Open access peer-reviewed chapter

Catalysis for Glycerol Production and Its Applications

Written By

Anele Sibeko, Lethiwe D. Mthembu, Rishi Gupta and Nirmala Deenadayalu

Submitted: 11 October 2022 Reviewed: 15 December 2022 Published: 29 September 2023

DOI: 10.5772/intechopen.109553

From the Edited Volume

Ethanol and Glycerol Chemistry - Production, Modelling, Applications, and Technological Aspects

Edited by Rampal Pandey, Israel Pala-Rosas, José L. Contreras and José Salmones

Chapter metrics overview

78 Chapter Downloads

View Full Metrics

Abstract

Globally, there is a climate change due to greenhouse gases, hence the production processes for chemicals should comply with green chemistry principles to decrease the impact it has on the climate. This book chapter focuses on the catalytic production of glycerol, which is a platform chemical that is widely used in the manufacture of various industrially important chemicals and derivatives, namely 2,3-dihydroxypropanal, glycerol ether, glycerol ester, acrolein, 1,2-propanediol and glycidol. The literature reviewed compares the production of glycerol using homogeneous and heterogeneous catalysts, to determine efficient and environmentally benign glycerol catalysts and to study glycerol as a platform chemical and its value in application.

Keywords

  • catalysis
  • homogenous catalysts
  • heterogeneous catalysts
  • value-added compounds
  • glycerol production

1. Introduction

The enormous growth in demand for fuels, along with growing environmental concerns and limited raw oil sources has increased the use of renewable energy. Biodiesel is one of the potential alternatives, and renewable fuels, has gained popularity in recent years, and their production capacity have grown significantly.

It is produced through various methods such as the transesterification of non-edible and waste vegetable oils with methanol and efforts are also being made to utilise the glycerol by-product to compensate the production cost of biodiesel to make it commercially viable, yielding quite significant percentage of a glycerol by-product which lowers the production cost and makes it commercially available. For every 4 litres of biodiesel generated [1].

Around 500 grams of glycerol is made, this equates to approximately 11,500 tons of 99.9% pure glycerine produced by a plant with a capacity of 113,562,354 million litres per year. The resulting oversupply of raw glycerol from biodiesel production can influence the purified glycerol market significantly as glycerol is a high-value and commercial chemical with thousands of applications [2].

Although extensive research has been carried out on the use of glycerol for various industrial applications, however, a compilation review on different approaches of glycerol production using homogenous and heterogeneous catalysts is scarce. The present chapter focuses on compiling different state-of-the-art in glycerol manufacturing techniques with a special emphasis on homogeneous and heterogeneous catalysis approaches. Moreover, an attempt has also been made to review the application of glycerol in the production of various platform chemicals preferably using microbial pathways. A section has been dedicated on reviewing the application of glycerol in animal feed.

Advertisement

2. Glycerol production

Glycerol can be manufactured using a variety of chemical synthesis feedstocks. It can be produced, for example, by propylene synthesis by several methods [3], such as oil hydrolysis, or transesterification of fatty acids or oils. The following sections describe briefly about different glycerol production processes.

2.1 Glycerol production by propylene

As previously stated, several methods for producing glycerol from propylene can be used [4, 5]. In Figure 1, one of the major processes is shown, which includes the use of chlorination (Cl2) [6].

Figure 1.

Flow diagram illustrating the production of glycerol from propylene [6].

2.1.1 Glycerol production via chlorination process

Propylene chlorination (Figures 1 and 2) produces allyl chloride at a temperature of 510°C in the presence of hypochlorous acid at 38°C. Glycerine dichlorohydrin is formed when allyl chloride reacts. The glycerol dichlorohydrin is then hydrolysed by sodium carbon oxide in a 6% sodium carbonate solution at 96°C or directly to glycerine, the epichlorohydrin being removed as an overhead in a stripping column. Finally, the epichlorohydrin is hydrated to glycerine using sodium hydroxide [4], resulting in a final glycerol yield of around 90% [6].

Figure 2.

Reaction of the propylene chlorination process.

2.1.2 Glycerol production via oxygenation process

Figure 3 illustrates two paths to produce glycerol from propylene via oxygenation. Oxygen (O2) reacts with propylene to produce acrolein, adding an aldehyde (HC=O). Acrolein can be converted to allyl alcohol with a reducing agent sodium borohydride (NaBH4) in a presence of isopropanol as a solvent; peroxide is added to allyl alcohol to produce glycerol. In the other reactions, peroxide is added to acrolein which results in the formation of glyceraldehyde; the glyceraldehyde reacts with hydrogen to produce glycerol.

Figure 3.

Production of glycerol from propylene oxygenation reaction.

2.2 Saponification

In this reaction, sodium hydroxide (base) reacts with triglyceride as an ester to form glycerol and soap molecules. This method has been employed since 2800, and the first industrial factory was developed in 1860 [7]. As demonstrated in Figure 4, this reaction occurs between triglyceride and sodium hydroxide (caustic soda), producing glycerol and soap [6, 8].

Figure 4.

Illustrates the saponification reaction between triglyceride and sodium hydroxide (caustic soda) for the glycerol production [6].

2.3 Transesterification of the beaver oil

The transesterification reaction of beaver oil with ethanol to produce glycerol was carried out in 1864 [9, 10]. Figure 5 shows the reaction in which methyl-esters from triglycerides (oils) and methanol (alcohol) combine to form glycerol and fatty esters (or biodiesel) [5, 6, 11, 12].

Figure 5.

Glycerol production by a transesterification reaction [6].

Interestingly, the glycerol yield from transesterification was not only found to be dependent on different types of processes but also on the different type of oil feedstocks (Table 1) [13, 14, 15, 16, 17, 18].

FeedstockGlycerol concentration (w/w %)Methanol concentration (w/w %)Soap concentration (w/w %)Impurities (w/w %)Ref.
Palm oil waste876[13]
Oil of Jatropha19–2214.52911–21[14]
Soybean oil636.2[15]
Soybean oil2210.926.223.5[15]
Soybean oil3312.626.122.3[15]
Vegetable oil waste2892139[15]
Palm oil8112.0[16]
Seed oils63–77[17]
Used frying oil8515[18]

Table 1.

Different glycerol streams depending on initial feedstocks and production reactions.

Advertisement

3. Glycerol catalysis

Transesterification of oil is accomplished using both homogenous as well as heterogeneous catalysts. Table 2 depicts glycerol production advantages and disadvantages of using different type of catalysts.

Catalysts groupType of catalystAdvantagesDisadvantages
Homogeneous base catalystNaOH/KOH
  • Fast reaction rate, mild condition and less energy intensity.

  • Catalysts are widely available and economical.

  • If the usage limit for oil is less than 0.5 wt. % free fatty acid. Soap formation occurs as well if the free fatty acid content in the oil is more than 2 wt. %.

  • Excessive soap formation reduces the glycerol yield and causes problems during product purification.

Heterogeneous base catalystCaO/MgO
  • Reaction conditions are mild and less energy-intensive, reuse and regenerating of a catalyst.

  • Mild reaction condition and less energy intensive.

  • Sensitive to free fatty acid content in the oil due to its basicity property.

  • Excessive soap formation decreases the glycerol yield and causes problems during product refining.

Homogenous Acid catalystH2SO4/HCl
  • Affordable than base catalysed process.

  • Very slow reaction rate.

Not easy to separate the catalyst from products.

Table 2.

Advantages and disadvantages of glycerol catalysts [19].

Moreover, depending on the type of catalyst used, the transesterification process can be categorised as homogeneous and heterogeneous catalysis to make biodiesel and subsequently glycerol.

3.1 Homogeneous catalysis

During homogenous catalysis, the first stage comprises the reaction of vegetable oils with methanol in the presence of a catalyst, and then the separation of glycerol from the resultant mixture using a settler unit follows. The remaining flow is sent to a chamber that uses mineral acids to remove the catalytic component, resulting in two paths: a glycerol recovery chamber and an evaporator that separates biodiesel from the other products. The unit for purifying comprises three output units: the first with 80–95% glycerol; the second one with water, dissolved salts and unreacted methanol (it is then recycled back to the reactor); and one with fatty esters [12]. Figure 6 depicts the glycerol manufacturing process employing homogeneous catalysts (namely, sodium hydroxide or sodium methylate) [6, 20, 21].

Figure 6.

The production plant for biodiesel is based on a homogenous catalyst.

3.2 Heterogeneous catalysis

This type of catalysis procedure envisions two reaction phases to improve vegetable oil conversion; reactor 1 is supplied by vegetable oil and methanol. The product stream is sent through a heat exchanger to evaporate some of the residual methanol, and the remaining stream is directed to a decanter to separate polar and non-polar components such as glycerol and mainly vegetable oil and biodiesel, respectively. While, in reactor 2, the non-polar stream is reacted for the second time to boost biodiesel synthesis and recover methanol. The product stream travels through the heat exchanger, which takes out all unreacted methanol, and the decanter, which separates the biodiesel from polar components.

The polar streams from the first and second polar decanters are directed to another heat exchanger to recover the remaining methanol in the mixture, while the leftover fraction is delivered to a final decanter to separate vegetable oil and residual glycerol. Figure 7 is a flowchart of triglyceride transesterification using heterogeneous catalysts such as aluminium and zinc oxide [6].

Figure 7.

A heterogeneous catalysis-based production plant flowchart [6].

Advertisement

4. Glycerol: a platform chemical

Synthesis of glycerol following microbial route has been known for over a century, however, new improvements in the biodiesel business have resulted in the production of large amounts of glycerol. During the biodiesel production process, approximately 10% of glycerol is produced, accounting for about 90% of total glycerol produced [22]. Glycerol has gathered substantial interest in its conversion to higher value-added compounds due to its availability and potential to operate as a key building block in a biorefinery (Figure 8) [23].

Figure 8.

Glycerol as a platform chemical.

Glycerol oxidation produces a wide range of compounds, such as glyceric acid dihydroxyacetone, glyceraldehyde, hydroxy-pyruvic acid, glycolic acid and others. Controlling reaction selectivity is a critical challenge in obtaining the desired molecules.

For example, glyceric acid is a crucial intermediary for more extensively oxidised compounds such as tartronic acid and mesoxalic acid. The catalytic aerobic oxidation of glycerol in a basic media has been extensively studied using monometallic or bimetallic catalysts such as Au, Pt and Pd.

Table 3 lists some of the most prominent catalysts used in this field [24, 25, 26, 27, 30, 38, 39, 40]. Another approach for producing value-added compounds from glycerol is the reduction process. Lactic acid is produced by a reduction of glycerol in the presence of hydroxide bases [41]. This reaction is frequently carried out at medium to high pressures and temperatures ranging from 100 to 240°C using Cu- and Zn-based catalysts enhanced by sulphide Ru [28].

Reaction typeReactantCatalystPressure (bar)Temperature (°C)ProductRef.
Glycerol oxidationO2Pd–Ag/C380Dihydroxyacetone[24]
O2Pt/NCNT60[25]
O2Pt/MCN340Glyceraldehyde[26]
O2Pt/SiO21100[27]
O2Pt/MCN340Glyceric acid[26]
O2Pt/SiO21100[27]
Glycerol reductionH2Ru/Al2O3251801,2-propanediol[28]
H2Ru/Al2O325200Ethylene glycol[29]
H2Ru/ZrO280240[30]
Glycerol dehydrogenationAlPO4–4501190–230Acrolein[31]
HY(5.2)1170–230[31]
12 wt. % V2O5, V/P molar ratio of 0.21325[32]
Glycerol halogenationHClAspartic acid4.51001,3-dichloropropanol[33]
HClGlutamic acid4.5100[33]
Glycerol esterificationAcetic acidSb2O5180–120Monoglicerides[34]
Palmitic acidZrSBA-151160–180Diacylglicerol[35]
Acetic acidGraphene oxide1120[36]
Acetic acidZSM-51120[36]
Glycerol pyrolysisBituminous carbon1400–900Syngas[37]
Coconut shell1400–900[37]

Table 3.

Compounds derived from traditional glycerol conversion under similar operating conditions.

Glycerol carbonate is another derivative of glycerol that is formed by reaction between glycerol and urea, ethylene or propylene carbonate [22] or carbon dioxide [42]. It is also used in the commercial manufacture of epichlorohydrin. Epichlorohydrin is produced in a similar manner by Solvay and Dow Chemical Company [43]. When the principal hydroxyl groups in glycerol are selectively oxidised, the economically valuable chemicals glyceraldehyde [44], glyceric acid [45] and tartronic acid [46] are formed. Dihydroxyacetone (DHA) is produced by oxidation of the secondary hydroxyl group, whereas ketomalonic acid is produced by oxidation of all three hydroxyl groups [47].

Another glycerol derivative, glycidol, offers immense potential for the synthesis of industrially useful compounds such as epoxy resins, polyurethanes and polyglycerol esters. A bio-based technique for producing glycidol from glycerol was recently published [48]. The manufacture of acrolein from glycerol is an innovative, eco-friendly technology that has several advantages, such as less oil extraction and a minimal environmental impact [31]. In general, acrolein is synthesised from glycerol by acid-catalysed dehydrogenation over synthetic aluminium phosphate (AlPO4), zeolites with varied channel configurations (HY and H-ZSM-5) and SiO2/Al2O3 ratio [31, 49].

A novel synthetic approach for the synthesis of chlorohydrin was proposed, which involved reacting a polyhydroxy aliphatic hydrocarbon with a chlorination agent. Vitiello et al. [33] focused on the activity and selectivity of homologous chlorinated series of catalysts for glycerol halogenation, such as acetic acid, monochloro, dichloro and trichloroacetic acid.

Table 3 also includes information on one of the most significant glycerol conversion processes, esterification with acetic acid, which produces monoacylglycerol, diacylglycerol and glycerol carbonate. These materials are often used in cryogenics, biodegradable polyester and cosmetics [35, 36]. Sulphated-based superacids, heteropoly acid-based catalysts, tin chloride, zeolite, ZrO2-based solid acids and other significant acid catalysts can be used for glycerol esterification [35, 36, 37, 50, 51, 52].

Finally, pyrolysis of glycerol to produce syngas is another method of converting glycerol. The pyrolysis of biomass has been extensively studied in the specialist literature, although in most cases, only metal-based catalysts have been used. The microwave-assisted pyrolysis of glycerol over a carbonaceous catalyst is a unique approach for syngas generation in which the heating method and operating temperature (between 400 and 900°C) can impact the catalytic action of the activated carbons to optimise syngas production [37]. Table 3 outlines some of the products derived from glycerol that may be transformed into other compounds with high added value.

Advertisement

5. Value-added products from glycerol via biological conversions

From 2004 to 2008, the global production of crude glycerol from biodiesel conversion increased from 200 thousand tonnes to 1.224 million tonnes [2, 19]. Meanwhile, in 2005, the global market for purified glycerol was anticipated to be over 900,000 tonnes [53]. This provided a chance for scientists to discover new uses for refined and crude glycerol. Multiple publications on the direct use of crude glycerol from biodiesel synthesis have been published.

5.1 1,3-Propanediol

The most promising alternative for the biological conversion of glycerol in anaerobic fermentative production is 1,3-propanediol [54], which indicates that crude glycerol could be employed directly for the manufacture of 1,3-propanediol in fed-batch cultures of pneumoniae.

Raw glycerol composition had less influence on the biological conversion and, therefore, a low fermentation cost could be predicted. However, using a response surface approach, the generation of 1,3-propanediol by Klebsiella pneumoniae was optimised. The highest concentration of 1,3-propanediol produced However, statistical optimisation along with genetic engineering approaches may be utilised to improve the 1,3-propanediol production [14, 55]. K. pneumoniae ATCC 15380 recently improved the synthesis of 1,3-propanediol from crude glycerol from Jatropha biodiesel. The yield, purity and recovery of 1,3-propanediol obtained were 56 g/L, 99.7% and 34%, respectively [14]. In addition, a hollow fibre membrane was used to produce an integrated bioprocess that linked biodiesel generation by lipase with microbial production of 1,3-propanediol by K. pneumoniae [56].

5.2 Citric acid

Citric acid synthesis by Yarrowia lipolytica ACA-DC 50109 from raw glycerol was not only comparable to that obtained from sugar-based standard media [57] but also single-cell oil and citric acid were produced simultaneously [58, 59]. When acetate-negative mutants of the Y. lipolytica Wratislavia AWG7 strain were employed in a fed-batch fermentation to ferment crude glycerol, the final concentration of citric acid was 131.5 g/L, which was similar to that produced from pure glycerol (139 g/L). Similarly, Y. lipolytica LGAM S(7)1 has also shown the ability to convert crude glycerol to citric acid [60]. Interestingly, another strain Y. lipolytica N15 could produce large levels of citric acid, namely up to 98 g/L of citric acid and 71 g/L of citric acid from pure glycerol medium and crude glycerol medium, respectively [61].

5.3 Hydrogen and other lower molecule fuels

The photo-fermentative conversion of crude glycerol to hydrogen is one of the most fascinating approach to utilise glycerol. Both crude glycerol and pure glycerol can produce up to 6 moles of H2 per mole of glycerol, representing 75% of the theoretical value. However, significant technological challenges, such as increasing the efficiency of light use by organisms and building effective photobioreactors, must be overcome before a viable method can be developed [62]. When Enterobacter aerogenes HU-101 was used, hydrogen and ethanol were synthesised at high yields and rates. However, in order to improve the rate of glycerol use, the crude glycerol should be diluted with a synthetic medium [63]. While Jitrwung and Yargeau [64] modified several media compositions of the E. aerogenes ATCC 35029 fermented crude glycerol procedure to maximise hydrogen generation.

5.4 Polyhydroxyalkanoates (PHB)

As an estimate, a biodiesel facility with a capacity of 10 million gallons per year could produce 20.9 tons of PHB [65]. The feasibility of using crude glycerol for PHB manufacture was investigated using Paracoccus denitrificans and Cupriavidus necator JMP134, and the resultant polymers were shown to be remarkably comparable to those generated from glucose. However, a high osmotic (sodium chloride-contaminated) crude glycerol was found to have harmful impact on PHB synthesis and needs to be taken care of. One way to handle the issue is combining crude glycerol from various producers to reduce the harmful effect of NaCl contamination [66]. In addition, for a large-scale PHB synthesis, a technique based on the C. necator DSM 545 fermentation of crude glycerol was developed [67]. Following this in the presence of NaCl, Zobellella denitrificans MW1 could use crude glycerol for growth and PHB synthesis at high concentrations. As a result, it was recommended as an appealing alternative for large-scale PHB manufacturing using crude glycerol [68]. Furthermore, when mixed microbial consortia (MMC) were utilised to produce PHA from crude glycerol, it was shown that methanol in the crude glycerol was converted to PHB by MMC.

5.5 Lipids as the sole carbon source

Crude glycerol might be used to manufacture lipids, which could be utilised to make a sustainable biodiesel feedstock. For example, raw glycerol might be used to culture Schizochytrium limacinum SR21 and Cryptococcus curvatus. However, the glycerol content over a certain threshold may prevent the rapid reproduction of cells. The best glycerol content for batch culturing of crude glycerol obtained from yellow grease were 25 and 35 g/L for untreated and treated crude glycerol, respectively, which may subsequently lead to cellular lipid content of approximately 75%. Methanol residues in crude glycerol may cause damage to the development of S. limacinum SR21 [69].

For lipid synthesis in C. curvatus yeast, fed-batch was preferable to batch; however, the addition of ammonium sulphate and Tween 20 improved the accumulation of lipids and carotenoids Saenge et al. [70] demonstrated that the oleaginous red yeast Rhodotorula glutinis TISTR 5159 generated lipids and carotenoids when grown on crude glycerol. Chlorella protothecoides was also capable of converting crude glycerol to lipids.

The lipid yield was 0.31 g lipids/g substrate [71]. Similarly, using C. protothecoides and crude glycerol (62% purity), Furthermore, Chatzifragkou et al. [72] did research, to investigate the ability of 15 eukaryotic micro-organisms to change crude glycerol to metabolic products. The results showed that yeast accumulated limited lipids (up to 22 wt.% in the case of Rhodotorula), whereas fungi collected greater levels of lipids in their mycelia (range between 18.1 and 42.6 wt.% of dry biomass). Interestingly, Chen and Walker [73] found that a fed-batch operation yielded a maximum lipid productivity of 3 g/L per day, which was greater than that generated by a batch procedure.

Tables 4 and 5 outline an overview of the conversions of crude glycerol to potential chemical through biological and catalytic conversions.

ProductReactionYieldRef.
1,3-PropanediolFed-batch cultures of Klebsiella pneumoniae strain1.7 g/L/h[54]
Maximum 1,3-propanediol production from K. pneumonia13.8 g/L[55]
Citric acidYarrowia lipolytica strain ACA-DC 50109 (process modelling)NA[57]
Acetate mutants of Y. lipolytica Wratislavia AWG7 strain; fed-batch operation139 g/L[74]
Y. lipolytica strain LGAM S (7)135 g/L[60]
HydrogenPhotofermentative conversion process; Rhodopseudomonas palustris strain6 mol/mol glycerol[62]
Enterobacter aerogenes strain HU-101; continuous culture; porous ceramics as a support material to fix cells63 mmol/L/h[63]
Poly(hydroxyalkanoates) (PHAs)Pseudomonas oleovorans NRRL B-14682 and P. corrugata 388 grew and synthesised PHB and mcl-PHA, respectivelyNA[75]
Producing PHB; Paracoccus denitrificans and Cupriavidus necator JMP 134 strains48%[66]
LipidSchizochytrium limacinum SR21; batch culture73.3%[69]
Cryptococcus curvatus; two-stage fed-batch process52%[76]

Table 4.

Biological conversion of crude glycerol.

ProductReactionYieldRef.
AcroleinFluidised bed, tungsten-doped zirconia catalyst21%[77]
MonoglycerideTwo-step process, purification of the monoglyceride produced from glycerolysis of palm stearin~99% purity[78]
Glycerolysis of soybean oil~42%[79]
Gaseous productsSteam gasification with in situ CO2 removal88 vol.% H2 purity[80]
Hydrothermal reforming of crude glycerol~90 vol.% H2 purity[81]

Table 5.

Conventional catalytic conversions of crude glycerol.

Advertisement

6. Application of crude glycerol in animal feedstock

Glycerol has been used as an animal feed additive since the 1970s [82]. However, the availability of glycerol has limited its application in diets [83], because of the rising corn prices and the oversupply of crude glycerol, the possibility of using crude glycerol from biodiesel in feeds has recently been examined.

6.1 Crude glycerol in non-ruminant diets

Crude glycerol is an excellent energy source due to its high absorption rates for non-ruminants such as broilers. Once ingested, the enzyme glycerol kinase converts it to glucose for energy generation in the liver of mammals [83]. Its samples from various biodiesel manufacturers were tested as energy sources. The digestible energy (DE) values for 85% of the crude glycerol samples ranged from 14.9 to 15.3 MJ/kg, with metabolisable energy (ME) values ranging from 13.9 to 14.7 MJ/kg [84]. Overall, the use of crude glycerol derived from biodiesel process as an animal feed component offers significant potential for replacing maize in diets and is gaining popularity. However, the existence of potentially dangerous contaminants in biodiesel crude glycerol needs to be taken into consideration [85].

6.2 Crude glycerol in ruminant diets

Besides, the non-ruminants, crude glycerol may play a very significant role in the diets of ruminant animals as well. However, to improve its edibility, more emphasis should be placed on the crude glycerol produced by small-scale biodiesel plants that employ basic batch distillation or evaporation processes. There are several reports where use of crude glycerol has shown significant improvement in the overall performance of ruminants. Crude glycerol, at up to 15% dry matter in finishing lamb diets, might increase feedlot performance, particularly during the first 14 days, but had little influence on carcass attributes [86]. Diets for meat goats containing up to 5% crude glycerol were shown to be superior to medium-quality hay [87]. Nursing dairy cows can also be fed up to 15% of their dry matter diet without affecting feed intake, milk output or yield [88, 89]. When crude glycerol was added at 8% or less of dry matter in cow-finishing diets, its weight growth and feed efficiency were increased [90].

Advertisement

7. Summary and conclusions

Glycerol may be produced using various techniques and feedstocks, such as propylene synthesis by various routes, hydrolysis of fatty acid triglycerides, or transesterification of fatty acids or oils. The efficient use of crude glycerol is critical to the commercialisation and advancement of biodiesel synthesis. In the long run, using biomass-derived glycerol will not only help to reduce society’s reliance on non-renewable resources, but it will also encourage the development of integrated biorefineries. This review focuses on the value-added prospects for crude glycerol derived from biodiesel production, primarily as a feed ingredient for animal feed and as a feedstock for chemicals.

For example, crude glycerol can be converted into 1,3-propanediol, citric acid, poly(hydroxyalkanoates), butanol, hydrogen, docosahexaenoic acid-rich algae, monoglycerides, lipids and syngas. Though many of the processes discussed have already been employed by the industries, they require additional research to minimise the manufacturing cost and be operationally practical for inclusion into biorefineries.

Furthermore, contaminants in crude glycerol can have a noticeable impact on the conversion of glycerol into other products. Pollutants in crude glycerol hinder cell and fungi’s rapid reproduction, resulting in less production rates and product yields in many biological conversion processes (compared with pure or commercial glycerol under the same culture conditions). Contaminants, on the other hand, poison the catalysts in traditional catalytic conversions, boosting char generation and affecting product yield.

Many technologies need to be better understood and refined, such as optimising reaction parameters, production yields and fermentation conditions; generating mutant strains and efficient bioreactors for stable cultures and enhancing the activity and selectivity of catalysts.

Researchers have also obtained promising results on utilisation of crude glycerol as animal feed, particularly with non-ruminant animals such as pigs, laying hens and broilers. But various precautions must be taken before this biomass-derived chemical may be used on a large scale in animal diets. To begin with, animal producers must exercise caution when deciding to incorporate crude glycerol as a component of animal feed diets, since the chemical composition of crude glycerol varies greatly depending on the processes and feedstocks used to manufacture biodiesel. Secondly, contaminants in crude glycerol affect feed performance to some extent. Finally, the amount of crude glycerol in feed formulations must be considered. It is advised that a crude glycerol feed standard be established so that it would be uniform for all producers, the resulting “standard” crude glycerol would have greater value.

There is a need to develop improved processes as well as other important value-added products. For example, among other renewable and bio-derived sources, glycerol has come up as an appealing possibility since it represents a relevant and alternative solution for producing hydrogen via reforming processes that may be carried out in both traditional and novel reactors.

Besides, catalytic process, though it is not yet introduced, the transesterification reaction using supercritical fluids has also gained noticeable attention. As one or two reaction stages are possible in a single-step supercritical fluid transesterification, the reaction occurs only once reactants are heated to critical temperatures and pressures with triglycerides [20, 21]. Triglycerides are initially transformed to free fatty acids and by-products in the hydrolysis reaction during the two-step subcritical-supercritical fluid transesterification. The acquired free fatty acids undergo esterification reaction, yielding fatty acid methyl esters in a supercritical fluid process [91, 92].

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Yang F, Hanna MA, Sun R. Value-added uses for crude glycerol--a byproduct of biodiesel production. Biotechnology for Biofuels. 2012;5(1):1-10
  2. 2. Kerr BJ, Dozier WA III, Bregendahl K. Nutritional value of crude glycerin for nonruminants. In: Proceedings of the 23rd Annual Carolina Swine Nutrition Conference. Journal of Animal Science. Raleigh, NC; 2007. pp. 6-18
  3. 3. Shukla K, Srivastava VC. Synthesis of organic carbonates from alcoholysis of urea: A review. Catalysis Reviews. 2017;59(1):1-43
  4. 4. Ciriminna R, Pina CD, Rossi M, Pagliaro M. Understanding the glycerol market. European Journal of Lipid Science and Technology. 2014;116(10):1432-1439
  5. 5. Matar S, Hatch LF. Chemistry of Petrochemical Processes. Elsevier; 2001
  6. 6. Bagnato G, Iulianelli A, Sanna A, Basile A. Glycerol production and transformation: A critical review with particular emphasis on glycerol reforming reaction for producing hydrogen in conventional and membrane reactors. Membranes. 2017;7(2):17
  7. 7. Hunt JA. A short history of soap. The Pharmaceutical Journal. 1999;263(7076):985-989
  8. 8. Marchetti J, Miguel V, Errazu A. Possible methods for biodiesel production. Renewable and Sustainable Energy Reviews. 2007;11(6):1300-1311
  9. 9. Formo M, Jungermann E, Norris F, Sonntag N. In: Swern D, editor. Bailey's Industrial Oil and Fat Products. New York, NY: John Wiley and Sons; 1979
  10. 10. Quispe CA, Coronado CJ, Carvalho JA Jr. Glycerol: Production, consumption, prices, characterization and new trends in combustion. Renewable and Sustainable Energy Reviews. 2013;27:475-493
  11. 11. Wang Z, Zhuge J, Fang H, Prior BA. Glycerol production by microbial fermentation: A review. Biotechnology Advances. 2001;19(3):201-223
  12. 12. Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterification of oils. Journal of Bioscience and Bioengineering. 2001;92(5):405-416
  13. 13. Hidawati EN, Sakinah AM. "treatment of glycerin pitch from biodiesel production," international journal of chemical and environmental. Engineering. 2011;2(5):309-313
  14. 14. Hiremath A, Kannabiran M, Rangaswamy V. 1,3-Propanediol production from crude glycerol from Jatropha biodiesel process. New Biotechnology. 2011;28(1):19-23
  15. 15. Hu S, Luo X, Wan C, Li Y. Characterization of crude glycerol from biodiesel plants. Journal of Agricultural and Food Chemistry. 2012;60(23):5915-5921
  16. 16. Liu Y-P, Sun Y, Tan C, Li H, Zheng X-J, Jin K-Q , et al. Efficient production of dihydroxyacetone from biodiesel-derived crude glycerol by newly isolated Gluconobacter frateurii. Bioresource Technology. 2013;142:384-389
  17. 17. Tan H, Aziz AA, Aroua M. Glycerol production and its applications as a raw material: A review. Renewable and Sustainable Energy Reviews. 2013;27:118-127
  18. 18. Bouaid A, Vázquez R, Martinez M, Aracil J. Effect of free fatty acids contents on biodiesel quality. Pilot plant studies. Fuel. 2016;174:54-62
  19. 19. Nomanbhay S, Hussain R, Rahman MM, Palanisamy K. Review paper integration of biodiesel and bioethanol processes: Convertion of low cost waste glycerol to bioethanol. Advances in Natural and Applied Sciences. 2012;6(6):802-818
  20. 20. Ilham Z, Saka S. Dimethyl carbonate as potential reactant in non-catalytic biodiesel production by supercritical method. Bioresource Technology. 2009;100(5):1793-1796
  21. 21. Goembira F, Saka S. Optimization of biodiesel production by supercritical methyl acetate. Bioresource Technology. 2013;131:47-52
  22. 22. Okoye P, Hameed B. Review on recent progress in catalytic carboxylation and acetylation of glycerol as a byproduct of biodiesel production. Renewable and Sustainable Energy Reviews. 2016;53:558-574
  23. 23. Schultz EL, de Souza DT, Damaso MCT. The glycerol biorefinery: A purpose for Brazilian biodiesel production. Chemical and Biological Technologies in Agriculture. 2014;1(1):1-9
  24. 24. Hirasawa S, Watanabe H, Kizuka T, Nakagawa Y, Tomishige K. Performance, structure and mechanism of Pd–Ag alloy catalyst for selective oxidation of glycerol to dihydroxyacetone. Journal of Catalysis. 2013;300:205-216
  25. 25. Ning X, Li Y, Yu H, Peng F, Wang H, Yang Y. Promoting role of bismuth and antimony on Pt catalysts for the selective oxidation of glycerol to dihydroxyacetone. Journal of Catalysis. 2016;335:95-104
  26. 26. Wang F-F, Shao S, Liu C-L, Xu C-L, Yang R-Z, Dong W-S. Selective oxidation of glycerol over Pt supported on mesoporous carbon nitride in base-free aqueous solution. Chemical Engineering Journal. 2015;264:336-343
  27. 27. Li Y, Zaera F. Sensitivity of the glycerol oxidation reaction to the size and shape of the platinum nanoparticles in Pt/SiO2 catalysts. Journal of Catalysis. 2015;326:116-126
  28. 28. Lee S-H, Moon DJ. Studies on the conversion of glycerol to 1, 2-propanediol over Ru-based catalyst under mild conditions. Catalysis Today. 2011;174(1):10-16
  29. 29. Soares AV, Salazar JB, Falcone DD, Vasconcellos FA, Davis RJ, Passos FB. A study of glycerol hydrogenolysis over Ru–Cu/Al2O3 and Ru–Cu/ZrO2 catalysts. Journal of Molecular Catalysis A: Chemical. 2016;415:27-36
  30. 30. Olmos CM, Chinchilla LE, Rodrigues EG, Delgado JJ, Hungría AB, Blanco G, et al. Synergistic effect of bimetallic Au-Pd supported on ceria-zirconia mixed oxide catalysts for selective oxidation of glycerol. Applied Catalysis B: Environmental. 2016;197:222-235
  31. 31. Estevez R, Lopez-Pedrajas S, Blanco-Bonilla F, Luna D, Bautista F. Production of acrolein from glycerol in liquid phase on heterogeneous catalysts. Chemical Engineering Journal. 2015;282:179-186
  32. 32. Cecilia J, García-Sancho C, Mérida-Robles J, Santamaría-González J, Moreno-Tost R, Maireles-Torres P. V and V–P containing Zr-SBA-15 catalysts for dehydration of glycerol to acrolein. Catalysis Today. 2015;254:43-52
  33. 33. Vitiello R, Russo V, Turco R, Tesser R, Di Serio M, Santacesaria E. Glycerol chlorination in a gas-liquid semibatch reactor: New catalysts for chlorohydrin production. Chinese Journal of Catalysis. 2014;35(5):663-669
  34. 34. Hu W, Zhang Y, Huang Y, Wang J, Gao J, Xu J. Selective esterification of glycerol with acetic acid to diacetin using antimony pentoxide as reusable catalyst. Journal of Energy Chemistry. 2015;24(5):632-636
  35. 35. Yusoff MHM, Abdullah AZ. Catalytic behavior of sulfated zirconia supported on SBA-15 as catalyst in selective glycerol esterification with palmitic acid to monopalmitin. Journal of the Taiwan Institute of Chemical Engineers. 2016;60:199-204
  36. 36. Gao X, Zhu S, Li Y. Graphene oxide as a facile solid acid catalyst for the production of bioadditives from glycerol esterification. Catalysis Communications. 2015;62:48-51
  37. 37. Fernández Y, Arenillas A, Díez M, Pis J, Menéndez J. Pyrolysis of glycerol over activated carbons for syngas production. Journal of Analytical and Applied Pyrolysis. 2009;84(2):145-150
  38. 38. Rodrigues EG, Pereira MF, Delgado JJ, Chen X, Órfão JJ. Enhancement of the selectivity to dihydroxyacetone in glycerol oxidation using gold nanoparticles supported on carbon nanotubes. Catalysis Communications. 2011;16(1):64-69
  39. 39. Skrzyńska E, Zaid S, Girardon J-S, Capron M, Dumeignil F. Catalytic behaviour of four different supported noble metals in the crude glycerol oxidation. Applied Catalysis A: General. 2015;499:89-100
  40. 40. Gil S, Marchena M, Fernández CM, Sánchez-Silva L, Romero A, Valverde JL. Catalytic oxidation of crude glycerol using catalysts based on Au supported on carbonaceous materials. Applied Catalysis A: General. 2013;450:189-203
  41. 41. Maris EP, Ketchie WC, Murayama M, Davis RJ. Glycerol hydrogenolysis on carbon-supported PtRu and AuRu bimetallic catalysts. Journal of Catalysis. 2007;251(2):281-294
  42. 42. Ma J, Song J, Liu H, Liu J, Zhang Z, Jiang T, et al. One-pot conversion of CO 2 and glycerol to value-added products using propylene oxide as the coupling agent. Green Chemistry. 2012;14(6):1743-1748
  43. 43. Bell BM, Briggs JR, Campbell RM, Chambers SM, Gaarenstroom PD, Hippler JG, et al. Glycerin as a renewable feedstock for epichlorohydrin production. The GTE process. CLEAN–Soil, Air, Water. 2008;36(8):657-661
  44. 44. Kim HJ, Lee J, Green SK, Huber GW, Kim WB. Selective glycerol oxidation by electrocatalytic dehydrogenation. ChemSusChem. 2014;7(4):1051-1056
  45. 45. Kondamudi N, Misra M, Banerjee S, Mohapatra S, Mohapatra S. Simultaneous production of glyceric acid and hydrogen from the photooxidation of crude glycerol using TiSi2. Applied Catalysis B: Environmental. 2012;126:180-185
  46. 46. Behr A, Eilting J, Irawadi K, Leschinski J, Lindner F. Improved utilisation of renewable resources: New important derivatives of glycerol. Green Chemistry. 2008;10(1):13-30
  47. 47. Ciriminna R, Pagliaro M. One-pot homogeneous and heterogeneous oxidation of glycerol to ketomalonic acid mediated by TEMPO. Advanced Synthesis & Catalysis. 2003;345(3):383-388
  48. 48. Bai R, Zhang H, Mei F, Wang S, Li T, Gu Y, et al. Retraction: One-pot synthesis of glycidol from glycerol and dimethyl carbonate over a highly efficient and easily available solid catalyst NaAlO 2. Green Chemistry. 2016;18(22):6144-6144
  49. 49. Jiang T, Zhou Y, Liang S, Liu H, Han B. Hydrogenolysis of glycerol catalyzed by Ru-Cu bimetallic catalysts supported on clay with the aid of ionic liquids. Green Chemistry. 2009;11(7):1000-1006
  50. 50. Ishak ZI, Sairi NA, Alias Y, Aroua MKT, Yusoff R. Production of glycerol carbonate from glycerol with aid of ionic liquid as catalyst. Chemical Engineering Journal. 2016;297:128-138
  51. 51. Munshi MK, Biradar PS, Gade SM, Rane VH, Kelkar AA. Efficient synthesis of glycerol carbonate/glycidol using 1,8-diazabicyclo [5.4. 0] undec-7-ene (DBU) based ionic liquids as catalyst. RSC Advances. 2014;4(33):17124-17128
  52. 52. Liu J, Li Y, Zhang J, He D. Glycerol carbonylation with CO2 to glycerol carbonate over CeO2 catalyst and the influence of CeO2 preparation methods and reaction parameters. Applied Catalysis A: General. 2016;513:9-18
  53. 53. Nilles D. Combating the Glycerin Glut. Biodiesel Magazine; 2006 Available: https://biodieselmagazine.com/articles/1123/combating-the-glycerin-glut/
  54. 54. Mu Y, Teng H, Zhang D-J, Wang W, Xiu Z-L. Microbial production of 1,3-propanediol by Klebsiella pneumoniae using crude glycerol from biodiesel preparations. Biotechnology Letters. 2006;28(21):1755-1759
  55. 55. Oh B-R, Seo J-W, Choi MH, Kim CH. Optimization of culture conditions for 1,3-propanediol production from crude glycerol by Klebsiella pneumoniae using response surface methodology. Biotechnology and Bioprocess Engineering. 2008;13(6):666-670
  56. 56. Mu Y, Xiu Z-L, Zhang D-J. A combined bioprocess of biodiesel production by lipase with microbial production of 1,3-propanediol by Klebsiella pneumoniae. Biochemical Engineering Journal. 2008;40(3):537-541
  57. 57. Papanikolaou S, Aggelis G. Modelling aspects of the biotechnological valorization of raw glycerol: Production of citric acid by Yarrowia lipolytica and 1,3-propanediol by clostridium butyricum. Journal of Chemical Technology & Biotechnology: International Research in Process, Environmental & Clean Technology. 2003;78(5):542-547
  58. 58. Papanikolaou S, Fakas S, Fick M, Chevalot I, Galiotou-Panayotou M, Komaitis M, et al. Biotechnological valorisation of raw glycerol discharged after bio-diesel (fatty acid methyl esters) manufacturing process: Production of 1, 3-propanediol, citric acid and single cell oil. Biomass and Bioenergy. 2008;32(1):60-71
  59. 59. Papanikolaou S, Aggelis G. Biotechnological valorization of biodiesel derived glycerol waste through production of single cell oil and citric acid by Yarrowia lipolytica. Lipid Technology. 2009;21(4):83-87
  60. 60. Papanikolaou S, Muniglia L, Chevalot I, Aggelis G, Marc I. Yarrowia lipolytica as a potential producer of citric acid from raw glycerol. Journal of Applied Microbiology. 2002;92(4):737-744
  61. 61. Kamzolova SV, Fatykhova AR, Dedyukhina EG, Anastassiadis SG, Golovchenko NP, Morgunov IG. Citric acid production by yeast grown on glycerol-containing waste from biodiesel industry. Food Technology and Biotechnology. 2011;49(1):65-74
  62. 62. Sabourin-Provost G, Hallenbeck PC. High yield conversion of a crude glycerol fraction from biodiesel production to hydrogen by photofermentation. Bioresource Technology. 2009;100(14):3513-3517
  63. 63. Ito T, Nakashimada Y, Senba K, Matsui T, Nishio N. Hydrogen and ethanol production from glycerol-containing wastes discharged after biodiesel manufacturing process. Journal of Bioscience and Bioengineering. 2005;100(3):260-265
  64. 64. Jitrwung R, Yargeau V. Optimization of media composition for the production of biohydrogen from waste glycerol. International Journal of Hydrogen Energy. 2011;36(16):9602-9611
  65. 65. Dobroth ZT, Hu S, Coats ER, McDonald AG. Polyhydroxybutyrate synthesis on biodiesel wastewater using mixed microbial consortia. Bioresource Technology. 2011;102(3):3352-3359
  66. 66. Mothes G, Schnorpfeil C, Ackermann JU. Production of PHB from crude glycerol. Engineering in Life Sciences. 2007;7(5):475-479
  67. 67. Cavalheiro JM, de Almeida MCM, Grandfils C, Da Fonseca M. Poly (3-hydroxybutyrate) production by Cupriavidus necator using waste glycerol. Process Biochemistry. 2009;44(5):509-515
  68. 68. Ibrahim MH, Steinbüchel A. Poly (3-hydroxybutyrate) production from glycerol by Zobellella denitrificans MW1 via high-cell-density fed-batch fermentation and simplified solvent extraction. Applied and Environmental Microbiology. 2009;75(19):6222-6231
  69. 69. Liang Y, Sarkany N, Cui Y, Blackburn JW. Batch stage study of lipid production from crude glycerol derived from yellow grease or animal fats through microalgal fermentation. Bioresource Technology. 2010;101(17):6745-6750
  70. 70. Saenge C, Cheirsilp B, Suksaroge TT, Bourtoom T. Potential use of oleaginous red yeast Rhodotorula glutinis for the bioconversion of crude glycerol from biodiesel plant to lipids and carotenoids. Process Biochemistry. 2011;46(1):210-218
  71. 71. O’Grady J, Morgan JA. Heterotrophic growth and lipid production of Chlorella protothecoides on glycerol. Bioprocess and Biosystems Engineering. 2011;34(1):121-125
  72. 72. Chatzifragkou A, Makri A, Belka A, Bellou S, Mavrou M, Mastoridou M, et al. Biotechnological conversions of biodiesel derived waste glycerol by yeast and fungal species. Energy. 2011;36(2):1097-1108
  73. 73. Chen Y-H, Walker TH. Biomass and lipid production of heterotrophic microalgae Chlorella protothecoides by using biodiesel-derived crude glycerol. Biotechnology Letters. 2011;33(10):1973-1983
  74. 74. Rywinska A, Rymowicz W, Zarowska B, Wojtatowicz M. Biosynthesis of citric acid from glycerol by acetate mutants of Yarrowia lipolytica in fed-batch fermentation. Food Technology and Biotechnology. 2009;47(1):1-6
  75. 75. Ashby RD, Solaiman DK, Foglia TA. Bacterial poly (hydroxyalkanoate) polymer production from the biodiesel co-product stream. Journal of Polymers and the Environment. 2004;12(3):105-112
  76. 76. Liang Y, Cui Y, Trushenski J, Blackburn JW. Converting crude glycerol derived from yellow grease to lipids through yeast fermentation. Bioresource Technology. 2010;101(19):7581-7586
  77. 77. Sereshki BR, Balan S-J, Patience G, Dubois J-L. Reactive vaporization of crude glycerol in a fluidized bed reactor. Industrial & Engineering Chemistry Research. 2010;49(3):1050-1056
  78. 78. Chetpattananondh P, Tongurai C. "synthesis of high purity monoglycerides from crude glycerol and palm stearin," Songklanakarin. Journal of Science & Technology. 2008;30(4):515-521
  79. 79. Echeverri DA, Cardeño F, Rios LA. Glycerolysis of soybean oil with crude glycerol c containing residual alkaline catalysts from biodiesel production. Journal of the American Oil Chemists' Society. 2011;88(4):551-557
  80. 80. Dou B, Dupont V, Williams PT, Chen H, Ding Y. Thermogravimetric kinetics of crude glycerol. Bioresource Technology. 2009;100(9):2613-2620
  81. 81. Dou B, Rickett GL, Dupont V, Williams PT, Chen H, Ding Y, et al. Steam reforming of crude glycerol with in situ CO2 sorption. Bioresource Technology. 2010;101(7):2436-2442
  82. 82. Fisher L, Erfle J, Lodge G, Sauer F. Effects of propylene glycol or glycerol supplementation of the diet of dairy cows on feed intake, milk yield and composition, and incidence of ketosis. Canadian Journal of Animal Science. 1973;53(2):289-296
  83. 83. Mario Pagliaro MR. Royal Society of Chemistry (Great Britain), the Future of Glycerol: New Uses of a Versatile Raw Material. 2nd ed. Cambridge: Royal Society of Chemistry; 2008
  84. 84. Dasari M. Crude glycerol potential described. Feedstuffs. 2007;79(43):1-3
  85. 85. Cerrate S, Yan F, Wang Z, Coto C, Sacakli P, Waldroup P. Evaluation of glycerine from biodiesel production as a feed ingredient for broilers. International Journal of Poultry Science. 2006;5(11):1001-1007
  86. 86. Gunn P, Neary M, Lemenager R, Lake S. Effects of crude glycerin on performance and carcass characteristics of finishing wether lambs. Journal of Animal Science. 2010;88(5):1771-1776
  87. 87. Hampy K, Kellogg D, Coffey K, Kegley E, Caldwell J, Lee M, et al. Glycerol as a supplemental energy source for meat goats. AAES Research Series. 2008;553:63-64
  88. 88. Donkin SS. Glycerol from biodiesel production: The new corn for dairy cattle. Revista Brasileira de Zootecnia. 2008;37:280-286
  89. 89. Donkin S, Koser S, White H, Doane P, Cecava M. Feeding value of glycerol as a replacement for corn grain in rations fed to lactating dairy cows. Journal of Dairy Science. 2009;92(10):5111-5119
  90. 90. Parsons G, Shelor M, Drouillard J. Performance and carcass traits of finishing heifers fed crude glycerin. Journal of Animal Science. 2009;87(2):653-657
  91. 91. Saka S, Isayama Y, Ilham Z, Jiayu X. New process for catalyst-free biodiesel production using subcritical acetic acid and supercritical methanol. Fuel. 2010;89(7):1442-1446
  92. 92. Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic catalysis for transesterification of high free fatty acid oil (waste cooking oil) to biodiesel: A review. Biotechnology Advances. 2010;28(4):500-518

Written By

Anele Sibeko, Lethiwe D. Mthembu, Rishi Gupta and Nirmala Deenadayalu

Submitted: 11 October 2022 Reviewed: 15 December 2022 Published: 29 September 2023