Next Article in Journal
Evaluation, Driving Mechanism and Spatial Correlation Analysis of Atmospheric Environmental Efficiency in the “2+26” Cities Based on the Nonradial MEA Model
Next Article in Special Issue
Sustainable Co-Production of Xylanase, Cellulase, and Pectinase through Agroindustrial Residue Valorization Using Solid-State Fermentation: A Techno-Economic Assessment
Previous Article in Journal
Navigating the Challenges of Environmental, Social, and Governance (ESG) Reporting: The Path to Broader Sustainable Development
Previous Article in Special Issue
Broccoli Leaves (Brassica oleracea var. italica) as a Source of Bioactive Compounds and Chemical Building Blocks: Optimal Extraction Using Dynamic Maceration and Life Cycle Assessment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Perspectives and Progress in Bioethanol Processing and Social Economic Impacts

by
Mario Alberto Yaverino-Gutiérrez
1,
Alán Yazid Chávez-Hita Wong
1,
Lizbeth Alejandra Ibarra-Muñoz
1,
Ana Cristina Figueroa Chávez
1,
Jazel Doménica Sosa-Martínez
1,
Ana Sofia Tagle-Pedroza
1,
Javier Ulises Hernández-Beltran
1,
Salvador Sánchez-Muñoz
2,
Julio César dos Santos
2,
Silvio Silvério da Silva
2,* and
Nagamani Balagurusamy
1,*
1
Bioremediation Laboratory, Faculty of Biological Sciences, Autonomous University of Coahuila, Torreón-Matamoros Highways km 7.5, Torreón 27276, Mexico
2
Bioprocesses and Sustainable Products Laboratory, Department of Biotechnology, Engineering School of Lorena, University of São Paulo (EEL-USP), Lorena 12602-810, Brazil
*
Authors to whom correspondence should be addressed.
Sustainability 2024, 16(2), 608; https://doi.org/10.3390/su16020608
Submission received: 30 November 2023 / Revised: 26 December 2023 / Accepted: 28 December 2023 / Published: 10 January 2024

Abstract

:
The liquid biofuel bioethanol is widely produced worldwide via fermenting sugars extracted from a variety of raw materials, including lignocellulose biomass, one of the world’s most abundant renewable resources. Due to its recalcitrant character, lignocellulose is usually pretreated by mechanical, chemical, and biological methods to maximize sugar recovery. Pretreated lignocellulose biomass undergoes a fermentation process performed sequentially or simultaneously to saccharification. The different fermentation strategies (e.g., separate or simultaneous hydrolysis and fermentation or co-fermentation) and conditions (e.g., inoculum type load, agitation, temperature, and pH) affect ethanol yield. Genetic modification of the inoculum has been focused recently to improve ethanol tolerance and as well as to use different sugars to enhance the performance of the microorganisms involved in fermentation. Nonetheless, these improvements result in a substantial increase in costs and have certain environmental costs. This review offers an overview of advancements in bioethanol production, with a primary focus on lignocellulosic feedstock, while also considering other feedstocks. Furthermore, it provides insights into the economic, social, and environmental impacts associated with bioethanol production.

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.
This review article deals with the recent advancements in different processes involved in the production of second-generation bioethanol such as pretreatment steps for LCB, process operating conditions, and genetic approaches to improve the ethanol yield. Additionally, this review presents a summary of the socio-economic and environmental impacts of bioethanol production based on the analyses reported in the literature.

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.
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.
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.

4.1. Stress Factors and Inhibitors Affecting Ethanol Yield

Most of the factors that directly affect ethanol production are associated with the optimal growing conditions of the strain, which depends on the yeast or microbe used for fermentation. S. cerevisiae growth conditions may vary from 30 to 40 °C and pH 4 to 6. Disturbing these conditions lowers ethanol production by causing stress in microorganisms or inhibiting their metabolism [108].
Osmotic stress, heat stress, and chaotropic stress are the main stress factors affecting ethanol yield during fermentation. Osmotic stress occurs due to the high concentration of fermentable sugars or ethanol present in the medium. Concentrations of 150 mg of sugars/L have been reported for the Crabtree effect [83]. Osmotic stress also results in the formation of reactive oxygen species, resulting in oxidative stress by altering the redox state of key fermenting molecules. Heat stress occurs in most organisms and is characterized by denaturalization of proteins, nucleic acids, and membrane phospholipids [109].
Chaperone proteins help to fold both nascent proteins and unfolded proteins that were denatured due to heat stress [109]. Another consequence of high temperatures in the reactor is an increase in rigidity in yeasts’ cell membrane. Yeast starts replacing the short and unsaturated fatty acid chains in phospholipids with long and less saturated fatty acid chains, which have higher heat tolerance. Ergosterol is the predominant sterol found in yeast membranes; it oversees maintaining membrane stability [110]. Chaotropes are chemical compounds such as n-butanol or ethanol that promote disorder in macromolecules by disrupting the hydrogen bonds between them, inhibiting their role. Other factors that directly affect yeast fermentation are induced by the previously mentioned stress factors. Oxidative stress is the result of both osmotic and chaotropic stress by the production of reactive oxygen species [111]. Chaotropic stress occurs due to the accumulation of chaotropes which are compounds that disorder proteins and nucleic acids as well as disrupt the cell membrane. Ethanol itself is considered a chaotrope; thus, a high concentration of it inside the reactor results in chaotropic stress [112].
One of the key factors for ethanol production is the substrate concentration. Theoretically, a higher substrate concentration will result in higher production of ethanol. However, practically, increasing the number of substrate derivatives inhibits ethanol production. Wild-type strains of S. cerevisiae tolerate around 150 g/L of sugars. Going over this concentration starts inhibiting the fermentation process [112]. Ethanol is an antimicrobial compound that can diffuse through the cell membrane and directly reduce the activity of enzymes that act in glycolysis or even denature them [111].
Also, the presence of ethanol lowers water activity. A concentration of 20% (w/v) reduces water activity by 5% and reduces intracellular pH. Some inhibitors are carried in the liquid fraction after chemical hydrolysis such as furfural. The presence of 1 g/L of furfural lowers the activity of different growth enzymes, and dehydrogenases are found to be more sensitive to furfurals. The formation of carboxylic acids occurs by the deacetylation of hemicellulose. It decreases the yeast’s pH, causing acidic stress. Lignin degradation releases a wide range of phenolic compounds that commonly inhibit microbial growth by directly affecting membrane activity. Phenolic compounds seem to affect the rate of ethanol production but not the overall ethanol yield [112].

4.2. Different Setups and Optimal Conditions of the Fermentation Process

Selecting a method for bioethanol production comes with advantages and disadvantages depending on the configuration chosen. In this discussion, various fermentation setups and their respective pros and cons will be presented. Liquid-state fermentation (LSF), also known as submerged fermentation, is the most common way to obtain bioethanol. During LSF, the substrate is solubilized in a liquid medium where the microbes will grow to perform fermentation. Advantages of submerged fermentation include high homogeneity and short periods of fermentation. Nonetheless, LSF has several disadvantages such as high overall energy and water usage, high waste generation, high cost for culture medium, and a large space requirement for bioreactors [113].
Solid-state fermentation (SSF) utilizes solid wastes as a substrate to obtain various value-added products, including ethanol. SSF has been shown to offer high substrate concentrations, resulting in a high ethanol yield. The SSF process has no free water or low concentrations of it inside the reactor. SSF is paired with simultaneous saccharification, where enzymes are added to solid pretreated biomass to release the sugars inside it [114]. The enzymes act on the surface and then go down to the bottom of the biomass, hydrolyzing it.
The resulting solid-state hydrolysate acts as the same natural environment of different filamentous fungi, such as Trichoderma and Aspergillus. However, SSF allows the growth of single-cell fermentative organisms like yeast or bacteria. The most important condition in solid-state fermentation is moisture; high moisture promotes agglomeration of particles, reducing heat dissipation and gas transfer. On the other hand, low moisture decreases substrate solubility, thereby reducing substrate availability [113]. Different advantages of SSF include the reduction in catabolite repression and substrate inhibition, low energy consumption, higher ethanol concentration, no discharge of wastewater, and better enzymatic reaction with the substrate. Nonetheless, to date, solid-state fermentation still faces several challenges to be implemented on an industrial level, such as controlling pH and temperature, uneven distribution of enzymes and microorganisms, varying moisture content inside the biomass, difficulty in the purification process, long fermentation time, and limited availability of bioreactors [115].
Both SSF and LSF can be utilized synergistically. Chu et al. [115] used biomass from corn stover to produce ethanol. They centrifuged the hydrolysate from enzymatic hydrolysis and fermented the supernatant in a submerged fermentation, as well as the pellet using solid-state fermentation. The ethanol yield for LSF was 94.98%, with a total of 91.31 g/L of ethanol from 216.7 g/L of sugars. On the other hand, SSF was performed on hydrolysate residue containing 5.89% (w/w) of cellulose. After 24 h, cellulose was reduced to 3.78% (w/w), indicating simultaneous saccharification and fermentation occurred during the process. This resulted in the production of 1.51 g of glucose after 64 h, presenting an ethanol yield of 0.195 g/g.
The concentration of sugar for ethanol production can be categorized as normal gravity (<180 g/L of sugars), high gravity (180–240 g/L of sugars), and very high gravity (VHG) (>240 g/L of sugars). VHG fermentation can produce more than 15% (v/v) of ethanol, in contrast to the 10–12% (v/v) found in most processes. During VHG, microorganisms are put under several stress due to the increase in nutrients, metals, and sodium ions, as well as elevated temperatures, pH, osmotic stress, and ethanol concentration. After VHG, yeast produces other by-products such as trehalose, glycogen, succinic acid, and glycerol, which function as coping mechanisms for the stress factors. Various yeast strains have been shown to tolerate high ethanol concentrations without requiring adaptation or genetic modification [116]. Cruz et al. [117] inoculated an industrial strain of S. cerevisiae Y-904 into a broth containing 300 g/L of sugars for up to 30 h and reported an ethanol yield of 90% with a final concentration of 135 g/L of ethanol.
In addition to strain robustness, temperature and nutrient supplementation are crucial factors that prevent the implementation of VHG fermentation in industries. High temperatures directly affect the stress-coping mechanisms of yeast against ethanol tolerance. A decrease in temperature from 30 to 27 °C during VHG fermentation has been reported to increase the ethanol yield by 2.1%. Nutrient supplementation involves the addition of key components needed by cells besides the carbon source. Increasing substrate concentration often results in an imbalance of other nutrients, which lowers microbial growth and therefore reduces ethanol yield. Another drawback of VHG fermentation is the requirement of molasses addition to the medium, and the use of molasses in VHG competes with the food industry [116].
Optimizing both the enzymatic hydrolysis and fermentation processes results in increased ethanol yield. To achieve this, different parameters need to be considered. Going over or under a pH of 4.5–5.5 and a temperature of 45–50 °C can negatively affect enzymatic hydrolysis. Stoichiometrically, a molecule of glucose results in two molecules of ethanol; thus, 1 g of glucose produces 0.511 g of ethanol. Not all sugars are transformed by microorganisms during fermentation; around 10% are utilized for new biomass generation and other metabolic needs. Nevertheless, there are no fixed ideal conditions for cellulases to achieve maximum efficiency [118].
Optimization of enzymatic hydrolysis and fermentation involves the use of statistical analysis to determine the ideal parameters that approach 0.511 g of ethanol in a bioprocess. Response surface methodology (RSM) is the most reported analysis for optimization. RSM involves factorial analysis from the results of different factorial levels, with three-level analysis being the most efficient [119]. For instance, Assabjeu et al. [120] optimized the hydrolysis of pretreated sawdust from Triplochiton scleroxylon using RSM. The pretreated biomass contained 53.2 ± 0.3% cellulose. The researchers found that loading 9.07% of the substrate and using 21.36 FPU/gDM of the enzyme for 72 h were the most efficient parameters for ethanol fermentation, resulting in an ethanol yield of 21.88 mg/mL.
Another study aimed to optimize the fermentation of microalgae biomass for ethanol production. El-Mekkawi et al. [121] found that the best ethanol yield for the microalgal biomass was 18.57 g/L. To achieve this, they supplied 98.7 g/L of microalgal biomass to 15.09% of immobilized yeast for 43.6 h. The disadvantage of using response surface methodology is that a large number of experiments are required to find the optimal parameters. Microfluid reactions (MFRs) have recently been studied as a solution to this issue. MFRs simulate bioreactor conditions under microliter scales, allowing for a large screening of different conditions and the ability to scale up the results [122]. Beltrán et al. [123] used MFRs to test the impact of particle size during enzymatic hydrolysis. To optimize the conditions of the hydrolysis, they tested three levels of pH (4.5, 5, and 5.5) and temperature (45, 50, and 55 °C) and four levels of enzyme loading (1.0, 3.0, 5.0, and 7.0 μL/0.015 g-biomass), resulting in a total of 108 experiments. They concluded that the optimal conditions for the hydrolysis were an enzyme loading of 7 μL/0.015 g-biomass, a pH of 5.5, and a temperature of 46.31 °C.
Consolidated bioprocessing refers to the integration of various steps in bioprocessing, including the production of hydrolytic enzymes, saccharification of biomass, and fermentation of five- and six-carbon sugars. While there is no single microorganism capable of performing all these bioprocess steps, both fungi and bacteria have useful properties that can be exploited. Clostridium bacteria possess a highly efficient enzymatic complex capable of hydrolyzing cellulose, while fungi such as Fusarium and Trichoderma have demonstrated a large complex of cellulolytic enzymes [124]. However, the anaerobic microorganisms mentioned above also produce several by-products that directly reduce ethanol yield [125]. To date, there is no naturally viable method for consolidating bioethanol production. Nevertheless, there are different configurations that reduce the number of steps in the process.
One such configuration is simultaneous hydrolysis and fermentation (SHF), which involves the integration of biological saccharification and the fermentation process. This allows for the fermentation of sugar monomers by hexose-fermenting microorganisms (HFMs) as soon as cellulose is transformed, resulting in reduced production time and less material used for the bioprocess. However, optimizing this process can be challenging since the cellulase complex and fermenting microorganisms have different parameters where they work best. High efficiency thermo-tolerant strains must be used to achieve both enzyme and strain optimum parameters [114].
Another configuration used to lower the cost of ethanol production and utilize all sugars produced during pretreatment is co-fermentation (CF). Co-fermentation involves the use of different strains in the same medium that can ferment both hexose and pentose. This method is often combined with simultaneous hydrolysis to maintain a constant production of pentose and hexose sugars. Hexose fermentation is performed first in the bioreactor, and then ethanol is separated from the medium. The medium without ethanol is then inoculated with PFMs to ferment five-carbon sugars. By adding both hexose- and pentose-fermenting microorganisms, ethanol production increases and substrates are better leveraged, which speeds up the process and results in fewer reactors used for saccharification and both fermentations [114].
There are two major obstacles to co-fermentation. The first limiting factor is the same as that for simultaneous hydrolysis and fermentation, where the optimal conditions for different microorganisms and enzymes are not the same, resulting in less efficient ethanol yield. The second obstacle is the difference in microbial growth between pentose-utilizing microorganisms and hexose-utilizing microorganisms; HFMs grow considerably faster than PFMs and utilize most of the nutrients in the medium. A solution that has been widely studied is the implementation of a strain that can ferment both five- and six-carbon sugars. However, this solution does not address the problem of different optimal points, but it does help to improve microbial growth. Nonetheless, microorganisms capable of fermenting both pentoses and hexoses use five-carbon sugars when a six-carbon one is not present in the medium. This presents a disadvantage for continuous fermentations where ethanol production increases the time needed to produce it [126].
Response surface methodology (RSM) is widely employed to identify the optimal parameters for co-fermentation and simultaneous saccharification and co-fermentation process. Based on RSM, Derman et al. [127] employed the simultaneous saccharification and fermentation process of empty fruit bunches (EFBs) by employing a microbial consortium consisting of Trichoderma harzianum with S. cerevisiae. The highest bioethanol yield of 0.29 g/g of EFB was achieved with 4% (w/v) of biomass at 30 °C for 72 h. The parameters for ethanol production in different fermentation processes are summarized in Table 3.

4.3. Downstream Process: Ethanol Upgrading

The final crucial step in ethanol production is downstream processing. Ethanol recovery from the fermentation broth involves three different steps: solid-liquid separation, primary purification, and fine purification. While centrifugation is a well-established technique for solid–liquid separation, other techniques such as membrane filtration can also be employed. In membrane filtration, the broth is pressurized through a membrane with small pores, allowing solubilized compounds to pass through while retaining bigger molecules like biomass. However, membrane technologies have a disadvantage compared to centrifugation as the membranes need to be changed after several uses and require frequent maintenance, adding extra costs to the process [136]. For primary purification, the aqueous part of the centrifugation is distilled using two columns: a crude column that removes non-volatile molecules and most of the water and a polishing column that produces an azeotropic mixture of ethanol and remaining water. Fine purification separates the ethanol and water azeotrope, and this can be achieved using molecular sieves, which have fine porosity, allowing ethanol molecules to pass through but blocking the flow of water molecules. Another method to purify the azeotrope is by pressure-swing distillation, where the polishing column operates under high vacuum conditions to separate the mixture between water and ethanol. The azeotrope point of water and ethanol can also be displaced by adding an extra compound like toluene or cyclohexane, allowing the mixture to be traditionally distilled [137].

5. Genetic Engineering Approaches to Improve Bioethanol Yield

To ensure efficient bioethanol production, the selection of microorganisms that can grow and develop under stressful conditions is a crucial step. These microorganisms must have good tolerance to ethanol and be capable of utilizing a variety of sugar-rich substrates to obtain a high yield of biofuel [138]. Normally, S. cerevisiae only ferments glucose to ethanol and produces carbon dioxide by repressing unnecessary metabolic pathways [139,140]. Although, the fermentation process is well understood for metabolizing glucose as the main substrate. However, when other sugars such as xylose, a pentose sugar derived from lignocellulosic materials, are present, ethanol production can become challenging. Therefore, it is essential to consider some aspects of ethanol production from lignocellulosic materials, as the bioconversion process may result in a high release of xylose [141]. Moreover, xylose can only be naturally fermented by other yeasts, such as Scheffersomyces stipitis and Spathaspora passalidarum, for conversion to ethanol [140]. Hence, genetic engineering of microorganisms is a promising approach for improving bioethanol production, as it can reduce raw material usage and associated costs. Several strains of S. cerevisiae have been genetically modified to convert xylose to ethanol, resulting in improved metabolic flux for efficient utilization of this substrate (see Table 4).
In recent years, studies have been carried out to construct recombinant S. cerevisiae that use xylose transporters through the XR-XDH (xylose reductase-xylitol dehydrogenase) or XI (xylose isomerase), pentose phosphate, and glycolysis pathways to produce pyruvate, which is then converted to ethanol with the introduction of Xyl (xyloglucokinase) and PPP (pentose phosphate pathway) genes [142,143]. It has been reported that in the diploid wild-type strain S. cerevisiae BSIF, two heterologous genes were integrated into the genome, the MGT05196N360F mutant of Meyerozyma guilliermondii, which encodes a xylose-specific and glucose-insensitive transporter, and Ru-xylA, which encodes a xylose isomerase (XI), resulting in the final strain S. cerevisiae LF1, with excellent fermentation of xylose as the sole carbon source and an ethanol yield of 0.475 g−1 in 12 h, which is more than 93% of the theoretical yield [144]. Demeke et al. [143] studied the robust strain S. cerevisiae GS1.11-26, in which an expression cassette containing 13 genes, including Clostridium phytofermentans XylA, encoding D-xylose isomerase (XI), and enzymes of the pentose phosphate pathway (PPP), was inserted into its genome in two copies, resulting in their ability to ferment D-xylose and tolerate inhibitors for bioethanol production from lignocellulose hydrolysates showing a maximum specific consumption rate of 1.1 g/g DW/h in a synthetic medium with a complete attenuation of 35 g/L of D-xylose in about 17 h. On the other hand, metabolic and evolutionary engineering was also performed in S. cerevisiae WXY74, an acetate-tolerant strain for the rapid utilization of xylose from wheat straw residues. In this study, a six-gene expression cluster was optimized to obtain better ethanol yields from mixtures consisting of 80 g/L glucose and 40 g/L xylose with the presence of 3 g/L acetate in the medium, which approached 99% of the theoretical yield. The genetically modified strain achieved an ethanol of 58.4 g/L [144].
Table 4. Engineered yeast strains for hexose and pentose fermentation for bioethanol production.
Table 4. Engineered yeast strains for hexose and pentose fermentation for bioethanol production.
GeneGene SourceModified StrainSubstrateBioethanol Concentration, g/LBioethanol Yield, g/gReference
XYL1, XYL2, XKS1, TAL1Scheffersomyce stipitisS. cerevisiae MEC1122Glucose, xylose2.5 0.12 [145]
PPP genes (TAL1, TKL1, RKI1, RPE1)S. cerevisiae and Kluyveromyces marxianusS. cerevisiae
YK246
Glucose, xylose48.6 0.45 [146]
BvuXylA, XIqXylA, TAAXylA, araA, araB, araDS. cerevisiae JUK36αS. cerevisiae 36aS1.10.4Glucose, xylose54.11 0.44 [141]
HXK2, and RSP5S. cerevisiaeS. cerevisiae IMS0629Glucose, xylose-0.18 [147]
PHO4S. cerevisiae MC15S. cerevisiae MF01-PHO4Glucose114.71 -[148]
PMA1, VMA1, VMA2, VMA4-8, VMA22S. cerevisiaeS. cerevisiae XUSAE57Glucose, xylose-0.49[149]
GPD2, FPS1, ADH2, DLD3S. cerevisiaeS. cerevisiae SCGFADGlucose23,29 -[150]
SESTCAmpullaria gigas SpixS. cerevisiae wild typeGlucose7.53 0.377[151]
ADH2 and hygS. cerevisiaeS. cerevisiae ΔADH2 As2.4Glucose14.6 -[152]
noxELactococcus lactiswasS. cerevisiae JX123_noxEGlucose, xylose55.5 0.433[153]
EG1, CBH1, BglcTrichoderma reeseiS. cerevisiae BY4743-4AGlucose, xylose32.60.42[154]
EGII, CBHII, XynII, BGL, XylATrichoderma reesei, Aspergillus aculeatus and Aspergillus oryzaeS. cerevisiae Y5Xylose1.61 0.33[155]
BGL1, XYLA, XYNAspergillus aculeatus, Aspergillus oryzae and Trichoderma reeseiS. cerevisiae industrial strainGlucose, xylose11.1 0.328[156]
XylA and XynAspergillus niger, Saccharophagus degradansS. cerevisiaeXylose6.0-[157]
XYNII, XylA, BGL1Aspergillus aculeatusS. cerevisiae MN8140/XBXXGlucose, xylose8.20.32[158]
Genetic tools play an important role in the modification of the metabolism of different strains for the improvement of biofuel production. Through gene editing based on CRISPR/cas9 technology, genes can be modified to improve the tolerance and activity of yeast strains to obtain better ethanol production [159]. For example, in S. cerevisiae Y1H, a wild-type strain, the ADH2 gene was first disrupted and completely deleted using CRISPR-Cas9-mediated double-strand breaks (DSBs), resulting in a significant improvement in ethanol yield by 74.7% [160]. Meanwhile, Liu et al. [161] studied the mutant strain S. cerevisiae YS-6 from wild-type Y1H to alter ADH2 and, for the first time, ALD4 genes in ethanol fermentation. Y1H used 33.779 g/L of glucose and produced 15.431 g/L of ethanol, while strain YS-6 consumed 45.790 g/L of glucose to produce 21.711 g/L of ethanol, obtaining high productivity and yield. Moreover, this genetic tool has also demonstrated that the regulation of gene expression improves tolerance to multiple stresses.
In S. cerevisiae KF-7 and E-158, overexpression of the ENA5 gene of E. coli DH5α significantly improved multiple stress tolerance. The E-158-ENA5 engineered strain resulting from S. cerevisiae E-158 achieved higher ethanol accumulation with a concentration of 138.43 g/L when cassava was used as raw material [162]. Moreover, the simultaneous knockdown of GPD2, FPS1, and ADH2 genes indicated that S. cerevisiae SCGFA produced 23.1 g/L of ethanol compared to the wild-type strain that obtained 19.6 g/L with 50 g/L of glucose as substrate at 72 h and showed the highest ethanol conversion rate of 0.462 g/g of glucose [163].
The CRISPR/Cas9 technology that assisted with determining the genes for α-amylase (amyA) and glucoamylase (glaA) from Aspergillus tubingensis was cloned and expressed in a multicopy plasmid in the S. cerevisiae L20 strain, from which its recombinant strain L20 dT8 produced an ethanol concentration of 0.67 g/L; therefore, it obtained great potential for application in the industry of corn starch transformation into bioethanol [164].
In addition to evaluating the conventional yeast S. cerevisiae, several studies have considered other microorganisms as good candidates for bioethanol production based on their genetic modifications. Such is the case of Zymomonas mobilis, a bacterium that has been considered a model for bioethanol production due to its high sugar uptake, yield, and tolerance in which CRISPR/Cas9 technology has been used to facilitate the characterization of important genes and to construct new metabolic pathways. For example, a comparative transcriptome analysis between two ZM4 and ZM1 strains of Z. mobilis identified four highly expressed green fluorescent proteins (ORFs) potentially related to higher rates of glucose uptake and thus high bioethanol production [165].
On the other hand, Bacillus subtilis has an innate ability to grow on a diversity of carbohydrates and produce bioethanol from plant biomass. In the study conducted by Maleki et al. [125], the genes encoding alcohol dehydrogenases, adhZ and adhS from Z. mobilis and S. cerevisiae, respectively, along with the pdcZ gene (pyruvate decarboxylase from Z. mobilis) were used to create ethanol operons in a lactate-deficient B. subtilis, giving rise to strains NZ, NZS, and the recombinant NS:Z, where ethanol production by NZS and NS:Z using potatoes resulted in 16.3 g/L and 21.5 g/L during 96 h of fermentation, respectively. Another bacterium that has high potential to be a biocatalyst in ethanol production is Lactiplantibacillus casei, which is distinguished by its strong carbohydrate metabolism capacity and high tolerance to ethanol.
In a study, the feasibility of producing bioethanol using the ethanologenic engineered strain L. casei E1 as a starter culture and sugarcane molasses as a substrate medium was evaluated, and the results showed an ethanol yield of 13.77 g/L in 36 h and a carbohydrate percentage of 78.60% [166]. Moreover, the unconventional yeast Kluyveromyces marxianus is a promising eukaryotic microbe for bioethanol production and other environmental applications. Similarly, the modified K. marxianus strain expresses the major cellulosome complex OlpB of H. thermocellum on its cell surface, producing an ethanol concentration of 8.61 g/L [167]. Although various strains of different microorganisms have been modified by genetic tools to improve performance and tolerance to ethanol, it has not yet been reported if any are commercially viable. Therefore, S. cerevisiae is the model yeast that has been studied the most to be able to ferment hexoses and pentoses to produce bioethanol.
Genetic tools can also be used to develop feedstock for enhancing bioethanol yield. Modification of specific genes is essential to either increase or decrease the amount of plant cell wall components, making it an effective strategy to increase productivity, improve biomass yields, and reduce the high costs of bioethanol production [168]. Genetic engineering, coupled with pretreatment and fermentation methods, can optimize the process for increasing the mobility of degrading enzymes and achieving higher saccharification yield [169].
It is well known that in all lignocellulosic crops, lignin is one of the main targets in transgenic studies due to its complex structural characteristics [170]. Therefore, in vitro techniques and gene manipulation using DNA mutations have been used to reduce the degree of lignification in transgenic plants, achieving a reduction in lignin content by 10–50% [171]. Some studies have used RNA interference (RNAi), a mechanism that helps to tailor endogenous gene expression in plants to target genes of pests and pathogens [172]. Bewg et al. [170] used RNAi to individually knock down the expression of key genes in the lignin biosynthetic pathway in sugarcane (Saccharum officinarum L.): caffeoyl-CoA O-methyltransferase (CCoAOMT), ferulate 5-hydroxylase (F5H), and caffeic acid O-methyltransferase (COMT), respectively. The CCoAOMT gene transgenic lines did not have a reduction in lignin content but showed a significant release of glucose in one plant. The same occurred with the F5H gene lines. In contrast, for COMT, only one line reduced lignin content and obtained the highest levels of glucose during enzymatic hydrolysis.
Likewise, to improve sugarcane straw saccharification, it is possible to perform genetic RNAi silencing of the BAHD01 gene, which is responsible for the feruloylation of the cell walls of grasses. Research by de Souza et al. [173] identified six BAHD genes in the sugarcane genome (Saccharum officinarum L.) (SacBAHDs) and generated five lines with decreased expression of BAHD01. However, only three lines showed an improvement in saccharification efficiency of 24% and obtained a lower cell wall ferulate content but without any significant change in cellulose, hemicellulose, and lignin composition. On the other hand, in the study carried out by Poovaiah et al. [174], MYB31 and MYB42 genes from a maize plant (Zea mays) were cloned and expressed in sugarcane (Saccharum officinarum L.). MYB31 expression decreased lignin content in some plants, while all plants expressed with MYB42 had significant increases in glucose release by enzymatic hydrolysis, which makes MYB42 a transcription factor of interest for improving the production of second-generation bioethanol from sugarcane bagasse.
Besides sugarcane genetic modification to improve the amount of ethanol produced, other types of plants have been modified, such as rice (Oryza sativa L.). Huang et al. [168], who studied two GH9B subclass genes (OsGH9B1 and OsGH9B3) and analyzed cell wall changes and biomass saccharification in transgenic rice plants, demonstrated that the transgenic plants showed little alteration in cell wall composition (cellulose, hemicellulose, and lignin), stem mechanical strength, and biomass yield, resulting in high sugar efficiency after alkaline pretreatments and subsequent enzymatic hydrolysis, yielding 22.5% dry matter bioethanol yield. The RNAi approach was also used for the knockdown of OsSEX4 gene expression in rice straw, which caused a large accumulation of starch, resulting in a 50% increase in bioethanol production yield compared to that of wild-type straw [175]. In a study by Ai et al. [176], they worked with transgenic rice lines overexpressing AtCesA6, a gene involved in cellulose biosynthesis of primary cell walls in Arabidopsis thaliana. Enzymatic saccharification was significantly improved and was almost complete; thus, the transgenic rice samples achieved bioethanol yields of over 20% dry matter and bioethanol concentrations of 18.3 and 19.1 g/L. It has been suggested that, in rice and maize crops, overexpression of GA20-OX1, a key gene in gibberellic acid (GA) biosynthesis, is responsible for differences in plant growth rates. Voorend et al. [177] examined the biomass yield and quality characteristics of maize plants overexpressing GA20-OX1 and determined that the stalks of these plants were longer but also thinner, revealing that cellulose and lignin were deposited earlier in stalk development.
Furthermore, high saccharification efficiency per unit dry weight was obtained when the biomass was pretreated with NaOH. Therefore, the stem biomass yields as well as the conversion efficiency of cell wall polysaccharides into fermentable sugars are important for bioethanol production. Thus, genetic modification of different crops promises to incorporate desired characteristics to make the process more cost-effective in obtaining renewable fuels.

6. Challenges in Ethanol Conversion: Social, Economic, and Environmental Aspects

Regardless of bioethanol’s energetic potential, every process must be analyzed and verified to approach sustainable production and development. As new strategies for environmental assessment emerge, a methodology has been developed to assess the performance of products and services considering their environmental impacts over their life cycle, called life cycle assessment (LCA).
LCAs are tools that help develop and test environmentally and socio-economically improved scenarios. Furthermore, generated emissions and discharges are quantified, which represent environmental challenges such as global warming potential, eutrophication, and acidification [178,179]. According to Soleymani Angili [180], in a compilation of several studies on bioethanol production, 80% of studies revealed a reduction in global warming potential due to bioethanol production. It was also observed that an increase in acidification, eutrophication, and photochemical oxidation formation was due to feedstocks. They concluded that LCA studies should consider agrochemicals used in feedstock production and processing into the bioethanol process, giving importance to the chemicals generated, such as fertilizers and pesticides that increase acidification, eutrophication, and land change. New strategies are suggested to generate feedstock with fewer amounts of chemicals. However, even if LCA studies demonstrated the effectiveness of bioethanol to reduce GHG emissions, they have failed to report the impact of land usage.
Institutions like FAO suggest deep analysis of land usage, meaning that land should be categorized and utilized for specific crops for increased sustainability. To achieve this, benefits and input must be considered to assess its productive potential. The analysis requires a multi-disciplinary approach, studying it through biophysical, economic, social, and political aspects [181]. Using the LCA tool prevents the burden shifting between the production stages, originates from efforts to reduce environmental impact, and unintentionally creates impacts in other processes [182]. This procedure involves a meticulously planned supply chain system to ensure the quality and the efficient flow of materials, information, and finances. It is suggested that having croplands near the operation helps with supply chain optimization and cost reduction. Rural zones are suggested to avoid space limitations and to generate social impact, generating employment and development in these areas, which lead to academic improvement, so people from rural zones can match the demands and abilities required for the new job market. Additionally, this functions as a favorable factor to gain social acceptance and make the installation of the operation smoother.
Furthermore, the technical aspects must be studied further, especially pretreatment technologies, because they represent economic and environmental concerns; also, the inputs and outputs in bioethanol production must be considered. Wang et al. [183] conducted an LCA analysis in China comparing first- and second-generation bioethanol. The first generation of bioethanol production has a water-use intensity of 2545 m3/104 CNY (USD 1460), with 73% green and 27% blue water, and a farmland-use and CO2 emission intensity of 0.34 ha/104 CNY and 1.86 t/104 CNY, respectively. In contrast, second-generation bioethanol production has a water-use intensity of 1281 m3/104 CNY, farmland-use intensity of 0.13 ha/104 CNY, and a CO2 emission intensity of 1.21 t/104 CNY. On the other hand, Hossain et al. [184] used algal biomass as feedstock and produced 110 tons per year of algal biomass, which resulted in a bioethanol production of 57,087.5 gallons/year, with an annual CO2 consumption of 366.66 tons (1.83 CO2 kg consumed per kg of biomass) and an annual CO2 emission of 147.8 tons. Therefore, the annual net CO2 balance is 218.86 tons. However, due to energy needed throughout the process, there is an energy balance of -2749.6 GJ per year, despite generating 1290 GJ per year.
A circular economy (CE) is another concept used to evaluate the impact of the entire process, not only the main product generated in biofuels (e.g., ethanol, methane, diesel) but also the by-products generated and their application in real-life scenarios [185]. The CE approach for bioethanol production involves obtaining value-added products other than fuel products under a biorefinery approach. Although these products can inhibit fermentation, their management is crucial for effectively converting biomass into ethanol and other income sources, such as furfural, HMF, and lignin residues. HMF is the result of the dehydration of sugar monomers, mainly hexoses, and is used as a key intermediate in the synthesis of polymers from biomass, while furfural is obtained from xylose sugar and has industrial applications, such as making plastics, fertilizers, adhesives, fungicides, and antacids as shown in Figure 3. Moreover, furfural has been reported as a precursor for D-proline, which is used as a precursor in pharmaceuticals [186,187]. On the other hand, lignin is a natural raw material with higher aromatic content that can be used to produce dispersants, adhesives, and surfactants. Furthermore, lignin residues have been used in the rubber industry [188].
Some countries have evaluated the economic and environmental aspects of bioethanol production to predict suitable scenarios by considering several factors that play an important role in ethanol production, such as Lopez-Ortega et al. [189]. They proposed a scenario for Mexico in which a sugarcane factory is transformed into a biorefinery to estimate a techno-economic analysis. Table 5 summarizes several parameters in a techno-economic analysis and lists the inputs and outputs under a scenario of bioethanol production from the sugarcane industry under a biorefinery concept with 1G bioethanol production. Table 5 shows the techno-economic analysis and the main inputs and outputs of bioethanol production in Mexico.
On the other hand, Table 6 provides a comparison of 1G, 2G, and 3G bioethanol production in different countries, analyzing inputs and outputs throughout the process. While 3G has some environmental advantages compared to 1G and 2G, such as CO2 capture during biomass cultivation and the use of wastewater to cultivate microalgae, as well as a relatively lower water and land footprint compared to other feedstocks such as sorghum, cassava, and sugarcane [184], there are also some factors that are not as favorable for 3G, as can be seen in Table 6 below.

7. Future Perspectives

The present challenges in bioethanol technologies for 1G environmental issues include unfavorable climate conditions such as droughts that lead to a decrease in food crop production and interfere with feed and food chains around the world. Additionally, the use of fertilizers and pesticides results in an increase in air, water, and soil pollution. Low government investment and an increase in taxes may cause some bioethanol plants to disappear [15]. Meanwhile, 2G faces problems in the pretreatment and hydrolysis processes that may not be suitable on an industrial scale. In 3G, microalgae cultivation (including strain selection and cultivation conditions) and cell collection are critical steps that influence bioethanol production. A large volume of water and nutrients are required, but this can be resolved by using wastewater. Biomass represents 40% to 70% of the total production cost in both 1G and 2G bioethanol production, making it an important economic consideration. Further, in the case of 1G, biomass is obtained directly, while in 2G, additional steps such as pretreatment are required for the biomass used in the process. In some cases, biological hydrolysis represents 20% to 40% of the total cost in the second generation of bioethanol, which is the main challenge in the 1G and 2G bioethanol production process [193].
Biorefineries require high investments, which is why businesses look for stakeholders. Stakeholders could be individuals or a group of people that hold rightful interest in the operation. They can provide financial support or human resources. Stakeholders may be governmental or non-governmental, and they play crucial roles in economic and social aspects. Non-governmental stakeholders help to increase the social acceptance of these kinds of projects and obtain funding and resources. Governmental stakeholders fasten the land-use approval, provide subsidies, and facilitate the generation of policies [194]. Policies for biofuels around the world dictate that global governments should promote initiatives to enhance the development of the biofuels market.
The most important initiatives include creating a global partnership to support the production of biofuels from biomass, especially in developing countries, taking into consideration the predominant use of biomass. Global actions to fight climate change and reduce carbon footprint are also necessary, along with boosting the development of clean energy and innovating strategies to accelerate development in energy. Several countries aspire to grow their production of biofuels, such as some Latin American countries, specifically Brazil, which is projected to increase its biofuel production by 18% by 2030. Additionally, the use of genetic engineering to improve strain yield and manipulate biosynthesis has shown promising advances. Moreover, genetic engineering can manipulate cellulose and hemicellulose content, leading to advances in bioethanol production. On the other hand, the use of modified crops has raised ethical debates about their employment due to concerns about changes in nature and plant integrity [195,196]. To conclude, though various strategies are being developed to improve biofuel yield by employing sustainable practices, there is an urgent need to focus on the evaluation of the sustainability and the impact of new technologies on the environment or on society in the mid- and long-term before adopting them at a large scale.

8. Conclusions

The effective conversion of LCB raw materials into bioethanol depends on appropriate pretreatment (if required) and adequate operation conditions that maximize fermentable sugar content and the purification of final products. New technologies, such as genetic engineering, offer tools for developing microorganisms that are more competent in the direct conversion of fermentable sugars into ethanol, thereby improving the bioethanol process. It is critical to ensure the availability of raw materials without interfering with society’s development and necessities.
Effective pretreatment is necessary to prepare feedstock for convenient conversion into bioethanol. Furthermore, conditions during hydrolysis and fermentation play a key role in the conversion of substances from which ethanol can actually be obtained, and in subsequent downstream processes ultimately improve the quality of final products. Different technological perspectives, such as genetic tools, are enhancing bioethanol production through various metabolic pathways and developing robust strains that can be applied in the industrial sector, not just at the laboratory scale. Although bioethanol is an attractive alternative for combating the global energy crisis and climate change, life cycle assessment (LCA) evaluates the environmental risks that could be associated with bioethanol production, such as greenhouse gas emissions and energy consumption. LCA, in combination with the circular economy concept under a biorefinery approach, assesses the inputs required to obtain all final products (outputs) and their utilization in different industrial sectors and economic perspectives for the sustainable development society.

Author Contributions

Conceptualization, N.B.; methodology, N.B.; investigation, M.A.Y.-G., L.A.I.-M., A.Y.C.-H.W., A.C.F.C., J.D.S.-M., A.S.T.-P. and N.B.; validation, S.S.-M., S.S.d.S., J.C.d.S., J.U.H.-B. and N.B.; formal analysis, M.A.Y.-G., L.A.I.-M., A.Y.C.-H.W. and S.S.-M.; resources, N.B.; data curation, M.A.Y.-G. and S.S.-M.; writing—original draft preparation, M.A.Y.-G., L.A.I.-M., A.Y.C.-H.W., A.C.F.C., J.D.S.-M. and A.S.T.-P.; writing—review and editing, S.S.d.S., S.S.-M., J.C.d.S., J.U.H.-B. and N.B.; supervision, N.B.; project administration, N.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that their work was conducted in the absence of any commercial or financial relationships that could be interpreted as a potential conflict of interest.

References

  1. Sadigov, R. Rapid Growth of the World Population and Its Socioeconomic Results. Sci. World J. 2022, 2022, 8110229. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  7. Renewable Fuels Association. 2023 Ready. Set. Go! Renewable Fuels Association: Washington, DC, USA, 2023. [Google Scholar]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  13. Broda, M.; Yelle, D.J.; Serwańska, K. Bioethanol Production from Lignocellulosic Biomass—Challenges and Solutions. Molecules 2022, 27, 8717. [Google Scholar] [CrossRef]
  14. Maity, S.; Mallick, N. Trends and advances in sustainable bioethanol production by marine microalgae: A critical review. J. Clean. Prod. 2022, 345, 131153. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  27. Ramachandra, T.; Hebbale, D. Bioethanol from macroalgae: Prospects and challenges. Renew. Sustain. Energy Rev. 2020, 117, 109479. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  29. Dave, N.; Selvaraj, R.; Varadavenkatesan, T.; Vinayagam, R. A critical review on production of bioethanol from macroalgal biomass. Algal Res. 2019, 42, 101606. [Google Scholar] [CrossRef]
  30. Chowdhury, H.; Loganathan, B. Third-generation biofuels from microalgae: A review. Curr. Opin. Green Sustain. Chem. 2019, 20, 39–44. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  33. Alalwan, H.A.; Alminshid, A.H.; Aljaafari, H.A. Promising evolution of biofuel generations. Subject review. Renew. Energy Focus 2019, 28, 127–139. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  40. Azeez, M.A.; Orege, J.I. Bamboo, Its Chemical Modification and Products. In Bamboo—Current and Future Prospects; IntechOpen: London, UK, 2018. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  45. Soni, S.K.; Sharma, A.; Soni, R. Microbial Enzyme Systems in the Production of Second Generation Bioethanol. Sustainability 2023, 15, 3590. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  53. Brigham, C. Biopolymers. In Green Chemistry; Elsevier: Amsterdam, The Netherlands, 2018; pp. 753–770. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  58. Zhou, X.; Li, W.; Mabon, R.; Broadbelt, L.J. A Critical Review on Hemicellulose Pyrolysis. Energy Technol. 2017, 5, 52–79. [Google Scholar] [CrossRef]
  59. Bajpai, P. Wood and Fiber Fundamentals. In Biermann’s Handbook of Pulp and Paper; Elsevier: Amsterdam, The Netherlands, 2018; pp. 19–74. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  61. Sharma, S.; Kumar, A. Lignin; Springer International Publishing: Cham, Switzerland, 2020. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  66. Kim, D. Physico-Chemical Conversion of Lignocellulose: Inhibitor Effects and Detoxification Strategies: A Mini Review. Molecules 2018, 23, 309. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  93. Uzuner, S.; Cekmecelioglu, D. Enzymes in the Beverage Industry. In Enzymes in Food Biotechnology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 29–43. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, H. Lignocellulose biorefinery feedstock engineering. In Lignocellulose Biorefinery Engineering, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 37–86. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  97. Vasić, K.; Knez, Ž.K.; Leitgeb, M. Bioethanol Production by Enzymatic Hydrolysis from Different Lignocellulosic Sources. Molecules 2021, 26, 753. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  101. Modenbach, A.A.; Nokes, S.E. Enzymatic hydrolysis of biomass at high-solids loadings—A review. Biomass Bioenergy 2013, 56, 526–544. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef]
  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. [Google Scholar] [CrossRef] [PubMed]
  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. [Google Scholar] [CrossRef] [PubMed]
  106. Godbey, W.T. Fermentation, Beer, and Biofuels. In An Introduction to Biotechnology; Elsevier: Amsterdam, The Netherlands, 2014; pp. 331–351. [Google Scholar] [CrossRef]
  107. Spector, M.P. Metabolism, Central (Intermediary). In Encyclopedia of Microbiology; Elsevier: Amsterdam, The Netherlands, 2009; pp. 242–264. [Google Scholar] [CrossRef]
  108. D’Amore, T.; Panchal, C.J.; Russell, I.; Stewart, G.G. A Study of Ethanol Tolerance in Yeast. Crit. Rev. Biotechnol. 1989, 9, 287–304. [Google Scholar] [CrossRef]
  109. Tse, T.J.; Wiens, D.J.; Reaney, M.J.T. Production of Bioethanol—A Review of Factors Affecting Ethanol Yield. Fermentation 2021, 7, 268. [Google Scholar] [CrossRef]
  110. Li, C.; Liu, Q.; Wang, Y.; Yang, X.; Chen, S.; Zhao, Y.; Wu, Y.; Li, L. Salt stress improves thermotolerance and high-temperature bioethanol production of multi-stress-tolerant Pichia kudriavzevii by stimulating intracellular metabolism and inhibiting oxidative damage. Biotechnol. Biofuels 2021, 14, 222. [Google Scholar] [CrossRef]
  111. Taherzadeh, M.J.; Karimi, K. Fermentation Inhibitors in Ethanol Processes and Different Strategies to Reduce Their Effects. In Biofuels; Elsevier: Amsterdam, The Netherlands, 2011; pp. 287–311. [Google Scholar] [CrossRef]
  112. Varize, C.S.; Bücker, A.; Lopes, L.D.; Christofoleti-Furlan, R.M.; Raposo, M.S.; Basso, L.C.; Stambuk, B.U. Increasing Ethanol Tolerance and Ethanol Production in an Industrial Fuel Ethanol Saccharomyces cerevisiae Strain. Fermentation 2022, 8, 470. [Google Scholar] [CrossRef]
  113. Rastogi, M.; Shrivastava, S. Recent advances in second generation bioethanol production: An insight to pretreatment, saccharification and fermentation processes. Renew. Sustain. Energy Rev. 2017, 80, 330–340. [Google Scholar] [CrossRef]
  114. Sharma, S.; Nair, A.; Sarma, S.J. Biorefinery concept of simultaneous saccharification and co-fermentation: Challenges and improvements. Chem. Eng. Process. Process Intensif. 2021, 169, 108634. [Google Scholar] [CrossRef]
  115. Chu, Q.; Li, X.; Ma, B.; Xu, Y.; Ouyang, J.; Zhu, J.; Yu, S.; Yong, Q. Bioethanol production: An integrated process of low substrate loading hydrolysis-high sugars liquid fermentation and solid state fermentation of enzymatic hydrolysis residue. Bioresour. Technol. 2012, 123, 699–702. [Google Scholar] [CrossRef] [PubMed]
  116. Gomes, D.; Cruz, M.; de Resende, M.; Ribeiro, E.; Teixeira, J.; Domingues, L. Very High Gravity Bioethanol Revisited: Main Challenges and Advances. Fermentation 2021, 7, 38. [Google Scholar] [CrossRef]
  117. Cruz, M.L.; de Resende, M.M.; Ribeiro, E.J. Improvement of ethanol production in fed-batch fermentation using a mixture of sugarcane juice and molasse under very high-gravity conditions. Bioprocess Biosyst. Eng. 2021, 44, 617–625. [Google Scholar] [CrossRef]
  118. Chang, Y.-H.; Chang, K.-S.; Chen, C.-Y.; Hsu, C.-L.; Chang, T.-C.; Jang, H.-D. Enhancement of the Efficiency of Bioethanol Production by Saccharomyces cerevisiae via Gradually Batch-Wise and Fed-Batch Increasing the Glucose Concentration. Fermentation 2018, 4, 45. [Google Scholar] [CrossRef]
  119. Kumari, M.; Gupta, S.K. Response surface methodological (RSM) approach for optimizing the removal of trihalomethanes (THMs) and its precursor’s by surfactant modified magnetic nanoadsorbents (sMNP)—An endeavor to diminish probable cancer risk. Sci. Rep. 2019, 9, 18339. [Google Scholar] [CrossRef]
  120. Assabjeu, A.C.; Noubissié, E.; Desobgo, S.C.Z.; Ali, A. Optimization of the enzymatic hydrolysis of cellulose of triplochiton scleroxylon sawdust in view of the production of bioethanol. Sci. Afr. 2020, 8, e00438. [Google Scholar] [CrossRef]
  121. El-Mekkawi, S.A.; Abdo, S.M.; Samhan, F.A.; Ali, G.H. Optimization of some fermentation conditions for bioethanol production from microalgae using response surface method. Bull. Natl. Res. Cent. 2019, 43, 164. [Google Scholar] [CrossRef]
  122. Žnidaršič-Plazl, P. The Promises and the Challenges of Biotransformations in Microflow. Biotechnol. J. 2019, 14, e1800580. [Google Scholar] [CrossRef]
  123. Hernández-Beltrán, J.U.; Fontalvo, J.; Hernández-Escoto, H. Fed-batch enzymatic hydrolysis of plantain pseudostem to fermentable sugars production and the impact of particle size at high solids loadings. Biomass Convers. Biorefin. 2021, 11, 2975–2982. [Google Scholar] [CrossRef]
  124. Van Zyl, W.H.; Lynd, L.R.; den Haan, R.; McBride, J.E. Consolidated Bioprocessing for Bioethanol Production Using Saccharomyces cerevisiae. In Biofuels; Springer: Berlin/Heidelberg, Germany, 2007; pp. 205–235. [Google Scholar] [CrossRef]
  125. Maleki, F.; Changizian, M.; Zolfaghari, N.; Rajaei, S.; Noghabi, K.A.; Zahiri, H.S. Consolidated bioprocessing for bioethanol production by metabolically engineered Bacillus subtilis strains. Sci. Rep. 2021, 11, 13731. [Google Scholar] [CrossRef]
  126. Pinaki, D.; Lhakpa, W.; Joginder, S. Simultaneous Saccharification and Fermentation (SSF), An Efficient Process for Bio-Ethanol Production: An Overview. Biosci. Biotechnol. Res. Asia 2015, 12, 87–100. [Google Scholar] [CrossRef]
  127. Derman, E.; Abdulla, R.; Marbawi, H.; Sabullah, M.K.; Gansau, J.A.; Ravindra, P. Simultaneous Saccharification and Fermentation of Empty Fruit Bunches of Palm for Bioethanol Production Using a Microbial Consortium of S. cerevisiae and T. harzianum. Fermentation 2022, 8, 295. [Google Scholar] [CrossRef]
  128. Adekunle, A.E.; Rabeya, T.; Jehadin, F.; Asad, M.A.; Ayodele, O.O.; Islam, S. Compressed hot water pretreatment enhanced bioethanol production from corn stalk. Bioresour. Technol. Rep. 2020, 12, 100595. [Google Scholar] [CrossRef]
  129. Fetyan, N.A.H.; El-Sayed, A.E.-K.B.; Ibrahim, F.M.; Attia, Y.A.; Sadik, M.W. Bioethanol production from defatted biomass of Nannochloropsis oculata microalgae grown under mixotrophic conditions. Environ. Sci. Pollut. Res. 2022, 29, 2588–2597. [Google Scholar] [CrossRef] [PubMed]
  130. Jugwanth, Y.; Sewsynker-Sukai, Y.; Kana, E.G. Valorization of sugarcane bagasse for bioethanol production through simultaneous saccharification and fermentation: Optimization and kinetic studies. Fuel 2020, 262, 116552. [Google Scholar] [CrossRef]
  131. Chohan, N.A.; Aruwajoye, G.; Sewsynker-Sukai, Y.; Kana, E.G. Valorisation of potato peel wastes for bioethanol production using simultaneous saccharification and fermentation: Process optimization and kinetic assessment. Renew. Energy 2020, 146, 1031–1040. [Google Scholar] [CrossRef]
  132. Kumar, D.; Surya, K.; Verma, R. Bioethanol production from apple pomace using co-cultures with Saccharomyces cerevisiae in solid-state fermentation. J. Microbiol. Biotechnol. Food Sci. 2020, 9, 742–745. [Google Scholar] [CrossRef]
  133. Efunwoye, O.O.; Oluwole, O.R. Bioethanol Production from Pineapple Waste by Solid State Fermentation Method. 2019. Available online: www.nsmjournal.org (accessed on 19 December 2023).
  134. Prasoulas, G.; Gentikis, A.; Konti, A.; Kalantzi, S.; Kekos, D.; Mamma, D. Bioethanol Production from Food Waste Applying the Multienzyme System Produced On-Site by Fusarium oxysporum F3 and Mixed Microbial Cultures. Fermentation 2020, 6, 39. [Google Scholar] [CrossRef]
  135. Camargos, C.V.; Moraes, V.D.; de Oliveira, L.M.; Guidini, C.Z.; Ribeiro, E.J.; Santos, L.D. High Gravity and Very High Gravity Fermentation of Sugarcane Molasses by Flocculating Saccharomyces cerevisiae: Experimental Investigation and Kinetic Modeling. Appl. Biochem. Biotechnol. 2021, 193, 807–821. [Google Scholar] [CrossRef]
  136. Granjo, J.F.; Nunes, D.S.; Duarte, B.P.; Oliveira, N.M. A comparison of process alternatives for energy-efficient bioethanol downstream processing. Sep. Purif. Technol. 2020, 238, 116414. [Google Scholar] [CrossRef]
  137. Bisgaard, T.; Mauricio-Iglesias, M.; Huusom, J.K.; Gernaey, K.V.; Dohrup, J.; Petersen, M.A.; Abildskov, J. Adding Value to Bioethanol through a Purification Process Revamp. Ind. Eng. Chem. Res. 2017, 56, 5692–5704. [Google Scholar] [CrossRef]
  138. Jhariya, U.; Dafale, N.A.; Srivastava, S.; Bhende, R.S.; Kapley, A.; Purohit, H.J. Understanding Ethanol Tolerance Mechanism in Saccharomyces cerevisiae to Enhance the Bioethanol Production: Current and Future Prospects. BioEnergy Res. 2021, 14, 670–688. [Google Scholar] [CrossRef]
  139. Wu, M.; Li, H.; Wei, S.; Wu, H.; Wu, X.; Bao, X.; Hou, J.; Liu, W.; Shen, Y. Simulating Extracellular Glucose Signals Enhances Xylose Metabolism in Recombinant Saccharomyces cerevisiae. Microorganisms 2020, 8, 100. [Google Scholar] [CrossRef]
  140. Kwak, S.; Jin, Y.-S. Production of fuels and chemicals from xylose by engineered Saccharomyces cerevisiae: A review and perspective. Microb. Cell Factories 2017, 16, 82. [Google Scholar] [CrossRef] [PubMed]
  141. Huang, S.; Liu, T.; Peng, B.; Geng, A. Enhanced ethanol production from industrial lignocellulose hydrolysates by a hydrolysate-cofermenting Saccharomyces cerevisiae strain. Bioprocess Biosyst. Eng. 2019, 42, 883–896. [Google Scholar] [CrossRef]
  142. Wu, Y.; Xu, S.; Gao, X.; Li, M.; Li, D.; Lu, W. Enhanced protopanaxadiol production from xylose by engineered Yarrowia lipolytica. Microb. Cell Factories 2019, 18, 83. [Google Scholar] [CrossRef]
  143. Cunha, J.T.; Soares, P.O.; Baptista, S.L.; Costa, C.E.; Domingues, L. Engineered Saccharomyces cerevisiae for lignocellulosic valorization: A review and perspectives on bioethanol production. Bioengineered 2020, 11, 883–903. [Google Scholar] [CrossRef]
  144. Li, H.; Shen, Y.; Wu, M.; Hou, J.; Jiao, C.; Li, Z.; Liu, X.; Bao, X. Engineering a wild-type diploid Saccharomyces cerevisiae strain for second-generation bioethanol production. Bioresour. Bioprocess. 2016, 3, 51. [Google Scholar] [CrossRef] [PubMed]
  145. Romaní, A.; Pereira, F.; Johansson, B.; Domingues, L. Metabolic engineering of Saccharomyces cerevisiae ethanol strains PE-2 and CAT-1 for efficient lignocellulosic fermentation. Bioresour. Technol. 2015, 179, 150–158. [Google Scholar] [CrossRef]
  146. Kobayashi, Y.; Sahara, T.; Ohgiya, S.; Kamagata, Y.; Fujimori, K.E. Systematic optimization of gene expression of pentose phosphate pathway enhances ethanol production from a glucose/xylose mixed medium in a recombinant Saccharomyces cerevisiae. AMB Express 2018, 8, 139. [Google Scholar] [CrossRef]
  147. Papapetridis, I.; Verhoeven, M.D.; Wiersma, S.J.; Goudriaan, M.; van Maris, A.A.J.; Pronk, J.T. Laboratory evolution for forced glucose-xylose co-consumption enables identification of mutations that improve mixed-sugar fermentation by xylose-fermenting Saccharomyces cerevisiae. FEMS Yeast Res. 2018, 18, foy056. [Google Scholar] [CrossRef]
  148. Wu, R.; Chen, D.; Cao, S.; Lu, Z.; Huang, J.; Lu, Q.; Chen, Y.; Chen, X.; Guan, N.; Wei, Y.; et al. Enhanced ethanol production from sugarcane molasses by industrially engineered Saccharomyces cerevisiae via replacement of the PHO4 gene. RSC Adv. 2020, 10, 2267–2276. [Google Scholar] [CrossRef] [PubMed]
  149. Ko, J.K.; Enkh-Amgalan, T.; Gong, G.; Um, Y.; Lee, S. Improved bioconversion of lignocellulosic biomass by Saccharomyces cerevisiae engineered for tolerance to acetic acid. GCB Bioenergy 2020, 12, 90–100. [Google Scholar] [CrossRef]
  150. Yang, P.; Jiang, S.; Lu, S.; Jiang, S.; Jiang, S.; Deng, Y.; Lu, J.; Wang, H.; Zhou, Y. Ethanol yield improvement in Saccharomyces cerevisiae GPD2 Delta FPS1 Delta ADH2 Delta DLD3 Delta mutant and molecular mechanism exploration based on the metabolic flux and transcriptomics approaches. Microb. Cell Factories 2022, 21, 160. [Google Scholar] [CrossRef] [PubMed]
  151. Yang, P.; Wu, Y.; Zheng, Z.; Cao, L.; Zhu, X.; Mu, D.; Jiang, S. CRISPR-Cas9 Approach Constructing Cellulase sestc-Engineered Saccharomyces cerevisiae for the Production of Orange Peel Ethanol. Front. Microbiol. 2018, 9, 2436. [Google Scholar] [CrossRef] [PubMed]
  152. Ye, W.; Zhang, W.; Liu, T.; Tan, G.; Li, H.; Huang, Z. Improvement of Ethanol Production in Saccharomyces cerevisiae by High-Efficient Disruption of the ADH2 Gene Using a Novel Recombinant TALEN Vector. Front. Microbiol. 2016, 7, 1067. [Google Scholar] [CrossRef]
  153. Lee, Y.-G.; Jin, Y.-S.; Cha, Y.-L.; Seo, J.-H. Bioethanol production from cellulosic hydrolysates by engineered industrial Saccharomyces cerevisiae. Bioresour. Technol. 2017, 228, 355–361. [Google Scholar] [CrossRef]
  154. Zhang, Y.; Wang, C.; Wang, L.; Yang, R.; Hou, P.; Liu, J. Direct bioethanol production from wheat straw using xylose/glucose co-fermentation by co-culture of two recombinant yeasts. J. Ind. Microbiol. Biotechnol. 2017, 44, 453–464. [Google Scholar] [CrossRef]
  155. Chen, L.; Du, J.-L.; Zhan, Y.-J.; Li, J.-A.; Zuo, R.-R.; Tian, S. Consolidated bioprocessing for cellulosic ethanol conversion by cellulase–xylanase cell-surfaced yeast consortium. Prep. Biochem. Biotechnol. 2018, 48, 653–661. [Google Scholar] [CrossRef]
  156. Cunha, J.T.; Romaní, A.; Inokuma, K.; Johansson, B.; Hasunuma, T.; Kondo, A.; Domingues, L. Consolidated bioprocessing of corn cob-derived hemicellulose: Engineered industrial Saccharomyces cerevisiae as efficient whole cell biocatalysts. Biotechnol. Biofuels 2020, 13, 138. [Google Scholar] [CrossRef]
  157. Sasaki, Y.; Takagi, T.; Motone, K.; Kuroda, K.; Ueda, M. Enhanced direct ethanol production by cofactor optimization of cell surface-displayed xylose isomerase in yeast. Biotechnol. Prog. 2017, 33, 1068–1076. [Google Scholar] [CrossRef] [PubMed]
  158. Sakamoto, T.; Hasunuma, T.; Hori, Y.; Yamada, R.; Kondo, A. Direct ethanol production from hemicellulosic materials of rice straw by use of an engineered yeast strain codisplaying three types of hemicellulolytic enzymes on the surface of xylose-utilizing Saccharomyces cerevisiae cells. J. Biotechnol. 2012, 158, 203–210. [Google Scholar] [CrossRef]
  159. Chin, Y.-W.; Kang, W.-K.; Jang, H.W.; Turner, T.L.; Kim, H.J. CAR1 deletion by CRISPR/Cas9 reduces formation of ethyl carbamate from ethanol fermentation by Saccharomyces cerevisiae. J. Ind. Microbiol. Biotechnol. 2016, 43, 1517–1525. [Google Scholar] [CrossRef]
  160. Xue, T.; Liu, K.; Chen, D.; Yuan, X.; Fang, J.; Yan, H.; Huang, L.; Chen, Y.; He, W. Improved bioethanol production using CRISPR/Cas9 to disrupt the ADH2 gene in Saccharomyces cerevisiae. World J. Microbiol. Biotechnol. 2018, 34, 154. [Google Scholar] [CrossRef]
  161. Liu, K.; Yuan, X.; Liang, L.; Fang, J.; Chen, Y.; He, W.; Xue, T. Using CRISPR/Cas9 for multiplex genome engineering to optimize the ethanol metabolic pathway in Saccharomyces cerevisiae. Biochem. Eng. J. 2019, 145, 120–126. [Google Scholar] [CrossRef]
  162. Wang, L.; Li, B.; Su, R.-R.; Wang, S.-P.; Xia, Z.-Y.; Xie, C.-Y.; Tang, Y.-Q. Screening novel genes by a comprehensive strategy to construct multiple stress-tolerant industrial Saccharomyces cerevisiae with prominent bioethanol production. Biotechnol. Biofuels Bioprod. 2022, 15, 11. [Google Scholar] [CrossRef] [PubMed]
  163. Yang, P.; Jiang, S.; Jiang, S.; Lu, S.; Zheng, Z.; Chen, J.; Wu, W.; Jiang, S. CRISPR-Cas9 Approach Constructed Engineered Saccharomyces cerevisiae with the Deletion of GPD2, FPS1, and ADH2 to Enhance the Production of Ethanol. J. Fungi 2022, 8, 703. [Google Scholar] [CrossRef] [PubMed]
  164. Gronchi, N.; De Bernardini, N.; Cripwell, R.A.; Treu, L.; Campanaro, S.; Basaglia, M.; Foulquié-Moreno, M.R.; Thevelein, J.M.; Van Zyl, W.H.; Favaro, L.; et al. Natural Saccharomyces cerevisiae Strain Reveals Peculiar Genomic Traits for Starch-to-Bioethanol Production: The Design of an Amylolytic Consolidated Bioprocessing Yeast. Front. Microbiol. 2022, 12, 768562. [Google Scholar] [CrossRef]
  165. Zhang, K.; Lu, X.; Li, Y.; Jiang, X.; Liu, L.; Wang, H. New technologies provide more metabolic engineering strategies for bioethanol production in Zymomonas mobilis. Appl. Microbiol. Biotechnol. 2019, 103, 2087–2099. [Google Scholar] [CrossRef]
  166. Wang, S.; Tian, R.; Liu, B.; Wang, H.; Liu, J.; Li, C.; Li, M.; Evivie, S.E.; Li, B. Effects of carbon concentration, oxygen, and controlled pH on the engineering strain Lactiplantibacillus casei E1 in the production of bioethanol from sugarcane molasses. AMB Express 2021, 11, 95. [Google Scholar] [CrossRef]
  167. Ha-Tran, D.M.; Nguyen, T.T.M.; Huang, C.-C. Kluyveromyces marxianus: Current State of Omics Studies, Strain Improvement Strategy and Potential Industrial Implementation. Fermentation 2020, 6, 124. [Google Scholar] [CrossRef]
  168. Huang, J.; Xia, T.; Li, G.; Li, X.; Li, Y.; Wang, Y.; Wang, Y.; Chen, Y.; Xie, G.; Bai, F.-W.; et al. Overproduction of native endo-β-1,4-glucanases leads to largely enhanced biomass saccharification and bioethanol production by specific modification of cellulose features in transgenic rice. Biotechnol. Biofuels 2019, 12, 11. [Google Scholar] [CrossRef]
  169. Schmatz, A.A.; Tyhoda, L.; Brienzo, M. Sugarcane biomass conversion influenced by lignin. Biofuels Bioprod. Biorefin. 2020, 14, 469–480. [Google Scholar] [CrossRef]
  170. Bewg, W.P.; Poovaiah, C.; Lan, W.; Ralph, J.; Coleman, H.D. RNAi downregulation of three key lignin genes in sugarcane improves glucose release without reduction in sugar production. Biotechnol. Biofuels 2016, 9, 270. [Google Scholar] [CrossRef] [PubMed]
  171. Rocha-Meneses, L.; Ferreira, J.A.; Mushtaq, M.; Karimi, S.; Orupõld, K.; Kikas, T. Genetic modification of cereal plants: A strategy to enhance bioethanol yields from agricultural waste. Ind. Crops Prod. 2020, 150, 112408. [Google Scholar] [CrossRef]
  172. Mezzetti, B.; Smagghe, G.; Arpaia, S.; Christiaens, O.; Dietz-Pfeilstetter, A.; Jones, H.; Kostov, K.; Sabbadini, S.; Opsahl-Sorteberg, H.-G.; Ventura, V.; et al. RNAi: What is its position in agriculture? J. Pest Sci. 2004, 93, 1125–1130. [Google Scholar] [CrossRef]
  173. De Souza, W.R.; Pacheco, T.F.; Duarte, K.E.; Sampaio, B.L.; Molinari, P.A.d.O.; Martins, P.K.; Santiago, T.R.; Formighieri, E.F.; Vinecky, F.; Ribeiro, A.P.; et al. Silencing of a BAHD acyltransferase in sugarcane increases biomass digestibility. Biotechnol. Biofuels 2019, 12, 111. [Google Scholar] [CrossRef] [PubMed]
  174. Poovaiah, C.R.; Bewg, W.P.; Lan, W.; Ralph, J.; Coleman, H.D. Sugarcane transgenics expressing MYB transcription factors show improved glucose release. Biotechnol. Biofuels 2016, 9, 143. [Google Scholar] [CrossRef]
  175. Huang, L.; Liu, Y.; Su, S.; Lai, C.; Wu, C.; Chao, T.; Yang, Y. Genetic engineering of transitory starch accumulation by knockdown of OsSEX4 in rice plants for enhanced bioethanol production. Biotechnol. Bioeng. 2020, 117, 933–944. [Google Scholar] [CrossRef]
  176. Ai, Y.; Feng, S.; Wang, Y.; Lu, J.; Sun, M.; Hu, H.; Hu, Z.; Zhang, R.; Liu, P.; Peng, H.; et al. Integrated genetic and chemical modification with rice straw for maximum bioethanol production. Ind. Crops Prod. 2021, 173, 114133. [Google Scholar] [CrossRef]
  177. Voorend, W.; Nelissen, H.; Vanholme, R.; De Vliegher, A.; Van Breusegem, F.; Boerjan, W.; Roldán-Ruiz, I.; Muylle, H.; Inzé, D. Overexpression of GA20-OXIDASE1 impacts plant height, biomass allocation and saccharification efficiency in maize. Plant Biotechnol. J. 2016, 14, 997–1007. [Google Scholar] [CrossRef]
  178. Morales-Vera, R.; Vásquez-Ibarra, L.; Scott, F.; Puettmann, M.; Gustafson, R. Life Cycle Assessment of Bioethanol Production: A Case Study from Poplar Biomass Growth in the U.S. Pacific Northwest. Fermentation 2022, 8, 734. [Google Scholar] [CrossRef]
  179. Ingrao, C.; Matarazzo, A.; Gorjian, S.; Adamczyk, J.; Failla, S.; Primerano, P.; Huisingh, D. Wheat-straw derived bioethanol production: A review of Life Cycle Assessments. Sci. Total. Environ. 2021, 781, 146751. [Google Scholar] [CrossRef]
  180. Angili, T.S.; Grzesik, K.; Rödl, A.; Kaltschmitt, M. Life Cycle Assessment of Bioethanol Production: A Review of Feedstock, Technology and Methodology. Energies 2021, 14, 2939. [Google Scholar] [CrossRef]
  181. Food and Agriculture Organization of the United Nations—FAO. Land Evaluation towards a Revised Framework; Food and Agriculture Organization of the United Nations—FAO: Rome, Italy, 2007. [Google Scholar]
  182. Hauschild, M.Z.; Rosenbaum, R.K.; Olsen, S.I. (Eds.) Life Cycle Assessment—Theory and Practice; Springer: Cham, Switzerland, 2018. [Google Scholar] [CrossRef]
  183. Li, W.-C.; Zhang, S.-J.; Xu, T.; Sun, M.-Q.; Zhu, J.-Q.; Zhong, C.; Li, B.-Z.; Yuan, Y.-J. Fractionation of corn stover by two-step pretreatment for production of ethanol, furfural, and lignin. Energy 2020, 195, 117076. [Google Scholar] [CrossRef]
  184. Hossain, N.; Zaini, J.; Mahlia, T.M.I. Life cycle assessment, energy balance and sensitivity analysis of bioethanol production from microalgae in a tropical country. Renew. Sustain. Energy Rev. 2019, 115, 109371. [Google Scholar] [CrossRef]
  185. Devi, A.; Bajar, S.; Kour, H.; Kothari, R.; Pant, D.; Singh, A. Lignocellulosic Biomass Valorization for Bioethanol Production: A Circular Bioeconomy Approach. BioEnergy Res. 2022, 15, 1820–1841. [Google Scholar] [CrossRef] [PubMed]
  186. Kong, Q.-S.; Li, X.-L.; Xu, H.-J.; Fu, Y. Conversion of 5-hydroxymethylfurfural to chemicals: A review of catalytic routes and product applications. Fuel Process. Technol. 2020, 209, 106528. [Google Scholar] [CrossRef]
  187. Yao, H.; Wang, Y.; Liu, J.; Xu, M.; Ma, P.; Ji, J.; You, Z. Review on Applications of Lignin in Pavement Engineering: A Recent Survey. Front. Mater. 2022, 8, 803524. [Google Scholar] [CrossRef]
  188. López-Ortega, M.G.; Guadalajara, Y.; Junqueira, T.L.; Sampaio, I.L.; Bonomi, A.; Sánchez, A. Sustainability analysis of bioethanol production in Mexico by a retrofitted sugarcane industry based on the Brazilian expertise. Energy 2021, 232, 121056. [Google Scholar] [CrossRef]
  189. Bonomi, A.; Cavalett, O. Virtual Biorefinery; da Cunha, P., Lima, M.A.P., Eds.; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
  190. Souza, A.; Watanabe, M.D.B.; Cavalett, O.; Ugaya, C.M.L.; Bonomi, A. Social life cycle assessment of first and second-generation ethanol production technologies in Brazil. Int. J. Life Cycle Assess. 2018, 23, 617–628. [Google Scholar] [CrossRef]
  191. García, C.A.; Manzini, F. Environmental and economic feasibility of sugarcane ethanol for the Mexican transport sector. Sol. Energy 2012, 86, 1063–1069. [Google Scholar] [CrossRef]
  192. Wang, C.; Malik, A.; Wang, Y.; Chang, Y.; Pang, M.; Zhou, D. Understanding the resource-use and environmental impacts of bioethanol production in China based on a MRIO-based hybrid LCA model. Energy 2020, 203, 117877. [Google Scholar] [CrossRef]
  193. Niphadkar, S.; Bagade, P.; Ahmed, S. Bioethanol production: Insight into past, present and future perspectives. Biofuels 2018, 9, 229–238. [Google Scholar] [CrossRef]
  194. Gupta, S.; Chen, H.; Hazen, B.T.; Kaur, S.; Gonzalez, E.D.S. Circular economy and big data analytics: A stakeholder perspective. Technol. Forecast. Soc. Chang. 2019, 144, 466–474. [Google Scholar] [CrossRef]
  195. Tsegaye, B.; Balomajumder, C.; Roy, P. Microbial delignification and hydrolysis of lignocellulosic biomass to enhance biofuel production: An overview and future prospect. Bull. Natl. Res. Cent. 2019, 43, 51. [Google Scholar] [CrossRef]
  196. Castillo, R.M.; Feng, K.; Sun, L.; Guilhoto, J.; Pfister, S.; Miralles-Wilhelm, F.; Hubacek, K. The land-water nexus of biofuel production in Brazil: Analysis of synergies and trade-offs using a multiregional input-output model. J. Clean. Prod. 2019, 214, 52–61. [Google Scholar] [CrossRef]
Figure 1. Graphical representation of the different generations in bioethanol production.
Figure 1. Graphical representation of the different generations in bioethanol production.
Sustainability 16 00608 g001
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.
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.
Sustainability 16 00608 g002
Figure 3. Circular economy: the formation of value-added compounds during bioethanol production from LCB.
Figure 3. Circular economy: the formation of value-added compounds during bioethanol production from LCB.
Sustainability 16 00608 g003
Table 1. Lignocellulosic biomass composition.
Table 1. Lignocellulosic biomass composition.
Lignocellulosic BiomassCellulose, %Hemicellulose, %Lignin, %Reference
Bamboo 472328[40]
Banana waste 28.9225.2310.56[41]
Barley straw 38.4328.5516.26[42]
Corn cobs 33.637.219.3[43]
Corn stover35–4017–357–18[44]
Corn stalk 34.527.68.7[45]
Grass25–4035–5010–30[46]
Hardwood stems40–5524–4018–25[47]
Miscanthus38.3824.2317.66[48]
Municipal solid waste33–499–1610–14[21]
Oat straw31–3727–3814–19[49]
Pinewood 4028.527.7[50]
Rice straw28–36 23–28 12–15 [44]
Rye straw30.9 21.5 22.1 [51]
Softwood stems45–50 25–35 25–35 [41]
Sugarcane bagasse 42–48 19–25 20–42 [41]
Sunflower stalk 34.6 21 30 [52]
Sweet sorghum bagasse 34–45 18–27 14–21 [44]
Wheat straw33–38 26–32 17–19 [43]
Table 2. Summary on the efficiency of different pretreatments on bioethanol production from different second-generation feedstocks.
Table 2. Summary on the efficiency of different pretreatments on bioethanol production from different second-generation feedstocks.
PretreatmentFeedstocksPretreatment ConditionsPretreatment ResultsLimitationsBioethanol YieldReference
MechanicalRice strawDry ball milling
Loading charge biomass 50 g
60 min
1700 rpm
Sugar yields:
Glucose (331 mg/g rice straw)
Xylose (74 mg/g rice straw)
Arabinose (14 mg/g rice straw)
High consumption of energy.
No solubilization of lignin.
*[70]
Oil palm biomassBall milling
Loading charge biomass 20 g 50 min
250 rpm
Size reduction (88.2%)
Crystallinity reduction (46.6%)
Sugar yield:
Glucose 36.1%
Xylose 56.4%
Inhibitors produced as acetic acid (1.3 mg/ g of biomass).
High energy consumption.
*[71]
PhysicalWood biomassMicrowave treatment/DES
Choline chloride/oxalic acid
80 °C
800 W
3 min
Lignin removed 80%
Glucose yield 22.3%.
Inhibitors such as HMF are produced.*[72]
Sweet lime peel2% v/v H2SO4
Loading charge of 6%
60 min
750 W
20 kHz.
181.5 mg sugars/g biomassRelative low loading biomass charge
High biomass content increases viscosity resulting in low chemical reaction.
64% in acidic pH[73]
SwitchgrassSteam explosion
200 °C
10 min
Sugar yield:
Glucose 50.9%
Xylose 28%
Lignin removed 50%
Inhibitors such as formic acid, acetic acid, HMF, and furfural.*[74]
ChemicalSugarcane bagasseAlkaline
15% NaOH
140 °C
1 h
Reducing sugars 5.29 g/L
Lignin removed 90%
Glucose released is negatively affected by the amount of hemicellulose and lignin present.0.1 g ethanol/g biomass and 0.88 g/Lh
Sugarcane bagasseOrganosolv
Glycerol–acid 15%
130 °C
60 min
Glucan 65.8% High concentration of glycerol–acid is required to improve hydrolysis.
No significant influence on the lignin content.
0.38 g ethanol/g biomass and 0.57 g/Lh[75]
Rice strawAcid
1%H2SO4
10%(w/v) rice straw
100 °C
2 h
Reducing sugars 14 g/L
Lignin removed 11.7%
Xylose 83%
More recycling times for acid treatment hydrolysate decrease xylose yield.Concentration of 40.6 g/L yield of 86.4%[76]
Rice strawDES: choline chloride-ethylene glycol
150 °C
24 h
Biomass loading 5% w/w
Lignin removed 74%
Glucan digestibility 87%
Residual DES decreases the enzymatic hydrolysis of cellulose. *[77]
Wheat strawAlkaline-Microwave
2.75% NaOH
Solid loading 10% (w/v)
23 min
100 °C
Lignin removed 60%
Total carbohydrates 82%
High energy consumption.
Rapid oxidation of carbohydrates in alkaline conditions.
48 g ethanol/g sugar consumed[78]
Wheat strawUltrasound-assisted ionic liquid
Triethylamine hydrogen sulfate ([TBA][H2SO4])
Sonicated at 24 kHz
130 °C
30 min
Saccharification yield 76.1%
Lignin removed 74.9%
Sodium azide is used and represents a metabolic inhibitor.
Ultrasound is not effective on a bigger scale.
42 g/L[79]
Corn stalksIonic liquid
150 °C
11.5 h
2.5% arginine
420 MPa
Lignin removed 92%
Purity of cellulose reach 85%
Arginine can inhibit cellulose degradation.
Good method for fabrication of cellulose fiber but not ethanol.
*[80]
Green coconut shellsOrganosolv
80% (w/w) glycerol
1%(w/w) sulfuric acid
121 °C
Glucose 49 g/L
Lignin removed 60%
Inhibitors such as furfural and HMF are formed. 29.6 g/L[81]
Olive tree biomassCombined acid–alkaline
2.4% H2SO4
130 °C
84 min
Peroxide: 7% H2O2
80 °C
90 min
NaOH until 11.5 pH
Solubilization of hemicellulose 71%
Lignin removed 80%
Cellulose highly accessible 72%
Overliming method is needed for removing degradation products from lignocellulosic hydrolysates.
Presence of acetic acid and furfural as potential inhibitors.
15 g ethanol/100 g biomass[82]
Oil palm trunkAlkaline
Peroxide 3% H2O2
70 °C
30 min
Lignin removed 58%
Cellulose extraction 74%.
High phenolic compounds released.
Black liquor released with high amounts of tannins and gallic acid.
*[83]
Corn stoverAmmonia Recycle Percolation Process
170 °C
10% ammonia (v/w)
1 h
Lignin removed 70%
Xylan removed 47%
High temperature and energy required to improve pretreatment performance.
High operation cost.
19.4 g/L[83]
Miscanthus (Miscanthus giganteus)Acid diluted; 1% H2SO4 (v/v), 1:10 solid ratio (w/v), 121 °C
for 30 min.
Xylose 24 g/LPresence of furfural and HMF. 13.58 g/L; 0.148 g bioethanol/g dry biomass[84]
BiologicalPaddy strawWhite rot fungi
Pleurotus florida 5% inoculum 25–29 °C
Biomass loading 10% (w/v)
28 days
Saccharification efficiency of 75%High time required.
No convenience in an industrial stage.
Fungal residues limit enzymatic hydrolysis.
*[85]
Chlorella sp. KR-1Polygalacturonase, amyloglucosidase, cellulase, and β-glucanase
(simultaneous)
pH 5.5; and 45 °C; 0.3 N HCl.
28.5 g of sugar released/L of hydrolysate Inhibitors such as furfurals, HMF, and formic acid might be produced. Ethanol yield of 0.4 g/g of fermentable sugar[86]
Scenedesmus abundansH2SO4/amyloglucosidase, α-amylase (simultaneous)10.752 g of total sugars/L and 5.730 g of glucose/L of the hydrolysateSugar content released depends on cultivation and pretreatment performance.Ethanol yield of 0.1 g/g of algal biomass.[87]
* Not reported. HMF: hydroxymethylfurfural. DES: deep eutectic solvent.
Table 3. Summary of varying process and parameter conditions for different types of fermentation.
Table 3. Summary of varying process and parameter conditions for different types of fermentation.
FermentationMicroorganism/EnzymeSubstrateSubstrate ConcentrationInoculum VolumeParametersEthanol YieldTotal EthanolReference
LSFS.cerevisiae X19G2Corn Stalk 5% (w/v) 1% (w/v) 30 °C, 0 rpm, 96 h -16.48 g/L [128]
LSF S.cerevisiae NRRLY-2034Defatted Nannochloropsis oculate biomass100% (v/v) 2% (v/v) 30 °C, 0 rpm, 96 h 64.5% 5.70–6.64 g/L [129]
SScF S.cerevisiae BY4743; cellulase (Celluclast 1.5 L)Sugarcane bagasse 10% (w/v) 100 (U/g) cellulases
10% (v/v) S.cerevisiae
40 °C, 120 rpm, 4.8 pH0.49 g/g 4.88 g/L [130]
SSCF S.cerevisiae X19G2; glucoamylasePotato peel wastes 12.25% (w/v) 10% (v/v) 41.5 °C, 120 rpm, 5.5 pH, 24 h 0.26–0.33% 23.09–29.86 g/L [131]
SSCF S.cerevisiae (YSCII), T.harzianum, cellulase, β-glucosidaseEmpty fruit bunches from palm45 (w/v) 6.79%(v/v)30 °C, pH 4.8, 150 rpm, 72 h 0.46 (g/g) 9.65 g/L [125]
SSCF and SSF S.cerevisiae MTCC173, soil inoculumApple waste - 1% (w/v) S.cerevisiae;
1% (w/v) inoculum
30 °C, 6 pH, 72 h, 70.89% (w/w) moisture 49.64 g/L - [132]
SSF and SSCF S.cerevisiae; A.nigerPineapple waste 15 g/mL 1.64 × 105 (spores/mL) A.niger; 1.9 × 108 (cells/mL) S.cerevisiae30 °C, 4.1 pH, 7 days11.20% -[133]
SSF and SSCF F.oxyspurum; C.cerevisiae; glucoamylaseFood waste 0.1% (w/w) F.oxyspurum; 15 mg/g S.cerevisiae30 °C, 80 rpm, 94 h-30.8 g/L [134]
VHG S.cerevisiae C2/00Sugarcane molasses 250.34 g/L 30% (v/v) 32 °C, 4.5 pH63.91% 87.37 g/L [135]
VHG; SSCF L. sacchari LP175; Kluyveromyces marxianus DMKU-KS07; glucoamylasesSugarcane bagasse250 g/L 15% (v/v) L.sacchari and K.marxianus50 °C, 200 rpm, 6 h98.6 g/L 118 g/L [130]
Liquid-state fermentation (LSF); simultaneous saccharification and fermentation (SScF); simultaneous saccharification and co-fermentation (SSCF); solid-state fermentation (SSF); very-high-gravity fermentation (VHG).
Table 5. Techno-economic analysis and main inputs and outputs in Mexico of bioethanol production.
Table 5. Techno-economic analysis and main inputs and outputs in Mexico of bioethanol production.
ParameterCollected DataReference
Plan lifetime25 years[190]
Income tax rate30%[190]
Sugarcane price34.5USD/TC (metric ton)[190]
Ethanol price 0.52 USD/L[191]
Electricity price67.8 USD/MWh[191]
Investment 139 USD × 106[192]
Return of investment 16% per year[189]
Net present value 66 USD × 106[190]
Co-products sold (sugar in a concept of biorefinery)0.49 USD/ kg[190]
Greenhouse gases emitted 1.67 × 105 gCO2 eq/ TC (metric ton of sugarcane)[190]
Table 6. Comparison of inputs and outputs of bioethanol production in different countries.
Table 6. Comparison of inputs and outputs of bioethanol production in different countries.
Parameter1G Mexico2G China3G BruneiReference
Feedstock typeSugarcaneWheat strawAlgal biomass (Chlorella vulgaris)[183,184,189]
Bioethanol produced 77 L/metric ton of sugarcane-57,087.5805 gallons per year/210 tons biomass per year[184,189]
CO2 emitted 1.67 × 102 kgCO2 eq/metric ton of sugarcane (GHG not only CO2)1.21 tons/USD 1420 of ethanol produced 2.59 kg/gallon of bioethanol[183,184,189]
Energy input16 kWh/metric ton of sugarcane -3822 GJ per year[183,184]
Water usage 55.6 L/metric ton of sugarcane1281 m3/USD 1420 of ethanol produced2 m3/GJ−1[183,184,189,190]
Land usage -0.13 ha/USD 1420 of ethanol produced2 m2/GJ−1[183,184]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yaverino-Gutiérrez, M.A.; Wong, A.Y.C.-H.; 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. Perspectives and Progress in Bioethanol Processing and Social Economic Impacts. Sustainability 2024, 16, 608. https://doi.org/10.3390/su16020608

AMA Style

Yaverino-Gutiérrez MA, Wong AYC-H, Ibarra-Muñoz LA, Chávez ACF, Sosa-Martínez JD, Tagle-Pedroza AS, Hernández-Beltran JU, Sánchez-Muñoz S, Santos JCd, da Silva SS, et al. Perspectives and Progress in Bioethanol Processing and Social Economic Impacts. Sustainability. 2024; 16(2):608. https://doi.org/10.3390/su16020608

Chicago/Turabian Style

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, and et al. 2024. "Perspectives and Progress in Bioethanol Processing and Social Economic Impacts" Sustainability 16, no. 2: 608. https://doi.org/10.3390/su16020608

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop