Come From?” Our World in Data. Available online: https://ourworldindata.org/co2-emissions-from-transport (accessed on 28 December 2023).
  3. Huang, Y.; Zhang, Y.; Deng, F.; Zhao, D.; Wu, R. Impacts of Built-Environment on Carbon Dioxide Emissions from Traffic: A Systematic Literature Review. Int. J. Environ. Res. Public Health 2022, 19, 16898.
  4. Prasad, R.K.; Chatterjee, S.; Mazumder, P.B.; Gupta, S.K.; Sharma, S.; Vairale, M.G.; Datta, S.; Dwivedi, S.K.; Gupta, D.K. Bioethanol production from waste lignocelluloses: A review on microbial degradation potential. Chemosphere 2019, 231, 588–606.
  5. Sharma, G.; Kaur, M.; Punj, S.; Singh, K. Biomass as a sustainable resource for value-added modern materials: A review. Biofuels Bioprod. Biorefin. 2020, 14, 673–695.
  6. Islam, K.; Wang, H.; Rehman, S.; Dong, C.; Hsu, H.-Y.; Lin, C.S.K.; Leu, S.-Y. Sustainability metrics of pretreatment processes in a waste derived lignocellulosic biomass biorefinery. Bioresour. Technol. 2020, 298, 122558.
  7. Renewable Fuels Association. 2023 Ready. Set. Go! Renewable Fuels Association: Washington, DC, USA, 2023.
  8. Gonçalves, F.d.O.; Perna, R.F.; Lopes, E.S.; Tovar, L.P.; Filho, R.M.; Lopes, M.S. Strategies to Ensure Fuel Security in Brazil Considering a Forecast of Ethanol Production. Biomass 2023, 3, 1–17.
  9. De Almeida, M.A.; Colombo, R. Production Chain of First-Generation Sugarcane Bioethanol: Characterization and Value-Added Application of Wastes. BioEnergy Res. 2023, 16, 924–939.
  10. Carrillo-Nieves, D.; Alanís, M.J.R.; Quiroz, R.d.l.C.; Ruiz, H.A.; Iqbal, H.M.; Parra-Saldívar, R. Current status and future trends of bioethanol production from agro-industrial wastes in Mexico. Renew. Sustain. Energy Rev. 2019, 102, 63–74.
  11. Santos, C.I.; Silva, C.C.; Mussatto, S.I.; Osseweijer, P.; van der Wielen, L.A.; Posada, J.A. Integrated 1st and 2nd generation sugarcane bio-refinery for jet fuel production in Brazil: Techno-economic and greenhouse gas emissions assessment. Renew. Energy 2018, 129, 733–747.
  12. Aghaei, S.; Alavijeh, M.K.; Shafiei, M.; Karimi, K. A comprehensive review on bioethanol production from corn stover: Worldwide potential, environmental importance, and perspectives. Biomass Bioenergy 2022, 161, 106447.
  13. Broda, M.; Yelle, D.J.; Serwańska, K. Bioethanol Production from Lignocellulosic Biomass—Challenges and Solutions. Molecules 2022, 27, 8717.
  14. Maity, S.; Mallick, N. Trends and advances in sustainable bioethanol production by marine microalgae: A critical review. J. Clean. Prod. 2022, 345, 131153.
  15. Ayodele, B.V.; Alsaffar, M.A.; Mustapa, S.I. An overview of integration opportunities for sustainable bioethanol production from first- and second-generation sugar-based feedstocks. J. Clean. Prod. 2020, 245, 118857.
  16. Zabed, H.; Sahu, J.; Suely, A.; Boyce, A.; Faruq, G. Bioethanol production from renewable sources: Current perspectives and technological progress. Renew. Sustain. Energy Rev. 2017, 71, 475–501.
  17. Duque, A.; Álvarez, C.; Doménech, P.; Manzanares, P.; Moreno, A.D. Advanced Bioethanol Production: From Novel Raw Materials to Integrated Biorefineries. Processes 2021, 9, 206.
  18. Morales, M.; Arvesen, A.; Cherubini, F. Integrated process simulation for bioethanol production: Effects of varying lignocellulosic feedstocks on technical performance. Bioresour. Technol. 2021, 328, 124833.
  19. El Akkari, M.; Ben Fradj, N.; Gabrielle, B.; Djomo, S.N. Spatially-explicit environmental assessment of bioethanol from miscanthus and switchgrass in France. Clean. Circ. Bioecon. 2023, 6, 100059.
  20. Singh, A.; Chen, C.-W.; Patel, A.K.; Dong, C.-D.; Singhania, R.R. Subcritical Water Pretreatment for the Efficient Valorization of Sorghum Distillery Residue for the Biorefinery Platform. Bioengineering 2022, 10, 38.
  21. Patel, A.; Shah, A.R. Integrated lignocellulosic biorefinery: Gateway for production of second generation ethanol and value added products. J. Bioresour. Bioprod. 2021, 6, 108–128.
  22. Mujtaba, M.; Fraceto, L.F.; Fazeli, M.; Mukherjee, S.; Savassa, S.M.; de Medeiros, G.A.; Pereira, A.D.E.S.; Mancini, S.D.; Lipponen, J.; Vilaplana, F. Lignocellulosic biomass from agricultural waste to the circular economy: A review with focus on biofuels, biocomposites and bioplastics. J. Clean. Prod. 2023, 402, 136815.
  23. Mat Aron, N.S.; Khoo, K.S.; Chew, K.W.; Show, P.L.; Chen, W.; Nguyen, T.H.P. Sustainability of the four generations of biofuels—A review. Int. J. Energy Res. 2020, 44, 9266–9282.
  24. Müller, C.; Scapini, T.; Rempel, A.; Abaide, E.R.; Camargo, A.F.; Nazari, M.T.; Tadioto, V.; Bonatto, C.; Tres, M.V.; Zabot, G.L.; et al. Challenges and opportunities for third-generation ethanol production: A critical review. Eng. Microbiol. 2023, 3, 100056.
  25. Anto, S.; Mukherjee, S.S.; Muthappa, R.; Mathimani, T.; Deviram, G.; Kumar, S.S.; Verma, T.N.; Pugazhendhi, A. Algae as green energy reserve: Technological outlook on biofuel production. Chemosphere 2020, 242, 125079.
  26. Jambo, S.A.; Abdulla, R.; Azhar, S.H.M.; Marbawi, H.; Gansau, J.A.; Ravindra, P. A review on third generation bioethanol feedstock. Renew. Sustain. Energy Rev. 2016, 65, 756–769.
  27. Ramachandra, T.; Hebbale, D. Bioethanol from macroalgae: Prospects and challenges. Renew. Sustain. Energy Rev. 2020, 117, 109479.
  28. Sarwer, A.; Hamed, S.M.; Osman, A.I.; Jamil, F.; Al-Muhtaseb, A.H.; Alhajeri, N.S.; Rooney, D.W. Algal biomass valorization for biofuel production and carbon sequestration: A review. Environ. Chem. Lett. 2022, 20, 2797–2851.
  29. Dave, N.; Selvaraj, R.; Varadavenkatesan, T.; Vinayagam, R. A critical review on production of bioethanol from macroalgal biomass. Algal Res. 2019, 42, 101606.
  30. Chowdhury, H.; Loganathan, B. Third-generation biofuels from microalgae: A review. Curr. Opin. Green Sustain. Chem. 2019, 20, 39–44.
  31. Li, S.; Li, X.; Ho, S.-H. Microalgae as a solution of third world energy crisis for biofuels production from wastewater toward carbon neutrality: An updated review. Chemosphere 2022, 291, 132863.
  32. Bhatia, L.; Bachheti, R.K.; Garlapati, V.K.; Chandel, A.K. Third-generation biorefineries: A sustainable platform for food, clean energy, and nutraceuticals production. Biomass Convers. Biorefin. 2022, 12, 4215–4230.
  33. Alalwan, H.A.; Alminshid, A.H.; Aljaafari, H.A. Promising evolution of biofuel generations. Subject review. Renew. Energy Focus 2019, 28, 127–139.
  34. Sarwer, A.; Hussain, M.; Al-Muhtaseb, A.H.; Inayat, A.; Rafiq, S.; Khurram, M.S.; Ul-Haq, N.; Shah, N.S.; Din, A.A.; Ahmad, I.; et al. Suitability of Biofuels Production on Commercial Scale from Various Feedstocks: A Critical Review. ChemBioEng Rev. 2022, 9, 423–441.
  35. Olguin-Maciel, E.; Singh, A.; Chable-Villacis, R.; Tapia-Tussell, R.; Ruiz, H.A. Consolidated Bioprocessing, an Innovative Strategy towards Sustainability for Biofuels Production from Crop Residues: An Overview. Agronomy 2020, 10, 1834.
  36. Mehmood, M.A.; Shahid, A.; Malik, S.; Wang, N.; Javed, M.R.; Haider, M.N.; Verma, P.; Ashraf, M.U.F.; Habib, N.; Syafiuddin, A.; et al. Advances in developing metabolically engineered microbial platforms to produce fourth-generation biofuels and high-value biochemicals. Bioresour. Technol. 2021, 337, 125510.
  37. Osman, A.I.; Qasim, U.; Jamil, F.; Al-Muhtaseb, A.H.; Abu Jrai, A.; Al-Riyami, M.; Al-Maawali, S.; Al-Haj, L.; Al-Hinai, A.; Al-Abri, M.; et al. Bioethanol and biodiesel: Bibliometric mapping, policies and future needs. Renew. Sustain. Energy Rev. 2021, 152, 111677.
  38. Abdullah, B.; Muhammad, S.A.F.S.; Shokravi, Z.; Ismail, S.; Kassim, K.A.; Mahmood, A.N.; Aziz, M.A. Fourth generation biofuel: A review on risks and mitigation strategies. Renew. Sustain. Energy Rev. 2019, 107, 37–50.
  39. Dietrich, K.; Dumont, M.-J.; Del Rio, L.F.; Orsat, V. Sustainable PHA production in integrated lignocellulose biorefineries. New Biotechnol. 2019, 49, 161–168.
  40. Azeez, M.A.; Orege, J.I. Bamboo, Its Chemical Modification and Products. In Bamboo—Current and Future Prospects; IntechOpen: London, UK, 2018.
  41. Gabhane, J.; William, S.P.; Gadhe, A.; Rath, R.; Vaidya, A.N.; Wate, S. Pretreatment of banana agricultural waste for bio-ethanol production: Individual and interactive effects of acid and alkali pretreatments with autoclaving, microwave heating and ultrasonication. Waste Manag. 2014, 34, 498–503.
  42. Díaz, M.J.; Moya, M.; Castro, E. Bioethanol Production from Steam-Exploded Barley Straw by Co-Fermentation with Escherichia coli SL100. Agronomy 2022, 12, 874.
  43. Nigam, P.S.N.; Gupta, N.; Anthwal, A. Pre-treatment of Agro-Industrial Residues. In Biotechnology for Agro-Industrial Residues Utilisation; Springer: Dordrecht, The Netherlands, 2009; pp. 13–33.
  44. Saini, J.K.; Saini, R.; Tewari, L. Lignocellulosic agriculture wastes as biomass feedstocks for second-generation bioethanol production: Concepts and recent developments. 3 Biotech 2015, 5, 337–353.
  45. Soni, S.K.; Sharma, A.; Soni, R. Microbial Enzyme Systems in the Production of Second Generation Bioethanol. Sustainability 2023, 15, 3590.
  46. Chandra, R.; Takeuchi, H.; Hasegawa, T. Methane production from lignocellulosic agricultural crop wastes: A review in context to second generation of biofuel production. Renew. Sustain. Energy Rev. 2012, 16, 1462–1476.
  47. Kim, M.; Day, D.F. Composition of sugar cane, energy cane, and sweet sorghum suitable for ethanol production at Louisiana sugar mills. J. Ind. Microbiol. Biotechnol. 2011, 38, 803–807.
  48. Xu, P.; Cheng, S.; Han, Y.; Zhao, D.; Li, H.; Wang, Y.; Zhang, G.; Chen, C. Natural Variation of Lignocellulosic Components in Miscanthus Biomass in China. Front. Chem. 2020, 8, 595143.
  49. Rajendran, K.; Drielak, E.; Varma, V.S.; Muthusamy, S.; Kumar, G. Updates on the pretreatment of lignocellulosic feedstocks for bioenergy production—A review. Biomass Convers. Biorefin. 2018, 8, 471–483.
  50. Du, B.; Sharma, L.N.; Becker, C.; Chen, S.; Mowery, R.A.; van Walsum, G.P.; Chambliss, C.K. Effect of varying feedstock–pretreatment chemistry combinations on the formation and accumulation of potentially inhibitory degradation products in biomass hydrolysates. Biotechnol. Bioeng. 2010, 107, 430–440.
  51. García-Cubero, M.T.; González-Benito, G.; Indacoechea, I.; Coca, M.; Bolado, S. Effect of ozonolysis pretreatment on enzymatic digestibility of wheat and rye straw. Bioresour. Technol. 2009, 100, 1608–1613.
  52. Monlau, F.; Barakat, A.; Steyer, J.P.; Carrere, H. Comparison of seven types of thermo-chemical pretreatments on the structural features and anaerobic digestion of sunflower stalks. Bioresour. Technol. 2012, 120, 241–247.
  53. Brigham, C. Biopolymers. In Green Chemistry; Elsevier: Amsterdam, The Netherlands, 2018; pp. 753–770.
  54. Damayanti, D.; Supriyadi, D.; Amelia, D.; Saputri, D.R.; Devi, Y.L.L.; Auriyani, W.A.; Wu, H.S. Conversion of Lignocellulose for Bioethanol Production, Applied in Bio-Polyethylene Terephthalate. Polymers 2021, 13, 2886.
  55. Frassoldati, A.; Ranzi, E. Modeling of Thermochemical Conversion of Biomasses. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2019.
  56. Hokkanen, S.; Bhatnagar, A.; Sillanpää, M. A review on modification methods to cellulose-based adsorbents to improve adsorption capacity. Water Res. 2016, 91, 156–173.
  57. Banu, J.R.; Preethi; Kavitha, S.; Tyagi, V.K.; Gunasekaran, M.; Karthikeyan, O.P.; Kumar, G. Lignocellulosic biomass based biorefinery: A successful platform towards circular bioeconomy. Fuel 2021, 302, 121086.
  58. Zhou, X.; Li, W.; Mabon, R.; Broadbelt, L.J. A Critical Review on Hemicellulose Pyrolysis. Energy Technol. 2017, 5, 52–79.
  59. Bajpai, P. Wood and Fiber Fundamentals. In Biermann’s Handbook of Pulp and Paper; Elsevier: Amsterdam, The Netherlands, 2018; pp. 19–74.
  60. Malik, K.; Sharma, P.; Yang, Y.; Zhang, P.; Zhang, L.; Xing, X.; Yue, J.; Song, Z.; Nan, L.; Yujun, S.; et al. Lignocellulosic biomass for bioethanol: Insight into the advanced pretreatment and fermentation approaches. Ind. Crops Prod. 2022, 188, 115569.
  61. Sharma, S.; Kumar, A. Lignin; Springer International Publishing: Cham, Switzerland, 2020.
  62. Akpan, E.I. Chemistry and Structure of Lignin. In Sustainable Lignin for Carbon Fibers: Principles, Techniques, and Applications; Springer International Publishing: Cham, Switzerland, 2019; pp. 1–50.
  63. Lu, Y.; Lu, Y.-C.; Hu, H.-Q.; Xie, F.-J.; Wei, X.-Y.; Fan, X. Structural Characterization of Lignin and Its Degradation Products with Spectroscopic Methods. J. Spectrosc. 2017, 2017, 8951658.
  64. Hou, J.; Tang, J.; Chen, J.; Zhang, Q. Quantitative Structure-Toxicity Relationship analysis of combined toxic effects of lignocellulose-derived inhibitors on bioethanol production. Bioresour. Technol. 2019, 289, 121724.
  65. Kordala, N.; Lewandowska, M.; Bednarski, W. Effect of the method for the elimination of inhibitors present in Miscanthus giganteus hydrolysates on ethanol production effectiveness. Biomass Convers. Biorefin. 2023, 13, 2089–2097.
  66. Kim, D. Physico-Chemical Conversion of Lignocellulose: Inhibitor Effects and Detoxification Strategies: A Mini Review. Molecules 2018, 23, 309.
  67. Jeswani, H.K.; Chilvers, A.; Azapagic, A. Environmental sustainability of biofuels: A review. Proc. R Soc. A Math. Phys. Eng. Sci. 2020, 476, 20200351.
  68. Rezania, S.; Oryani, B.; Cho, J.; Talaiekhozani, A.; Sabbagh, F.; Hashemi, B.; Rupani, P.F.; Mohammadi, A.A. Different pretreatment technologies of lignocellulosic biomass for bioethanol production: An overview. Energy 2020, 199, 117457.
  69. Das, N.; Jena, P.K.; Padhi, D.; Mohanty, M.K.; Sahoo, G. A comprehensive review of characterization, pretreatment and its applications on different lignocellulosic biomass for bioethanol production. Biomass Convers. Biorefin. 2023, 13, 1503–1527.
  70. Hideno, A.; Inoue, H.; Tsukahara, K.; Fujimoto, S.; Minowa, T.; Inoue, S.; Endo, T.; Sawayama, S. Wet disk milling pretreatment without sulfuric acid for enzymatic hydrolysis of rice straw. Bioresour. Technol. 2009, 100, 2706–2711.
  71. Zakaria, M.R.; Fujimoto, S.; Hirata, S.; Hassan, M.A. Ball Milling Pretreatment of Oil Palm Biomass for Enhancing Enzymatic Hydrolysis. Appl. Biochem. Biotechnol. 2014, 173, 1778–1789.
  72. Liu, Y.; Chen, W.; Xia, Q.; Guo, B.; Wang, Q.; Liu, S.; Liu, Y.; Li, J.; Yu, H. Efficient Cleavage of Lignin–Carbohydrate Complexes and Ultrafast Extraction of Lignin Oligomers from Wood Biomass by Microwave-Assisted Treatment with Deep Eutectic Solvent. ChemSusChem 2017, 10, 1692–1700.
  73. John, I.; Pola, J.; Appusamy, A. Optimization of Ultrasonic Assisted Saccharification of Sweet Lime Peel for Bioethanol Production Using Box–Behnken Method. Waste Biomass Valorization 2019, 10, 441–453.
  74. Bonfiglio, F.; Cagno, M.; Rey, F.; Torres, M.; Böthig, S.; Menéndez, P.; Mussatto, S.I. Pretreatment of switchgrass by steam explosion in a semi-continuous pre-pilot reactor. Biomass Bioenergy 2019, 121, 41–47.
  75. De Carvalho, D.M.; de Queiroz, J.H.; Colodette, J.L. Assessment of alkaline pretreatment for the production of bioethanol from eucalyptus, sugarcane bagasse and sugarcane straw. Ind. Crops Prod. 2016, 94, 932–941.
  76. Hilares, R.T.; Swerts, M.P.; Ahmed, M.A.; Ramos, L.; da Silva, S.S.; Santos, J.C. Organosolv Pretreatment of Sugar Cane Bagasse for Bioethanol Production. Ind. Eng. Chem. Res. 2017, 56, 3833–3838.
  77. Hossain, A.; Rahaman, M.S.; Yelle, D.; Shang, H.; Sun, Z.; Renneckar, S.; Dong, J.; Tulaphol, S.; Sathitsuksanoh, N. Effects of polyol-based deep eutectic solvents on the efficiency of rice straw enzymatic hydrolysis. Ind. Crops Prod. 2021, 167, 113480.
  78. Singh, A.; Bishnoi, N.R. Enzymatic hydrolysis optimization of microwave alkali pretreated wheat straw and ethanol production by yeast. Bioresour. Technol. 2012, 108, 94–101.
  79. Ziaei-Rad, Z.; Pazouki, M.; Fooladi, J.; Azin, M.; Gummadi, S.N.; Allahverdi, A. Investigation of a robust pretreatment technique based on ultrasound-assisted, cost-effective ionic liquid for enhancing saccharification and bioethanol production from wheat straw. Sci. Rep. 2023, 13, 446.
  80. Yang, J.; Lu, X.; Zhang, Y.; Xu, J.; Yang, Y.; Zhou, Q. A facile ionic liquid approach to prepare cellulose fiber with good mechanical properties directly from corn stalks. Green Energy Environ. 2020, 5, 223–231.
  81. Padilha, C.E.d.A.; Nogueira, C.d.C.; Souza, D.F.d.S.; de Oliveira, J.A.; dos Santos, E.S. Valorization of green coconut fibre: Use of the black liquor of organolsolv pretreatment for ethanol production and the washing water for production of rhamnolipids by Pseudomonas aeruginosa ATCC 27583. Ind. Crops Prod. 2019, 140, 111604.
  82. Martínez-Patiño, J.C.; Ruiz, E.; Romero, I.; Cara, C.; López-Linares, J.C.; Castro, E. Combined acid/alkaline-peroxide pretreatment of olive tree biomass for bioethanol production. Bioresour. Technol. 2017, 239, 326–335.
  83. Kim, T.H.; Lee, Y.Y.; Sunwoo, C.; Kim, J.S. Pretreatment of Corn Stover by Low-Liquid Ammonia Recycle Percolation Process. Appl. Biochem. Biotechnol. 2006, 133, 41–58.
  84. Turner, W.; Greetham, D.; Mos, M.; Squance, M.; Kam, J.; Du, C. Exploring the Bioethanol Production Potential of Miscanthus Cultivars. Appl. Sci. 2021, 11, 9949.
  85. Kumar, M.N.; Ravikumar, R.; Sankar, M.K.; Thenmozhi, S. New insight into the effect of fungal mycelia present in the bio-pretreated paddy straw on their enzymatic saccharification and optimization of process parameters. Bioresour. Technol. 2018, 267, 291–302.
  86. Lee, O.K.; Oh, Y.-K.; Lee, E.Y. Bioethanol production from carbohydrate-enriched residual biomass obtained after lipid extraction of Chlorella sp. KR-1. Bioresour. Technol. 2015, 196, 22–27.
  87. Guo, H.; Daroch, M.; Liu, L.; Qiu, G.; Geng, S.; Wang, G. Biochemical features and bioethanol production of microalgae from coastal waters of Pearl River Delta. Bioresour. Technol. 2013, 127, 422–428.
  88. Kumar, V.; Yadav, S.K.; Kumar, J.; Ahluwalia, V. A critical review on current strategies and trends employed for removal of inhibitors and toxic materials generated during biomass pretreatment. Bioresour. Technol. 2020, 299, 122633.
  89. Zhao, J.; Lee, J.; Wang, D. A Critical Review on Water Overconsumption in Lignocellulosic Biomass Pretreatment for Ethanol Production through Enzymic Hydrolysis and Fermentation. Energy Fuels 2023, 37, 2667–2680.
  90. Ranjithkumar, M.; Uthandi, S.; Kumar, P.S.; Muniraj, I.; Thanabal, V.; Rajarathinam, R. Highly crystalline cotton spinning wastes utilization: Pretreatment, optimized hydrolysis and fermentation using Pleurotus florida for bioethanol production. Fuel 2022, 308, 122052.
  91. Li, F.; Liu, Y.; Jia, J.; Yu, H. Combination of biological pretreatment with deep eutectic solvent pretreatment for enhanced enzymatic saccharification of Pinus massoniana. Bioresour. Technol. 2023, 380, 129110.
  92. Sun, E.; Zhang, Y.; Yong, C.; Qu, P.; Huang, H.; Xu, Y. Biological fermentation pretreatment accelerated the depolymerization of straw fiber and its mechanical properties as raw material for mulch film. J. Clean. Prod. 2021, 284, 124688.
  93. Uzuner, S.; Cekmecelioglu, D. Enzymes in the Beverage Industry. In Enzymes in Food Biotechnology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 29–43.
  94. Suganuma, S.; Nakajima, K.; Kitano, M.; Yamaguchi, D.; Kato, H.; Hayashi, S.; Hara, M. Hydrolysis of Cellulose by Amorphous Carbon Bearing SO3H, COOH, and OH Groups. J. Am. Chem. Soc. 2008, 130, 12787–12793.
  95. Chen, H. Lignocellulose biorefinery feedstock engineering. In Lignocellulose Biorefinery Engineering, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 37–86.
  96. Wada, M.; Ike, M.; Tokuyasu, K. Enzymatic hydrolysis of cellulose I is greatly accelerated via its conversion to the cellulose II hydrate form. Polym. Degrad. Stab. 2010, 95, 543–548.
  97. Vasić, K.; Knez, Ž.K.; Leitgeb, M. Bioethanol Production by Enzymatic Hydrolysis from Different Lignocellulosic Sources. Molecules 2021, 26, 753.
  98. Deshmukh, M.; Pande, A.; Marathe, A. Different particle size study of castor deoiled cake for biofuel production with an environmental sustainability perspective. Heliyon 2022, 8, e09710.
  99. Kapoor, M.; Semwal, S.; Satlewal, A.; Christopher, J.; Gupta, R.P.; Kumar, R.; Puri, S.K.; Ramakumar, S. The impact of particle size of cellulosic residue and solid loadings on enzymatic hydrolysis with a mass balance. Fuel 2019, 245, 514–520.
  100. Gao, W.; Li, Z.; Liu, T.; Wang, Y. Production of high-concentration fermentable sugars from lignocellulosic biomass by using high solids fed-batch enzymatic hydrolysis. Biochem. Eng. J. 2021, 176, 108186.
  101. Modenbach, A.A.; Nokes, S.E. Enzymatic hydrolysis of biomass at high-solids loadings—A review. Biomass Bioenergy 2013, 56, 526–544.
  102. Amândio, M.S.T.; Rocha, J.M.S.; Xavier, A.M.R.B. Enzymatic Hydrolysis Strategies for Cellulosic Sugars Production to Obtain Bioethanol from Eucalyptus globulus Bark. Fermentation 2023, 9, 241.
  103. Constantino, A.; Rodrigues, B.; Leon, R.; Barros, R.; Raposo, S. Alternative chemo-enzymatic hydrolysis strategy applied to different microalgae species for bioethanol production. Algal Res. 2021, 56, 102329.
  104. Wang, W.; Zhang, C.; Tong, S.; Cui, Z.; Liu, P. Enhanced Enzymatic Hydrolysis and Structural Features of Corn Stover by NaOH and Ozone Combined Pretreatment. Molecules 2018, 23, 1300.
  105. Ostadjoo, S.; Hammerer, F.; Dietrich, K.; Dumont, M.-J.; Friščić, T.; Auclair, K. Efficient Enzymatic Hydrolysis of Biomass Hemicellulose in the Absence of Bulk Water. Molecules 2019, 24, 4206.
  106. Godbey, W.T. Fermentation, Beer, and Biofuels. In An Introduction to Biotechnology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 331–351.
  107. Spector, M.P. Metabolism, Central (Intermediary). In Encyclopedia of Microbiology; Elsevier: Amsterdam, The Netherlands, 2009; pp. 242–264.
More
Information
Subjects: Energy & Fuels
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , , , , ,
View Times: 99
Revisions: 2 times (View History)
Update Date: 26 Jan 2024
1000/1000
Video Production Service