Next Article in Journal
Targeting of Nrf2/PPARγ/NLRP3 Signaling Pathway by Stevia rebudiana Bertoni Extract Provides a Novel Insight into Its Protective Effect against Acute Gouty Arthritis-Induced Synovial Inflammation, Oxidative Stress and Apoptosis in a Rat Model
Next Article in Special Issue
Miniaturized Solid Phase Extraction Techniques Applied to Natural Products
Previous Article in Journal
Investigation on Spectral Characteristics of Gliding Arc Plasma Assisted Ammonia Lean Combustion
Previous Article in Special Issue
A Systematic Review on Waste as Sustainable Feedstock for Bioactive Molecules—Extraction as Isolation Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extraction, Isolation, and Purification of Value-Added Chemicals from Lignocellulosic Biomass

by
Tanmay Chaturvedi
*,
Laura Sini Sofia Hulkko
,
Malthe Fredsgaard
and
Mette Hedegaard Thomsen
Department of Energy, Aalborg University, 6700 Esbjerg, Denmark
*
Author to whom correspondence should be addressed.
Processes 2022, 10(9), 1752; https://doi.org/10.3390/pr10091752
Submission received: 14 July 2022 / Revised: 27 July 2022 / Accepted: 3 August 2022 / Published: 2 September 2022

Abstract

:
This review covers the operating conditions for extracting top value-added chemicals, such as levulinic acid, lactic acid, succinic acid, vanillic acid, 3-hydroxypropionic acid, xylitol, 2,5-furandicarboxylic acid, 5-hydroxymethyl furfural, chitosan, 2,3-butanediol, and xylo-oligosaccharides, from common lignocellulosic biomass. Operating principles of novel extraction methods, beyond pretreatments, such as Soxhlet extraction, ultrasound-assisted extraction, and enzymatic extraction, are also presented and reviewed. Post extraction, high-value biochemicals need to be isolated, which is achieved through a combination of one or more isolation and purification steps. The operating principles, as well as a review of isolation methods, such as membrane filtration and liquid–liquid extraction and purification using preparative chromatography, are also discussed.

1. Introduction

Lignocellulosic biomass is the most widely available feedstock for biofuel production. As a second-generation feedstock, lignocellulosic biomass does not compete with crops used for food products, such as corn, sugarcane, beetroot, and others. Traditionally, large-scale biomass processing facilities have focused on two main bio-based products, i.e., biofuels and bioenergy. Biofuels include ethanol, butanol, biodiesel, etc. [1,2]. Methane from biogas plants and syngas constitute the primary bioenergy products obtained from biomass processing [3,4]. Whereas these products have spurred the quest for cleaner fuels, they have fallen short of presenting a sustainable business model for production of high-value biomass-derived products. Moving away from single feedstock to a single-product approach leads toward the contemporary biorefinery approach, whereby multi feedstock processing leads to multiple bio-based products, including but not limited to biofuel, bioenergy, biochemicals, proteins, and other high-value bioproducts.
In 2004, the U.S. Department of Energy (USDOE) identified the top 12 platform chemicals (based on their market potential) that can be derived from biomass [5]. Four carbon (C4) 1,4-dicarboxylic acids (succinic, fumaric, and malic acid), 2,5-furandicarboxylic acid, 3-hydroxypropionic acid, aspartic acid, glucaric acid, glutamic acid, itaconic acid, 3-hydroxybutyrolactone, glycerol, sorbitol, and xylitol were identified as the top 12 candidates that can subsequently be converted into numerous high-value biomass-derived products. ‘Top Value Added Chemicals from Biomass’ was the first volume of a two-volume report, with the second volume published in 2007, in which the list was updated to include lignin-based derivatives, such as vanillin, vanillic acid, syringaldehyde, aromatic diacids, and quinones, amongst others [6]. Several promising technologies to obtain these building block/intermediate chemicals are identified in this report based on technological maturity and reported product yields.

2. Methodology

In this paper, we build on the foundation laid by these two reports and aim to consolidate the literature with respect to the effectiveness of extracting some of these value-added chemicals from the most common lignocellulosic biomasses. The most common lignocellulosic biomasses include corn stover, sugarcane bagasse, pine, wheat straw, rice straw, softwood, aspen wood, etc. This review identifies which value-added chemicals can be derived from the most common lignocellulosic biomass and presents their respective yields. Furthermore, we evaluate the technologies and the operating conditions used to extract value-added chemicals from the most common lignocellulosic biomasses. Subsequently, methods for the isolation of chemicals are also discussed.
Three factors govern the processing of lignocellulosic biomass: availability of the type of biomass, accessibility and maturity of processing technology, and the intended bioproduct and its market demand. We begin this review begins by identifying the critical value-added chemicals based on their availability and potential to be extracted from lignocellulosic biomass. Next, the literature-reported concentrations, yield, and productivity of extracting value-added chemicals from some of the most common lignocellulosic biomasses are presented. Lastly, the most promising technologies for the isolation of bioproducts are reported.
Lignocellulosic biomass can be broken down into three major building blocks: cellulose, hemicellulose, and lignin. Value-added chemicals can be categorized according to their functional groups of origin, and based on processing routes as shown in Figure 1.
In this paper, we will discuss the following derivatives of cellulose:
  • Levulinic acid;
  • Lactic acid;
  • 3-hydroxypropionic acid (3-HP);
  • Succinic acid;
  • Vanillic acid and vanillin;
  • Itaconic acid;
  • Adipic acid;
  • 2,5-furandicarboxylic acid (FDCA); and
  • 5-hydroxymethylfurfural (HMF).
The value-added chemicals extracted from hemicellulose and lignin discussed in this paper are:
  • Xylitol;
  • Furfural;
  • Chitosan;
  • 2,3-butanediol (2,3-BD); and
  • Xylo-oligosaccharides (XOs)

3. Top Value-Added Chemicals

3.1. Levulinic Acid

Levulinic acid can be extracted by the dehydration of sugars, hydration of hydroxymethylfurfural (HMF) or hydrolysis of furfuryl alcohol, both of which are derived from xylose, which is a hemicellulose sugar [5]. A simple structure of levulinic acid is seen in Figure 2. Levulinic acid has applications in additives, pharmaceutical, and plastic industries [7]. Biofine Technology (Boston, MA, USA), GFBiochemicals (Paris, France), and Avantium (Amsterdam, The Netherlands) are companies that are invested in the commercial production of levulinic acid. Table 1 summarizes the levulinic acid yield extracted from lignocellulosic biomass under optimized pretreatment conditions (acid concentration, time, and operating temperature). Levulinic acid can serve as a precursor for succinic acid, diphenolic acid, valeric acid, γ-valerolactone, acetyl acrylic acid, 1,4-butanediol, and other value-added chemicals [8,9]. The ease of deriving levulinic acid from a variety of lignocellulosic crops and its important position in the supply chain as an intermediate for the production of resins, herbicides, plasticizers, solvents, fuels, food, flavoring, and fragrance components makes it one of the top value-added chemicals that can be derived in a lignocellulosic biorefinery.

3.2. Lactic Acid

Lactic acid (structure in Figure 3) has become increasingly popular as a biomass-derived chemical due to its utility for production of polylactic acid (PLA). Currently, Corbin (Amsterdam, The Netherlands), Futerro (Escanaffles, Belgium), NatureWorks (Minnetonka, MN, USA), and Myriant (Quincy, MA, USA) are commercial-scale users of lactic acid [7]. Some of the most common lactic acid bacteria used for lactic acid production from lignocellulosic-derived sugars include L. planterum, L. pentosus, L. delbrueckii, L. casei, L. brevis, E. mundtii, E. faecalis, L. coryniformis, L. rhamnosus, L. salivarius, L. amylovorans, and L. amylophilus, amongst others [26]. Lactic acid also serves as a critical platform chemical for production of lactide (the intermediate for polylactic acid production), propanoic acid, 1,2-propanediol, polyurethanes, pyruvic acid, acrylic acid, 2,3-pentanedione, and others. Lactic acid concentration, yield, and productivity from some of the most common lignocellulosic biomasses are shown in Table 2.

3.3. 3-Hydroxypropionic Acid

3-hydroxypropionic acid (3-HP) is an important C3 platform chemical, primarily due to its contribution as precursor for the production of 1,3-propanediol. 3-HP is also a platform chemical used for the production of malonic acid, acrylic acid, acrylonitrile, polyamides, and 3-hydroxypropionate esters and its structure can be seen in Figure 4. BASF-Cargill-Novozymes (Ludwigshafen, Germany/Wayzata, MN, USA/Bagsværd, Denmark), and Dow (Midland, MI, USA) are commercial-scale producers of 3-HP [7].
3-HP can be produced via microbial fermentation pathways from two major substrates: glucose and glycerol [39]. Matsakas et al. [40] presented a comprehensive review on 3-HP production and its production pathway mechanisms. Table 3 provides an overview of how various microorganism compare in terms of their concentrations and productivity of 3-HP, with glycerol and glucose as the two main substrate sources.

3.4. Succinic Acid

Succinic acid is a four-carbon dicarboxylic acid (structure in Figure 5) that has been produced via chemical routes in the past but is gaining popularity for production via the biological route. Myriant Technologies now PTT Global Chemical (Bangkok, Thailand) and Reverdia (Utrecht, the Netherlands) have commercial-scale facilities for the production of bio-based succinic acid [7].
Succinic acid is a key intermediate chemical used for the production of several derivatives of industrial importance, such as 1,4-butanediol, tetrahydrofuran, N-methylpyrrolidone, and γ-butyrolactone. These derivatives are, in turn, utilized in the production of polyurethanes, polyesters, and polyvinylpyrrolidone (PVP). Succinic acid can be extracted from various lignocellulosic biomasses. Table 4 provides an overview of succinic acid concentrations, alongside the concentrations of other value-added chemicals that can be derived simultaneously under varying pretreatment conditions, from some of the most common lignocellulosic biomasses.

3.5. Vanillic Acid and Vanillin

Vanillin is a widely used food flavoring agent that is most typically extracted from Vanilla spp.; however, it is currently produced inexpensively via petrochemical routes (structure of vanillic acid and vanillin in Figure 6). Besides the food industry, vanillin finds applications in the pharmaceutical and fragrance industries. The increasing demand for this molecule has propelled the search for biomass-derived pathways for vanillin production.
Vanillin can be extracted from lignin, ferulic acid, glucose, vanillic acid, aromatic amino acids, isoeugenol, waste residue, and other substrates [72]. Currently, Borregaard (Sarpsborg, Norway), a Norwegian company, claims to be the only producer of biovanillin, which it produces from wood. Table 4 summarizes some lignocellulosic biomasses that can be used to derive vanillin and vanillic acid.

3.6. Itaconic Acid

Itaconic acid is a C5 dicarboxylic acid (see Figure 7) that does not have as large a market share as the likes of succinic acid, levulinic acid, and lactic acid. However, it remains an intermediate building block of interest due to its significance in producing other value-added chemicals with larger market shares. Itaconic acid is a precursor for polymethyl methacrylate (PMMA), 3-methyltetrahydrofuran, polyitaconic acid, and styrene-butadiene rubber latex. Additionally, itaconic acid can be converted to methyl pyrrolidones, 2-methylbutanediol, 3-methyltetrahydrofuran, 4-methyl-γ-butyrolactone, and 4-methyl-γ-butyrolactone. Itaconic acid is recommended as a replacement for maleic acid/anhydride and sodium tripolyphosphate, which are, in turn, used for the production of polyester resin and detergent [5].
Qingdao Kehai Biochemistry (Jiaonan City, China), Zhejiang Guoguang Biochemistry (Quzhou City, China), Jinan Huaming Biochemistry (Mingshui Zhangqiu City, China), and Itaconix (Stratham, NH, USA) are producers of itaconic acid on a commercial scale [7]. Table 4 provides a summary of lignocellulosic biomasses and their respective pretreatment conditions for extraction of itaconic acid.

3.7. Adipic Acid

Adipic acid is a C6 dicarboxylic acid (see Figure 8) with applications in the production of nylon -6,6 fibers, resins, plasticizers, polyester polyols, food ingredients, and lubricants [5].
Adipic acid is also a precursor for the production of muconic acid and glucaric acid. Biomass-derived adipic acid can be obtained from glucose as a starting substrate and E. coli, S. cerevisiae, and P. putida as the chassis [73]. Lignin, lipids, xylose, and amino acids can also serve as substrates to obtain bio-derived adipic acid. Genomatica (San Diego, CA, USA) and DSM (Heerlen, The Netherlands) are developing strategies to produce commercial-scale quantities of biomass-derived adipic acid [74].

3.8. Furfural

Furfural is a C5 molecule (as shown in Figure 9) most often derived from the hemicellulose fraction in lignocellulosic biomass. Furfural is a top value-added chemical with several commercial-scale facilities operating across China, South Africa, and the Dominican Republic [75]. The largest fractions of furfural are used in the production of furfuryl alcohol, which, in turn, is used in resin production.
Furfural is the precursor for the production of furoic acid, tetrahydrofuran, fumaric acid, 2-methyltetrahydrofuran, tetrahydrofurfuryl alcohol, and furfurylamine [7]. The highest reported furfural yields from lignocellulosic biomass vary from 34 to 87 weight% (wt%) [76,77,78]. Table 4 highlights the concentrations of furfural obtained after pretreatment from some of the most common lignocellulosic biomasses.

3.9. 5-Hydroxymethylfurfural (HMF)

5-hydroxymethylfurfural (HMF) is a C6 derivative obtained through dehydration of glucose and fructose, a structure of which can be seen in Figure 10 [79]. For lignocellulosic biomasses, cellulose is the main contributor to HMF production. HMF production, in principle, is possible from all biomasses containing hexoses and its oligomers, providing a wide range of possible feedstock for HMF production [75].
HMF is a precursor for the production of adipic acid, levulinic acid, 2,5-dimethylfuran, caprolactone, polyamide 6, 2,5-furandicarboxylic acid, and others. Despite the presence of hydroxyl and aldehyde functional groups in this molecule, industrial-scale production of HMF remains unfeasible, predominantly due to the high costs of fructose and low reactivity of cellulose. AVA Biochem (Muttenz, Switzerland), is one of the few companies close to commercial-scale production of HMF [7]. Table 4 shows HMF concentrations obtained for different lignocellulosic biomasses after pretreatment.
Table 4 provides an overview of succinic acid, glutaric acid, vanillin, vanillic acid, itaconic acid, adipic acid, furfural, and 5-HMF concentrations, alongside the concentrations of other value-added chemicals that can be derived simultaneously under varying pretreatment conditions from some of the most common lignocellulosic biomasses. Table 4 does not present the highest possible concentrations of each biochemical but, rather presents the biorefinery view, wherein multiple value-added products can be obtained from lignocellulosic biomass.
Table 4. Pretreatment conditions and concentrations of selected furans, organic acids, and phenolics.
Table 4. Pretreatment conditions and concentrations of selected furans, organic acids, and phenolics.
BiomassFeedstock ContentPretreatment/CatalystOperating Temp. (°C)Time (min)ConcentrationsUnitRef.
FuransOrganic AcidsPhenolics
Furoic AcidFurfural5 HMFLactic AcidSuccinic AcidGlutaric AcidItaconic AcidAdipic AcidVanillinVanillic AcidFerulic Acid
Corn Stover34.4% glucan, 22.8% xylan, and 11% lignin0.7% H2SO4 (w/w)18082.42204420 2.90.577.20.1143.36.6g/L[80]
0.07% H2SO4 (w/w)1.1261117.81.70.2420.142.81.52.6g/L
Liquid hot water0.8882.35.52.20.231.20.152.62.62.2g/L
Deionized water saturated with oxygen at 174 psi1.26.52.8245.20.652.10.26.74.31g/L
Aqueous ammonia 0.1% (w/w)1.10.40.89386.51.23.20.182.63.24.2g/L
34.4% glucan, 22.4% xylan, 4.2% arabinan, 0.6% mannan, 1.4% galactan, 11% lignin, 2.3% protein, 6.1% ash, and 3.8% uronic acidsAmmonia fiber expansion (AFEX)130150.0060.0030.6420.3180.5960.0080.0220.0030.1950.0460.103g/g DM[81]
30% H2SO4 (w/w)190 0.1557.9415.71.50.260.0120.580.0050.2810.1241.314g/g DM
N/A1% H2SO4 (w/w)1608N/A18.70.70141N/AN/AN/AN/A0.060.034N/AmM[82]
Poplar43.8% glucan, 14.85% xylan, 3.94% mannan, and 29.12% lignin0.7% H2SO4 (w/w)18083.122064292.50.610.110.0575.55.90.19g/L[80]
0.07% H2SO4 (w/w)1.7314190.930.260.130.15.65.70.46g/L
Liquid hot water0.942.60.451.82.30.230.0930.0483.14.10.23g/L
Deionized water saturated with oxygen at 12 bar0.762.10.39222.40.250.170.149.15.30.07g/L
Aqueous ammonia 0.1% (w/w)0.490.50.079261.70.350.0880.132.82.50.13g/L
48.9% glucan, 15.7% xylan, 27.7% lignin, and 1.2% ashSteam explosion 2146N/A5.92.6N/AN/AN/AN/AN/A0.035N/AN/Amg/g DM[83]
Pine40% glucan, 8.9% xylan, 16% mannan, and 27.7% lignin0.7% H2SO4 (w/w)18081.11901703.70.730.370.070.0764.65.20.12g/L[80]
0.07% H2SO4 (w/w)0.8139.54.50.340.180.0320.095.83.60.22g/L
Liquid hot water0.832.51.38.70.750.160.090.0542.42.30.31g/L
Deionized water saturated with oxygen at 12 bar0.911.90.64181.80.310.240.187.14.80.14g/L
Aqueous ammonia 0.1% (w/w)0.550.650.16362.390.660.0990.133.24.80.16g/L
Spruce41.6% glucan, 11.5% mannan, 4.7% xylan, 2% galactan, 1.1% arabinan, 25.7% lignin, and 5.4% extractives0.5% H2SO4 (w/w)2227N/A15.9N/AN/AN/AN/AN/A0.120.034N/Ag/L[84]
Wheat Straw36.3% cellulose, 30.9% hemicellulose, and 7.1% lignin6.5g/L Na2CO318510N/AN/AN/A0.4610.899N/AN/AN/A0.0080.0040.009g/100 g DM[85]
2g/L Na2CO3195150.0170.1460.016N/A0.447N/AN/AN/A0.0960.0840.015
Barley Straw33% glucan, 20% xylan, 3.8% arabinan, 1% galactan, 16.1% lignin, 7.6% ash, and 13.8% extractivesN/A2105N/A0.280.08N/AN/AN/AN/AN/A254.410mg/100 g DM[86]
DM: dry matter. Adapted from [87].

3.10. 2,5-Furandicarboxylic Acid

2,5-Furandicarboxylic acid (FDCA) is considered a promising alternative for petroleum-derived terephthalic acid for the production of bioplastics, such as polyamides, polyesters, and polyethylene furandicarboxylate. FDCA can be synthesized from 5-HMF or 2-furoic acid derived from lignocellulose-based C6 and C5 sugars, respectively. The structure of FDCA can be seen in Figure 11. Besides the required pretreatment for lignocellulose saccharification, the low efficiency of the dehydration process from hexoses to 5-HMF has been an obstacle for commercial production [88,89]. For example, reported yields from the conversion of fructose to HMF vary between 26 and 92% [90].
No commercial process has been established for FDCA production from lignocellulosic biomass, and the production of sugar-derived FDCA has only been tested at the pilot scale [91]. Some studies have presented direct conversion (so-called one-pot, two-step method) for 5-HMF synthetized from sugars and oxidation to FDCA [92,93,94,95]. Another challenge has been the need for a cost-efficient catalyst for the oxidation of furans, as utilization of noble-metal-assisted catalysts without improved recycling can be costly [89,91]. Zhou et al. [89] developed a partly lignin-derived catalyst, which showed high selectivity, as well as promising substrate conversion and FDCA yields. Some research has been conducted with consideration of enzymatic oxidation, which is suggested as a renewable alternative for catalytic conversion, although the area should be further explored [91,96]. Table 5 provides a summary of the most common oxidation processes used to derive 5-HMF and 2-furoic acid.

3.11. Xylitol

Xylitol is featured as a top value-added product from biorefineries in both reports published by the USDOE and is one of the most studied molecules due to its applications in the pharmaceutical, cosmetic, and food industries [100]. The Asia Pacific region (China in particular) represents a disproportionally large share of xylitol as compared to the rest of the world, with chewing gum being the major market, representing 80–90% of the total demand in Asia [100,101]. Xylitol demand grew from 6000 tons in 1978 to 190 thousand metric tons in 2016, which was valued at USD 725.9 million [102].
Xylitol (structure in Figure 12) can be derived through the hydrogenation of xylose-by-xylose reductase; however, microbial production of xylitol using yeast fungi and bacteria has proven to be the more promising route for production. Candia spp. is the most studied fungus with respect to xylitol production [103]. Xylitol serves as a precursor for the production xylaric acid, ethylene glycol, and propylene glycol. Table 6 summarizes reported xylitol production from lignocellulosic biomass. Whereas the pretreatment conditions are quite diverse, the hydrolysate concentration is a good indicator of the relative effectiveness of pretreatment.

3.12. Chitosan

Chitosan (deacetylated chitin) is biopolymer used in food, pharmaceutical, and cosmetics industries due to its non-toxicity, biocompatibility, biodegradability, and antimicrobial properties [104,105]. It mainly consists of amino sugar D-glucosamine bounded to small amounts of N-acetyl-D-glucosamine [105]. Satari et al. [106] approximated the amount of chitosan in a biomass sample by measuring the amount of D-glucosamine. A representative structure of chitosan can be seen in Figure 13.
Traditionally, chitosan is produced from waste exoskeletons of shellfish, but as a high concentration (45–60%) NaOH is needed to extract the chitosan from chitin, fungus-derived chitosan is a suitable alternative, as the more diluted alkaline required for the extraction decreases the environmental pollution caused by the process [105]. Sigma-Aldrich (St. Louis, MO, USA) and ChitoLytic (Toronto, ON, Canada) are companies producing non-animal-derived chitosan at the commercial level. Chitosan can be found in the cell wall of Zygomycetes fungi, and strains from Rhizopus and Mucor genera have been tested for fermentation of lignocellulose prehydrolysates [104,105,106,107]. Xylose, which is digested by microorganisms, is released from lignocellulose by treating the biomass using hydrothermal or acid-assisted pretreatment at moderate temperatures, which can be followed by enzymatic hydrolysis [108]. Despite being inhibitors, some sugar degradation compounds, such as formic acid and acetic acid, in moderate concentrations have been shown to stimulate the fungal growth and increase the accumulation of protective chitosan in the fungal cell wall [104]. However, severe pretreatment conditions must be avoided to prevent the excessive production of inhibitory compounds for fungal growth, such as furfural. Chitosan can be extracted from fungal biomass by first separating the alkaline, insoluble material, which is subsequently extracted using diluted acetic acid or acetate [104,105,106,107]. Tai et al. [104] demonstrated fermentation in hemicellulose-based hydrolysate to enhance the fungal growth and chitosan production compared to fermentation in synthetic glucose and xylose-containing medium. Table 7 provides a summary of fungal chitosan production from lignocellulosic biomass.

3.13. 2,3 Butanediol (2,3-BD)

Long-chain alcohol 2,3-butanediol (2,3-BD), structure seen in Figure 14, is an important platform bulk chemical that is used as a fuel additive and in various other industries, including chemicals, plastic manufacturing, pharmaceuticals, cosmetics, and even food [109,110,111]. Currently, it is mainly sourced from the petrochemical industry, but the increasing interest in biorefining and sustainable bio-based chemicals has made microbial production of 2,3-BD a desirable alternative [112,113,114].
2,3-BD can be produced by fermentation of lignocellulose-derived sugars using bacterial strains that can simultaneously utilize glucose and xylose. Many industrial strains can biologically produce 2,3-BD and have been shown to be suitable for fermentation. However, to enhance the production and ensure sufficient yields, metabolic engineering of the strains has been studied [109,110,112,114,115]. The pathogenic nature of most robust strains in Klebsiella and Enterobacter genera remains an obstacle to commercialization of microbial 2,3-BD production. On the other hand, generally recognized as safe (GRAS) organisms, such strains in the Bacillus genus, have shown lower fermentation efficiency [109,115]. Lignocellulose is pretreated and hydrolyzed before fermentation, usually by sodium hydroxide or sulfuric acid treatment followed by enzymatic hydrolysis. Joo et al. [116] studied the effect of inhibitory sugar degradation compounds and reported that formic acid, furans, and phenolic compounds have negative effects on cell growth and 2,3-BD production, which enhances the importance of appropriate pretreatment conditions. Table 8 provides a summary of fermentation techniques used to derive 2,3-BD from lignocellulosic biomass.

3.14. Xylo-Oligosaccharides (XOs)

Xylo-oligosaccharides (XOs) are non-digestible carbohydrates with prebiotic and other beneficial health properties that have gained commercial interest due to their potential for use as nutraceuticals; they can be produced by hydrolyzing xylan, the main component of hemicellulose. [117,118]. This provides a route to produce high-value compounds from abundant and inexpensive lignocellulosic feedstock, such as agricultural residues [119,120]. A representative structure of XOs can be seen in Figure 15.
Several methods have been tested for XO production from lignocellulose, the most common of which is hydrothermal or alkaline pretreatment followed by enzymatic or acid hydrolysis. Hydrothermal pretreatment [120], acid hydrolysis [121], and enzymatic hydrolysis [122] without pretreatment have also been tested. However, the high temperatures required for hydrothermal pretreatment and autohydrolysis increase the energy consumption of the process. Therefore, it is suggested that alkaline pretreatment at moderate temperatures, together with enzymatic hydrolysis, could be the most sustainable processing route [118]. Downstream processing is required to purify XO-containing liquors (75–90% purity required for food applications) from unwanted toxic and unhealthy compounds, such as furfural and 5-HMF, and this process directly impacts the production costs [118]. Production of XOs from agricultural residues has been extensively studied in recent years, and in the reviewed studies, the reported yields, depending on the xylan extraction and hydrolysis method, vary between 10.2% from wheat straw [119] up to >99% from sugarcane bagasse [123]. One of the most promising feedstocks is corn cob, which has shown high yields in various studies [124,125,126,127]. Table 9 summarizes XO production from lignocellulose.
Table 6. Review of extraction methods for xylitol from lignocellulosic biomass.
Table 6. Review of extraction methods for xylitol from lignocellulosic biomass.
BiomassFeedstock ContentAcid Conc.Operating Temp. (°C)Time (min)HydrolysateStrainXylitol ReportedReference
Corncob20.91 g/L of Xylose1% H2SO4 (v/v)1213040.16 g/L and 52.71 g/L of xylose (hydrolysate concentrated through rotavapor and microwave, respectively)Candida tropicalis1.2-fold increase in yield and 1.8-fold increase in productivity[128]
31.2 g/L of xylose and 3.3 g/L of glucose1% H2SO4 (v/v)12140160 g/L of xyloseCandida tropicalis As 2.1776Yield of 0.83 g/g and productivity of 1.01 g/L/h. Maximum xylitol production of 96.5 g/L[129]
41.2% cellulose, 33.4% hemicellulose, and 18.7% lignin1% H2SO4 (v/v)1256021.67 g/L of xyloseCandida tropicalis CCTCC M2012462Maximal xylitol concentration of 38.8 g/L. Yield of 0.7 g/g of xylose and a productivity of 0.46 g/L/h[130]
42.7% cellulose, 34.3% hemicellulose, and 17.5% lignin1% H2SO4 (v/v)1206028.7 g/L of xyloseCandida tropicalis W103Maximal xylitol concentration of 68.4 g/L. Yield of 0.7 g/g xylose and a productivity of 0.95 g/L/h[131]
32% cellulose, 35% hemicellulose, 20% lignin, and 4% ash and others1% H2SO4 (v/v)1216024.9 g/L of XyloseCandida magnoliaeProduction rate of 0.51 g/L/h and 18.7 g xylitol/L[132]
Sugarcane bagasse----65% xylose, 15% arabinose, and 8% glucoseDebaryomyces hanseniiMaximum yield of 0.76 and 0.82 g/g using free and immobilized cells, respectively, with corresponding volumetric productivities of 0.44 and 0.46 g/L/h at 100 g/L initial xylose concentration[133]
17.5% DM loading1% H2SO41503015.73 g/L of D-xyloseCandida guilliermondii FTI 20037Maximal xylitol production of 50.5 g/L. Yield of 0.81 g/g of xylose and productivity of 0.6 g/L/h[134]
9.3 xylose, 15.2 glucose, and 8.5 lignin (% of DM)100 mg of sulfuric acid per g of bagasse (dry wt)12110Hydrolysate 64.7 g/L xylose, 3.08 g/L glucose, 4.23 g/L arabinose, and 1.84  g/ L acetic acidCandida guilliermondiiMaximum xylitol concentration of 28.7 g/L, xylitol yield on consumed xylose of 0.49 g/g, and a xylitol volumetric productivity of 0.24 g/L/h[135]
10% DM loading1% H2SO4 (v/v)12160Sugar composition in hydrolysate xylose 56%, glucose 15%, and arabinose 24%Candida tropicalisXylitol yield was 0.65 g/g of xylose[136]
Corn Fiber20% DM loading1% H2SO4 (v/v)12160Sugar composition in hydrolysate xylose 30%, glucose 38%, arabinose 22%, and galactose 4%.Candida tropicalisXylitol yield was 0.65 g/g of xylose
DM: dry matter.
Table 7. Review of production methods for fungal chitosan from lignocellulose.
Table 7. Review of production methods for fungal chitosan from lignocellulose.
BiomassTreatment ConditionsXylose (g/L)Fungal StrainChitosan ExtractionBiomass
Production (g/L/day)
Chitosan Content
(g/g
Biomass)
CommentRef.
Corn Stover2% H2SO4
100 °C 2 h
22.4Rhizopus oryzae ME-F121 M NaOH at 121 °C for 15 min + 2% acetic acid at 95 °C for 24 h5.20.09Total production[104]
Autoclave at
121 °C for 20 min
N/AAspergillus niger1 M NaOH at 121 °C for 20 min + 2% acetic acid at 95 °C for 6–8 h15.86.8(g/kg)
Solid-state fermentation
[107]
Autoclave
at 21 °C 20 for min
N/ARhizopus oryzae1 M NaOH at 121 °C for 20 min + 2% acetic acid at 95 °C for 6-8 h14.68.6
Acid-assisted steam explosion, 0.8 MPa30Rhizopus oryzae
AS 3.819
1 M NaOH at 121 °C for 15 min + 2% acetate at 95 °C for 24 h3.70.09N/A[105]
Elm Wood85% H3PO4 at 60 ° for C 45 min + enzymaticNDMucor indicus CCUG 224240.5 M NaOH at 121 °C for 20 min (alkali-insoluble material)3.30.06Determined as the amount of glucosamine[106]
Pine Wood85% H3PO4 at 60 °C for 45 min + enzymatic6.9Mucor indicus CCUG 224240.5 M NaOH at 121 °C for 20 min (alkali-insoluble material)2.80.06
Rice Straw85% H3PO4 at 60 °C for 45 min + enzymaticNDMucor indicus CCUG 224240.5 M NaOH at 121 °C for 20 min (alkali-insoluble material)3.10.06
Wheat StrawNMMO * 120 °C
3 h + enzymatic
19.8Mucor indicus CCUG 22424Autolysis + NaOH treatment + extractionN/A0.13N-methylmorpholine
-N-oxide
[137]
* N-methylmorpholine-N-oxide.
Table 8. Review of fermentation methods for 2,3-butanediol (2,3-BD) from lignocellulose.
Table 8. Review of fermentation methods for 2,3-butanediol (2,3-BD) from lignocellulose.
BiomassPretreatmentHydrolysisStrainProductivity (g/L/h)2,3-BD Conc. (g/L)Yield (%)Ref.
Corn cob2% NaOH at 80 °C for 2 hEnzymaticEnterobacter cloacae CICC 100110.9N/A42[138]
Corn stover0.1 M NaOH at 80 °CEnzymaticZymomonas mobilisN/A10N/A[114]
N/AEnzymaticPaenibacillus polymyxa1.118.831[139]
N/AN/ABacillus licheniformis2.3119.495[112]
Jerusalem Artichoke Stalk1% H2SO4 at 130 °C for 90 minEnzymaticKlebsiella pneumoniaeN/A80.516.8[111]
Oil Plan Frond3% NaOH at 121 °C for 20 minEnzymaticEnterobacter cloacae SG1 0.330.7N/A[113]
Pine TreeN/AN/AKlebsiella oxytoca CHA0060.75.830[110]
Rice Straw0.375 M NaOH at 120 °C for 20 minEnzymaticKlebsiella sp. Zmd302.4N/A62[140]
Rice WasteNa2CO3 + NaHCO3 + Na2SO4 at 100 °C for 3 hEnzymaticKlebsiella pneumoniae KMK-050.4811.538.4[141]
Sorghum Stalk1.25% NaOH at 121 °C for 30 minEnzymaticBacillus licheniformis DSM 87851N/A45[109]
Sugarcane Bagasse0.375 M NaOH at 120 °C for 20 minEnzymaticKlebsiella sp. Zmd300.7N/A15[140]
10% NaOH at 90 °C for 90 minEnzymaticKlebsiella pneumoniae CGMCC 1.9131N/A9N/A[142]
1% H2SO4 at 121 °C for 30 minN/AEnterobacter aerogenes EMY-220.866.442[110]
5% Na2CO3 + 5% Na2SO3 at 100 °C for 4 hN/AEnterobacter aerogenes EMY-22N/AN/A39.5[143]
Sunflower StalkN/AN/AKlebsiella oxytoca CHA0060.84.334[110]
Wood24 N H2SO4 at 30 °C for 60 min + diluted acid at 105 °C for 60 minN/AEnterobacter aerogenesN/A9.9N/A[116]
N/AN/ABacillus licheniformis DSM 87851.6N/A40[109]
Table 9. Review of xylo-oligosaccharide (XO) production from lignocellulose.
Table 9. Review of xylo-oligosaccharide (XO) production from lignocellulose.
BiomassHemicellulose (%)Xylan (%)Extraction/PretreatmentHydrolysisXO Conc. (g/L)YieldUnitRef.
Barley HuskN/A26.8N/AAutohydrolysis
at 220 °C for 0.75 h
N/A27.1%[120]
Corn CobN/A30.6N/ANon-isotherm autohydrolysis at 202 °CN/A78.7gXOs/100 g xylan[126]
38.8N/A12% NaOH + steam at 121 °C for 45 min0.25 M H2SO4
at 90 °C for 60 h
0.9N/AN/A[144]
38.9N/A4–16% NaOH + steam at 121 °C for 45 minEnzymatic at 40.9–41.4 °C for 16.6–17.3 h<2.0N/AN/A[145]
N/AN/A1.25 M NaOH
at 37 °C for 180 min
Enzymatic
at 45 °C for 8 h
6.760%[124]
N/A31.3N/AAutohydrolysis
at 220 °C for 0.75 h
N/A24.8%[120]
N/A31.92% NaOH at 20 °C for 6 hEnzymatic
at 50 °C for 24–36 h
8.286.7%[125]
N/A34.81.0 g/L H2SO4 + steam at 135 °C for 30 minEnzymatic
at 50 °C for 24 h
N/A67.7gXOs/100 g xylan[127]
Eucalyptus WoodN/A16.6N/AAutohydrolysis
at 220 °C for 0.75 h
N/A15.4%[120]
Maize Silage35.1N/A1 M NaOH + steam at 121 °C for 15 minEnzymatic
at 50 °C for 24 h
3.5N/AN/A[117]
Oil Palm Frond30.4N/ASteam at 121 °C for 60 minEnzymatic
at 40 °C for 24 h
N/A17.5w/w%[146]
ReedN/A21.2Steam at 170 °C for 30 minEnzymatic
at 50 °C for 48 h
N/A68.1gXOs/100 g xylan[147]
Rice HuskN/A15.6N/AAutohydrolysis
at 220 °C for 0.75 h
N/A18%[120]
25N/A18% NaOH + steam at 120 °C for 45 minEnzymatic
at 50 °C for 9 h
N/A34.7gXOs/100 g xylan[148]
11.2N/AN/AEnzymatic
at 50 °C for 24 h
N/A69gXOs/100 g xylan[122]
Ryegrass Silage36.6N/A1 M NaOH + steam at 121 °C for 15 minEnzymatic
at 50 °C for 24 h
2.4N/AN/A[117]
Sugar Cane BagasseN/A20.6Aqueous ammonia + steam at 121 °C for 30 minEnzymatic
at 50 °C for 30 h
N/A>99%[123]
N/AN/AN/A0.1% H2SO4
at 140 °C for 1 h
N/A92.28gXOs/100 g xylan[121]
N/AN/A6% Alkaline peroxide
at 20 °C for 180 min
Enzymatic
at 50 °C for 96 h
N/A31.5%[149]
N/AN/A10% Acetic acid + steam at 150 °C for 45 minEnzymatic
at 30 °C for 1.25 h
N/A39.1gXOs/100 g xylan[150]
23.2N/A12% NaOH + steam at 121 °C for 15 minEnzymatic
at 40 °C for 8 h
1.72N/AN/A[151]
Sunflower StalkN/A18.924% KOH
at 35 °C for 120 min
Enzymatic
at 40 °C for 24 h
3.2N/AN/A[152]
N/A19.124% KOH
at 35 °C for 120 min
0.25 M H2SO4
at 100 °C for 30 h
N/A12.6gXOs/100 g xylan[119]
Wheat StrawN/A20.624% KOH
at 35 °C for 120 min
Enzymatic 4
at 0 °C for 24 h
2.3N/AN/A[152]
N/A20.924% KOH
at 35 °C for 120 min
0.25 M H2SO4
at 100 °C for 30 h
N/A10.2gXOs/100 g xylan[119]
N/AN/A2% NaOH
at 80 °C for 90 min
Enzymatic
at 60 °C for 15 h
N/A39.8gXOs/100 g xylan[153]

4. Extraction Methods

Physical pretreatment is one of the first steps to open the molecular structure of lignocellulosic biomass. Several physical, chemical, and physiochemical methods have been studied and optimized to provide access to cellulose and hemicellulose by opening up the binding lignin structure [154,155]. Whereas the primary goal of pretreatment is to provide accessibility to cellulose and hemicellulose sugars, with time and increased interest in biomass-derived production of value-added chemicals, molecules that were once considered pretreatment inhibitors, such as HMF and furfural, are now the primary compounds of interest. This newfound interest prioritizes the secondary goal, which is opening the lignin structure. Whereas accessibility to polysaccharides still holds relevance, if high-value compounds can be extracted via low-cost extraction methods, then this step precedes the step of extraction of polysaccharides into monomeric sugars. As a result of this shift, new extractions methods are continually being investigated with respect to their efficacy in obtaining high-value biochemicals. Conventional pretreatment operating conditions, as described in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8 and Table 9 are often energy-intensive, requiring reactor temperatures above 150 °C and reactor pressure between 1 and 20 atm and are frequently performed in the presence of a catalyst. Novel extraction methods include less energy-intensive alternatives that allow for recovery of high-value chemicals. The attractive prices and wide-ranging applications of specialty chemicals derived from biomass justify the development of new extraction methods. In the following sections, extraction methods for targeted high-value compounds from lignocellulosic biomass are reviewed. The isolation and encapsulation techniques used to stabilize high-value compounds are also discussed and reviewed.
For efficient extraction of high-value compounds from lignocellulosic biomass, physiochemical pretreatment is followed by extraction methods. The extraction method is governed by the characteristics of the targeted compounds identified for extraction. Some compounds are thermolabile and therefore prone to thermal degradation. Therefore, prolonged extractions using high temperatures should be chosen with caution [156,157]. For a solvent to dissolve the solute, the diffusion, solubilization, and/or transfer are governed by the thermodynamic properties of the solute and the solvent. Solvents are widely used for extraction of targeted molecules, partly due to the simplicity of the method, its scalability, and low cost [158]. Solvents used to extract high-value compounds discussed in this review include water, ethanol (EtOH), methanol (MeOH), ethyl acetate (EtOAc), dichloromethane (DCM), acetone, dimethyl sulfoxide (DMSO), deep eutectic and ionic solvents (DESs), acids and bases, and supercritical CO2 [159,160,161,162,163,164,165,166].
Phenolics are secondary metabolites with a chemical structure comprising one or more aromatic rings attached to one or more hydroxyl groups. Phenolics can be derived from food plants and lignocellulosic plants. With more than known 8,000 phenolics, there is certainly a growing interest in the ability to extract these compounds from a variety of biomass sources [158]. In an in silico study, Galanakis et al. [167] investigated the ability of solvents to solubilize phenolics, e.g., hydroxycinnamic acids (HCA), flavonoids, phenolic aldehydes, or hydroxybenzoic acids, all of which are molecules with at least one hydroxyl group (polar group) connected to an aromatic ring (non-polar group). The activity coefficient and polarity, which can be predicted by computer models, were the main parameters used to determine the solubility of a solute in a solvent. The authors stated that different solvents target different phenolic groups [167]. Of the 15 investigated phenolic groups, the solvents with the best properties for extracting phenolics were EtOH, MeOH, EtOAc, and DCM. However, these solvents do not consider the liberation of phenolics from the lignocellulosic matrix.

4.1. Hansen Solubility Parameters

Hansen solubility parameters (HSPs) are associated with the method of theoretically calculating a solute’s solubility in a solvent based on the thermodynamical properties of dispersion, polarity, and hydrogen bonding. Compounds with similar HSP values have high miscibility, solubility, diffusivity, and affinity for each other. Such compounds have similar HSP values because they have comparable atomic dispersion forces, molecular dipole forces, and electron exchanges, denoted by δD, δP, and δH, respectively.
A compound is defined as soluble in a solvent if the solvent lies inside in three-dimensional solubility parameter spheroid in the Hansen space [168,169]. Hansen emphasizes the impracticality of using water to predict solubility behaviors with HPS. Water is a small molecule with strong polar interactions, as well as strong hydrogen-bonding and hydrogen-donor capabilities. Therefore, the HPS values depend on the local environment, and Hansen does not recommend water predictions of solubility with the HPS method [168,169]. Ionic compounds are not well-described by the HSP method. Table 10 ad Table 11 consolidate HSP for common solvents and solutes.
By applying the solvent HSP values to the solute HSP values using Equation (1), the distance in Hansen space (RAB) can be calculated to determine the theoretically best solvents for individual solutes of interest, as shown in Table 12, where the solvents with the lowest distance to the solute in the Hansen space as defined by R1, the second-lowest is defined by R2, etc.
R AB = 4 Δ δ D 2 + Δ δ P 2 + Δ δ H 2
As shown in Table 12, the best solvent for chitosan is water, according to calculations based on HSP values. However, this does not mean that chitosan is soluble in water, as chitosan is a polymer, and a second chemical is therefore needed to change the ionic charge of the solute and solvent, e.g., by the addition of a weak acid [172].
If a solvent with good HSP values compared to the compound of interest is unfit for handling due to safety concerns, high cost, or due to processing inability or environmental restrictions, other miscible solvents with desirable characteristics can be chosen and mixed in ratios that will result in similar HSP values. If mixed on the basis of a volume-weighted average, a new RAB can be calculated such that the mixture of solvents might be low-cost, safer, or more environmentally accepted, with a possibly lower RAB distance in Hansen-space. By applying Equations (2)–(4), new HSP values can be calculated for a mix of solvents. This technique can also shift the solubility of a solute in the mix of solvents to induce crystallization of the solute without extensive and expensive downstream processing.
δ D = δ D 1 V 1 + δ D 2 V 2 + δ D n V n V t o t
δ P = δ P 1 V 1 + δ P 2 V 2 + δ P n V n V t o t
δ H = δ H 1 V 1 + δ H 2 V 2 + δ H n V n V t o t

4.2. Solvent Extraction

Maceration extraction (ME) is the most common and easy solvent extraction technique. This method is often called simple solvent extraction or conventional extraction [173,174]. Compounds can be extracted by simply submerging the biomass in a solvent and heating below the boiling point. As ME relies on the diffusion transfer of compounds, physical pretreatment is often required for efficient extraction to increase the surface area of the biomass and possibly to open the plant cell walls of lignocellulosic material. Such physical pretreatment can be achieved by milling or crushing. ME can also happen before, during, or after fermentation. This is the case for winemaking, where the solvent slowly changes from sugar-rich water to a mix of water–sugar–ethanol, shifting the solvent properties, e.g., polarity, dielectric constant, and surface tension [175]. As many organic solvents are perfectly miscible with water, the effect of shifting solvent properties can be well-controlled.
Decoction extraction (DE) involves boiling the plant material in a solvent to extract compounds of interest. This method is also commonly referred to as hot maceration or boiling maceration. Silva et al. [176] investigated DE and microwave-assisted extraction of bioactive compounds from lignocellulosic halophyte Salicornia ramosissima. The biomass was milled to open the lignocellulosic structure, and particles of 1 mm were obtained. DE was conducted by boiling 300 mg biomass in 10 mL distilled water for 5 min and leaving it to cool for 25 min. The extract was filtered and freeze-dried into a powder. The extracts were analyzed by a Folin–Ciocalteu total phenolic assay, with DE showing 80% higher extraction of total phenolics compared to microwave-assisted extraction.
Most analysis methods require liquid samples, as analytical equipment cannot handle solid samples. For analysis, a solid–liquid extraction method is needed to solubilize the solute of interest. Soxhlet extraction is a commonly used method for solid–liquid extraction, as the method is well-described, reliable, safe, and easy to operate. Soxhlet extraction works on the principle of continuous evaporation and simultaneous condensation of a solvent. The biomass is gradually submerged in the solvent. Once the extraction chamber containing the biomass is full of condensed solvent at near-boiling temperature, the solvent is siphoned off, and a new condensed solvent can fill up the extraction chamber. The transfer equilibrium is thereby shifted and does not determine the mass transfer of the desired compound into the solvent. This allows for an unmonitored operation of the Soxhlet apparatus, and the operator can stop the operation when desired. The continuous and cyclic nature of this extraction method ensures a concentrated extract. The solvent is usually determined by the polarity of the solute, but other factors, such as flammability, toxicity, and price, can be used as parameters for the selection of a solvent [177].
Soxhlet extraction can also be applied for the extraction of oils, fats, waxes, sterols, and other non-polar compounds from lignocellulosic biomass using a non-polar solvent. More polar compounds, e.g., phenolic compounds, can be extracted using ethanol, methanol, or ethyl acetate. A schematic of this setup is shown in Figure 16. Cascade extraction using a Soxhlet apparatus can achieve relatively pure extract phases and yield a lignocellulosic fraction with a low concentration of residual material, such as lipids or ash [178,179,180].
Medini et al. [178] studied the phytochemical composition, antioxidant capacity, anti-inflammatory effect, and anticancer activities of the halophyte Limonium densiflorum. The tested extracts were extracted by Soxhlet extraction using a successive cascade with hexane, dichloromethane, ethanol, and methanol. These solvents were carefully chosen to extract as much non-lignocellulosic material as possible and thereby afford a pure lignocellulosic fraction after extraction [178].

4.3. Subcritical Water Extraction

Subcritical water extraction (SWE) utilizes increasing temperatures and pressure of water of 100–374 °C and 0.1–22.1 MPa, respectively, while remaining below the vapor pressure. SWE utilizes the shift in chemical properties, such as viscosity, dielectric constant, surface tension, diffusivity characteristics, polarity, and pH value. Changing these properties can cause the water assimilate acids and organic solvents, allowing the operator to select the compounds of interest and yielding an aqueous extract free of harmful acids and solvents [181]. Zhang et al. [181] reviewed the extraction of phenolic compounds by SWE and found the optimal extraction conditions for total phenolic content (TPC) and total flavonoid content (TFC) of lignocellulosic biomass to be approximately 160–220 °C for 16–45 min at 2–6 MPa. Table 13 summarizes the parameters and raw materials reviewed by Zhang et al. [181]. The optimized extraction parameters are changed by increasing the pressure, and the extraction temperature can be lowered significantly to 100–140 °C for 5–130 min at 10–15 MPa.

4.4. Extraction at Varying pH Values

Changing pH during an extraction process can purify the product from other compounds that will dissolve in acidic or alkali solutions and hence extract unwanted compounds. Examples include polymers and resin chitosan, 2,5-furandicarboxylic acid polymers, and furfural resins.
Treating biomass using alkalis is a common method for pretreatment or extraction of certain compounds [85,187]. To avoid irreversible degradation of many compounds with low pKa values, such as organic acids, the stabilities of compounds of interest should be reviewed. Friedman et al. [187] demonstrated the irreversible degradation of caffeic, chlorogenic, and gallic acids at high pH, and degradation was shown to occur even at acidic pH levels, which emphasizes the necessity of low pH values during storage and extraction of phenolic compounds. Chlorogenic acids were shown to degrade into their subcomponent monophenolic compounds, caffeic acid and quinic acid. Flavonoids were shown to be more stable than monophenolic compounds or heterodimers, such as chlorogenic acid. Friedman et al. explained that this was due to the exposed carboxyl groups on the less complex phenolic compounds [187]. However, Peanparkdee et al. contested that protocatechuic acid, similar in molecular structure to gallic acid and thus considered to be more stable, exhibited degradation at lower pH levels [188]. Table 14 summarizes the stability and degradation of phenolic compounds at varying pH.

4.5. Ultrasound Extraction

Ultrasound extraction (USE) has been studied extensively with the purpose of providing enhanced accessibility to cellulose and hemicellulose. Bussemaker and Zhang [193] reviewed the application of ultrasound on various lignocellulosic biomasses, such as sorghum, corn stover, sugarcane bagasse, rice hull, and rice straw. However, the focus, like that of other reviews, was the extraction of lignin from biomass or the extraction of cellulose or hemicellulose [194,195,196]. Fang et al.’s [197] book, ‘Production of biofuels and chemicals with ultrasound’ covers the extraction of biofuels and enhanced biogas production using ultrasound, as well as the extraction of chemicals from algae. The use of ultrasound to extract value-added chemicals from lignocellulosic biomass remains scarcely researched. This expands the review slightly to accommodate other biomasses that are not necessarily lignocellulosic but have been utilized for the extraction of high-value chemicals using ultrasound.
Corbin et al. [198] investigated the USE of flax seeds for efficient extraction of phenolic compounds. The phenolics extracted by Corbin et al. [198] were bound to a glucoside group, as the phenolics in flaxseed are bound in the seed coat, with high content of glucosidic bonds. The molecules are still considered HCA, despite the glucosidic bonds. Corbin et al. [198] used slightly alkaline operation conditions (0.2 N NaOH) for the compounds bound in the seed coat matrix to release the polyphenolic lignans, polyphenolic flavonoids, and monophenolic hydroxycinnamic acids. The study by Corbin et al. [198] also compared phenolic extractions in flax seeds by optimized microwave extraction (MWE), USE, enzymatic-assisted extraction (EASE), and heated reflux. This shows MWE to be superior for the extraction of ferulic acid glucoside (76 w% higher than USE), USE to be superior for the extraction of p-coumaric acid glucoside (20 w% higher than MWE), and MWE and USE to show similar properties in the extraction of caffeic acid glucoside. The extraction of other polyphenolic compounds was also dominated by MWE and USE extractions in flaxseed, with USE showing superior extraction properties [198].
USE has also been investigated with respect to the extraction of phenolics from saline lignocellulosic biomass and halophytes. Padalino et al. [199] extracted phenolics from vacuum-dried fresh Salicornia europaea to increase the phenolic content and antioxidant capacity of freshly made pasta. USE was executed at an extraction temperature of 50 °C with an ethanol/water ratio of 40/60 v/v% and a DM loading of 1:30 w/v. The researchers achieved an increase in antioxidant capacity of 148% [199].

4.6. Enzymatic Extraction

Arabinoxylan–lignin and glucan–lignin linkage with the HCAs ferulic acid, p-coumaric acid, and flavonoid tricin, amongst others, were shown to be present in wheat straw by Zikeli et al. [200]. Of the carbohydrate–lignin linkages in wheat straw, the ferulic acid linkage is the most predominant, with high amounts of tricin linkages also shown between glucan and lignin. These intermolecular bonds are too strong to be broken solvents; hence, enzymatic hydrolysis could be investigated as part of an extraction cascade, as different extraction methods and solvents have been shown to target the extraction of different phenolics [161,167,198].
Zhu et al. [161] demonstrated that p-coumaric acid, ferulic acid, and caffeic acid in dehulled barley cannot be (or are poorly) extracted by regular solvent extraction using acetone but very easily extracted after digestion in 2 M NaOH in the presence of nitrogen. It should be noted that Zikeli et al. [200] found both ferulic acid and p-coumaric acid in the carbohydrate–lignin linkage, which is supported by Bartolomé and Gómez-Cordovés’ [201] characterization of the purified enzymes from microbial cultures, ferulic acid, and p-coumaric acid esterases. Extraction of barley using enzymatic digestion with pepsin, pancreatin, Pronase E, and Viscozyme L, compared to solvent extraction of free and bound phenolics, increased the extraction of (+)-catechin by 232–239 wt% and that of p-coumaric acid 29–82 w%, with no significant increase in the extraction of ferulic acid or caffeic acid [162].
Torres-Mancera et al. [202] described enzymatic extraction methods, whereby the majority of the phenolics in ground coffee pulp were recorded to be bound in the plant cell wall. Pectinase was used as a coenzyme to break the structure of the cell wall, and Rhizomucor pusillus strain 23aIV was used in solid-state fermentation to extract the HCAs. The phenolics were extracted downstream with solvents [202].
In the case of high protein content in the lignified biomass, protein removal should be considered, as high protein content can inhibit enzymatic hydrolysis, as described by Faulds et al. [203]. For low-protein biomasses, a hydrothermal pretreatment with a low severity factor, followed by enzymatic hydrolysis using commercial enzymatic blend DEPOL 740 L containing ferulic acid and p-coumaric acid esterases and subsequent Soxhlet extraction, can be considered. Table 15 describes extraction methods for phenols and hydroxycinnamic acids from lignocellulosic biomass.

5. Isolation and Purification Methods

Once the value-added compounds have been extracted from the biomass, the next step is to isolate and purify said compounds. Here, we discuss membrane filtration, liquid–liquid extraction, and purification using preparative high-performance liquid chromatography (Prep-HPLC).

5.1. Membrane Filtration

Membrane filtration can be seen as an easily scalable and inexpensive method for the filtration of extracts. Amongst the various membrane filtration technologies, micro-(MF), ultra-(UF), and nano-(NF) filtration are pressure-driven technologies. These technologies have the benefits of a low energy input, high separation efficiency, simple operation, no use of expensive solvents or effluents, and scalability [205]. A disadvantage of membrane separation is the inability to separate specific compounds of similar polarities and molecular weights (MWs), as the membranes only retain compounds above a certain molecular size or approximate MW.
Galanakis and Castro-Muñoz et al. [206,207,208] reviewed the separation of functional macro- and micromolecules using ultra- and nanofiltration (NF). NF with a pore size of 120 Da was used to separate phenolic compounds, achieving a separation efficiency of 99%. The smallest possible phenolic acid, benzoic acid, has an MW of 122 Da, which means all phenolics should be retained by the 120 Da membrane filter. Most phenolics with three or fewer aromatic rings have an MW of <650, which means that an initial filtration to remove larger molecules and particles, such as bacteria, hemicellulose, cellulose, lignin, proteins, and starch, is necessary, as such particles would clog the NF membrane. Even if the range of MW is established as 141–650 Da for membrane separation of phenolic compounds, this does not imply that the molecules will not be retained in larger pores. Galanakis et al. [206] investigated the separation of the phenolic compounds of HCA derivatives and flavanol from olive mill wastewater using UF and NF. Four UF pore sizes and one size of NF (100, 25, 10, 2 kDa, and 120 Da, respectively) were found to retain <1, 32, 44, 53, and 99% HCA, respectively, and 10, 37, 56, 62, and 99% flavonols, respectively [206]. This indicates the ability of these molecules to attach to other molecules or in lignocellulosic structures. Furthermore, 79, 98, 98, 99, 99% of pectin was retained with a pore size of 100, 25, 10, 2 kDa, and 120 Da, respectively, indicating that the optimal UF pore size for retention of pectin and purification of a phenolic-rich fraction is between 25 and 100 kDa, with a secondary NF membrane filtration with a pore size of 120 Da.
Whereas membrane filtration is one of the most promising technologies for recovery of macro- and micromolecules derived from lignocellulosic biomass, it is associated with some drawbacks. Some of the major drawbacks of membrane filtration are membrane fouling, leading to a decrease in permeate flux, as well as reduced efficiency of the process, and the high cost associated with cleaning and maintenance of membranes.

5.2. Liquid–Liquid Extraction

Many organic solvents used for liquid–liquid purposes have lower relative polarity than water, and these might coextract lipids, waxes, and other non-polar compounds, along with the targeted compounds. Therefore, a prior liquid–liquid extraction using a non-polar non-water-soluble solvent, such as hexane, dichloromethane, or chloroform, can positively affect the isolation and purification of the targeted compounds, yielding a more concentrated product with fewer contaminates.
Stiger-Pouvreau et al. [209] originally developed a cascade liquid–liquid extraction for the isolation of phlorotannin, a specific phenolic compound group in macroalgae Sargassacaea spp. crude extract (Figure 17). A hydroethanolic maceration obtained the extract (v/v, 1:1). The method implies the use of various consequent washings with organic solvents by liquid–liquid extraction in a separatory funnel, evaporation of solvents, and resuspension of the extracted material in an aqueous phase. Liquid–liquid extraction using dichloromethane extracted the lipids from the crude aqueous extract. Sugars and proteins were separated by low-temperature acetone and ethanol washings. Liquid–liquid extraction using ethyl acetate purified the phenolic compounds and isolated them in the organic solvent [209]. Kim et al. [210] resuspended a powdered methanolic extract of Salicornia herbacea in water and successively partitioned it with n-hexane, chloroform, EtOAc, and n-butanol. These compounds are immiscible with water or have a low solubility in water, which can be further decreased by introducing salt [210,211]. The solvent layers were easily separated from the aqueous layers and concentrated in a vacuum at 38 °C. Both the ethyl acetate and the n-butanol phases showed high free-radical scavenging activity, and the compound’s two layers were further isolated by column chromatography.
Similar to membrane filtration, liquid–liquid extraction also has its drawbacks. Bokhary et al. [212] highlighted these challenges, some of which include poor biocompatibility of solvents with microbial species, energy-intensive processes for solvent recovery, solvent toxicity, high cost of solvents, extensive safety protocols, and environmental risks associated with using large quantities of solvents at high temperatures.

5.3. Purification Using Preparative-HPLC

Prep-HPLC, like analytical HPLC, is a method of isolating chemical compounds by attracting force retention of compounds in the column stationary phase. While the mixture is passing through the column, the compounds separate in retention time, as detected by ultraviolet (UV) reflection at specific wavelengths, but in prep-HPLC, unlike analytical HPLC, the compounds are quantitatively separated and collected after detection. After collecting the separated compounds, i.e., extract purification, the separated compounds can be further processed and stabilized using encapsulation methods if necessary. Compared to analytical HPLC, Prep-HPLC usually has a larger column diameters and stationary-phase particle diameters, and the operator aims for maximum allowed sample weight, which requires more attention with respect to injection flow rate. To compare the retention times of the separated compounds, the retention factor of compound n (k’n) can be calculated. This is a dimensionless number representing how long a compound is retained in the stationary phase [213,214,215].
k n = t n t 0   t 0
where tn is the retention time for compound n, and t0 is the column dead time, i.e., the shortest retention time.
Making the peak width as narrow as possible is important to determine the fraction concentration. Unlike a narrow peak with the same area, a broad peak decreases the concentration of the compound of interest, as a more mobile phase dilutes the concentration of the separated compound. The peak width is determined at half the peak height and denoted as w½. The peak number of compound n (Nn) is derived from fractional distillation theory and is a height equivalent to a theoretical plate. The higher the peak number, the narrower the peak in the chromatogram and the higher the concentration of the separated compound [213,214,215]. The peak number is calculated as:
N n = 5.54 · ( t n w ½ )
Prep-HPLC can be scaled-up to allow for the separation of larger volumes with a column diameter of up to 30 cm, whereas industrial prep-HPLC can separate up to 3 kg h−1 of the analyte, although overloading effects are common. Overloading volume or mass in the column can result in non-optimal peak characteristics, leading to a diluted collected sample, so the maximum injection volume must determined [213,215].
Some of the challenges of using prep-HPLC for large-scale extraction of high-value compounds include the high cost of solvents, challenges in scaling up stationary phase chemistry, and the relationship between the quantity of material recovered and the size of the column [216,217]. However, many such challenges have been successfully addressed in the pharmaceutical sector, which suggests that prep-HPLC is viable for extraction of high-value products [218].

6. Conclusions

Research and development activity with respect to the extraction of high-value biomass-derived chemicals have increased considerably since the first report published by USDOE in 2004, highlighting the future of biochemicals. Although the list of top contenders has been altered and appended in the last two decades, several of the top 12 featured compounds have already reached commercial-scale production, such as succinic acid, xylitol, 2,5-FDCA, itaconic acid, levulinic acid, and furfural. Advances in genetically modified microbial strains have also boosted derivation of these compounds using biological routes as an alternative to petrochemical pathways. The considerable advances with respect to processing of lignocellulosic biomasses has enabled researchers to explore cost-effective extraction, isolation, and purification methods that directly target high-value molecules. Whereas pretreatment remains one of the primary steps in opening up the lignocellulosic structure, the severity of pretreatment can be reduced in lieu of the possibility of protecting the unstable phenolic groups, which have a high market value. Although there are still many challenges with respect to the extraction and isolation of high-value compounds, shifting the biorefinery approach to maximize the utility of lignocellulosic biomasses will aid in the cost-effective discovery of novel methods to produce biochemicals from lignocellulosic biomasses.

Author Contributions

T.C. and M.H.T. conceptualized and planned the structure of the manuscript. L.S.S.H., M.F. and T.C. performed the literature search and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This review was funded by the Aqua Combine Project of the European Union’s Horizon 2020 research and innovation program under Grant Agreement No. 862834.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dutta, K.; Daverey, A.; Lin, J.G. Evolution Retrospective for Alternative Fuels: First to Fourth Generation. Renew. Energy 2014, 69, 114–122. [Google Scholar] [CrossRef]
  2. Cherubini, F.; Strømman, A.H. Principles of Biorefining. In Biofuels: Alternative Feedstocks and Conversion Processes; Pandey, A., Larroche, C., Ricke, S.C., Dussap, C.-G., Gnansounou, E., Dussap, C.-G., Eds.; Academic Press: Cambridge, MA, USA, 2011; pp. 3–24. ISBN 9780123850997. [Google Scholar]
  3. Cherubini, F.; Jungmeier, G.; Wellisch, M.; Willke, T.; Skiadas, I.; Van Ree, R.; de Jong, E. Toward a Common Classification Approach for Biorefinery Systems. Biofuels Bioprod. Biorefining 2009, 3, 534–546. [Google Scholar] [CrossRef]
  4. Henrich, E.; Dahmen, N.; Dinjus, E. Cost Estimate for Biosynfuel Production via Biosyncrude Gasification. Biofuels, Bioprod. Biorefining 2009, 3, 28–41. [Google Scholar] [CrossRef]
  5. Werpy, T.; Petersen, G. Top Value Added Chemicals from Biomass Volume I—Results of Screening for Potential Candidates from Sugars and Synthesis Gas; U.S. Department of Energy: Washington, DC, USA, 2004; Volume 1. [Google Scholar]
  6. Holladay, J.E.; White, J.F.; Bozell, J.J.; Johnson, D. Top Value Added Chemicals from Biomass-Volume II, Results of Screening for Potential Candidates from Biorefinery Lignin; Pacific Northwest National Laboratory (PNNL): Richland, WA, USA; National Renewable Energy Laboratory (NREL): Golden, CO, USA, 2007; Volume 2. [Google Scholar]
  7. Choi, S.; Song, C.W.; Shin, J.H.; Lee, S.Y. Biorefineries for the Production of Top Building Block Chemicals and Their Derivatives. Metab. Eng. 2015, 28, 223–239. [Google Scholar] [CrossRef]
  8. Girisuta, B. Levulinic Acid from Lignocellulosic Biomass. Ph.D. Thesis, University of Gronigen, Groningen, The Netherlands, 2007. [Google Scholar]
  9. Rackemann, D.W.; Doherty, W.O.S. The Conversion of Lignocellulosics to Levulinic Acid. Biofuels Bioprod. Biorefining 2011, 5, 198–214. [Google Scholar] [CrossRef]
  10. Fitzpatrick, S.W. Production of Levulinic Acid from Carbohydrate-Containing Materials. U.S. Patent 5,608,105, 4 March 1997. [Google Scholar]
  11. Chang, C.; Cen, P.; Ma, X. Levulinic Acid Production from Wheat Straw. Bioresour. Technol. 2007, 98, 1448–1453. [Google Scholar] [CrossRef]
  12. Yan, L.; Yang, N.; Pang, H.; Liao, B. Production of Levulinic Acid from Bagasse and Paddy Straw by Liquefaction in the Presence of Hydrochloride Acid. CLEAN–Soil Air Water 2008, 36, 158–163. [Google Scholar] [CrossRef]
  13. Efremov, A.A.; Pervyshina, G.G.; Kuznetsov, B.N. Production of Levulinic Acid from Wood Raw Material in the Presence of Sulfuric Acid and Its Salts. Chem. Nat. Compd. 1998, 34, 182–185. [Google Scholar] [CrossRef]
  14. Binder, J.B.; Raines, R.T. Fermentable Sugars by Chemical Hydrolysis of Biomass. Proc. Natl. Acad. Sci. USA 2010, 107, 4516–4521. [Google Scholar] [CrossRef]
  15. Sah, P.P.T.; Ma, S.-Y. Levulinic Acid and Its Esters. J. Am. Chem. Soc. 1930, 52, 4880–4883. [Google Scholar] [CrossRef]
  16. Thomas, R.W.; Schuette, H.A. Studies on Levulinic Acid. I. Its Preparation from Carbohydrates by Digestion with Hydrochloric Acid under Pressure. J. Am. Chem. Soc. 1931, 53, 2324–2328. [Google Scholar] [CrossRef]
  17. Redmon, B.C. Process for the Production of Levulinic Acid. U.S. Patent 2,738,367, 13 March 1956. [Google Scholar]
  18. McKibbins, S.W. Kinetics of the Acid Catalyzed Conversion of Glucose to 5-Hydroxymethyl-2-Furaldehyde and Levulinic Acid; The University of Wisconsin-Madison: Madison, WI, USA, 1958; ISBN 1084403552. [Google Scholar]
  19. Carlson, L.J. Process for the Manufacture of Levulinic Acid. U.S. Patent 3,065,263, 20 November 1962. [Google Scholar]
  20. Lourvanij, K.; Rorrer, G.L. Dehydration of Glucose to Organic Acids in Microporous Pillared Clay Catalysts. Appl. Catal. A Gen. 1994, 109, 147–165. [Google Scholar] [CrossRef]
  21. Lourvanij, K.; Rorrer, G.L. Reaction Rates for the Partial Dehydration of Glucose to Organic Acids in Solid-acid, Molecular-sieving Catalyst Powders. J. Chem. Technol. Biotechnol. Int. Res. Process. Environ. Clean Technol. 1997, 69, 35–44. [Google Scholar] [CrossRef]
  22. Sumiki, Y. Levulinic Acid. Japanese Patent 176,438, 30 June 1948. [Google Scholar]
  23. Sassenrath, C.P.; Shilling, W.L. Preparation of Levulinic Acid from Hexose-Containing Material. U.S. Patent 3,258,481, 28 June 1966. [Google Scholar]
  24. Kin, Z.; Stawecka, J. Use of Waste Lignocellulose for the Production of Levulinic Acid. Prz. Pap. 1968, 303–306. [Google Scholar]
  25. Efremov, A.A.; Pervyshina, G.G.; Kuznetsov, B.N. Thermocatalytic Transformations of Wood and Cellulose in the Presence of HCl, HBr, and H2SO4. Chem. Nat. Compd. 1997, 33, 84–88. [Google Scholar] [CrossRef]
  26. 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. [Google Scholar] [CrossRef]
  27. Wee, Y.-J.; Yun, J.-S.; Park, D.-H.; Ryu, H.-W. Biotechnological Production of L (+)-Lactic Acid from Wood Hydrolyzate by Batch Fermentation of Enterococcus Faecalis. Biotechnol. Lett. 2004, 26, 71–74. [Google Scholar] [CrossRef]
  28. Guo, W.; Jia, W.; Li, Y.; Chen, S. Performances of Lactobacillus Brevis for Producing Lactic Acid from Hydrolysate of Lignocellulosics. Appl. Biochem. Biotechnol. 2010, 161, 124–136. [Google Scholar] [CrossRef]
  29. Shen, X.; Xia, L. Lactic Acid Production from Cellulosic Material by Synergetic Hydrolysis and Fermentation. Appl. Biochem. Biotechnol. 2006, 133, 251–262. [Google Scholar] [CrossRef]
  30. Moldes, A.B.; Torrado, A.; Converti, A.; Dominguez, J.M. Complete Bioconversion of Hemicellulosic Sugars from Agricultural Residues into Lactic Acid by Lactobacillus Pentosus. Appl. Biochem. Biotechnol. 2006, 135, 219–227. [Google Scholar] [CrossRef]
  31. Zhu, Y.; Lee, Y.Y.; Elander, R.T. Conversion of Aqueous Ammonia-Treated Corn Stover to Lactic Acid by Simultaneous Saccharification and Cofermentation. Appl. Biochem. Biotechnol. 2007, 137, 721–738. [Google Scholar] [PubMed]
  32. Cui, F.; Li, Y.; Wan, C. Lactic Acid Production from Corn Stover Using Mixed Cultures of Lactobacillus Rhamnosus and Lactobacillus Brevis. Bioresour. Technol. 2011, 102, 1831–1836. [Google Scholar] [CrossRef] [PubMed]
  33. Garde, A.; Jonsson, G.; Schmidt, A.S.; Ahring, B.K. Lactic Acid Production from Wheat Straw Hemicellulose Hydrolysate by Lactobacillus Pentosus and Lactobacillus Brevis. Bioresour. Technol. 2002, 81, 217–223. [Google Scholar] [CrossRef]
  34. Iyer, P.V.; Thomas, S.; Lee, Y.Y. High-Yield Fermentation of Pentoses into Lactic Acid. In Proceedings of the Twenty-First Symposium on Biotechnology for Fuels and Chemicals, Fort Collins, CO, USA, 2–6 May 1999; Walt, D.R., Ed.; Springer: Berlin/Heidelberg, Germany, 2000; pp. 665–677. [Google Scholar]
  35. Adsul, M.; Khire, J.; Bastawde, K.; Gokhale, D. Production of Lactic Acid from Cellobiose and Cellotriose by Lactobacillus Delbrueckii Mutant Uc-3. Appl. Environ. Microbiol. 2007, 73, 5055–5057. [Google Scholar] [CrossRef] [PubMed]
  36. Laopaiboon, P.; Thani, A.; Leelavatcharamas, V.; Laopaiboon, L. Acid Hydrolysis of Sugarcane Bagasse for Lactic Acid Production. Bioresour. Technol. 2010, 101, 1036–1043. [Google Scholar] [CrossRef]
  37. Yun, J.-S.; Wee, Y.-J.; Kim, J.-N.; Ryu, H.-W. Fermentative Production of Dl-Lactic Acid from Amylase-Treated Rice and Wheat Brans Hydrolyzate by a Novel Lactic Acid Bacterium, Lactobacillus sp. Biotechnol. Lett. 2004, 26, 1613–1616. [Google Scholar] [CrossRef]
  38. Mussatto, S.I.; Fernandes, M.; Mancilha, I.M.; Roberto, I.C. Effects of Medium Supplementation and PH Control on Lactic Acid Production from Brewer’s Spent Grain. Biochem. Eng. J. 2008, 40, 437–444. [Google Scholar] [CrossRef]
  39. Kumar, V.; Ashok, S.; Park, S. Recent Advances in Biological Production of 3-Hydroxypropionic Acid. Biotechnol. Adv. 2013, 31, 945–961. [Google Scholar] [CrossRef]
  40. Matsakas, L.; Hrůzová, K.; Rova, U.; Christakopoulos, P. Biological Production of 3-Hydroxypropionic Acid: An Update on the Current Status. Fermentation 2018, 4, 13. [Google Scholar] [CrossRef]
  41. Kildegaard, K.R.; Jensen, N.B.; Schneider, K.; Czarnotta, E.; Özdemir, E.; Klein, T.; Maury, J.; Ebert, B.E.; Christensen, H.B.; Chen, Y.; et al. Engineering and Systems-Level Analysis of Saccharomyces Cerevisiae for Production of 3-Hydroxypropionic Acid via Malonyl-CoA Reductase-Dependent Pathway. Microb. Cell Factories 2016, 15, 1–13. [Google Scholar] [CrossRef]
  42. Borodina, I.; Kildegaard, K.R.; Jensen, N.B.; Blicher, T.H.; Maury, J.; Sherstyk, S.; Schneider, K.; Lamosa, P.; Herrgård, M.J.; Rosenstand, I.; et al. Establishing a Synthetic Pathway for High-Level Production of 3-Hydroxypropionic Acid in Saccharomyces Cerevisiae via β-Alanine. Metab. Eng. 2015, 27, 57–64. [Google Scholar] [CrossRef] [Green Version]
  43. Kildegaard, K.R.; Wang, Z.; Chen, Y.; Nielsen, J.; Borodina, I. Production of 3-Hydroxypropionic Acid from Glucose and Xylose by Metabolically Engineered Saccharomyces Cerevisiae. Metab. Eng. Commun. 2015, 2, 132–136. [Google Scholar] [CrossRef] [PubMed]
  44. Cheng, Z.; Jiang, J.; Wu, H.; Li, Z.; Ye, Q. Enhanced Production of 3-Hydroxypropionic Acid from Glucose via Malonyl-CoA Pathway by Engineered Escherichia coli. Bioresour. Technol. 2016, 200, 897–904. [Google Scholar] [CrossRef]
  45. Liu, C.; Ding, Y.; Zhang, R.; Liu, H.; Xian, M.; Zhao, G. Functional Balance between Enzymes in Malonyl-CoA Pathway for 3-Hydroxypropionate Biosynthesis. Metab. Eng. 2016, 34, 104–111. [Google Scholar] [CrossRef]
  46. Jung, I.Y.; Lee, J.W.; Min, W.K.; Park, Y.C.; Seo, J.H. Simultaneous Conversion of Glucose and Xylose to 3-Hydroxypropionic Acid in Engineered Escherichia coli by Modulation of Sugar Transport and Glycerol Synthesis. Bioresour. Technol. 2015, 198, 709–716. [Google Scholar] [CrossRef]
  47. Suyama, A.; Higuchi, Y.; Urushihara, M.; Maeda, Y.; Takegawa, K. Production of 3-Hydroxypropionic Acid via the Malonyl-CoA Pathway Using Recombinant Fission Yeast Strains. J. Biosci. Bioeng. 2017, 124, 392–399. [Google Scholar] [CrossRef]
  48. Chen, Z.; Huang, J.; Wu, Y.; Wu, W.; Zhang, Y.; Liu, D. Metabolic Engineering of Corynebacterium Glutamicum for the Production of 3-Hydroxypropionic Acid from Glucose and Xylose. Metab. Eng. 2017, 39, 151–158. [Google Scholar] [CrossRef]
  49. Su, M.; Li, Y.; Ge, X.; Tian, P. 3-Hydroxypropionaldehyde-Specific Aldehyde Dehydrogenase from Bacillus Subtilis Catalyzes 3-Hydroxypropionic Acid Production in Klebsiella Pneumoniae. Biotechnol. Lett. 2014, 37, 717–724. [Google Scholar] [CrossRef]
  50. Huang, Y.; Li, Z.; Shimizu, K.; Ye, Q. Co-Production of 3-Hydroxypropionic Acid and 1,3-Propanediol by Klebseilla Pneumoniae Expressing AldH under Microaerobic Conditions. Bioresour. Technol. 2013, 128, 505–512. [Google Scholar] [CrossRef] [PubMed]
  51. Ko, Y.; Seol, E.; Sundara Sekar, B.; Kwon, S.; Lee, J.; Park, S. Metabolic Engineering of Klebsiella Pneumoniae J2B for Co-Production of 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol: Reduction of Acetate and Other by-Products. Bioresour. Technol. 2017, 244, 1096–1103. [Google Scholar] [CrossRef]
  52. Li, Y.; Wang, X.; Ge, X.; Tian, P. High Production of 3-Hydroxypropionic Acid in Klebsiella Pneumoniae by Systematic Optimization of Glycerol Metabolism. Sci. Reports 2016, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, K.; Tian, P. Engineering Plasmid-Free Klebsiella Pneumoniae for Production of 3-Hydroxypropionic Acid. Curr. Microbiol. 2016, 74, 55–58. [Google Scholar] [CrossRef]
  54. Ko, Y.; Ashok, S.; Zhou, S.; Kumar, V.; Park, S. Aldehyde Dehydrogenase Activity Is Important to the Production of 3-Hydroxypropionic Acid from Glycerol by Recombinant Klebsiella Pneumoniae. Process Biochem. 2012, 47, 1135–1143. [Google Scholar] [CrossRef]
  55. Kumar, V.; Sankaranarayanan, M.; Jae, K.; Durgapal, M.; Ashok, S.; Ko, Y.; Sarkar, R.; Park, S. Co-Production of 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol Using Resting Cells of Recombinant Klebsiella Pneumoniae J2B Strain Overexpressing Aldehyde Dehydrogenase. Appl. Microbiol. Biotechnol. 2012, 96, 373–383. [Google Scholar] [CrossRef]
  56. Kumar, V.; Sankaranarayanan, M.; Durgapal, M.; Shengfang, Z.; Yeounjoo, K.; Ashok, S.; Sarkar, R.; Park, S. Simultaneous Production of 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol Using Resting Cells of the Lactate Dehydrogenase-Deficient Recombinant Klebsiella Pneumoniae Overexpressing an Aldehyde Dehydrogenase. Bioresour. Technol. 2013, 135, 555–563. [Google Scholar] [CrossRef]
  57. Ashok, S.; Raj, S.M.; Rathnasingh, C.; Park, S. Development of Recombinant Klebsiella Pneumoniae ∆dhaT Strain for the Co-Production of 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol. Appl. Microbiol. Biotechnol. 2011, 90, 1253–1265. [Google Scholar] [CrossRef]
  58. Ashok, S.; Sankaranarayanan, M.; Ko, Y.; Jae, K.-E.; Ainala, S.K.; Kumar, V.; Park, S. Production of 3-Hydroxypropionic Acid from Glycerol by Recombinant Klebsiella Pneumoniae ΔdhaTΔyqhD Which Can Produce Vitamin B12 Naturally. Biotechnol. Bioeng. 2013, 110, 511–524. [Google Scholar] [CrossRef]
  59. Zhao, L.; Lin, J.; Wang, H.; Xie, J.; Wei, D. Development of a Two-Step Process for Production of 3-Hydroxypropionic Acid from Glycerol Using Klebsiella Pneumoniae and Gluconobacter Oxydans. Bioprocess Biosyst. Eng. 2015, 38, 2487–2495. [Google Scholar] [CrossRef]
  60. Jung, W.S.; Kang, J.H.; Chu, H.S.; Choi, I.S.; Cho, K.M. Elevated Production of 3-Hydroxypropionic Acid by Metabolic Engineering of the Glycerol Metabolism in Escherichia coli. Metab. Eng. 2014, 23, 116–122. [Google Scholar] [CrossRef]
  61. Chu, H.S.; Kim, Y.S.; Lee, C.M.; Lee, J.H.; Jung, W.S.; Ahn, J.-H.; Song, S.H.; Choi, I.S.; Cho, K.M. Metabolic Engineering of 3-Hydroxypropionic Acid Biosynthesis in Escherichia coli. Biotechnol. Bioeng. 2015, 112, 356–364. [Google Scholar] [CrossRef]
  62. Lim, H.G.; Noh, M.H.; Jeong, J.H.; Park, S.; Jung, G.Y. Optimum Rebalancing of the 3-Hydroxypropionic Acid Production Pathway from Glycerol in Escherichia coli. ACS Synth. Biol. 2016, 5, 1247–1255. [Google Scholar] [CrossRef] [PubMed]
  63. Sankaranarayanan, M.; Somasundar, A.; Seol, E.; Chauhan, A.S.; Kwon, S.; Jung, G.Y.; Park, S. Production of 3-Hydroxypropionic Acid by Balancing the Pathway Enzymes Using Synthetic Cassette Architecture. J. Biotechnol. 2017, 259, 140–147. [Google Scholar] [CrossRef] [PubMed]
  64. Sankaranarayanan, M.; Ashok, S.; Park, S. Production of 3-Hydroxypropionic Acid from Glycerol by Acid Tolerant Escherichia Coli. J. Ind. Microbiol. Biotechnol. 2014, 41, 1039–1050. [Google Scholar] [CrossRef] [PubMed]
  65. Raj, S.M.; Rathnasingh, C.; Jung, W.-C.; Park, S. Effect of Process Parameters on 3-Hydroxypropionic Acid Production from Glycerol Using a Recombinant Escherichia coli. Appl. Microbiol. Biotechnol. 2009, 84, 649–657. [Google Scholar] [CrossRef]
  66. Rathnasingh, C.; Raj, S.M.; Jo, J.-E.; Park, S. Development and Evaluation of Efficient Recombinant Escherichia Coli Strains for the Production of 3-Hydroxypropionic Acid from Glycerol. Biotechnol. Bioeng. 2009, 104, 729–739. [Google Scholar] [CrossRef]
  67. Tsuruno, K.; Honjo, H.; Hanai, T. Enhancement of 3-Hydroxypropionic Acid Production from Glycerol by Using a Metabolic Toggle Switch. Microb. Cell Factories 2015, 14, 1–14. [Google Scholar] [CrossRef]
  68. Honjo, H.; Tsuruno, K.; Tatsuke, T.; Sato, M.; Hanai, T. Dual Synthetic Pathway for 3-Hydroxypropionic Acid Production in Engineered Escherichia coli. J. Biosci. Bioeng. 2015, 120, 199–204. [Google Scholar] [CrossRef]
  69. Dishisha, T.; Pereyra, L.P.; Pyo, S.-H.; Britton, R.A.; Hatti-Kaul, R. Flux Analysis of the Lactobacillus Reuteri Propanediol-Utilization Pathway for Production of 3-Hydroxypropionaldehyde, 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol. Microb. Cell Fact. 2014, 13, 1–11. [Google Scholar] [CrossRef]
  70. Zaushitsyna, O.; Dishisha, T.; Hatti-Kaul, R.; Mattiasson, B. Crosslinked, Cryostructured Lactobacillus Reuteri Monoliths for Production of 3-Hydroxypropionaldehyde, 3-Hydroxypropionic Acid and 1,3-Propanediol from Glycerol. J. Biotechnol. 2017, 241, 22–32. [Google Scholar] [CrossRef]
  71. Garai-Ibabe, G.; Ibarburu, I.; Berregi, I.; Claisse, O.; Lonvaud-Funel, A.; Irastorza, A.; Dueñas, M.T. Glycerol Metabolism and Bitterness Producing Lactic Acid Bacteria in Cidermaking. Int. J. Food Microbiol. 2008, 121, 253–261. [Google Scholar] [CrossRef]
  72. Kaur, B.; Chakraborty, D. Biotechnological and Molecular Approaches for Vanillin Production: A Review. Appl. Biochem. Biotechnol. 2013, 169, 1353–1372. [Google Scholar] [CrossRef] [PubMed]
  73. Kruyer, N.S.; Peralta-Yahya, P. Metabolic Engineering Strategies to Bio-Adipic Acid Production. Curr. Opin. Biotechnol. 2017, 45, 136–143. [Google Scholar] [CrossRef] [PubMed]
  74. Polen, T.; Spelberg, M.; Bott, M. Toward Biotechnological Production of Adipic Acid and Precursors from Biorenewables. J. Biotechnol. 2013, 167, 75–84. [Google Scholar] [CrossRef]
  75. Steinbach, D.; Kruse, A.; Sauer, J. Pretreatment Technologies of Lignocellulosic Biomass in Water in View of Furfural and 5-Hydroxymethylfurfural Production—A Review. Biomass Convers. Biorefinery 2017, 7, 247–274. [Google Scholar] [CrossRef]
  76. Zeitsch, K.J. The Chemistry and Technology of Furfural and Its Many by-Products; Elsevier: Amsterdam, The Netherlands, 2000; ISBN 0080528996. [Google Scholar]
  77. Ji, L.; Tang, Z.; Yang, D.; Ma, C.; He, Y.-C. Improved One-Pot Synthesis of Furfural from Corn Stalk with Heterogeneous Catalysis Using Corn Stalk as Biobased Carrier in Deep Eutectic Solvent–Water System. Bioresour. Technol. 2021, 340, 125691. [Google Scholar] [CrossRef]
  78. Kim, H.; Lee, S.; Lee, J.; Won, W. Simultaneous Production of 1,6-Hexanediol, Furfural, and High-Purity Lignin from White Birch: Process Integration and Techno-Economic Evaluation. Bioresour. Technol. 2021, 331, 125009. [Google Scholar] [CrossRef]
  79. Bozell, J.J.; Petersen, G.R. Technology Development for the Production of Biobased Products from Biorefinery Carbohydrates—the US Department of Energy’s “Top 10” Revisited. Green Chem. 2010, 12, 539. [Google Scholar] [CrossRef]
  80. Du, B.; Sharma, L.N.; Becker, C.; Chen, S.-F.; 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]
  81. Chundawat, S.P.S.; Vismeh, R.; Sharma, L.N.; Humpula, J.F.; da Costa Sousa, L.; Chambliss, C.K.; Jones, A.D.; Balan, V.; Dale, B.E. Multifaceted Characterization of Cell Wall Decomposition Products Formed during Ammonia Fiber Expansion (AFEX) and Dilute Acid Based Pretreatments. Bioresour. Technol. 2010, 101, 8429–8438. [Google Scholar] [CrossRef]
  82. Chen, S.F.; Mowery, R.A.; Castleberry, V.A.; Van Walsum, G.P.; Chambliss, C.K. High-Performance Liquid Chromatography Method for Simultaneous Determination of Aliphatic Acid, Aromatic Acid and Neutral Degradation Products in Biomass Pretreatment Hydrolysates. J. Chromatogr. A 2006, 1104, 54–61. [Google Scholar] [CrossRef]
  83. Cantarella, M.; Cantarella, L.; Gallifuoco, A.; Spera, A.; Alfani, F. Effect of Inhibitors Released during Steam-Explosion Treatment of Poplar Wood on Subsequent Enzymatic Hydrolysis and SSF. Biotechnol. Prog. 2004, 20, 200–206. [Google Scholar] [CrossRef]
  84. Larsson, S.; Reimann, A.; Nilvebrant, N.-O.; Jönsson, L.J. Comparison of Different Methods for the Detoxification of Lignocellulose Hydrolyzates of Spruce. Larssonal. Appl. Biochem. Biotechnol. 1999, 77, 77–79. [Google Scholar]
  85. Klinke, H.B.; Ahring, B.K.; Schmidt, A.S.; Thomsen, A.B. Characterization of Degradation Products from Alkaline Wet Oxidation of Wheat Straw. Bioresour. Technol. 2002, 82, 15–26. [Google Scholar] [CrossRef]
  86. García-Aparicio, M.P.; Ballesteros, I.; González, A.; Oliva, J.M.; Ballesteros, M.; Negro, M.J. Effect of Inhibitors Released During Steam-Explosion Pretreatment of Barley Straw on Enzymatic Hydrolysis. Twenty-Seventh Symp. Biotechnol. Fuels Chem. 2006, 129, 278–288. [Google Scholar] [CrossRef]
  87. van der Pol, E.C.; Bakker, R.R.; Baets, P.; Eggink, G. By-Products Resulting from Lignocellulose Pretreatment and Their Inhibitory Effect on Fermentations for (Bio) Chemicals and Fuels. Appl. Microbiol. Biotechnol. 2014, 98, 9579–9593. [Google Scholar] [CrossRef]
  88. Banerjee, A.; Dick, G.R.; Yoshino, T.; Kanan, M.W. Carbon Dioxide Utilization via Carbonate-Promoted C-H Carboxylation. Nature 2016, 531, 215–219. [Google Scholar] [CrossRef]
  89. Zhou, H.; Xu, H.; Wang, X.; Liu, Y. Convergent Production of 2,5-Furandicarboxylic Acid from Biomass and CO2. Green Chem. 2019, 21, 2923–2927. [Google Scholar] [CrossRef]
  90. Boisen, A.; Christensen, T.B.; Fu, W.; Gorbanev, Y.Y.; Hansen, T.S.; Jensen, J.S.; Klitgaard, S.K.; Pedersen, S.; Riisager, A.; Ståhlberg, T.; et al. Process Integration for the Conversion of Glucose to 2,5-Furandicarboxylic Acid. Chem. Eng. Res. Des. 2009, 87, 1318–1327. [Google Scholar] [CrossRef]
  91. Deshan, A.D.K.; Atanda, L.; Moghaddam, L.; Rackemann, D.W.; Beltramini, J.; Doherty, W.O.S. Heterogeneous Catalytic Conversion of Sugars Into 2,5-Furandicarboxylic Acid. Front. Chem. 2020, 8, 659. [Google Scholar] [CrossRef] [PubMed]
  92. Motagamwala, A.H.; Won, W.; Sener, C.; Alonso, D.M.; Maravelias, C.T.; Dumesic, J.A. Toward Biomass-Derived Renewable Plastics: Production of 2,5-Furandicarboxylic Acid from Fructose. Sci. Adv. 2018, 4, eaap9722. [Google Scholar] [CrossRef]
  93. Yang, Z.; Qi, W.; Su, R.; He, Z. Selective Synthesis of 2,5-Diformylfuran and 2,5-Furandicarboxylic Acid from 5-Hydroxymethylfurfural and Fructose Catalyzed by Magnetically Separable Catalysts. Energy Fuels 2017, 31, 533–541. [Google Scholar] [CrossRef]
  94. Yangyang, J.Z.; Zhou, F.; Ma, H.X.; Li, X.L.; Yuan, X.Z.; Liang, F.X.; Zhang, J. One Step Synthesis of 2,5-Furandicarboxylic Acid from Fructose Catalyzed by Ce Modified Ru/HAP. Ranliao Huaxue Xuebao/J. Fuel Chem. Technol. 2020, 48, 942–948. [Google Scholar] [CrossRef]
  95. Rathod, P.V.; Jadhav, V.H. Efficient Method for Synthesis of 2,5-Furandicarboxylic Acid from 5-Hydroxymethylfurfural and Fructose Using Pd/CC Catalyst under Aqueous Conditions. ACS Sustain. Chem. Eng. 2018, 6, 5766–5771. [Google Scholar] [CrossRef]
  96. Karich, A.; Kleeberg, S.; Ullrich, R.; Hofrichter, M. Enzymatic Preparation of 2,5-Furandicarboxylic Acid (FDCA)—A Substitute of Terephthalic Acid—By the Joined Action of Three Fungal Enzymes. Microorganisms 2018, 6, 5. [Google Scholar] [CrossRef] [PubMed]
  97. Hayashi, E.; Yamaguchi, Y.; Kamata, K.; Tsunoda, N.; Kumagai, Y.; Oba, F.; Hara, M. Effect of MnO2 Crystal Structure on Aerobic Oxidation of 5-Hydroxymethylfurfural to 2,5-Furandicarboxylic Acid. J. Am. Chem. Soc. 2019, 141, 899–900. [Google Scholar] [CrossRef] [PubMed]
  98. Casanova, O.; Iborra, S.; Corma, A. Biomass into Chemicals: Aerobic Oxidation of 5-Hydroxymethyl-2-Furfural into 2,5-Furandicarboxylic Acid with Gold Nanoparticle Catalysts. ChemSusChem 2009, 2, 1138–1144. [Google Scholar] [CrossRef]
  99. Mishra, D.K.; Lee, H.J.; Kim, J.; Lee, H.S.; Cho, J.K.; Suh, Y.W.; Yi, Y.; Kim, Y.J. MnCo2O4 Spinel Supported Ruthenium Catalyst for Air-Oxidation of HMF to FDCA under Aqueous Phase and Base-Free Conditions. Green Chem. 2017, 19, 1619–1623. [Google Scholar] [CrossRef]
  100. Venkateswar Rao, L.; Goli, J.K.; Gentela, J.; Koti, S. Bioconversion of Lignocellulosic Biomass to Xylitol: An Overview. Bioresour. Technol. 2016, 213, 299–310. [Google Scholar] [CrossRef]
  101. Insights, G.M. Alkyl Polyglucosides (APG) Biosurfactants Biosurfactants Market Size. 2022. Available online: https://www.Gminsights.Com/Industry-Analysis/Alkyl-Polyglucosides-Apg-Biosurfactants-Market (accessed on 10 July 2016).
  102. Delgado Arcaño, Y.; Valmaña García, O.D.; Mandelli, D.; Carvalho, W.A.; Magalhães Pontes, L.A. Xylitol: A Review on the Progress and Challenges of Its Production by Chemical Route. Catal. Today 2020, 344, 2–14. [Google Scholar] [CrossRef]
  103. Ikeuchi, T.; Azuma, M.; Kato, J.; Ooshima, H. Screening of Microorganisms for Xylitol Production and Fermentation Behavior in High Concentrations of Xylose. Biomass Bioenergy 1999, 16, 333–339. [Google Scholar] [CrossRef]
  104. Tai, C.; Li, S.; Xu, Q.; Ying, H.; Huang, H.; Ouyang, P. Chitosan Production from Hemicellulose Hydrolysate of Corn Straw: Impact of Degradation Products on Rhizopus Oryzae Growth and Chitosan Fermentation. Lett. Appl. Microbiol. 2010, 51, 278–284. [Google Scholar] [CrossRef]
  105. Yang, L.; Li, X.; Lai, C.; Fan, Y.; Ouyang, J.; Yong, Q. Fungal Chitosan Production Using Xylose Rich of Corn Stover Prehydrolysate by Rhizopus oryzae. Biotechnol. Biotechnol. Equip. 2017, 31, 1160–1166. [Google Scholar] [CrossRef] [Green Version]
  106. Satari, B.; Karimi, K.; Zamani, A. Oil, Chitosan, and Ethanol Production by Dimorphic Fungus Mucor Indicus from Different Lignocelluloses. J. Chem. Technol. Biotechnol. 2016, 91, 1835–1843. [Google Scholar] [CrossRef]
  107. Omogbai, B.A.; Ikenebomeh, M. Solid-State Fermentative Production and Bioactivity of Fungal Chitosan. J. Microbiol. Biotechnol. Food Sci. 2013, 3, 172–175. [Google Scholar]
  108. Cybulska, I.; Chaturvedi, T.; Thomsen, M.H. Lignocellulosic Thermochemical Pretreatment Processes. In Biorefinery; Bastidas-Oyanedel, J.R., Schmidt, J., Eds.; Springer: Cham, Switzerland, 2019; pp. 153–165. [Google Scholar]
  109. Guragain, Y.N.; Chitta, D.; Karanjikar, M.; Vadlani, P.V. Appropriate Lignocellulosic Biomass Processing Strategies for Efficient 2,3-Butanediol Production from Biomass-Derived Sugars Using Bacillus Licheniformis DSM 8785. Food Bioprod. Process. 2017, 104, 147–158. [Google Scholar] [CrossRef]
  110. Cha, J.W.; Jang, S.H.; Kim, Y.J.; Chang, Y.K.; Jeong, K.J. Engineering of Klebsiella oxytoca for Production of 2,3-butanediol Using Mixed Sugars Derived from Lignocellulosic Hydrolysates. GCB Bioenergy 2020, 12, 275–286. [Google Scholar] [CrossRef]
  111. Li, D.; Dai, J.Y.; Xiu, Z.L. A Novel Strategy for Integrated Utilization of Jerusalem Artichoke Stalk and Tuber for Production of 2,3-Butanediol by Klebsiella Pneumoniae. Bioresour. Technol. 2010, 101, 8342–8347. [Google Scholar] [CrossRef]
  112. Li, L.; Li, K.; Wang, K.; Chen, C.; Gao, C.; Ma, C.; Xu, P. Efficient Production of 2,3-Butanediol from Corn Stover Hydrolysate by Using a Thermophilic Bacillus Licheniformis Strain. Bioresour. Technol. 2014, 170, 256–261. [Google Scholar] [CrossRef] [PubMed]
  113. Hazeena, S.H.; Nair Salini, C.; Sindhu, R.; Pandey, A.; Binod, P. Simultaneous Saccharification and Fermentation of Oil Palm Front for the Production of 2,3-Butanediol. Bioresour. Technol. 2019, 278, 145–149. [Google Scholar] [CrossRef] [PubMed]
  114. Yang, S.; Mohagheghi, A.; Franden, M.A.; Chou, Y.C.; Chen, X.; Dowe, N.; Himmel, M.E.; Zhang, M. Metabolic Engineering of Zymomonas Mobilis for 2,3-Butanediol Production from Lignocellulosic Biomass Sugars. Biotechnol. Biofuels 2016, 9, 189. [Google Scholar] [CrossRef] [PubMed]
  115. Ji, X.J.; Huang, H.; Ouyang, P.K. Microbial 2,3-Butanediol Production: A State-of-the-Art Review. Biotechnol. Adv. 2011, 29, 351–364. [Google Scholar] [PubMed]
  116. Joo, J.; Lee, S.J.; Yoo, H.Y.; Kim, Y.; Jang, M.; Lee, J.; Han, S.O.; Kim, S.W.; Park, C. Improved Fermentation of Lignocellulosic Hydrolysates to 2,3-Butanediol through Investigation of Effects of Inhibitory Compounds by Enterobacter Aerogenes. Chem. Eng. J. 2016, 306, 916–924. [Google Scholar] [CrossRef]
  117. Quiñones, T.S.; Retter, A.; Hobbs, P.J.; Budde, J.; Heiermann, M.; Plöchl, M.; Ravella, S.R. Production of Xylooligosaccharides from Renewable Agricultural Lignocellulose Biomass. Biofuels 2015, 6, 147–155. [Google Scholar] [CrossRef]
  118. Santibáñez, L.; Henríquez, C.; Corro-Tejeda, R.; Bernal, S.; Armijo, B.; Salazar, O. Xylooligosaccharides from Lignocellulosic Biomass: A Comprehensive Review. Carbohydr. Polym. 2021, 251, 117118. [Google Scholar]
  119. Akpinar, O.; Erdogan, K.; Bostanci, S. Production of Xylooligosaccharides by Controlled Acid Hydrolysis of Lignocellulosic Materials. Carbohydr. Res. 2009, 344, 660–666. [Google Scholar] [CrossRef]
  120. Parajó, J.C.; Garrote, G.; Cruz, J.M.; Dominguez, H. Production of Xylooligosaccharides by Autohydrolysis of Lignocellulosic Materials. Trends Food Sci. Technol. 2004, 15, 115–120. [Google Scholar]
  121. Otieno, D.O.; Ahring, B.K. The Potential for Oligosaccharide Production from the Hemicellulose Fraction of Biomasses through Pretreatment Processes: Xylooligosaccharides (XOS), Arabinooligosaccharides (AOS), and Mannooligosaccharides (MOS). Carbohydr. Res. 2012, 360, 84–92. [Google Scholar]
  122. da Silva Menezes, B.; Rossi, D.M.; Squina, F.; Ayub, M.A.Z. Xylooligosaccharides Production by Fungi Cultivations in Rice Husk and Their Application as Substrate for Lactic Acid Bacteria Growth. Bioresour. Technol. Reports 2018, 2, 100–106. [Google Scholar] [CrossRef]
  123. Reddy, S.S.; Krishnan, C. Production of High-Pure Xylooligosaccharides from Sugarcane Bagasse Using Crude β-Xylosidase-Free Xylanase of Bacillus Subtilis KCX006 and Their Bifidogenic Function. LWT—Food Sci. Technol. 2016, 65, 237–245. [Google Scholar] [CrossRef]
  124. Chapla, D.; Pandit, P.; Shah, A. Production of Xylooligosaccharides from Corncob Xylan by Fungal Xylanase and Their Utilization by Probiotics. Bioresour. Technol. 2012, 115, 215–221. [Google Scholar] [CrossRef]
  125. Aachary, A.A.; Prapulla, S.G. Value Addition to Corncob: Production and Characterization of Xylooligosaccharides from Alkali Pretreated Lignin-Saccharide Complex Using Aspergillus Oryzae MTCC 5154. Bioresour. Technol. 2009, 100, 991–995. [Google Scholar] [CrossRef] [PubMed]
  126. Vázquez, M.J.; Alonso, J.L.; Domínguez, H.; Parajó, J.C. Enhancing the Potential of Oligosaccharides from Corncob Autohydrolysis as Prebiotic Food Ingredients. Ind. Crops Prod. 2006, 24, 152–159. [Google Scholar] [CrossRef]
  127. Yang, R.; Xu, S.; Wang, Z.; Yang, W. Aqueous Extraction of Corncob Xylan and Production of Xylooligosaccharides. LWT—Food Sci. Technol. 2005, 38, 677–682. [Google Scholar] [CrossRef]
  128. Misra, S.; Raghuwanshi, S.; Saxena, R.K. Evaluation of Corncob Hemicellulosic Hydrolysate for Xylitol Production by Adapted Strain of Candida Tropicalis. Carbohydr. Polym. 2013, 92, 1596–1601. [Google Scholar] [CrossRef]
  129. Li, M.; Meng, X.; Diao, E.; Du, F. Xylitol Production by Candida Tropicalis from Corn Cob Hemicellulose Hydrolysate in a Two-Stage Fed-Batch Fermentation Process. J. Chem. Technol. Biotechnol. 2012, 87, 387–392. [Google Scholar] [CrossRef]
  130. Ping, Y.; Ling, H.Z.; Song, G.; Ge, J.P. Xylitol Production from Non-Detoxified Corncob Hemicellulose Acid Hydrolysate by Candida Tropicalis. Biochem. Eng. J. 2013, 75, 86–91. [Google Scholar] [CrossRef]
  131. Cheng, K.K.; Zhang, J.A.; Ling, H.Z.; Ping, W.X.; Huang, W.; Ge, J.P.; Xu, J.M. Optimization of PH and Acetic Acid Concentration for Bioconversion of Hemicellulose from Corncobs to Xylitol by Candida Tropicalis. Biochem. Eng. J. 2009, 43, 203–207. [Google Scholar] [CrossRef]
  132. Tada, K.; Horiuchi, J.I.; Kanno, T.; Kobayashi, M. Microbial Xylitol Production from Corn Cobs Using Candida Magnoliae. J. Biosci. Bioeng. 2004, 98, 228–230. [Google Scholar] [CrossRef]
  133. Prakash, G.; Varma, A.J.; Prabhune, A.; Shouche, Y.; Rao, M. Microbial Production of Xylitol from D-Xylose and Sugarcane Bagasse Hemicellulose Using Newly Isolated Thermotolerant Yeast Debaryomyces Hansenii. Bioresour. Technol. 2011, 102, 3304–3308. [Google Scholar] [CrossRef]
  134. de Arruda, P.V.; de Cássia Lacerda Brambilla Rodrigues, R.; da Silva, D.D.V.; das Graças de Almeida Felipe, M. Evaluation of Hexose and Pentose in Pre-Cultivation of Candida Guilliermondii on the Key Enzymes for Xylitol Production in Sugarcane Hemicellulosic Hydrolysate. Biodegradation 2010, 22, 815–822. [Google Scholar] [CrossRef]
  135. da Cunha, M.A.A.; Converti, A.; Santos, J.C.; Ferreira, S.T.S.; da Silva, S.S. PVA-Hydrogel Entrapped Candida Guilliermondii for Xylitol Production from Sugarcane Hemicellulose Hydrolysate. Appl. Biochem. Biotechnol. 2008, 157, 527–537. [Google Scholar] [CrossRef] [PubMed]
  136. Rao, R.S.; Jyothi, C.P.; Prakasham, R.S.; Sarma, P.N.; Rao, L.V. Xylitol Production from Corn Fiber and Sugarcane Bagasse Hydrolysates by Candida Tropicalis. Bioresour. Technol. 2006, 97, 1974–1978. [Google Scholar] [CrossRef] [PubMed]
  137. Asachi, R.; Karimi, K. Enhanced Ethanol and Chitosan Production from Wheat Straw by Mucor Indicus with Minimal Nutrient Consumption. Process Biochem. 2013, 48, 1524–1531. [Google Scholar] [CrossRef]
  138. Ling, H.Z.; Cheng, K.K.; Ge, J.P.; Ping, W.X. Corncob Mild Alkaline Pretreatment for High 2,3-Butanediol Production by Spent Liquor Recycle Process. Bioenergy Res. 2017, 10, 566–574. [Google Scholar] [CrossRef]
  139. Ma, K.; He, M.; You, H.; Pan, L.; Wang, Z.; Wang, Y.; Hu, G.; Cui, Y.; Maeda, T. Improvement of (R,R)-2,3-Butanediol Production from Corn Stover Hydrolysate by Cell Recycling Continuous Fermentation. Chem. Eng. J. 2018, 332, 361–369. [Google Scholar] [CrossRef]
  140. Wong, C.L.; Huang, C.C.; Lu, W.B.; Chen, W.M.; Chang, J.S. Producing 2,3-Butanediol from Agricultural Waste Using an Indigenous Klebsiella Sp. Zmd30 Strain. Biochem. Eng. J. 2012, 69, 32–40. [Google Scholar] [CrossRef]
  141. Saratale, G.D.; Jung, M.Y.; Oh, M.K. Reutilization of Green Liquor Chemicals for Pretreatment of Whole Rice Waste Biomass and Its Application to 2,3-Butanediol Production. Bioresour. Technol. 2016, 205, 90–96. [Google Scholar] [CrossRef]
  142. Zhao, X.; Song, Y.; Liu, D. Enzymatic Hydrolysis and Simultaneous Saccharification and Fermentation of Alkali/Peracetic Acid-Pretreated Sugarcane Bagasse for Ethanol and 2,3-Butanediol Production. Enzyme Microb. Technol. 2011, 49, 413–419. [Google Scholar] [CrossRef]
  143. Um, J.; Kim, D.G.; Jung, M.Y.; Saratale, G.D.; Oh, M.K. Metabolic Engineering of Enterobacter Aerogenes for 2,3-Butanediol Production from Sugarcane Bagasse Hydrolysate. Bioresour. Technol. 2017, 245, 1567–1574. [Google Scholar] [CrossRef]
  144. Samanta, A.K.; Senani, S.; Kolte, A.P.; Sridhar, M.; Sampath, K.T.; Jayapal, N.; Devi, A. Production and in Vitro Evaluation of Xylooligosaccharides Generated from Corn Cobs. Food Bioprod. Process. 2012, 90, 466–474. [Google Scholar] [CrossRef]
  145. Samanta, A.K.; Jayapal, N.; Kolte, A.P.; Senani, S.; Sridhar, M.; Dhali, A.; Suresh, K.P.; Jayaram, C.; Prasad, C.S. Process for Enzymatic Production of Xylooligosaccharides from the Xylan of Corn Cobs. J. Food Process. Preserv. 2015, 39, 729–736. [Google Scholar] [CrossRef]
  146. Sabiha-Hanim, S.; Noor, M.A.M.; Rosma, A. Effect of Autohydrolysis and Enzymatic Treatment on Oil Palm (Elaeis Guineensis Jacq.) Frond Fibres for Xylose and Xylooligosaccharides Production. Bioresour. Technol. 2011, 102, 1234–1239. [Google Scholar] [CrossRef]
  147. Chen, M.; Lu, J.; Cheng, Y.; Li, Q.; Wang, H. Novel Process for the Coproduction of Xylo-Oligosaccharide and Glucose from Reed Scraps of Reed Pulp Mill. Carbohydr. Polym. 2019, 215, 82–89. [Google Scholar] [CrossRef]
  148. Khat-udomkiri, N.; Sivamaruthi, B.S.; Sirilun, S.; Lailerd, N.; Peerajan, S.; Chaiyasut, C. Optimization of Alkaline Pretreatment and Enzymatic Hydrolysis for the Extraction of Xylooligosaccharide from Rice Husk. AMB Express 2018, 8, 115. [Google Scholar] [CrossRef] [PubMed]
  149. Brienzo, M.; Carvalho, W.; Milagres, A.M.F. Xylooligosaccharides Production from Alkali-Pretreated Sugarcane Bagasse Using Xylanases from Thermoascus Aurantiacus. Appl. Biochem. Biotechnol. 2010, 162, 1195–1205. [Google Scholar] [CrossRef]
  150. Zhou, X.; Xu, Y. Integrative Process for Sugarcane Bagasse Biorefinery to Co-Produce Xylooligosaccharides and Gluconic Acid. Bioresour. Technol. 2019, 282, 81–87. [Google Scholar] [CrossRef] [PubMed]
  151. Jayapal, N.; Samanta, A.K.; Kolte, A.P.; Senani, S.; Sridhar, M.; Suresh, K.P.; Sampath, K.T. Value Addition to Sugarcane Bagasse: Xylan Extraction and Its Process Optimization for Xylooligosaccharides Production. Ind. Crops Prod. 2013, 42, 14–24. [Google Scholar] [CrossRef]
  152. Akpinar, O.; Erdogan, K.; Bostanci, S. Enzymatic Production of Xylooligosaccharide from Selected Agricultural Wastes. Food Bioprod. Process. 2009, 87, 145–151. [Google Scholar] [CrossRef]
  153. Faryar, R.; Linares-Pastén, J.A.; Immerzeel, P.; Mamo, G.; Andersson, M.; Stålbrand, H.; Mattiasson, B.; Karlsson, E.N. Production of Prebiotic Xylooligosaccharides from Alkaline Extracted Wheat Straw Using the K80R-Variant of a Thermostable Alkali-Tolerant Xylanase. Food Bioprod. Process. 2015, 93, 1–10. [Google Scholar] [CrossRef]
  154. Ruiz, H.A.; Thomsen, M.H.; Trajano, H.L. Hydrothermal Processing in Biorefineries; Ruiz, H.A., Thomsen, M.H., Trajano, H.L., Eds.; Springer: Cham, Switzerland, 2017; ISBN 9783319564562. [Google Scholar]
  155. Ruiz, H.A.; Conrad, M.; Sun, S.N.; Sanchez, A.; Rocha, G.J.M.; Romaní, A.; Castro, E.; Torres, A.; Rodríguez-Jasso, R.M.; Andrade, L.P.; et al. Engineering Aspects of Hydrothermal Pretreatment: From Batch to Continuous Operation, Scale-up and Pilot Reactor under Biorefinery Concept. Bioresour. Technol. 2020, 299, 122685. [Google Scholar] [CrossRef]
  156. Camel, V. Microwave-Assisted Solvent Extraction of Environmental Samples. TrAC—Trends Anal. Chem. 2000, 19, 229–248. [Google Scholar] [CrossRef]
  157. He, B.; Zhang, L.L.; Yue, X.Y.; Liang, J.; Jiang, J.; Gao, X.L.; Yue, P.X. Optimization of Ultrasound-Assisted Extraction of Phenolic Compounds and Anthocyanins from Blueberry (Vaccinium Ashei) Wine Pomace. Food Chem. 2016, 204, 70–76. [Google Scholar] [CrossRef] [PubMed]
  158. Dai, J.; Mumper, R.J. Plant Phenolics: Extraction, Analysis and Their Antioxidant and Anticancer Properties. Molecules 2010, 15, 7313–7352. [Google Scholar] [CrossRef] [PubMed]
  159. Pompeu, D.; Silva, E.; Rogez, H. Optimisation of the Solvent Extraction of Phenolic Antioxidants from Fruits of Euterpe Oleracea Using Response Surface Methodology. Bioresour. Technol. 2009, 100, 6076–6082. [Google Scholar] [CrossRef] [PubMed]
  160. Monroy, Y.M.; Rodrigues, R.A.; Sartoratto, A.; Cabral, F.A. Optimization of the Extraction of Phenolic Compounds from Purple Corn Cob (Zea mays L.) by Sequential Extraction Using Supercritical Carbon Dioxide, Ethanol and Water as Solvents. J. Supercrit. Fluids 2013, 116, 10–19. [Google Scholar] [CrossRef]
  161. Zhu, Y.; Li, T.; Fu, X.; Abbasi, A.M.; Zheng, B.; Liu, R.H. Phenolics Content, Antioxidant and Antiproliferative Activities of Dehulled Highland Barley (Hordeum vulgare L.). J. Funct. Foods 2015, 19, 439–450. [Google Scholar] [CrossRef]
  162. Zhu, Y.; Li, T.; Fu, X.; Brennan, M.; Abbasi, A.M.; Zheng, B.; Liu, R.H. The Use of an Enzymatic Extraction Procedure for the Enhancement of Highland Barley (Hordeum Vulgare L.) Phenolic and Antioxidant Compounds. Int. J. Food Sci. Technol. 2016, 51, 1916–1924. [Google Scholar] [CrossRef]
  163. Varun, T.K.; Senani, S.; Jayapal, N.; Chikkerur, J.; Roy, S.; Tekulapally, V.B.; Gautam, M.; Kumar, N. Extraction of Chitosan and Its Oligomers from Shrimp Shell Waste, Their Characterization and Antimicrobial Effect. Vet. World 2017, 10, 170–175. [Google Scholar] [CrossRef]
  164. Riveiro, E.; González, B.; Domínguez, Á. Extraction of Adipic, Levulinic and Succinic Acids from Water Using TOPO-Based Deep Eutectic Solvents. Sep. Purif. Technol. 2020, 241, 116692. [Google Scholar] [CrossRef]
  165. Marták, J.; Schlosser, Š. Extraction of Lactic Acid by Phosphonium Ionic Liquids. Sep. Purif. Technol. 2007, 57, 483–494. [Google Scholar] [CrossRef]
  166. Slak, J.; Pomeroy, B.; Kostyniuk, A.; Grilc, M.; Likozar, B. A Review of Bio-Refining Process Intensification in Catalytic Conversion Reactions, Separations and Purifications of Hydroxymethylfurfural (HMF) and Furfural. Chem. Eng. J. 2022, 429, 132325. [Google Scholar] [CrossRef]
  167. Galanakis, C.; Goulas, V.; Tsakona, S.; Manganaris, G.; Gekas, V. A Knowledge Base for The Recovery of Natural Phenols with Different Solvents. Int. J. food Prop. 2013, 16, 382–396. [Google Scholar] [CrossRef]
  168. Hansen, C.M. The Three Dimensional Solubility Parameter; Danish Technical Press: Copenhagen, Denmark, 1967. [Google Scholar]
  169. Hansen, C.M. Methods of Characterization—Surfaces. In Hansen Solubility Parameters: A Users Handbook, 2nd ed.; CRC Press: Raton, FL, USA, 2007; pp. 113–123. ISBN 9781420006834. [Google Scholar]
  170. Moulik, S.; Bukke, V.; Sajja, S.C.; Sridhar, S. Chitosan-Polytetrafluoroethylene Composite Membranes for Separation of Methanol and Toluene by Pervaporation. Carbohydr. Polym. 2018, 193, 28–38. [Google Scholar] [CrossRef] [PubMed]
  171. Huang, K.; Mohan, M.; George, A.; Simmons, B.A.; Xu, Y.; Gladden, J.M. Integration of Acetic Acid Catalysis with One-Pot Protic Ionic Liquid Configuration to Achieve High-Efficient Biorefinery of Poplar Biomass. Green Chem. 2021, 23, 6036–6049. [Google Scholar] [CrossRef]
  172. Qin, C.; Li, H.; Xiao, Q.; Liu, Y.; Zhu, J.; Du, Y. Water-Solubility of Chitosan and Its Antimicrobial Activity. Carbohydr. Polym. 2006, 63, 367–374. [Google Scholar] [CrossRef]
  173. Yan, X.; Zhang, G.; Zhao, J.; Ma, M.; Bao, X.; Zeng, Z.; Gong, X.; Yu, P.; Wen, X.; Gong, D. Influence of Phenolic Compounds on the Structural Characteristics, Functional Properties and Antioxidant Activities of Alcalase-Hydrolyzed Protein Isolate from Cinnamomum Camphora Seed Kernel. LWT 2021, 148, 111799. [Google Scholar] [CrossRef]
  174. Zuorro, A.; Iannone, A.; Lavecchia, R. Water–Organic Solvent Extraction of Phenolic Antioxidants from Brewers’ Spent Grain. Processes 2019, 7, 126. [Google Scholar] [CrossRef]
  175. Sacchi, K.L.; Bisson, L.F.; Adams, D.O. A Review of the Effect of Winemaking Techniques on Phenolic Extraction in Red Wines. Am. J. Enol. Vitic. 2005, 56, 197–206. [Google Scholar]
  176. Silva, A.M.; Lago, J.P.; Pinto, D.; Moreira, M.M.; Grosso, C.; Cruz Fernandes, V.; Delerue-Matos, C.; Rodrigues, F. Salicornia Ramosissima Bioactive Composition and Safety: Eco-Friendly Extractions Approach (Microwave-Assisted Extraction vs. Conventional Maceration). Appl. Sci. 2021, 11, 4744. [Google Scholar] [CrossRef]
  177. Luque de Castro, M.D.; Priego-Capote, F. Soxhlet Extraction: Past and Present Panacea. J. Chromatogr. A 2010, 1217, 2383–2389. [Google Scholar] [CrossRef]
  178. Medini, F.; Bourgou, S.; Lalancette, K.G.; Snoussi, M.; Mkadmini, K.; Coté, I.; Abdelly, C.; Legault, J.; Ksouri, R. Phytochemical Analysis, Antioxidant, Anti-Inflammatory, and Anticancer Activities of the Halophyte Limonium Densiflorum Extracts on Human Cell Lines and Murine Macrophages. S. Afr. J. Bot. 2015, 99, 158–164. [Google Scholar] [CrossRef]
  179. Fredsgaard, M.; Hulkko, L.S.S.; Chaturvedi, T.; Thomsen, M.H. Process Simulation and Techno-Economic Assessment of Salicornia Sp. Based Jet Fuel Refinery through Hermetia Illucens Sugars-to-Lipids Conversion and HEFA Route. Biomass Bioenergy 2021, 150, 106142. [Google Scholar] [CrossRef]
  180. Dalmeida Daffodil, E.; Rajalakshmi, K.; Mohan, V.R. Antioxidant Activity, Total Phenolics and Flavonoids of Salicornia Brachiata Rox. Leaf Extracts (Chenopodiaceae). World J. Pharm. Pharm. Sci. 2014, 2, 352–366. [Google Scholar]
  181. Zhang, J.; Wen, C.; Zhang, H.; Duan, Y.; Ma, H. Recent Advances in the Extraction of Bioactive Compounds with Subcritical Water: A Review. Trends Food Sci. Technol. 2020, 95, 183–195. [Google Scholar] [CrossRef]
  182. Yu, X.-M.; Zhu, P.; Zhong, Q.-P.; Li, M.-Y.; Ma, H.-R. Subcritical Water Extraction of Antioxidant Phenolic Compounds from XiLan Olive Fruit Dreg. J. Food Sci. Technol. 2015, 52, 5012–5020. [Google Scholar] [CrossRef]
  183. Aliakbarian, B.; Fathi, A.; Perego, P.; Dehghani, F. Extraction of Antioxidants from Winery Wastes Using Subcritical Water. J. Supercrit. Fluids 2012, 65, 18–24. [Google Scholar] [CrossRef]
  184. Duba, K.S.; Casazza, A.A.; Mohamed, H.B.; Perego, P.; Fiori, L. Extraction of Polyphenols from Grape Skins and Defatted Grape Seeds Using Subcritical Water: Experiments and Modeling. Food Bioprod. Process. 2015, 94, 29–38. [Google Scholar] [CrossRef]
  185. Mazzutti, S.; Salvador Ferreira, S.R.; Herrero, M.; Ibañez, E. Intensified Aqueous-Based Processes to Obtain Bioactive Extracts from Plantago Major and Plantago Lanceolata. J. Supercrit. Fluids 2017, 119, 64–71. [Google Scholar] [CrossRef]
  186. Pourali, O.; Asghari, F.S.; Yoshida, H. Production of Phenolic Compounds from Rice Bran Biomass under Subcritical Water Conditions. Chem. Eng. J. 2010, 160, 259–266. [Google Scholar] [CrossRef]
  187. Friedman, M.; Jürgens, H.S. Effect of PH on the Stability of Plant Phenolic Compounds. J. Agric. Food Chem. 2000, 48, 2101–2110. [Google Scholar] [CrossRef]
  188. Peanparkdee, M.; Patrawart, J.; Iwamoto, S. Physicochemical Stability and in Vitro Bioaccessibility of Phenolic Compounds and Anthocyanins from Thai Rice Bran Extracts. Food Chem. 2020, 329, 127157. [Google Scholar] [CrossRef] [PubMed]
  189. Dawidowicz, A.L.; Typek, R. The Influence of PH on the Thermal Stability of 5-O-Caffeoylquinic Acids in Aqueous Solutions. Eur. Food Res. Technol. 2011, 233, 223–232. [Google Scholar] [CrossRef]
  190. Wang, D.; Liu, J.; Qiu, S.; Wang, J.; Song, G.; Chu, B.; Li, L.; Xiao, G.; Gong, J.; Zheng, F. Ultrasonic Degradation Kinetics and Isomerization of 3- and 4-O-Caffeoylquinic Acid at Various PH: The Protective Effects of Ascorbic Acid and Epigallocatechin Gallate on Their Stability. Ultrason. Sonochem. 2021, 80, 105812. [Google Scholar] [CrossRef] [PubMed]
  191. Zhao, Y.K.; Cao, Q.E.; Liu, H.T.; Wang, K.T.; Yan, A.X.; Hu, Z.D. Determination of Baicalin, Chlorogenic Acid and Caffeic Acid in Traditional Chinese Medicinal Preparations by Capillary Zone Electrophoresis. Chromatographia 2000, 51, 483–486. [Google Scholar] [CrossRef]
  192. Narita, Y.; Inouye, K. Degradation Kinetics of Chlorogenic Acid at Various PH Values and Effects of Ascorbic Acid and Epigallocatechin Gallate on Its Stability under Alkaline Conditions. J. Agric. Food Chem. 2013, 61, 966–972. [Google Scholar] [CrossRef]
  193. Bussemaker, M.J.; Zhang, D. Effect of Ultrasound on Lignocellulosic Biomass as a Pretreatment for Biorefinery and Biofuel Applications. Ind. Eng. Chem. Res. 2013, 52, 3563–3580. [Google Scholar] [CrossRef]
  194. Ríos-González, L.J.; Medina-Morales, M.A.; Rodríguez-De la Garza, J.A.; Romero-Galarza, A.; Medina, D.D.; Morales-Martínez, T.K. Comparison of Dilute Acid Pretreatment of Agave Assisted by Microwave versus Ultrasound to Enhance Enzymatic Hydrolysis. Bioresour. Technol. 2021, 319, 124099. [Google Scholar] [CrossRef]
  195. Bussemaker, M.J.; Xu, F.; Zhang, D. Manipulation of Ultrasonic Effects on Lignocellulose by Varying the Frequency, Particle Size, Loading and Stirring. Bioresour. Technol. 2013, 148, 15–23. [Google Scholar] [CrossRef]
  196. Wang, Z.; Hou, X.; Sun, J.; Li, M.; Chen, Z.; Gao, Z. Comparison of Ultrasound-Assisted Ionic Liquid and Alkaline Pretreatment of Eucalyptus for Enhancing Enzymatic Saccharification. Bioresour. Technol. 2018, 254, 145–150. [Google Scholar] [CrossRef]
  197. Fang, Z.; Richard, L.; Smith, J.; Qi, X. Production of Biofuels and Chemicals with Ultrasound. In Biofuels and Biorefineries; Fang, Z., Smith, R.L., Qi, X., Eds.; Springer: Dordrecht, The Netherlands, 2015; Volume 4. [Google Scholar]
  198. Corbin, C.; Fidel, T.; Leclerc, E.A.; Barakzoy, E.; Sagot, N.; Falguiéres, A.; Renouard, S.; Blondeau, J.P.; Ferroud, C.; Doussot, J.; et al. Development and Validation of an Efficient Ultrasound Assisted Extraction of Phenolic Compounds from Flax (Linum usitatissimum L.) Seeds. Ultrason. Sonochem. 2015, 26, 176–185. [Google Scholar] [CrossRef]
  199. Padalino, L.; Costa, C.; Del Nobile, M.A.; Conte, A. Extract of Salicornia Europaea in Fresh Pasta to Enhance Phenolic Compounds and Antioxidant Activity. Int. J. Food Sci. Technol. 2019, 54, 3051–3057. [Google Scholar] [CrossRef]
  200. Zikeli, F.; Ters, T.; Fackler, K.; Srebotnik, E.; Li, J. Wheat Straw Lignin Fractionation and Characterization as Lignin-Carbohydrate Complexes. Ind. Crop. Prod. 2016, 85, 309–317. [Google Scholar] [CrossRef]
  201. Bartolomé, B.; Gómez-Cordovés, C. Barley Spent Grain: Release of Hydroxycinnamic Acids (Ferulic Andp-Coumaric Acids) by Commercial Enzyme Preparations. J. Sci. Food Agric. 1999, 79, 435–439. [Google Scholar] [CrossRef]
  202. Torres-Mancera, M.T.; Cordova-Lopez, J.; Rodriguez-Serrano, G.; Roussos, S.; Ramirez-Coronel, M.A.; Favela-Torres, E.; Saucedo-Castaneda, G. Enzymatic Extraction of Hydroxycinnamic Acids from Coffee Pulp. Food Technol. Biotechnol. 2011, 49, 369–373. [Google Scholar]
  203. Faulds, C.B.; Robertson, J.A.; Waldron, K.W. Effect of PH on the Solubilization of Brewers’ Spent Grain by Microbial Carbohydrases and Proteases. J. Agric. Food Chem. 2008, 56, 7038–7043. [Google Scholar] [CrossRef]
  204. Cristina, C.; Lucia, P.; Sara, S.; Francesco, S.; Nobile, D.; Alessandro, M.; Conte, A.; Amalia, C. Study of the Efficacy of Two Extraction Techniques from Crithmum Maritimum and Salicornia Europaea. J. Food Nutr. Res. 2018, 6, 456–463. [Google Scholar] [CrossRef]
  205. Huang, H.J.; Ramaswamy, S.; Tschirner, U.W.; Ramarao, B.V. A Review of Separation Technologies in Current and Future Biorefineries. Sep. Purif. Technol. 2008, 62, 1–21. [Google Scholar] [CrossRef]
  206. Galanakis, C.M.; Tornberg, E.; Gekas, V. Clarification of High-Added Value Products from Olive Mill Wastewater. J. Food Eng. 2010, 99, 190–197. [Google Scholar] [CrossRef]
  207. Galanakis, C.M. Separation of Functional Macromolecules and Micromolecules: From Ultrafiltration to the Border of Nanofiltration. Trends Food Sci. Technol. 2015, 42, 44–63. [Google Scholar] [CrossRef]
  208. Castro-Muñoz, R.; Yáñez-Fernández, J.; Fíla, V. Phenolic Compounds Recovered from Agro-Food by-Products Using Membrane Technologies: An Overview. Food Chem. 2016, 213, 753–762. [Google Scholar] [CrossRef]
  209. Stiger-Pouvreau, V.; Jégou, C.; Cérantola, S.; Guérard, F.; Lann, K. Le Phlorotannins in Sargassaceae Species from Brittany (France): Interesting Molecules for Ecophysiological and Valorisation Purposes. In Advances in Botanical Research; Sibout, R., Ed.; Academic Press Inc.: Cambridge, MA, USA, 2014; Volume 71, pp. 379–411. [Google Scholar]
  210. Kim, J.Y.; Cho, J.Y.; Ma, Y.K.; Park, K.Y.; Lee, S.H.; Ham, K.S.; Lee, H.J.; Park, K.H.; Moon, J.H. Dicaffeoylquinic Acid Derivatives and Flavonoid Glucosides from Glasswort (Salicornia herbacea L.) and Their Antioxidative Activity. Food Chem. 2011, 125, 55–62. [Google Scholar] [CrossRef]
  211. Xavier, L.; Freire, M.S.; Vidal-Tato, I.; González-Álvarez, J. Aqueous Two-Phase Systems for the Extraction of Phenolic Compounds from Eucalyptus (Eucalyptus globulus) Wood Industrial Wastes. J. Chem. Technol. Biotechnol. 2014, 89, 1772–1778. [Google Scholar] [CrossRef]
  212. Bokhary, A.; Leitch, M.; Liao, B.Q. Liquid–Liquid Extraction Technology for Resource Recovery: Applications, Potential, and Perspectives. J. Water Process Eng. 2021, 40, 101762. [Google Scholar] [CrossRef]
  213. Meyer, V.R. Practical High-Performance Liquid Chromatography, 3rd ed.; John Wiley & Sons: Hoboken, NJ, USA, 1999. [Google Scholar]
  214. Hanai, T. HPLC A Practical Guide. In Chromatography Monographs; Smith, R.M., Ed.; Royal Society of Chemistry: Cambridge, UK, 1999. [Google Scholar]
  215. Snyder, L.R.; Kirkland, J.J.; Glajch, J.L. Practical HPLC Method Development, 2nd ed.; Wiley-Interscience: Hoboken, NJ, USA, 1997. [Google Scholar]
  216. Teigiserova, D.A.; Hamelin, L.; Thomsen, M. Review of High-Value Food Waste and Food Residues Biorefineries with Focus on Unavoidable Wastes from Processing. Resour. Conserv. Recycl. 2019, 149, 413–426. [Google Scholar] [CrossRef]
  217. Valanciene, E.; Jonuskiene, I.; Syrpas, M.; Augustiniene, E.; Matulis, P.; Simonavicius, A.; Malys, N. Advances and Prospects of Phenolic Acids Production, Biorefinery and Analysis. Biomolecules 2020, 10, 874. [Google Scholar] [CrossRef] [PubMed]
  218. Latif, Z.; Sarker, S.D. Isolation of Natural Products by Preparative High Performance Liquid Chromatography (Prep-HPLC). Methods Mol. Biol. 2012, 864, 255–274. [Google Scholar] [CrossRef]
Figure 1. Biological and chemical routes for the production of top value-added chemicals derived from lignocellulosic biomass. Reprinted/adapted with permission from Ref. [7].
Figure 1. Biological and chemical routes for the production of top value-added chemicals derived from lignocellulosic biomass. Reprinted/adapted with permission from Ref. [7].
Processes 10 01752 g001
Figure 2. Structure of levulinic acid (C5H8O3).
Figure 2. Structure of levulinic acid (C5H8O3).
Processes 10 01752 g002
Figure 3. Structure of Lactic Acid (C3H6O3).
Figure 3. Structure of Lactic Acid (C3H6O3).
Processes 10 01752 g003
Figure 4. Structure of 3-hydroxypropionic acid (C3H6O3).
Figure 4. Structure of 3-hydroxypropionic acid (C3H6O3).
Processes 10 01752 g004
Figure 5. Structure of succinic acid (C4H6O4).
Figure 5. Structure of succinic acid (C4H6O4).
Processes 10 01752 g005
Figure 6. Structure of vanillic acid (C8H8O4) (left) and vanillin (C8H8O3) (right).
Figure 6. Structure of vanillic acid (C8H8O4) (left) and vanillin (C8H8O3) (right).
Processes 10 01752 g006
Figure 7. Structure of itaconic acid (C5H6O4).
Figure 7. Structure of itaconic acid (C5H6O4).
Processes 10 01752 g007
Figure 8. Structure of adipic acid (C6H10O4).
Figure 8. Structure of adipic acid (C6H10O4).
Processes 10 01752 g008
Figure 9. Structure of furfural (C5H4O2).
Figure 9. Structure of furfural (C5H4O2).
Processes 10 01752 g009
Figure 10. Structure of 5-HMF (C6H6O3).
Figure 10. Structure of 5-HMF (C6H6O3).
Processes 10 01752 g010
Figure 11. Structure of FDCA (C6H4O5).
Figure 11. Structure of FDCA (C6H4O5).
Processes 10 01752 g011
Figure 12. Structure of xylitol (C5H12O5).
Figure 12. Structure of xylitol (C5H12O5).
Processes 10 01752 g012
Figure 13. Representative structure of chitosan.
Figure 13. Representative structure of chitosan.
Processes 10 01752 g013
Figure 14. Structure of 2,3-BD (C4H10O2).
Figure 14. Structure of 2,3-BD (C4H10O2).
Processes 10 01752 g014
Figure 15. A representative structure of XOs (C5H12O5)n.
Figure 15. A representative structure of XOs (C5H12O5)n.
Processes 10 01752 g015
Figure 16. Experimental setup for experimental Soxhlet extraction. Adapted from [177].
Figure 16. Experimental setup for experimental Soxhlet extraction. Adapted from [177].
Processes 10 01752 g016
Figure 17. Procedure of liquid–liquid extraction of phenolic compounds from Sargassaceae spp. [209]. M:W: methanol:water; DCM: dichloromethane.
Figure 17. Procedure of liquid–liquid extraction of phenolic compounds from Sargassaceae spp. [209]. M:W: methanol:water; DCM: dichloromethane.
Processes 10 01752 g017
Table 1. Review of extraction methods for levulinic acid from lignocellulosic biomass.
Table 1. Review of extraction methods for levulinic acid from lignocellulosic biomass.
BiomassCellulose
Content %
Acid ConcentrationOperating
Temperature (°C)
Time (h)Theoretical Yield (mol%)Reference
Kraft Paper Pulp Residue801–5% H2SO41st Stage 210–230 2nd Stage 195–215N/A70–80[10]
Wheat Straw403.5% H2SO4210N/A68.8[11]
40.44.5% H2SO4220N/A79.6[12]
3.5% H2SO42100.6319.8[13]
Bagasse424.5% HCl220N/A82.7[12]
1.5% H2SO425–195217.5[14]
Glucose3220% HCl1002415[15]
296.5% HCl162124[16]
27Amberlite IR-120Room Temperature 1245.8[17]
5–200.1–4% H2SO4160–240N/A35.4[18]
106% HCl1600.2541.4[19]
123% Clay Catalyst (Fe-pillared montmorillonite)1502412[20]
123% HY Zeolite150246[21]
Rice HullN/A1% HCl160310.3[22]
Rice StrawN/A1% HCl16035.5
Corn StalksN/A1% HCl16037.5
Wood SawdustN/A1.5% HCl1900.59[23]
OakwoodN/A3% H2SO4180317.5[24]
Aspen, Pine, and SpruceN/A5% H2SO4200–2402–413–18[13]
CelluloseN/A1–5% H2SO4150–2502–7<25[25]
1–5% HCl<28
1–5% HBr<27
Aspen WoodN/A1–5% H2SO4150–2502–7<15.5
1–5% HCl<12.4
1–5% HBr<13
Table 2. Lactic acid’s concentration, yield, and productivity from some of the most common lignocellulosic biomasses.
Table 2. Lactic acid’s concentration, yield, and productivity from some of the most common lignocellulosic biomasses.
BiomassStrainConcentration (g/L)Yield a (g/g)Productivity b (g/L/h)Reference
Wood HydrolysateE. mundtii QU 25930.931.7[27]
Corn Cob/StoverLb. brevis39.10.70.81[28]
L. delbrueckii ZU-S248.7/44.20.95/0.921.01/5.7[29]
L. pentosus260.530.34[30]
L. pentosus ATCC 804174.80.65N/A[31]
L. rhamnosus and L. brevis20.950.70.58[32]
Wheat StrawL. brevis and L. pentosus7.10.95N/A[33]
SoftwoodL. casei subsp. rhamnosus21.1–23.750.74–0.830.15–0.23[34]
Sugarcane BagasseL. delbrueckii subsp. delbrueckii Mutant Uc-3670.830.93[35]
L. lactis IO-110.90.360.17[36]
Rice and Wheat BarnL. rhamnosus ATCC 9595 (CET288)1290.952.9[37]
Brewer’s Spent GrainL. delbrueckii UFV H2B2035.50.990.59[38]
a Ratio of the yield of lactic acid produced (g) to substrate consumed (g). b Lactic acid productivity.
Table 3. 3-hydroxypropionic acid concentration and productivity based on various microbes and two main substrate sources, i.e., glucose and glycerol.
Table 3. 3-hydroxypropionic acid concentration and productivity based on various microbes and two main substrate sources, i.e., glucose and glycerol.
SubstrateHost MicroorganismConcentration (g/L)Productivity (g/L/h)Reference
GlucoseS. cerevisiae9.80.1[41]
13.70.17[42]
7.40.06[43]
E. coli10.10.28[44]
40.60.56[45]
31.10.63[7]
29.70.54[46]
S. pombe7.60.25[47]
C. glutamicum62.60.87[48]
GlycerolK. pneumoniae180.77[49]
48.91.75[50]
430.9[51]
83.81.16[52]
0.90.04[53]
24.41.02[50]
160.3[54]
11.30.94[55]
22.70.38[56]
28.10.58[57]
220.46[58]
60.51.12[59]
E. coli42.11.32[60]
71.91.8[61]
40.51.35[62]
56.41.18[63]
41.50.86[64]
310.43[65]
38.70.54[66]
6.060.13[67]
5.050.105[68]
L. reuteri10.61.08[69]
3.30.09[70]
L. collinoides0.550.07[71]
Table 5. Review of oxidation processes from sugar-derived 5-hydroxymethylfurfural (5-HMF) and 2-furoic acid to 2,5-furandicarboxylic acid (FDCA).
Table 5. Review of oxidation processes from sugar-derived 5-hydroxymethylfurfural (5-HMF) and 2-furoic acid to 2,5-furandicarboxylic acid (FDCA).
SubstrateCatalystReagentsTemp. (°C)Time (h)Pressure (Bar)Substr. Conv. (%)FDCA Yield (%)Ref.
5-HMFLignin-derived Co SAs/N@CNa2CO3, O2853199.474.4[89]
858110099.5
MnO2NaHCO31002410>9991[97]
Au-TiO2N/A65810N/A>99[98]
5% Pt/CO21002040N/A94[92]
Magnetic ZnFe1.65Ru0.35ODimethyl sulfoxide13016N/AN/A91.2[94]
Ru/HAPN/A160420N/A34.2[94]
Pd/CCK2CO3, O214030N/AN/A85[95]
Ru (4%)/MnCo2O4N/A120102410099.1[99]
Fungal enzymes: aryl alcohol oxidase, peroxygenase, galactose oxidaseH2O2, phosphate bufferN/A>24N/AN/A80[96]
2-furoic acidN/ACs2CO2, CO220058N/A77[88]
Lignin-derived Co SAs/N@CCs2CO2, CO226036Flowing85.871.1[89]
Table 10. HSP values of solvents as described by Hansen [169].
Table 10. HSP values of solvents as described by Hansen [169].
CompoundδDδPδHCompoundδDδPδH
1,4-Dioxane191.87.4Iso-butanol15.15.715.9
1-Butanol165.715.8Methanol14.712.322.3
1-Propanol166.817.4Methyl cyclohexane 1601
2-Butanol15.85.714.5Methyl ethyl ketone1695.1
2-Propanol15.86.116.4Methyl isobutyl ketone15.36.14.1
Acetone15.510.47Methylene dichloride18.26.36.1
Acetonitrile15.3186.1N,N-dimethyl acetamide16.811.510.2
Benzene18.402N,N-dimethyl formamide17.413.711.3
Benzyl Alcohol18.46.313.7n-Butyl acetate15.83.76.3
Carbon tetrachloride17.800.6n-Heptane15.300
Chlorobenzene194.32n-Hexane14.900
Chloroform17.83.15.7n-Nonane15.700
Cyclohexane16.800.2n-Octane15.500
Cyclohexanone17.86.35.1n-Pentane15.600
Decalin (cis)1800sec-Butyl acetate153.77.6
Dichloromethane8.93.13Styrene18.614.1
Diethyl Ether14.52.94.6Tetralin19.622.9
Dimethyl Phthalate18.610.84.9Tetramethylene sulfoxide18.2119.1
Dimethyl Sulfoxide18.416.410.2Toluene181.42
Ethanol15.88.819.4Water18.117.116.9
Ethyl Acetate15.85.37.2Xylene17.613.1
Ethyl Benzene17.80.61.4γ-Butyrolactone1916.67.4
Ethylene Carbonate19.421.75.1
Units are in MPa1/2.
Table 11. HSP values of solutes [169,170,171].
Table 11. HSP values of solutes [169,170,171].
SoluteδDδPδH
Lactic acid17.08.328.4
Adipic acid17.19.014.6
Vanillin *19.49.811.2
Furfural18.614.95.1
Ferulic acid *19.06.615.1
4-Hydroxy cinnamic acid *19.16.715.9
Chitosan23.017.325.7
Xylo-oligosaccharides25.47.415.5
Units are in MPa1/2. * Phenolic compound.
Table 12. Chosen solutes of interest and the solvents closest to them in Hansen space.
Table 12. Chosen solutes of interest and the solvents closest to them in Hansen space.
SoluteR1R2R3R4
Lactic acidMeOHEtOH1-Propanol2-Propanol
Succinic acid
Adipic acid
EtOAcn-Butyl acetateSec-Butyl acetateChloroform
Vanillin *Tetramethylene sulfoxide Benzyl alcoholN,N-dimethyl acetamide N,N-dimethyl formamide
Furfuralγ-Butyrolactone Dimethyl phthalate Dimethyl sulfoxide Tetramethylene sulfoxide
Ferulic acid *Benzyl alcohol1-Butanol1-Propanol2-Butanol
4-Hydroxy cinnamic acid *Benzyl alcohol1-Butanol1-Propanol2-Propanol
ChitosanWaterMeOHEtOHDimethyl sulfoxide
Xylo-oligosaccharidesBenzyl alcohol1,4-Dioxane Tetramethylene sulfoxide Dichloromethane
R1 denotes the shortest distance between solute and solvent in Hansen space, R2 denotes the second shortest distance, etc. * Phenolic compound.
Table 13. Optimal extraction parameters by subcritical water extraction measured by Folin–Ciocalteu total phenolics assay. Literature reported in a review by Zhang et al. [181].
Table 13. Optimal extraction parameters by subcritical water extraction measured by Folin–Ciocalteu total phenolics assay. Literature reported in a review by Zhang et al. [181].
Detection MethodRaw MaterialTemperature (°C)Pressure (MPa)Time (min)Ref.
Folin–CiocalteuOleaceae europaea pulp residue160-30[182]
Vitis vinifera pomace14015130[183]
Vitis vinifera pomace12010120[184]
Plantago major100102[185]
Plantago lanceolata100102[185]
Japonica-type rice
(Oryza sativa)
100–3601810–30[186]
Table 14. Review of stability and degradation of phenolic compounds at varying pH levels.
Table 14. Review of stability and degradation of phenolic compounds at varying pH levels.
CompoundStable at pHHigh Degradation at pHRefs.
Gallic acid<7>10[187]
Protocatechuic acid<5>7[188]
Vanillic acid<5>7[188]
Caffiec acid<7>10[187]
Ferulic acid<9>11[187,188]
QuercetinN/A>6[53]
(-)-catechin<7>10[187]
Rutin<11N/A[187]
Neochlorogenic acid <6>9[189]
Cryptochlorogenic acid<5>9[190]
Chlorogenic acid<5>7[191,192]
Table 15. Review of extraction methods for phenolics.
Table 15. Review of extraction methods for phenolics.
BiomassMethodOptimal or Experimental ConditionsReferences
FlaxUSE0.2 M NaOH in water at 25 °C for 60 min at 30 kHz[198]
Crithmum maritimum and Salicornia europaeaUSEWater:ethanol, 40:60 v/v% at 50 °C for 20 min[199,204]
Wheat strawSolventWater:ethanol, 60:40 v/v%, 8 w% NaOH at 70 °C for 18 h[200]
Barley strawAlkaline + solventPretreatment: 2 M NaOH for 1 h, nitrogen atmosphere. Solvent: EtOAc.[161,162]
EnzymaticPepsin, Pancreatin, Pronase E, Viscozyme L
Used coffee bean pulpEnzymatic + fungi + solventPectinase, Rhizomucor pusillus, and EtOAc[202]
Brewer’s spent grainEnzymaticDEPOL 740 L, pH 8 at 50 °C[203]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chaturvedi, T.; Hulkko, L.S.S.; Fredsgaard, M.; Thomsen, M.H. Extraction, Isolation, and Purification of Value-Added Chemicals from Lignocellulosic Biomass. Processes 2022, 10, 1752. https://doi.org/10.3390/pr10091752

AMA Style

Chaturvedi T, Hulkko LSS, Fredsgaard M, Thomsen MH. Extraction, Isolation, and Purification of Value-Added Chemicals from Lignocellulosic Biomass. Processes. 2022; 10(9):1752. https://doi.org/10.3390/pr10091752

Chicago/Turabian Style

Chaturvedi, Tanmay, Laura Sini Sofia Hulkko, Malthe Fredsgaard, and Mette Hedegaard Thomsen. 2022. "Extraction, Isolation, and Purification of Value-Added Chemicals from Lignocellulosic Biomass" Processes 10, no. 9: 1752. https://doi.org/10.3390/pr10091752

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