Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3897 2024-01-18 20:35:28 |
2 Reference format revised. Meta information modification 3897 2024-01-19 04:33:50 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ojo, A.O. Importance of Lignocellulose in High-Value Product Production. Encyclopedia. Available online: https://encyclopedia.pub/entry/54069 (accessed on 08 May 2024).
Ojo AO. Importance of Lignocellulose in High-Value Product Production. Encyclopedia. Available at: https://encyclopedia.pub/entry/54069. Accessed May 08, 2024.
Ojo, Abidemi Oluranti. "Importance of Lignocellulose in High-Value Product Production" Encyclopedia, https://encyclopedia.pub/entry/54069 (accessed May 08, 2024).
Ojo, A.O. (2024, January 18). Importance of Lignocellulose in High-Value Product Production. In Encyclopedia. https://encyclopedia.pub/entry/54069
Ojo, Abidemi Oluranti. "Importance of Lignocellulose in High-Value Product Production." Encyclopedia. Web. 18 January, 2024.
Importance of Lignocellulose in High-Value Product Production
Edit

Lignocellulose consists of cellulose, hemicellulose, and lignin and is a sustainable feedstock for a biorefinery to generate marketable biomaterials like biofuels and platform chemicals. Enormous tons of lignocellulose are obtained from agricultural waste, but a few tons are utilized due to a lack of awareness of the biotechnological importance of lignocellulose. Underutilizing lignocellulose could also be linked to the incomplete use of cellulose and hemicellulose in biotransformation into new products. Utilizing lignocellulose in producing value-added products alleviates agricultural waste disposal management challenges. It also reduces the emission of toxic substances into the environment, which promotes a sustainable development goal and contributes to circular economy development and economic growth.

lignocellulose cellulose hemicellulose lignin value-added products

1. Introduction

Lignocellulose is a plant biomass available in large amounts, and it is a renewable resource. It is a complex structure primarily composed of the polymers cellulose and hemicellulose (polysaccharides), lignin (a phenolic macromolecule), and other components, such as proteins, lipids, and inorganic compounds [1][2][3]. These polymers contain cellulose ranging from 35 to 55%, hemicellulose from 20 to 40%, lignin (10–25%) by mass, and other polar and non-polar compounds [4]. The elemental compositions of most lignocellulosic biomass are classified as major elements (e.g., C, H, O, N, K, and Ca), minor elements (Mg, Al, Si, P, Cl, Na, S, and Fe), and trace elements (e.g., Mn and Ti) [5][6].
The cellulose in lignocellulose is a homopolysaccharide with chains of D-glucose monomers linked together via β-1-4 glycosyl units, stabilized by hydrogen bonds and van der Waals forces [1]. Cellulose comprises the repetitive structural unit called cellobiose (D-glucopyranosyl-β-1,4-D-glucopyranose) [7], and it is linked to lignin by hemicellulose via hydrogen and covalent bonds [8]. Hemicellulose is a branched heteropolysaccharide with two or more free monosaccharides, such as xylose and arabinose (five-carbon sugars), mannose, glucose, galactose (six-carbon sugars), and carboxylic acids (e.g., mannuronic acid and galacturonic acid) [8][9]. Moreover, hemicellulose bridges the deposition of lignin monomers in the secondary cell wall [9][10]. Meanwhile, lignin is a complex amorphous polymer that contains various monolignols, such as p-coumaryl alcohol, sinapyl alcohol, and coniferyl, and it is a crosslinked macromolecule formed via the polymerization of phenylpropanoid monomers (p-coumaryl alcohol, sinapyl alcohol, and coniferyl) [8][10]. Lignin is hydrophobic and highly resistant to hydrolysis; it binds hemicellulose to cellulose in the cell wall and acts as a barrier that limits cellulose accessibility [11].
Extensive studies on the physicochemical properties of lignocellulose have yet to be conducted; however, literature searches have indicated that lignocellulose physicochemical properties are assessed based on its particle size, density, flowability, moisture sorption, grindability, and thermal properties (physical properties), along with ash, volatile matter, moisture, and fixed carbon (chemical properties) [12].

2. Lignocellulosic Biomass Sources

Lignocellulosic biomass resources are widely available, and they are agricultural and forestry residues from plant wastes [13]. Industrial and food wastes are also sources of lignocellulose [14]. Several harvests from a single planting that reduce the average annual cost of managing energy crops compared to conventional crops make lignocellulosic biomass resources the most promising future resources to generate value-added products [15]. Rice, wheat, sugarcane, and maize are the major crops that generate a large amount of lignocellulosic biomass. The world’s first most important cereal crop is corn. In 2022/2023, around 1.2 billion metric tons of corn and nearly 783.8 million metric tons of wheat were produced, followed by 510 million metric tons of milled rice (the second-most important cereal crop) [16]. Approximately 177.3 million metric tons of sugarcane were also produced in 2022/2023 [16]. China and the United States of America account for more than half of worldwide corn production, while China is the world’s leading rice producer, followed by India and Bangladesh [16]. Similarly, China, followed by India, Russia, and the United States of America are the four largest wheat producers in the world, while India and Brazil are the world’s top two sugar producers [17].
Vast waste, such as rice straw, wheat straw, sugarcane bagasse, corn stover, etc., is generated annually via agricultural crop production. Rice straw (stems, leaf blades, and sheets) is generated from the rice harvest [18][19], and wheat straw is the waste obtained from wheat grain production [15]. Sugarcane waste or bagasse is obtained after sugarcane stalks are crushed for sugar [20], while corn stover (consisting of leaves, cobs, husks, and stalks) is the waste product obtained from corn kernel processing [21] from the maize plants. These wastes constitute a major portion of lignocellulosic biomass. Other agricultural wastes that contribute to a small amount of the total agricultural waste production include barley straw, cotton stalks, sweet sorghum straws, potato haulms (the tops, stems, and foliage of potato plants), and others [15].

3. Conversion of Lignocellulosic Biomass into Value-Added Products

3.1. Pretreatment Methods of Lignocellulose

The pretreatment of lignocellulose is a delignification process that makes lignocellulosic materials accessible to generate sugars. In many cases, lignocellulose was reported to be recalcitrant due to the complexity of the cell wall, lignin components, and crystalline structure of cellulose [13]. As such, selecting a suitable pretreatment method to generate sugars for the downstream application is essential. Lignocellulose can be pretreated using physical, chemical, physicochemical, biological, or nanotechnology methods [22][23][24].
Different processes used in chemical pretreatment methods include dilute acid, alkaline, ionic liquid, organosolv process, ozonolysis, and deep eutectic solvents [25]. In the dilute acid pretreatments, inorganic or organic acids, such as HCl, H2SO4, HNO3, and formic acid, break down the hydrogen and glycosidic bonds in cellulose/hemicellulose [26][27]. Bases such as NaOH, NH4OH, Ca(OH)2, and KOH are always used in alkaline pretreatment methods to solubilize lignin [3].
Physicochemical pretreatment methods such as ammonia fiber explosion (AFEX) or carbon dioxide explosion are employed where the milled or ground lignocellulose is treated with ammonia under a high temperature (e.g., 90 °C) or carbon dioxide pressure is released to disrupt the structure of cellulose [28][29]. The disruption of the structure of cellulose reduces cellulose crystallinity, enhances cellulose permeability, and increases its surface area, thereby increasing the accessibility of enzymes [30].
Biological pretreatments also offer capable commercially available microbial enzymes or crude enzymes in the delignification of lignocellulose. The different enzymes used in biological pretreatment include ligninolytic enzymes such as phenol oxidase (e.g., laccase) and heme peroxidase (e.g., lignin peroxidase). Additionally, fungi, such as white rot (e.g., Irpex lacteus, Ceriporiopsis subvermispora, and Lentinus edodes) [31][32][33], red rot (e.g., Fomitopsis annosa) [34], and brown rot (e.g., Neolentinus lepideus and Gloeophyllum trabeum) [35], have been used to attack lignin, hemicellulose and cellulose, lignin and hemicellulose, or cellulose and hemicellulose directly due to the lignolytic enzymes they produce [32][36].
Previous studies that had used white rot fungi in lignocellulose pretreatment established that: (i) two strains of Ceriporiopsis subvermispora used to pretreat wheat straw for seven weeks revealed that Ceriporiopsis subvermispora (CS), mostly CS1, showed a higher selectivity in lignin degradation than CS2, with higher laccase activity but lower manganese peroxide than C2 [37]; (ii) there was a selective degradation of lignin wheat straw and lignin oak wood chips when incubated with Ceriporiopsis subvermispora and Lentinus edodes, and alkylitaconic acids for delignification were produced by Ceriporiopsis subvermispora and Lentinus edodes [33]; (iii) there was degradation of 265 g·kg−1 of lignin and 320 g·kg−1 of neutral detergent soluble when eight different cultivars of wheat straw were incubated with Irpex lacteus for 56 days at 28 °C [38]; and (iv) the lignin content of wheat straw pretreated with Ceriporiopsis subvermispora, CS1 (CBS 347.63), at 24 °C reduced by 48.5% [39].
Lastly, the pretreatment method based on nanotechnology employs the ability of nanoparticles to penetrate the cell membrane of lignocellulose [23]. Recycling and reusing magnetic nanoparticles for subsequent cycles in lignocellulose pretreatment reduces the overall processing cost [13]. Examples of nanotechnology pretreatments are acid-functionalized magnetic nanoparticles and nano-scale shear hybrid alkaline methods. Acid-functionalized magnetic nanoparticles are strong acid nanocatalysts that effectively degrade lignocellulose [40]. In the nano-scale shear hybrid alkaline method, lignocellulose is degraded by combining chemical catalysts and the high-speed shear force [41].
Meanwhile, each pretreatment method has its pros and cons. For instance, physical methods do not generate inhibitory compounds. They can offer green pretreatments, in which the product (hydrolyzate) can be directly utilized to generate sugars. Still, physical methods, such as mechanical comminution and pyrolysis, have been considered to be too expensive for a full-scale process due to their high energy consumption; however, the main disadvantage of physical pretreatment methods is their inability to degrade the structure of lignin [24][42].
In chemical pretreatments, the hydrolysis of lignocellulose by acid alters the structure of lignin, thus resulting in high glucose yields and solubilizing hemicellulose to xylose and other sugars. The drawbacks of acid hydrolysis include the high cost of corrosive-resistant equipment and the generation of inhibitors, such as levulinic, formic, and acetic acids. Low inhibitors are produced under alkaline hydrolysis, but this process requires a long residence time and a high cost of alkaline catalysts [43].
Nanotechnology pretreatment methods have been considered the best option for delignification, as these pretreatment methods are cost-effective because the immobilized enzymes are easily retrievable and reusable [44][45]. Depending on the type of nanomaterial used, a few drawbacks of nanotechnology pretreatment methods include their potential poor dispersion abilities of some nanoparticles (due to the difficulty of dispersing in the aqueous solution, where hydronium ions are not effective) [46], and biocatalyst desorption could arise due to the weak bonds [47].

3.2. Hydrolysis of Lignocellulose

The pretreated lignocellulose is then subjected to hydrolysis. Hydrolysis is the process that liberates monomeric sugar molecules, viz. glucose, mannose, galactose, xylose, or arabinose, from structural polysaccharides, such as cellulose and hemicellulose in lignocellulose [48][49]. Cellulose hydrolysis using acids or enzymes has been reported. The first acid hydrolysis technology was developed in 1923, when a sulfuric acid solution was used to hydrolyze white spruce wood; the sugars obtained were glucose, mannose, galactose, xylose, and arabinose [50]. Inorganic acids (e.g., hydrochloric acid and hydrogen fluoride) and organic acids (e.g., citric, oxalic, and maleic acids) were also used in cellulose hydrolysis [51][52][53].
The hydrolysis of cellulose under room temperature using ca. 12 mol·L−1 of hydrochloric acid yielded approximately 32 percent of volume-reducing sugar [54], and cellulose hydrolysis with 6–7 mol·L−1 of hydrochloric acid at 90 °C in the presence of CaCl2 and LiCl as additives resulted in an 85% glucose yield [55]. When cellulose was hydrolyzed using hydrogen fluoride, the sugar yield was approximately 45% at 0 °C [53]. The most notable drawbacks of acid hydrolysis include problems in product/catalyst separation, catalyst recycling, corrosion of reactors, and waste effluent treatment that makes it environmentally unfriendly. 

3.3. Fermentation of Sugars

Fermentation is an enzyme-catalyzed biochemical process in which capable microorganisms convert sugars into new products [56], especially value-added products. Several fermentation products include biofuels like alcohol (e.g., ethanol), gases (such as methane and biogas), and organic acids (e.g., lactic, citric, succinic, and acetic acids) [15][57][58][59][60][61][62]. Producing these new products depends on the selected microorganisms and fermentation conditions. Different fermentation modes and methods have been employed in producing value-added products. These modes include batch fermentation, fed-batch fermentation, repeated-batch fermentation, and continuous fermentation [63].
Importantly, the systems are closed in batch mode, and all the required ingredients and microorganisms are added prior to fermentation. The pH is usually regulated during fermentation via an attached acid or alkaline system [63]. The fed-batch system contains the same required components as in the batch system, but during the fed-batch fermentation process, the depleted required components (e.g., carbon and nitrogen) are sequentially added at regular intervals to actively control microbial growth [64]. In repeated-batch fermentation, microbial cells are increased through repeated re-inoculation of microbial cells from one batch fermentation into the next batch [59][65].
Fermentation methods, such as separate hydrolysis and fermentation (SHF), simultaneous saccharification and fermentation (SSF), separate hydrolysis and co-fermentation (SHcoF), simultaneous saccharification and co-fermentation (SScoF), and consolidated bioprocessing (CBP), have been described in the literature [14][66] as methods that can be used during fermentation tasks. In SHF, lignocellulose is first pretreated, and following the degradation of lignin, the pretreated lignocellulose (hydrolyzate) is subjected to saccharification, followed by fermentation of the simple sugar. In SSF, the pretreated lignocellulose (hydrolyzate) is subjected to simultaneous saccharification and fermentation [14]. Separate hydrolysis and co-fermentation (SHcoF) is similar to SHF; the difference is the presence of at least two sugars for fermentation in SHcoF [67].

3.4. Purification of Value-Added Product

Purification is one of the most essential stages in value-added product production. Factors that elevate the difficulties of product recovery include, but are not limited to, the low concentration of the product, the presence of impurities, the product produced intracellularly, and heat-labile products. The extraction and purification of fermentation products depend on the specific product. The choice of purification process is based on the concentration of the product, intracellular or extracellular location of the product, physicochemical properties of the product, the impurities in the fermentation broth, acceptable standard of purity, and the product’s intended use.
The stages in the recovery/purification of products, such as organic acids (e.g., citric, lactic, and succinic acids) from fermentation broth (extracellular product), involve the removal of solid particles and microbial cells using filtration and centrifugation followed by broth extraction into different fractions [68]. Ultrafiltration, adsorption, precipitation, distillation, liquid–liquid extraction, supercritical fluid extraction, ion exchange, dialysis, electrodialysis, or membrane separation can be employed for broth extraction [68][69][70][71][72].
Moreover, biofuels can be recovered/purified using different methods. For instance, the conventional distillation process is the first step of ethanol recovery, followed by dehydration using azeotropic distillation, adsorption, pervaporation, or membrane processes [73]. Different methods like equilibrium-based separation (e.g., distillation, liquid–liquid extraction, and supercritical fluid extraction), affinity-based separation (adsorption and ion exchange), solid–liquid separation, and membrane-based separation have been employed in biodiesel purification [74]. Generally, the biodiesel purification process is wet washing using water and dry washing using adsorption, ion exchange, and membrane separation [75]

4. Value-Added Products from Lignocellulosic Biomass

4.1. Biofuels

Biofuels are an inexhaustible and biodegradable class of renewable energy obtained from living materials [76]. Biofuels are primarily used as transportation fuels and can be used to generate electricity and heat [77]. The three different generations of biofuel are: (i) first-generation biofuels (produced from edible crops); (ii) second-generation biofuels (produced from lignocellulose); and (iii) third-generation biofuels (produced from algae and microorganisms) [78]. In 2021, the United States of America produced 643,000 barrels of oil equivalent per day, followed by Brazil and Indonesia (which produced 376,000 and 140,000 barrels of oil equivalent per day, respectively) [79][80]. Additionally, biofuels are eco-friendly and capable of eliminating the emission of hazardous gases such as sulfur oxide and carbon monoxide, thereby maintaining a cleaner environment [81]. The most common biofuels include alcohols, biodiesel, biohydrogen, and biogas [15].

4.1.1. Alcohols

Alcohol is often used to denote ethanol or methanol. Since the development of the internal combustion engine, ethanol has been used as a motor fuel [82]. Bioethanol is produced from lignocellulose via pretreatment to break the recalcitrant structure of lignocellulose, followed by the enzymatic saccharification of cellulose and hemicellulose into simple sugars, and, lastly, fermentation of the generated simple sugars by microorganisms such as Saccharomyces cerevisiae, Zymomonas mobilis, and several genetically engineered microorganisms [83][84][85]. Notably, several recombinant microorganisms were developed to ferment hexose and pentose into ethanol [86][87][88]. The concentration and productivity of bioethanol depend on the lignocellulose source, the selected pretreatment method, and the microorganism(s) used in fermentation [88].

4.1.2. Biodiesel Production

Biodiesel or fatty acid methyl ester (FAME) with lower alkyl esters and long-chain fatty acids [89]. Biodiesel is a clean-burning, renewable substitute for petroleum diesel, and like petroleum diesel, it is used in diesel engines (e.g., generators and vehicles) and heating oil [90][91][92]. Pure biodiesel is called B100, and the most common blend is B20, which contains 20% biodiesel and 80% petroleum diesel [91]. Biodiesel increases energy security and improves air quality [90]. It was reported that a gallon of biodiesel (B100) produces 74% less carbon dioxide than petroleum diesel [91].
Biodiesel can be produced from second-generation biological materials, such as vegetable waste oil, non-edible vegetables, oleaginous microbes, and jatropha [89][93][94]. In biodiesel synthesis using lignocellulose, the pretreated hydrolyzate is saccharified, and oleaginous microbes such as Rhodosporidium toruloides, Gordonia sp., Yarrowia sp., Rhodotorula sp., etc., convert the generated simple sugar into pyruvate that will be further converted to lipids in the microbes [94][95]. The lipids are extracted via cell disruption using various methods that have been described by Khot (2020) [95]. The extracted lipids are then converted into biodiesel via transesterification, in which the lipid reacts with short-chain alcohols, such as methanol and ethanol, in the presence of a catalyst [89][94]. Different biodiesel production processes have been described [89][94]. Still, supercritical non-catalytic and enzymatic biodiesel production technologies are the best, as these technologies can process low-quality feedstock without pretreatment [96]. The production of biodiesel using second-generation feedstock is underdeveloped; as such, limited information on biodiesel production from lignocellulose is available.

4.1.3. Biohydrogen Production

Biohydrogen is an elementary substrate for ammonia, methane, methanol, synthesis gas, and olefin hydrogenation synthesis [97]. Commercial-scale biohydrogen production technologies are yet to be established; as such, more research focus should be directed towards biohydrogen production. Generally, biohydrogen can be produced using thermochemical, photoelectrochemical, electrolysis, and biological technologies, among which the biological method (dark fermentation) is eco-friendly and sustainable [98].

4.1.4. Biogas Production

Biogas (or biomethane) is a renewable pure energy source generated through biodigestion. Biogas has various applications in cooking, drying, cooling, and generating heat and electricity [99]. Biogas is produced through anaerobic digestion under a naturally occurring biological process that involves five steps. The steps are: (i) pretreatment of lignocellulose for easy accessibility to cellulose and hemicellulose to produce hydrolyzates; (ii) saccharification of hydrolyzate, resulting in monomers such as sugars, amino acids, and fatty acids; (iii) conversion of these monomers by acidogens into short-chain volatile fatty acids (acidogenesis); (iv) conversion of volatile fatty acids by acetogens into acetate, carbon dioxide, and hydrogen (acetogenesis); and (v) acetate, carbon dioxide, and hydrogen are converted into biomethane by methanogens (methanogenesis) [100].

4.2. Platform Chemicals

4.2.1. Fermentative Production of the Platform Chemicals from Lignocellulose

The traditional fermentative production of lactic, succinic, citric, and acetic acids from lignocellulose is completed through sequential pretreatment steps (to make cellulose and hemicellulose accessible to subsequent enzymatic hydrolysis), followed by enzymatic saccharification or hydrolysis (for the generation of simple sugars), and, finally, fermentation of the simple sugars by capable microorganisms. The microorganisms involved in simple sugar (hexose and pentose) fermentation to lactic acid include bacteria (LAB), Enterococcus faecalis, and Rhizopus sp. (for lactic acid production) [59][60][101]. Actinobaccilus succinogenes, Saccharomyces cerevisiae, and other engineered microorganisms were used to produce succinic acid [76][102].
Fermentable sugars (hexose and pentose) are metabolized to lactic, succinic, citric, and acetic acids through various microbial metabolic pathways. For instance, lactic acid can be produced via the (i) glycolytic, (ii) phosphoketolase, and (iii) pentose phosphate pathways [103]. In the glycolytic pathway, lactic acid bacteria, under anaerobic conditions, use glucose (a carbon source) to produce pyruvate, and lactate dehydrogenase catalyzes the conversion of pyruvate into lactate [103]. In the phosphoketolase pathway, glucose is converted into lactate, ethanol, and carbon dioxide, while bacteria, such as Leuconostoc sp., metabolize pentose to form lactate and acetate [103][104].
Succinic acid is biosynthesized from simple sugars via (i) reductive tricarboxylic acid (rTCA), (ii) the tricarboxylic acid (TCA) oxidation cycle; or (iii) glyoxylic pathways [105]. The rTCA pathway (the main succinic acid production pathway under anaerobic conditions) occurs by converting the simple sugar (e.g., glucose) into phosphoenolpyruvic (PEP) acid and PEP to oxaloacetic acid by PEP carboxykinase. Oxaloacetic acid is then reduced to succinic acid by malate dehydrogenase, fumarase, and fumarate reductase [57][58]. In the TCA cycle, glucose is converted into acetyl-CoA, citrate, isocitrate, and succinate by succinate dehydrogenase under aerobic conditions. The theoretical succinic acid yield of 1 mol mol−1 glucose with the release of 2 mol carbon dioxide is obtained in the TCA cycle, while in the glyoxylic pathway, the succinic acid yield is 1.71 mol mol−1 glucose due to carbon loss during the oxidative carboxylation reaction [57][58].
Citrate is produced from the aldol condensation of oxaloacetate and acetyl CoA in the Krebs cycle (known as the TCA cycle) by citrate synthase. Acetyl CoA may be derived from oxidative decarboxylation of pyruvate from glycolysis (where there is β-oxidation of fatty acids in the mitochondrial matrix) or by oxidative degradation of certain amino acids (e.g., leucine, isoleucine and threonine) [106].

4.2.2. Global Production and Market Values of the Platform Chemicals

The commercial production of lactic, succinic, citric, and acetic acids from lignocellulose has gained enormous attention. The demand for lactic acid (LA) in the past years has mainly increased due to polylactic acid (PLA) production, as LA serves as a building block for PLA production. Polylactic acid is used for drug delivery systems, prostheses, biodegradable packaging materials, and surgical suture production [107]. Furthermore, lactic acid is also used in the cosmetics, food, pharmaceutical, and chemical industries. For instance, LA is used in producing: (i) parenteral dialysis solutions, (ii) moisturizing and anti-acne creams, (iii) flavoring and preservatives, and (iv) acidulants and pH regulators [107]. In 2022, the lactic acid market volume was approximately 1.5 million metric tons, and its market value reached about USD 1.46 billion [108].
The demand for succinic acid (a dicarboxylic acid) is due to the global movement towards sustainability. In the organic and natural food industry, succinic acid is frequently used as a taste enhancer and food additive due to its ability to increase flavor and improve shelf life [109]. Succinic acid is also extensively used in the pharmaceutical industry as an excipient medicine formulation, and it is a precursor in the chemical industry to produce resins, polymers, solvents, plastics, fumaric acids, and glyoxylic acids [109][110][111]
Acetate is an anion form of acetic acid, and salts are formed by combining acetic acid with alkaline or other bases. Acetate is a vital building block in various industry applications [112]. Acetate is a coating solvent for paints and varnishes, printing inks, and nail polish [113]. It is used in the food industries as a food preservative (e.g., sodium acetate and potassium acetate) and a synthetic flavor enhancer (e.g., ethyl acetate, the ester of ethanol and acetic acid) [112]. Ethyl acetate is also used as a solvent for stains, fat, and dry cleaning [113][114], while vinyl acetate monomer (produced from the combination of acetic acid and ethene in the presence of oxygen), a building block of polyvinyl alcohol and polyvinyl acetate, is used to make packaging materials [115]. The increase in acetic acid demand results from its end-use applications, including vinyl acetate monomer and ethyl acetate. Vinyl acetate monomer accounts for 35% of global acetic acid consumption, and polyvinyl alcohol, polyvinyl acetate, and ethene vinyl acetate are the main downstream markets for vinyl acetate monomer [114]

5. Challenges and Alleviation Strategies in Upcycling Lignocellulose

Several difficulties must be mitigated to be able to fully utilize lignocellulose for economically feasible value-added product production. These difficulties include the high cost of pretreatment technology, production of inhibitors after delignification (which adversely affects the quality of the hydrolyzed sugars for fermentation), feedback inhibition, substrate inhibition, end-product inhibition, the high cost of hydrolytic enzymes, and challenges in developing efficient enzyme cocktails for the effective hydrolysis of cellulose and hemicellulose [63][116].
By-products (inhibitors) such as coumaric acid, acetic acid, formic acid, furfural, levulinic acid, and aldehyde formed during the chemical pretreatment of lignocellulose have been reported to affect microbial growth, substrate utilization, and fermentation adversely [117][118][119]. In saccharification of lignocellulose hydrolyzates, increased cellobiose and glucose concentrations could inhibit cellulase in breaking cellulose to cellobiose and glucose, thereby resulting in feedback inhibition [14]. In fermentation, challenges of substrate and end-product inhibition could occur. Substrate inhibition occurs when the fermentative microorganisms’ growth is inhibited due to a high feedstock concentration (glucose or pentose). Growth inhibition occurs due to low water activity, high osmotic pressure, and cell lysis [14].
The cost-effective operation of lignocellulose biorefinery will incline if these challenges are abated and all three constituents (cellulose, hemicellulose, and lignin) of lignocellulose are efficiently converted into value-added products. Nanotechnology, where enzyme immobilization is used in delignification, could be the best pretreatment technology for solving the bottlenecks of by-product inhibition [47]. Reports have shown that feedback inhibition can be minimized by removing sugars during saccharification using electrodialysis, avoiding cellobiose accumulation, and optimizing enzymatic activities during hydrolysis [120][121]

References

  1. Brethauer, S.; Shahab, R.L.; Studer, M.H. Impacts of Biofilms on the Conversion of Cellulose. Appl. Microbiol. Biotechnol. 2020, 104, 5201–5212.
  2. Tran, T.; Le, P.; Mai, P.; Nguyen, Q. Ethanol Production from Lignocellulosic Biomass; IntechOpen: Rijeka, Croatia, 2019; pp. 1–13.
  3. Kim, J.S.; Lee, Y.Y.; Kim, T.H. A Review on Alkaline Pretreatment Technology for Bioconversion of Lignocellulosic Biomass. Bioresour. Technol. 2016, 199, 42–48.
  4. Sharma, S.; Tsai, M.-L.; Sharma, V.; Sun, P.-P.; Nargotra, P.; Bajaj, B.K.; Chen, C.-W.; Dong, C.-D. Environment Friendly Pretreatment Approaches for the Bioconversion of Lignocellulosic Biomass into Biofuels and Value-Added Products. Environments 2022, 10, 6.
  5. Nigam, P.S.; Singh, A. Production of Liquid Biofuels from Renewable Resources. Prog. Energy Combust. Sci. 2011, 37, 52–68.
  6. Dahman, Y.; Syed, K.; Begum, S.; Roy, P.; Mohtasebi, B. Biofuels: Their Characteristics and Analysis. In Biomass, Biopolymer-Based Materials, and Bioenergy: Construction, Biomedical, and Other Industrial Applications; Elsevier: Amsterdam, The Netherlands, 2019; pp. 277–325.
  7. Zhao, Y.; Wu, B.; Yan, B.; Gao, P. Mechanism of Cellobiose Inhibition in Cellulose Hydrolysis by Cellobiohydrolase. Sci. China Ser. C Life Sci. 2004, 47, 18–24.
  8. Abo, B.O.; Gao, M.; Wang, Y.; Wu, C.; Ma, H.; Wang, Q. Lignocellulosic Biomass for Bioethanol: An Overview on Pretreatment, Hydrolysis and Fermentation Processes. Rev. Environ. Health 2019, 34, 1–12.
  9. Zhang, W.; Qin, W.; Li, H.; Wu, A.M. Biosynthesis and Transport of Nucleotide Sugars for Plant Hemicellulose. Front. Plant Sci. 2021, 12, 1–13.
  10. Doherty, W.O.S.; Mousavioun, P.; Fellows, C.M. Value-Adding to Cellulosic Ethanol: Lignin Polymers. Ind. Crops Prod. 2011, 13, 259–276.
  11. Zoghlami, A.; Paës, G. Lignocellulosic Biomass: Understanding Recalcitrance and Predicting Hydrolysis. Front. Chem. 2019, 7, 1–11.
  12. Cai, J.; He, Y.; Yu, X.; Banks, S.W.; Yang, Y.; Zhang, X.; Yu, Y.; Liu, R.; Bridgwater, A.V. Review of Physicochemical Properties and Analytical Characterization of Lignocellulosic Biomass. Renew. Sustain. Energy Rev. 2017, 76, 309–322.
  13. Singhvi, M.; Zinjarde, S.; Kim, B.S. Sustainable Strategies for the Conversion of Lignocellulosic Materials into Biohydrogen: Challenges and Solutions toward Carbon Neutrality. Energies 2022, 15, 8987.
  14. Ojo, A.O.; de Smidt, O. Lactic Acid: A Comprehensive Review of Production to Purification. Processes 2023, 11, 688.
  15. 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.
  16. Shahbandeh, M. Rice—Statistics and Facts. 2023. Available online: https://www.statista.com/topics/1443/rice/#topicOverview (accessed on 29 August 2023).
  17. World Population Review. Wheat Production by Country. 2023. Available online: https://worldpopulationreview.com/country (accessed on 29 August 2023).
  18. Goodman, B.A. Utilization of Waste Straw and Husks from Rice Production: A Review. J. Bioresour. Bioprod. 2020, 5, 143–162.
  19. Singh, R.B.; Sana, R.C.; Singh, M.; Chandra, D.; Shukla, S.G.; Walli, T.K.; Pradhan, P.K.; Kessels, H.P.P. Rice Straw—Its Production and Utilization in India. In Handbook for Straw Feeding Systems; Indian Council of Agricultural Research: New Delhi, Indian, 1995; pp. 325–337.
  20. Ajala, E.O.; Ighalo, J.O.; Ajala, M.A.; Adeniyi, A.G.; Ayanshola, A.M. Sugarcane Bagasse: A Biomass Sufficiently Applied for Improving Global Energy, Environment and Economic Sustainability. Bioresour. Bioprocess. 2021, 8, 87.
  21. Pennington, D. Corn Stover: What Is Its Worth? 2013. Available online: https://www.canr.msu.edu/news/corn_stover_what_is_its_worth (accessed on 27 August 2023).
  22. Nauman Aftab, M.; Iqbal, I.; Riaz, F.; Karadag, A.; Tabatabaei, M. Different Pretreatment Methods of Lignocellulosic Biomass for Use in Biofuel Production. In Biomass for Bioenergy—Recent Trends and Future Challenges; IntechOpen: Rijeka, Croatia, 2019; pp. 1–208.
  23. Chandel, H.; Kumar, P.; Chandel, A.K.; Verma, M.L. Biotechnological Advances in Biomass Pretreatment for Bio-Renewable Production through Nanotechnological Intervention. Biomass Convers. Biorefinery 2022, 4, 1–23.
  24. Kumar, P.; Barrett, D.M.; Delwiche, M.J.; Stroeve, P. Methods for Pretreatment of Lignocellulosic Biomass for Efficient Hydrolysis and Biofuel Production. Ind. Eng. Chem. Res. 2009, 48, 3713–3729.
  25. Baruah, J.; Nath, B.K.; Sharma, R.; Kumar, S.; Deka, R.C.; Baruah, D.C.; Kalita, E. Recent Trends in the Pretreatment of Lignocellulosic Biomass for Value-Added Products. Front. Energy Res. 2018, 6, 1–19.
  26. Sarip, H.; Sohrab Hossain, M.; Mohd Azemi, M.N.; Allaf, K. A Review of the Thermal Pretreatment of Lignocellulosic Biomass towards Glucose Production: Autohydrolysis with DIC Technology. BioResources 2016, 11, 10625–10653.
  27. Rasri, W.; Thu, V.T.; Corpuz, A.; Nguyen, L.T. Preparation and Characterization of Cellulose Nanocrystals from Corncob via Ionic Liquid Hydrolysis: Effects of Major Process Conditions on Dimensions of the Product. RSC Adv. 2023, 13, 19020–19029.
  28. Sun, Y.; Cheng, J. Hydrolysis of Lignocellulosic Materials for Ethanol Production: A Review. Bioresour. Technol. 2002, 83, 1–11.
  29. Sahay, S. Impact of Pretreatment Technologies for Biomass to Biofuel Production. In Clean Energy Production Technologies: Substrate Analysis for Effective Biofuels Production; Springer Nature: New York, NY, USA, 2020; pp. 173–216.
  30. Morais, A.R.C.; Da Costa Lopes, A.M.; Bogel-Łukasik, R. Carbon Dioxide in Biomass Processing: Contributions to the Green Biorefinery Concept. Chem. Rev. 2014, 115, 3–27.
  31. Niu, D.; Zuo, S.; Jiang, D.; Tian, P.; Zheng, M.; Xu, C. Treatment Using White Rot Fungi Changed the Chemical Composition of Wheat Straw and Enhanced Digestion by Rumen Microbiota in Vitro. Anim. Feed Sci. Technol. 2018, 237, 46–54.
  32. Wan, C.; Li, Y. Fungal Pretreatment of Lignocellulosic Biomass. Biotechnol. Adv. 2012, 30, 1447–1457.
  33. van Kuijk, S.J.A.; Sonnenberg, A.S.M.; Baars, J.J.P.; Hendriks, W.H.; del Río, J.C.; Rencoret, J.; Gutiérrez, A.; de Ruijter, N.C.A.; Cone, J.W. Chemical Changes and Increased Degradability of Wheat Straw and Oak Wood Chips Treated with the White Rot Fungi Ceriporiopsis Subvermispora and Lentinula Edodes. Biomass Bioenergy 2017, 105, 381–391.
  34. Fan, L.T.; Gharpuray, M.M.; Lee, Y.-H. Cellulose Hydrolysis; Aiba, S., Fan, L.T., Fiechter, A., Schiigerl, K.K., Eds.; Springer: Berlin, Germany, 1987.
  35. Zabel, R.A.; Morrell, J.J. Chemical Changes in Wood Caused by Decay Fungi. In Wood Microbiology; Academic Press: Cambridge, MA, USA, 2020; pp. 215–244.
  36. Abdel-Hamid, A.M.; Solbiati, J.O.; Cann, I.K.O. Insights into Lignin Degradation and Its Potential Industrial Applications. Adv. Appl. Microbiol. 2013, 82, 1–28.
  37. Nayan, N.; Sonnenberg, A.S.M.; Hendriks, W.H.; Cone, J.W. Differences between Two Strains of Ceriporiopsis Subvermispora on Improving the Nutritive Value of Wheat Straw for Ruminants. J. Appl. Microbiol. 2017, 123, 352–361.
  38. Niu, D.; Zuo, S.; Ren, J.; Li, C.; Zheng, M.; Jiang, D.; Xu, C. Effect of Wheat Straw Types on Biological Delignification and in Vitro Rumen Degradability of Wheat Straws during Treatment with Irpex Lacteus. Anim. Feed Sci. Technol. 2020, 267, 114558.
  39. Nayan, N.; Sonnenberg, A.S.M.; Hendriks, W.H.; Cone, J.W. Variation in the Solubilization of Crude Protein in Wheat Straw by Different White-Rot Fungi. Anim. Feed Sci. Technol. 2018, 242, 135–143.
  40. Arora, A.; Nandal, P.; Singh, J.; Verma, M.L. Nanobiotechnological Advancements in Lignocellulosic Biomass Pretreatment. Mater. Sci. Energy Technol. 2020, 3, 308–318.
  41. Sowmya Dhanalakshmi, C.; Madhu, P. Biofuel Production of Neem Wood Bark (Azadirachta Indica) through Flash Pyrolysis in a Fluidized Bed Reactor and Its Chromatographic Characterization. In Energy Sources, Part A: Recovery, Utilization and Environmental Effects; Bellwether Publishing, Ltd.: Columbia, MD, USA, 2021; Volume 43, pp. 428–443.
  42. Aslanzadeh, S.; Ishola, M.M.; Richards, T.; Taherzadeh, M.J. An Overview of Existing Individual Unit Operations. In Biorefineries: Integrated Biochemical Processes for Liquid Biofuels; Elsevier: Amsterdam, The Netherlands, 2014; pp. 3–36.
  43. Brodeur, G.; Yau, E.; Badal, K.; Collier, J.; Ramachandran, K.B.; Ramakrishnan, S. Chemical and Physicochemical Pretreatment of Lignocellulosic Biomass: A Review. Enzym. Res. 2011, 2011, 787532.
  44. Sánchez-Ramírez, J.; Martínez-Hernández, J.L.; Segura-Ceniceros, P.; López, G.; Saade, H.; Medina-Morales, M.A.; Ramos-González, R.; Aguilar, C.N.; Ilyina, A. Cellulases Immobilization on Chitosan-Coated Magnetic Nanoparticles: Application for Agave Atrovirens Lignocellulosic Biomass Hydrolysis. Bioprocess Biosyst. Eng. 2017, 40, 9–22.
  45. Srivastava, N.; Singh, R.; Srivastava, M.; Mohammad, A.; Harakeh, S.; Pratap Singh, R.; Pal, D.B.; Haque, S.; Tayeb, H.H.; Moulay, M.; et al. Impact of Nanomaterials on Sustainable Pretreatment of Lignocellulosic Biomass for Biofuels Production: An Advanced Approach. Bioresour. Technol. 2023, 369, 128471.
  46. Khoo, K.S.; Chia, W.Y.; Tang, D.Y.Y.; Show, P.L.; Chew, K.W.; Chen, W.H. Nanomaterials Utilization in Biomass for Biofuel and Bioenergy Production. Energies 2020, 13, 892.
  47. Tan, W.Y.; Gopinath, S.C.B.; Anbu, P.; Yaakub, A.R.W.; Subramaniam, S.; Chen, Y.; Sasidharan, S. Bio-Enzyme Hybrid with Nanomaterials: A Potential Cargo as Sustainable Biocatalyst. Sustainability 2023, 15, 7511.
  48. Xiros, C.; Topakas, E.; Christakopoulos, P. Hydrolysis and Fermentation for Cellulosic Ethanol Production. Wiley Interdiscip. Rev. Energy Environ. 2013, 2, 633–654.
  49. Stickel, J.J.; Elander, R.T.; Mcmillan, J.D.; Brunecky, R. Enzymatic Hydrolysis of Lignocellulosic Biomass. In Bioprocessing of Renewable Resources to Commodity Bioproducts; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; pp. 77–103.
  50. Sherrard, E.C.; Blanco, G.W. Some of the Products Obtained in the Hydrolysis of White Spruce Wood with Dilute Sulfuric Acid under Steam Pressure. Ind. Eng. Chem. 1923, 15, 611–616.
  51. Chen, X.; Zhang, K.; Xiao, L.P.; Sun, R.C.; Song, G. Total Utilization of Lignin and Carbohydrates in Eucalyptus Grandis: An Integrated Biorefinery Strategy towards Phenolics, Levulinic Acid, and Furfural. Biotechnol. Biofuels 2020, 13, 2.
  52. Lin, Q.; Li, H.; Ren, J.; Deng, A.; Li, W.; Liu, C.; Sun, R. Production of Xylooligosaccharides by Microwave-Induced, Organic Acid-Catalyzed Hydrolysis of Different Xylan-Type Hemicelluloses: Optimization by Response Surface Methodology. Carbohydr. Polym. 2017, 157, 214–225.
  53. Erickel, R.; Franz, R.; Woernle, R.; Riehm, T. Process for Hydrolyzing Cellulose-Material with Gaseous Hydrogen Fluoride. 1985. Available online: https://patentimages.storage.googleapis.com/7c/c3/7c/06812146dfef03/US4556431.pdf (accessed on 19 July 2023).
  54. Bergius, F. Conversion of Wood to Carbohydrates and Problems in the Industrial Use of Concentrated Hydrochloric Acid. Ind. Eng. Chem. 1937, 29, 247–253.
  55. Ragg, P.L.; Fields, P.R. The Development of a Process for the Hydrolysis of Lignocellulosic Waste. Math. Phys. Sci. 1987, 321, 537–547.
  56. Taveira, I.C.; Nogueira, K.M.V.; De Oliveira, D.L.G.; Silva, R.D.N. Preservation and Fermentation: Past, Present and Future. Front. Young Minds 2021, 79, 3–16.
  57. Dai, Z.; Guo, F.; Zhang, S.; Zhang, W.; Yang, Q.; Dong, W.; Jiang, M.; Ma, J.; Xin, F. Bio-Based Succinic Acid: An Overview of Strain Development, Substrate Utilization, and Downstream Purification. Biofuels Bioprod. Biorefining 2019, 14, 965–985.
  58. Cheng, K.K.; Wang, G.Y.; Zeng, J.; Zhang, J.A. Improved Succinate Production by Metabolic Engineering. Biomed Res. Int. 2013, 2013, 538790.
  59. Wee, Y.J.; Yun, J.S.; Kim, D.; Ryu, H.W. Batch and Repeated Batch Production of L(+)-Lactic Acid by Enterococcus Faecalis RKY1 Using Wood Hydrolyzate and Corn Steep Liquor. J. Ind. Microbiol. Biotechnol. 2006, 33, 431–435.
  60. Takano, M.; Hoshino, K. Lactic Acid Production from Paper Sludge by SSF with Thermotolerant Rhizopus Sp. Bioresour. Bioprocess. 2016, 3, 29.
  61. Li, Z.; Lu, J.K.; Yang, Z.X.; Han, L.; Tan, T. Utilization of White Rice Bran for Production of L-Lactic Acid. Biomass Bioenergy 2012, 39, 53–58.
  62. Kim, H.M.; Choi, I.S.; Lee, S.; Yang, J.E.; Jeong, S.G.; Park, J.H.; Ko, S.H.; Hwang, I.M.; Chun, H.H.; Wi, S.G.; et al. Biorefining Process of Carbohydrate Feedstock (Agricultural Onion Waste) to Acetic Acid. ACS Omega 2019, 4, 22438–22444.
  63. Abdel-Rahman, M.A.; Tashiro, Y.; Sonomoto, K. Lactic Acid Production from Lignocellulose-Derived Sugars Using Lactic Acid Bacteria: Overview and Limits. J. Biotechnol. 2011, 156, 286–301.
  64. Paulova, L.; Chmelik, J.; Branska, B.; Patakova, P.; Drahokoupil, M.; Melzoch, K. Comparison of Lactic Acid Production by L. Casei in Batch, Fed-Batch and Continuous Cultivation, Testing the Use of Feather Hydrolysate as a Complex Nitrogen Source. Braz. Arch. Biol. Technol. 2020, 63, 1–12.
  65. Zhao, B.; Wang, L.; Ma, C.; Yang, C.; Xu, P.; Ma, Y. Repeated Open Fermentative Production of Optically Pure L-Lactic Acid Using a Thermophilic Bacillus Sp. Strain. Bioresour. Technol. 2010, 101, 6494–6498.
  66. Rawoof, S.A.A.; Kumar, P.S.; Vo, D.-V.N.; Devaraj, K.; Mani, Y.; Devaraj, T.; Subramanian, S. Production of Optically Pure Lactic Acid by Microbial Fermentation: A Review. Environ. Chem. Lett. 2020, 19, 539–556.
  67. Raina, N.; Slathia, P.S.; Sharma, P. Experimental Optimization of Thermochemical Pretreatment of Sal (Shorea Robusta) Sawdust by Central Composite Design Study for Bioethanol Production by Co-Fermentation Using Saccharomyces Cerevisiae (MTCC-36) and Pichia Stipitis (NCIM-3498). Biomass Bioenergy 2020, 143, 105819.
  68. Mores, S.; Vandenberghe, L.P.d.S.; Magalhães Júnior, A.I.; de Carvalho, J.C.; de Mello, A.F.M.; Pandey, A.; Soccol, C.R. Citric Acid Bioproduction and Downstream Processing: Status, Opportunities, and Challenges. Bioresour. Technol. 2021, 320, 124426.
  69. Alexandri, M.; Schneider, R.; Mehlmann, K.; Venus, J. Recent Advances in D-Lactic Acid Production from Renewable Resources: Case Studies on Agro-Industrial Waste Streams. Food Technol. Biotechnol. 2019, 57, 293–304.
  70. Kumar, A.; Thakur, A.; Panesar, P.S. Lactic Acid and Its Separation and Purification Techniques: A Review. Rev. Environ. Sci. Biotechnol. 2019, 18, 823–853.
  71. Pleissner, D.; Neu, A.K.; Mehlmann, K.; Schneider, R.; Puerta-Quintero, G.I.; Venus, J. Fermentative Lactic Acid Production from Coffee Pulp Hydrolysate Using Bacillus Coagulans at Laboratory and Pilot Scales. Bioresour. Technol. 2016, 218, 167–173.
  72. Glassner, D.A.; Elankovan, P.; Beacom, D.R.; Berglund, K.A. Purification Process for Succinic Acid Produced by Fermentation. Appl. Biochem. Biotechnol. 1995, 51–52, 73–82.
  73. Saini, S.; Chandel, A.K.; Sharma, K.K. Past Practices and Current Trends in the Recovery and Purification of First Generation Ethanol: A Learning Curve for Lignocellulosic Ethanol. J. Clean. Prod. 2020, 268, 122357.
  74. Bateni, H.; Saraeian, A.; Able, C. A Comprehensive Review on Biodiesel Purification and Upgrading. Biofuel Res. J. 2017, 15, 668–690.
  75. Mma, S.; Jaber, R.; Shirazi, M.; Toufaily, J.; Hamieh, A.; Noureddin, A.; Ghanavati, H.; Ghaffari, A.; Zenouzi, A.; Karout, A.; et al. Biodiesel Wash-Water Reuse Using Microfiltration: Toward Zero-Discharge Strategy for Cleaner and Economized Biodiesel Production. Biofuel Res. J. 2015, 5, 148–151.
  76. BYJU’S. Types of Fermentation: Definition, Process, Advantages. 2023. Available online: https://byjus.com/biology/biofuel (accessed on 5 July 2023).
  77. U.S. Energy Information Administration. Biofuel Explained. 2021. Available online: https://www.eia.gov/energyexplained/biofuel/biodiesel-rd-other-use-supply.php (accessed on 15 September 2023).
  78. Bardhan, S.K.; Gupta, S.; Gorman, M.E.; Haider, M.A. Biorenewable Chemicals: Feedstocks, Technologies and the Conflict with Food Production. Renew. Sustain. Energy Rev. 2015, 51, 506–520.
  79. Aizarani, J. Fuel Ethanol Production Worldwide in 2022, by Country. 2023. Available online: https://www.statista.com/statistics/281606/ethanol-production-in-selected-countries/Source:https://www.statista.com/statistics/281606/ethanol-production-in-selected-countries/ (accessed on 25 August 2023).
  80. Gitnux. Biofuel Production: Statistics and Trends. 2023. Available online: https://blog.gitnux.com/biofuel-production-statistics/ (accessed on 28 August 2023).
  81. Suhud Shote, A. Biofuel: An Environmental Friendly Fuel. In Anaerobic Digestion; IntechOpen: Rijeka, Croatia, 2019; pp. 1–13.
  82. Ghosh, B.B.; Ahindra, N. Biofuels Refining and Performance. In Biofuels Refining and Performance, 1st ed.; McGraw-Hill: New York, NY, USA, 2008.
  83. Park, J.M.; Oh, B.R.; Seo, J.W.; Hong, W.K.; Yu, A.; Sohn, J.H.; Kim, C.H. Efficient Production of Ethanol from Empty Palm Fruit Bunch Fibers by Fed-Batch Simultaneous Saccharification and Fermentation Using Saccharomyces Cerevisiae. Appl. Biochem. Biotechnol. 2013, 170, 1807–1814.
  84. Braga, A.; Gomes, D.; Rainha, J.; Amorim, C.; Cardoso, B.B.; Gudiña, E.J.; Silvério, S.C.; Rodrigues, J.L.; Rodrigues, L.R. Zymomonas Mobilis as an Emerging Biotechnological Chassis for the Production of Industrially Relevant Compounds. Bioresour. Bioprocess. 2021, 8, 128.
  85. Piriya, P.S.; Vasan, P.T.; Padma, V.S.; Vidhyadevi, U.; Archana, K.; Vennison, S.J. Cellulosic Ethanol Production by Recombinant Cellulolytic Bacteria Harbouring Pdc and Adh II Genes of Zymomonas Mobilis. Biotechnol. Res. Int. 2012, 2012, 817549.
  86. Fernández-Sandoval, M.T.; Galíndez-Mayer, J.; Bolívar, F.; Gosset, G.; Ramírez, O.T.; Martinez, A. Xylose-Glucose Co-Fermentation to Ethanol by Escherichia coli Strain MS04 Using Single- and Two-Stage Continuous Cultures under Micro-Aerated Conditions. Microb. Cell Factories 2019, 18, 145.
  87. Saha, B.C.; Cotta, M.A. Ethanol Production from Lignocellulosic Biomass by Recombinant Escherichia coli Strain FBR5. Bioengineered 2012, 3, 197–202.
  88. Sierra-Ibarra, E.; Alcaraz-Cienfuegos, J.; Vargas-Tah, A.; Rosas-Aburto, A.; Valdivia-López, Á.; Hernández-Luna, M.G.; Vivaldo-Lima, E.; Martinez, A. Ethanol Production by Escherichia coli from Detoxified Lignocellulosic Teak Wood Hydrolysates with High Concentration of Phenolic Compounds. J. Ind. Microbiol. Biotechnol. 2022, 49, 77.
  89. Neupane, D. Biofuels from Renewable Sources, a Potential Option for Biodiesel Production. Bioengineering 2023, 10, 29.
  90. Alternative Fuels Data Centre (AFDC). Biodiesel Benefits and Considerations. 2020. Available online: https://afdc.energy.gov/fuels/biodiesel_benefits.html (accessed on 29 September 2023).
  91. Smoot, G. What Is the Carbon Footprint of Biofuel? A Life-Cycle Assessment. 2020. Available online: https://impactful.ninja/the-carbon-footprint-of-biofuel/ (accessed on 13 August 2023).
  92. Integrated Flow Solutions (IFS). Biodiesel Guide-Sources, Production, Uses, and Regulations. 2020. Available online: https://ifsolutions.com/category/power-generation/ (accessed on 29 September 2023).
  93. Climate Technology Centre and Network. Second and Third Generation of Biofuels. 2011. Available online: https://www.ctc-n.org/files/UNFCCC_docs/ref17x03_3.pdf (accessed on 18 September 2023).
  94. Chintagunta, A.D.; Zuccaro, G.; Kumar, M.; Kumar, S.P.J.; Garlapati, V.K.; Postemsky, P.D.; Kumar, N.S.S.; Chandel, A.K.; Simal-Gandara, J. Biodiesel Production From Lignocellulosic Biomass Using Oleaginous Microbes: Prospects for Integrated Biofuel Production. Front. Microbiol. 2021, 12, 1–23.
  95. Khot, M.; Raut, G.; Ghosh, D.; Alarcón-Vivero, M.; Contreras, D.; Ravikumar, A. Lipid Recovery from Oleaginous Yeasts: Perspectives and Challenges for Industrial Applications. Fuel 2020, 259, 116292.
  96. Pasha, M.K.; Dai, L.; Liu, D.; Guo, M.; Du, W. An Overview to Process Design, Simulation and Sustainability Evaluation of Biodiesel Production. Biotechnol. Biofuels 2021, 14, 192.
  97. Díaz-González, A.; Luna, M.Y.P.; Morales, E.R.; Saldaña-Trinidad, S.; Blanco, L.R.; de la Cruz-Arreola, S.; Pérez-Sariñana, B.Y.; Robles-Ocampo, J.B. Assessment of the Pretreatments and Bioconversion of Lignocellulosic Biomass Recovered from the Husk of the Cocoa Pod. Energies 2022, 15, 3544.
  98. Bhatia, S.K.; Jagtap, S.S.; Bedekar, A.A.; Bhatia, R.K.; Rajendran, K.; Pugazhendhi, A.; Rao, C.V.; Atabani, A.E.; Kumar, G.; Yang, Y.H. Renewable Biohydrogen Production from Lignocellulosic Biomass Using Fermentation and Integration of Systems with Other Energy Generation Technologies. Sci. Total Environ. 2021, 765, 144429.
  99. Environmental and Energy Study Institute (EESI). Biogas: Converting Waste to Energy. 2017. Available online: www.eesi.org (accessed on 23 August 2023).
  100. Xu, N.; Liu, S.; Xin, F.; Zhou, J.; Jia, H.; Xu, J.; Jiang, M.; Dong, W. Biomethane Production from Lignocellulose: Biomass Recalcitrance and Its Impacts on Anaerobic Digestion. Front. Bioeng. Biotechnol. 2019, 7, 191.
  101. Liu, W.; Bao, Q.; Jirimutu; Qing, M.; Siriguleng; Chen, X.; Sun, T.; Li, M.; Zhang, J.; Yu, J.; et al. Isolation and Identification of Lactic Acid Bacteria from Tarag in Eastern Inner Mongolia of China by 16S RRNA Sequences and DGGE Analysis. Microbiol. Res. 2012, 167, 110–115.
  102. Akhtar, J.; Hassan, N.; Idris, A.; Ngadiman, N.H.A. Optimization of Simultaneous Saccharification and Fermentation Process Conditions for the Production of Succinic Acid from Oil Palm Empty Fruit Bunches. J. Wood Chem. Technol. 2020, 40, 136–145.
  103. Wang, Y.; Wu, J.; Lv, M.; Shao, Z.; Hungwe, M.; Wang, J.; Bai, X.; Xie, J.; Wang, Y.; Geng, W. Metabolism Characteristics of Lactic Acid Bacteria and the Expanding Applications in Food Industry. Front. Bioeng. Biotechnol. 2021, 9, 1–19.
  104. Liu, S.-Q.; Holland, R. LEUCONOSTOC spp. In Encyclopedia of Dairy Sciences; Elsevier: Amsterdam, The Netherlands, 2002; pp. 1539–1543.
  105. Zhou, S.; Zhang, M.; Zhu, L.; Zhao, X.; Chen, J.; Chen, W.; Chang, C. Hydrolysis of Lignocellulose to Succinic Acid: A Review of Treatment Methods and Succinic Acid Applications. Biotechnol. Biofuels Bioprod. 2023, 16, 1.
  106. Williams, N.C.; O’Neill, L.A.J. A Role for the Krebs Cycle Intermediate Citrate in Metabolic Reprogramming in Innate Immunity and Inflammation. Front. Immunol. 2018, 9, 141.
  107. Cubas-Cano, E.; González-Fernández, C.; Ballesteros, M.; Tomás-Pejó, E. Biotechnological Advances in Lactic Acid Production by Lactic Acid Bacteria: Lignocellulose as Novel Substrate. Biofuels Bioprod. Biorefining 2018, 12, 290–303.
  108. Statista Research Department. Lactic Acid Global Market Volume. 2023. Available online: https://www.statista.com/statistics/1310495/lactic-acid-market-volume-worldwide/Source:https://www.statista.com/statistics/1310495/lactic-acid-market-volume-worldwide/ (accessed on 30 July 2023).
  109. Market.us Study. Succinic Acid Market Size and Value to Reach USD 359.8 Million in 2032. 2023. Available online: https://www.globenewswire.com/en/news-release/2023/05/09/2664413/0/en/Succinic-Acid-Market-Size-and-Value-to-Reach-USD-359-8-Million-in-2032-Growing-at-CAGR-of-7-3-Market-us-Study.html (accessed on 10 September 2023).
  110. Maleki, H.; Azimi, B.; Ismaeilimoghadam, S.; Danti, S. Poly(Lactic Acid)-Based Electrospun Fibrous Structures for Biomedical Applications. Appl. Sci. 2022, 12, 3192.
  111. Grand View Research. Succinic Acid Market Size, Share & Trends Analysis Report. 2021. Available online: https://www.grandviewresearch.com/industry-analysis/succinic-acid-market (accessed on 30 September 2023).
  112. Harahap, B.M.; Ahring, B.K. Acetate Production from Syngas Produced from Lignocellulosic Biomass Materials along with Gaseous Fermentation of the Syngas: A Review. Microorganisms 2023, 11, 995.
  113. Marino, D.J. Ethyl Acetate. 2005. Available online: https://www.sciencedirect.com/science/article/pii/B0123694000003902 (accessed on 14 September 2023).
  114. MC GROUP. Ethyl Acetate (ETAC): 2023 World Market Outlook and Forecast up to 2032. 2023. Available online: https://mcgroup.co.uk/researches/acetic-acid (accessed on 25 September 2023).
  115. McKeen, L.W. Polyolefins, Polyvinyls, and Acrylics. In Permeability Properties of Plastics and Elastomers; Elsevier: Amsterdam, The Netherlands, 2012; pp. 145–193.
  116. Singh, B.; Korstad, J.; Guldhe, A.; Kothari, R. Editorial: Emerging Feedstocks and Clean Technologies for Lignocellulosic Biofuel. Front. Energy Res. 2022, 10, 1–2.
  117. Zaldivar, J.; Martinez, A.; Ingram, L.O. Effect of Selected Aldehydes on the Growth and Fermentation of Ethanologenic Escherichia coli. Biotechnol. Bioeng. 1999, 65, 24–33.
  118. Van Der Pol, E.C.; Eggink, G.; Weusthuis, R.A. Production of L(+)-Lactic Acid from Acid Pretreated Sugarcane Bagasse Using Bacillus Coagulans DSM2314 in a Simultaneous Saccharification and Fermentation Strategy. Biotechnol. Biofuels 2016, 9, 248.
  119. Jönsson, L.J.; Martín, C. Pretreatment of Lignocellulose: Formation of Inhibitory by-Products and Strategies for Minimizing Their Effects. Bioresour. Technol. 2016, 199, 103–112.
  120. Shen, X.; Xia, L. Production and Immobilization of Cellobiase from Aspergillus Niger ZU-07. Process Biochem. 2004, 39, 1363–1367.
  121. Ou, M.S.; Mohammed, N.; Ingram, L.O.; Shanmugam, K.T. Thermophilic Bacillus Coagulans Requires Less Cellulases for Simultaneous Saccharification and Fermentation of Cellulose to Products than Mesophilic Microbial Biocatalysts. Appl. Biochem. Biotechnol. 2009, 155, 379–385.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 117
Revisions: 2 times (View History)
Update Date: 19 Jan 2024
1000/1000