14

Biodiesel and renewable diesel production methods

J.H. Van Gerpen and B.B. He,    University of Idaho, USA

Abstract:

Vegetable oils and animal fats have come to be recognized as important sources of renewable fuels. Fatty acid methyl esters are known as biodiesel, and are the leading source for non-petroleum diesel fuel. With hydrogenation and isomerization, oils and fats can be converted to renewable diesel and jet fuel that are drop-in replacements for petroleum diesel fuel and jet fuel. This chapter reviews the processes used to produce these fuels and the feedstocks that will provide the supplies of oils and fats needed to meet a growing world demand.

Key words

biodiesel; renewable diesel; biofuels; bioenergy; renewable energy

14.1 Introduction

Concern about the limited availability of fossil fuels and their impact on climate change is driving research into alternative energy sources. Replacing transportation fuels is particularly challenging because of the need for an energy dense fuel that will provide equivalent or better exhaust emissions and not contribute to corrosion or deposit formation in engines. Lignocellulosic biomass is seen as a plentiful and low cost feedstock for alternative transportation fuels (US DOE, 2011). However, starch, sugar and oil crops and animal fats are currently the most widely used feedstocks for alternative transportation fuels because the conversion processes are much simpler and more cost-effective.

Ethanol production from corn and other cereal grains is approaching 15 billion gallons per year, or about 10% of the current consumption of gasoline (EIA, 2012). Production of biomass-based diesel fuel in the US surpassed 1 billion gallons in 2011, which was about 3% of the on-highway diesel fuel consumption at the time (US EPA, 2012a). This biomass-based diesel fuel was composed primarily of fatty acid methyl esters (FAME), or biodiesel, along with a small quantity of non-ester-based fuel, which will be referred to here as renewable diesel fuel. Processes to produce diesel fuel from lignocellulosic biomass have been proposed (e.g., Huber et al., 2005), but there is no commercial production at this time. This chapter describes the characteristics, production, feedstock sources and current status of diesel fuels produced from vegetable oils and animal fats.

14.2 Overview of biodiesel and renewable diesel

Biodiesel is defined to be the monoalkyl esters of fatty acids from vegetable oils and animal fats (ASTM D 6751). It is produced via transesterification of the triacylglycerides in the oils and fats. The same triacylglyceride feedstock used for biodiesel can be hydrogenated to a hydrocarbon product called renewable diesel (Furimsky, 2000). Hydrogenation can also be used to upgrade a wide variety of different low-grade oils produced by thermochemical processes such as pyrolysis and liquefaction. While the paraffinic products of these processes bear some similarity to the hydrogenated products of triacylglycerides, they are not currently commercially viable fuels and are therefore not included in this discussion.

Carefully defining the meanings of the terms used for various fuels is important as the feedstock and processing options proliferate. This has led to the development of detailed fuel specifications that describe the performance properties of the fuel along with some limited description of its composition. For example, ASTM D 6751, the biodiesel standard used in the US (ASTM, 2011), is the source of the definition statement given above. The standard used in Europe, EN 14214, uses a similar definition. These standards do not include non-ester products such as the paraffinic product produced by hydrogenating the oils and fats. The wide range of possible feedstocks and processes that have been proposed for the production of alternative diesel fuels presents a major challenge to standard development, although a specification for non-petroleum jet fuel (ASTM D 7566) has been approved.

14.3 Renewable diesel production routes

The basic process to produce renewable diesel starts with hydrogenation which saturates the double bonds and removes the oxygen, either as H2O or CO2 depending on the availability of hydrogen, from the fatty acid chains of the triacylglyceride. Hydrogenation and decarboxylation are two of the basic reactions that occur during the production of renewable diesel and are shown in Fig. 14.1.

image
14.1 Decarboxylation and hydrogenation reactions for triolein.

In this case, a representative triacylglyceride molecule, triolein, is converted to the n-paraffin molecules heptadecane or octadecane by separate pathways that depend on the availability of hydrogen. Hydrogenation preserves more of the original carbon in the fuel but requires 2.5 times more hydrogen than decarboxylation. In either case, propane is produced as a byproduct, which can be reformed and used as a hydrogen source or sold directly as fuel. The ideal liquid fuel mass yields for the reactions are 0.815 kg C17H36/kg triolein and 0.862 kg C18H38/kg triolein for decarboxylation and hydrogenation, respectively. Most of the mass loss is as CO2, water, and propane but only the propane has a market value. Because the final product is typically sold on a volume basis, the volume yield is also of interest. Since the densities of n-paraffins are typically very low (0.777 kg/L for both C17H36 and C18H38), the volume yields are much higher, at 0.954 L C17H36/L triolein and 1.010 L C18H38/L triolein for decarboxylation and hydrogenation, respectively.

In the case shown in Fig. 14.1, the renewable diesel will consist of C17 and C18 n-paraffins. A real vegetable oil or animal fat will contain a diverse mixture of fatty acid chains in the triacylglycerides resulting in a mixture of n-paraffins suitable for blending with petroleum-based diesel fuel. One drawback of using real vegetable oils and animal fat is that the cloud point is usually found to be unacceptably high (Knothe, 2010). In this form, the renewable diesel is only suitable for use as a low-level blending component for No. 2, or heavier, diesel fuel. In a refinery, this paraffinic stream can be isomerized, cracked, and distilled to produce fuels in the boiling range of jet fuel or even gasoline. This is the product that is most commonly referred to as bio-jet fuel. The economics of bio-jet fuel are challenging because the price of diesel fuel and jet fuel are traditionally very similar, so there is no financial incentive for refiners to do the extra processing needed to upgrade the paraffinic renewable diesel stream to jet fuel specifications.

Hydrogenation of triacylglycerides may be done with dedicated processes or by co-feeding the triacylglycerides with petroleum-based products (Sebos et al., 2009). In the latter case, the n-paraffins produced from the triacylglycerides enhance some properties, such as the cetane number, but may also impact cold flow properties (Sebos et al., 2009). Common hydrotreatment catalysts such as NiMo or CoMo on γ-Al2O3 are used as opposed to zeolites, which tend to promote hydrocracking and lower the yield of fuel in the diesel boiling range. Although cracking and pyrolysis may provide molecules with better cold flow properties, these processes also produce substantial amounts of low-value materials with little value as transportation fuels. Therefore, isomerization is more commonly used to enhance the cold flow properties of renewable diesel (Knothe, 2010).

Although hydrogenation pathways consume more hydrogen than decarboxylation, water-gas-shift reactions with CO2 were indicated as significant sinks for hydrogen (Donnis et al., 2009) and may be more significant sources of CO in the product gases than decarbonylation reactions.

14.4 Biodiesel production routes

Biodiesel is most commonly produced via a transesterification reaction between a vegetable oil or animal fat and a simple alcohol such as methanol, as shown in Fig. 14.2 (Van Gerpen, 2005). The reaction is usually catalyzed with a strong base such as sodium methoxide although sodium or potassium hydroxides are also used.

image
14.2 Transesterification of triacylglyceride.

Another option is to use solid phase, or heterogeneous, catalysts currently under development. These will be discussed in more detail in a later section. The oil or fat is dried before the reaction so that moisture does not have a chance to enhance the saponification side reaction which consumes the catalyst and decreases yield. After drying takes place, the oil, alcohol, and catalyst are mixed and then agitated for a period of 10 minutes to an hour depending on the level and type of agitation. Simple stirred reactors are frequently used, although high shear mixers, ultrasonic mixers, and even co-solvents are used to accelerate the reaction. It is very common to conduct the reaction in two or more steps with a partial reaction allowed to occur in a first stage reactor, then removal of the glycerin that has been formed (which is accompanied by a significant fraction of the catalyst) and then adding additional methanol and catalyst for a second reaction to the final equilibrium.

Lower cost feedstocks frequently contain elevated levels of free fatty acids (FFAs). These can be removed by various techniques but can also be converted to methyl esters using an acid catalyzed pretreatment (Canakci and Van Gerpen, 2001). This pretreatment will be discussed later, but basically involves adding sulfuric acid and methanol to convert the FFAs to methyl esters so that the standard alkali-catalyzed process can be used to convert the triacylglyceride portion of the feedstock into biodiesel.

After the reaction is complete, the remaining glycerin is removed by either settling in a decanter, centrifugation, or possibly with a coalescer, although this approach is less common. Then the methanol is removed by flash evaporation and the small amounts of residual free glycerin, soaps, and methanol are removed by washing with deionized water, with an ion exchange resin, or with a solid adsorbent such as magnesium silicate. The latter two options are sometimes referred to as ‘dry washing.’ The final product may be further subjected to a cold filtration process whereby the fuel is cooled to near its freezing point, held at that temperature for sufficient time to allow the crystallization of minor impurities such as sterol glucosides and saturated monoglycerides, and then filtered. The filtration usually requires the addition of a filter aid such as diatomaceous earth and may follow a period of warming back to close to ambient temperature. Additives such as antioxidants and pour point depressants may be added before the fuel is sold. Finally, the fuel is analyzed to verify compliance with the ASTM specification and to ensure fuel quality (Van Gerpen, 2005).

The following discussion will focus in greater depth on recent developments in catalyst technology as well as ultrasonic and supercritical reactors and waterless purification techniques.

14.4.1 Heterogeneous catalysts for biodiesel production

A major concern with using a homogeneous alkali catalyst, such as sodium methylate, for biodiesel production is the large quantity of water used for washing the residual catalyst and soap from the biodiesel. The ratio of water used is typically 1:1, i.e., for every gallon of biodiesel produced, one gallon of washing water is needed. This ratio can be reduced to 10% of the volume of biodiesel produced if the soap is split by acidulation prior to or during washing (Van Gerpen, 2005). The consumption of wash water can be virtually eliminated if facilities are in place to recycle and reuse the wash water.

Another drawback of using a homogeneous base catalyst is soap formation due to the high FFA content (e.g., greater than 5 wt%) of the feedstocks, such as waste vegetable oils or microalgal oil. Soap formation in biodiesel production not only adversely affects the biodiesel yield but also makes biodiesel separation from the byproduct glycerin very difficult. When high concentrations of FFAs are present, a two-step process, i.e., a strong acid-catalyzed esterification followed by a base-catalyzed process, is needed to avoid the problem of significant soap formation (Canakci and Van Gerpen, 1999, 2001).

The advantages of using heterogeneous catalysts in biodiesel production include the elimination of water-washing or dry-washing of the post-reaction mixture for catalyst and soap removal. Other advantages include decreased production of waste water, reuse of the heterogeneous catalysts, high productivity per unit of reactor capacity, and easier scale-up for continuous-flow processes. It is also claimed that heterogeneous catalysts can catalyze both the esterification and transesterification reactions so that the two-step process is not required to deal with high FFA feedstocks (Furuta et al., 2004). A major disadvantage of using heterogeneous catalysts, however, is the high catalyst cost and higher operating cost due to the elevated temperatures usually required (Kiss et al., 2010; Sakai et al., 2009). To date, no cost-effective and highly efficient heterogeneous catalyst has been identified for practical commercial use.

Heterogeneous catalysts for biodiesel production are mainly metal hydroxides and oxides and other non-metal-based compounds (Borges and Diaz, 2012; Semwal et al., 2011; Chouhan and Sarma, 2011). The catalyst supports are typically aluminum, activated carbon, and organic resin. Table 14.1 provides an overview of the current efforts in exploring suitable heterogeneous catalysts for biodiesel production. Generally, all of the reported catalysts lack activity at low operating temperatures (65°C or lower). Even with extended periods of reaction time, the overall biodiesel yields or vegetable oil conversion rates catalyzed by heterogeneous catalysts are low (less than 98%) and are not comparable with those by homogeneous catalysts. This incomplete transesterification usually translates to poor quality biodiesel and it is likely that the fuel will not meet the standards specified by the ASTM.

Table 14.1

Examples of commonly researched heterogeneous catalysts and their effectiveness

No. Active ingredients/support Conditions of use Examples of effectiveness References
1 CaO, activated powder Molar ratio of methanol to oil 13:1, 60°C Yield: 94 wt% of oil after 90 min Granados et al., 2007
2 MgO, nano particles Molar ratio of methanol to oil 6 ∼ 36:1, 250°C and 24 MPa Yield: approx. 100 wt% of oil after 20 min with 2 ∼ 5 wt% catalyst Wang and Yang, 2007
3 CaO, activated powder Molar ratio of methanol to oil 12:1, 65°C, catalyst application 8 wt% Yield: 95 wt% after 3 h
Catalyst activity decreases with repeated uses
Liu et al., 2007
4 CaO, powder Ratio of methanol to oil 3.9:15 g, 60°C, catalyst application 0.1 g per 15 g of oil Yield: approx. 90 wt% after 3 h Kawashima et al., 2009
5 CaO, powder Ratio of methanol to oil 50:100 ml, ∼ 64°C (reflux), catalyst application 0.78 g per 100 ml of oil Yield: approx. 90 wt% after 1 h Kouzu et al., 2008
6 CaZrO3 and CaO–CeO2, powder Molar ratio of methanol to oil 6:1, 60°C, catalyst application 10 wt% of oil Yield: approx. 90 wt% after 10 h Kawashima et al., 2008
7 CaxMg2 -xO2 from MgO and Ca(NO3)2 Molar ratio of methanol to oil 12:1, catalyst application 6 wt%, approx. 65°C (reflux) Oil conversion of 91.3% achieved after 5 h Xie et al., 2012
8 Ca-Si oxides Ratio of 24 mL methanol to 1.0 g oil, catalyst application 20 wt% of oil, 65°C Yield: approx. 100% after 4 h Hsin et al., 2010
9 SrO, activated powder Molar ratio of methanol to oil 15:1–18:1, 65–70°C, catalyst application 2.5–3.0 wt% Yield: approx. 95 wt% after 30 min Catalyst activity decreases with repeated uses Liu et al., 2008
10 SrO/SiO2 Molar ratio of methanol to oil: 6:1, catalyst application 5 wt% of oil, 65°C Yield: approx. 80 wt% Catalyst activity decreases with repeated uses Chen et al., 2012
11 K2CO3 supported on MgO Ratio of methanol to oil 1.12:5 g, catalyst 50 mg/5 g of oil, 70°C Yield: approx. 99 wt% after 2 h Liang et al., 2009
12 KAlSiO4 (kalsilite) Ratio of methanol to oil: 150:300 g, catalyst application 5 wt% of oil, b120–180°C Yield: approx. 100% Wen et al., 2010
13 KI/mesoporous silica Molar ratio of methanol to oil 16:1, 70°C, a catalyst application 5.0 wt% of oil Approx. 90% of conversion achieved after 8 h Samart et al., 2009
14 Tetramethylguanidine on silica gel Ratio of methanol to oil 1.5:10.0 g, catalyst application 0.7 g of 10 g oil, 80°C Yield: approx. 86% after 3 h
Catalyst activity decreases with repeated uses
Faria et al., 2008
15 Sr(NO3)2 on ZnO Molar ratio of methanol to oil 12:1, approx. 65°C (reflux), catalyst application 5 wt% Conversion: 94.7% after 4 h Yang and Xie, 2007
16 LiNO3/Al2O3, NaNO3/Al2O3, and KNO3/Al2O3 Molar ratio of methanol to oil 65:1, catalyst application 10 wt% or more at 60°C Methyl esters content: approx. 94% achieved after 3 h Benjapornkulaphong et al., 2009
17 S-ZrO2 (sulfated zirconia) Molar ratio of methanol to oil 20:1, catalyst application 5 wt% at 120°C Yield: approx. 99% after 2 h Garcia et al., 2008
18 Al-MCM-41 mesoporous molecular sieves Molar ratio of methanol to oil 60:1, catalyst application 0.6 wt% at 130°C Conversion: approx. 79% after 2 h Carmo et al., 2009
19 La/zeolite from La(NO3)3 Molar ratio of methanol to oil 14.5:1, catalyst application 1.1 wt% at 60°C Conversion: approx. 49% after 4 h Shu et al., 2007
20 Na2MoO4 (sodium molybdate) Molar ratio of methanol to oil 54:1, catalyst application 5 wt% at 60°C Yield: approx. 96% after 3 h Nakagaki et al., 2008
21 SO42 −/TiO2–SiO2a Molar ratio of methanol to oil 9:1 to 12:1, catalyst application 3 wt% at 200–220°C Yield: approx. 95% after 3–4 h Peng et al., 2008
22 SO42 −/SnO2, powder, and SO42 −/ZrO2, powder Molar ratio of methanol to oil 6:1, catalyst application 0.5–3 wt%, 250°C and 24 MPa Yield: 90.3 wt% after 4 h Jitputti et al., 2006
23 SO42 −/ZrO2 Molar ratio of methanol to oil: 9:1, catalyst application 3.5 wt% of oil, 120°C Yield: approx. 98 wt% after 1 h Niu et al., 2012
24 Tungstated zirconia (W-Zr) Molar ratio of methanol to oil 6:1, catalyst application 2 wt% at 60°C Yield: approx. 95% after 3–4 h  
25 KAc (potassium acetate)-NaX zeolite interchanged Molar ratio of methanol to oil: 12:1 to 72:1, catalyst application 3–6 wt% of oil, 100–155°C, and 6–88 bar pressure Yield: approx. 96% at 48:1 methanol to oil ration, 6 wt% catalyst, 155°C for 3 h Borges et al., 2011
26 Mg-Al hydrotalcites Molar ratio of methanol to oil: 9:1 to 12:1, catalyst application 2.5 wt% of oil, 60 ∼ 65°C Yield: approx. 97% after 4 h Gomes et al., 2011
27 Na- and NH4-quntinites, bi-functional Molar ratio of methanol to oil: 15:1, catalyst application 10 wt% of oil, 75°C Yield: approx. 98% after 2 h Kondamudi et al., 2011
28 Nd2O3-K (neodymium oxide with potassium hydroxide) Molar ratio of methanol to oil: 14:1, catalyst application 6 wt% of oil, 60°C Yield: approx. 92% after 1.5 h Li et al., 2011
29 WO3/ZrO2 Molar ratio of methanol to oil: 9:1, catalyst application 10–20 wt% of oil, 75–200°C Yield: approx. 93–98% Park et al., 2010
30 Li2SiO3 Molar ratio of methanol to oil: 12:1 to 36:1, catalyst application 10 wt% of oil, 60°C Yield: approx. 92–96% after 3 h Wang et al., 2011

Image

aIt is claimed that a 10,000 tonnes/year biodiesel production demonstration plant using this catalyst has been built.

There are various methods to increase the catalyst activity, one of which is to raise the operating temperature to 120°C, or even to 200°C and higher, so that the metal catalysts start to show their advantageous catalytic activity (Jitputti et al., 2006; Garcia et al., 2008; Carmo et al., 2009: Peng et al., 2008; Park et al., 2010). However, this will obviously increase operating costs. Other cost considerations must be made when using heterogeneous catalysts in biodiesel production. This process requires a much higher methanol-to-vegetable oil molar ratio, as high as 60:1, in order to achieve a reasonable reactivity (Benjapornkulaphong et al., 2009; Garcia et al., 2008; Carmo et al., 2009; Nakagaki et al., 2008). Such a high methanol-to-oil molar ratio leads to additional effort in methanol recovery, re-purification, and reuse, and thus adds to the operating cost (Sakai et al., 2009; Kiss et al., 2010). Some catalyst producers recommend that the fuel be vacuum distilled to remove the products of incomplete reaction, which is an indication of low catalyst activity.

Most of the heterogeneous catalysts described in the literature were tested in a powder form or a slurry suspended in a vegetable oil and methanol mixture with high mixing intensity (up to 1500 rpm) in a batch mode. Keeping these catalysts suspended in large-scale, continuous-flow reactors would be a challenge. Pelletized or other shaped heterogeneous catalysts are typically not developed unless a catalyst has proved to be effective and worth the effort. At that stage, some engineering challenges must be overcome, such as the strength and pore sizes of the structure and ensuring adequate capability for mass transfer. Catalyst poisoning is also a serious consideration for commercialization. Another problematic phenomenon is the reusability of the heterogeneous catalysts. Some catalysts appear to lose their reactivity quickly due to leaching after being used for only a few batches (Liu et al., 2007, 2008; Faria et al., 2008; Di Serio et al., 2010; Chen et al., 2012).

14.4.2 Ultrasonic processing

Application of ultrasonication in chemical reaction systems has been extensively researched due to its effective production of radicals for initiating chemical reactions and its localized micro-scale cavitation and mixing (Shol, 1988; Flint and Suslick, 1991). The study of chemical changes induced by ultrasonication is now recognized as the science of sonochemistry and the chemical reactors that incorporate ultrasound are referred to as sonochemical reactors (Mason, 2000; Thompson and Doraiswamy, 1999).

The insolubility of methanol in vegetable oils is a limiting factor for the transesterification of triacylglycerides for biodiesel production (Van Gerpen, 2005). To overcome this problem, mechanical mixing is typically applied in conventional processes to improve the reaction rate. Recognizing its effectiveness in creating micro-scale cavitation and intensified local mixing, researchers have attempted to apply ultrasonication in transesterification of vegetable oils and animal fats for biodiesel production (e.g., Ji et al., 2006; Wu et al., 2007; Hanh et al., 2009).

The level of positive effect of ultrasonication on transesterification varies largely among reports. One consistent advantage is the shorter reaction time, which ranges from one minute (Teixeira et al., 2009) to a few minutes (Stavarache et al., 2005; Singh et al., 2007; Kumar et al., 2010a). Although most work has been done with conventional homogeneous catalysts, similar phenomena were observed in systems where heterogeneous catalysts were used (Kumar et al., 2010b; Yu et al., 2010).

Currently, most reports published on ultrasound-assisted transesterification are laboratory evaluations. The exact mechanism behind the enhancement by ultrasonication is not yet clearly understood. Based on experience with other ultrasound-assisted chemical systems, the ultrasonic cavitation combined with the radial motion of the ultrasonic cavitation bubbles are proposed to be the reason for the accelerated reaction rates (Mootabadi et al., 2010). The localized high intensity energy generates microbubbles that create increased interfacial areas to overcome the mass transfer limit. Such high energy density may also overcome the activation energy to initiate the transesterification (Singh et al., 2007). However, such a phenomenon is significant only in the close proximity to the ultrasound transducer (Monnier et al., 1999a, 1999b; Cintas et al., 2010). Another important finding under specific experimental conditions is that chemical radicals are not the reason for the enhanced transesterification under ultrasonication as claimed in other sonochemical systems (Kalva et al., 2009). This may be due to the fact that the moderate intensity of ultrasonication used for biodiesel production does not generate enough energy to produce radical formation.

Some use of ultrasound-assisted systems in smaller biodiesel production plants has been reported but no technical details are provided. According to the nature of ultrasonication and the experience obtained from other sonochemical systems, chemically effective and economically viable sonochemical reactors for biodiesel production still require additional research and development, especially regarding the engineering aspects of the sonochemical reactors (Thompson and Doraiswamy, 1999). The widespread commercial application of ultrasonication biodiesel production may take another decade. A robust yet highly productive transesterification system may be created through a combination of sonochemical reactors with heterogeneous catalysts for biodiesel production, which may also be more suitable for small producers (He and Van Gerpen, 2012a).

14.4.3 Supercritical processing

Conventional processes for biodiesel production use homogeneous catalysts, such as sodium methoxide or hydroxide which require water washing or ‘dry washing’ with absorbent materials to remove soaps from the crude biodiesel. This washing step requires extra equipment and consumes the homogeneous catalyst, which increases the overall operating cost. One non-catalyzed approach for biodiesel production is the processing of vegetable oils in supercritical methanol. Saka and his team first reported non-catalytic supercritical biodiesel preparation starting in the late 1990s and have published extensively in this field (e.g., Saka and Kusdiana, 1999, 2001a, 2001b; Kusdiana and Saka, 2001a, 2004a; Warabi et al., 2004a, 2004b; Saka et al., 2006).

Supercritical processing of vegetable oils can achieve the necessary transesterification without the need for a catalyst. Once in the supercritical stage, methanol has a much stronger solvent effect that dissolves the vegetable oils and thus overcomes the insolubility between the vegetable oils and methanol, allowing the reaction to proceed quickly in a homogeneous phase at a high reaction temperature. The enhanced solvent capability of supercritical methanol is attributed to its much weakened hydrogen bonds and highly reduced polarity (Yamaguchi et al., 2000). Most reported research uses supercritical methanol as the solvent. The operating conditions are close to or above the supercritical properties of methanol (i.e., 240°C and 8.1 MPa) and in the range of 270–430°C. Although a higher temperature provides a more effective supercritical fluid, considering the thermal stability of the unsaturated fatty acids and the increased operating cost due to high temperature and high pressures, a temperature range of 270–300°C is recommended (He et al., 2007b; Imahara et al., 2008).

In addition to the elimination of the catalyst, another advantage of supercritical methanol processing is its tolerance of a high FFA concentration in the feedstock. This is due to the fact that FFA can be esterified directly to methyl esters (Kusdiana and Saka, 2001b).

Generally water in the vegetable oil adversely affects the process (e.g., soap formation which requires extra catalyst) and the biodiesel quality (e.g., hydrolyte formation) in the conventional alkali-catalyzed transesterification process (He and Van Gerpen, 2012b). However, in supercritical methanol processing, the presence of water was found to have no negative effect on biodiesel (i.e., methyl esters) production (Kusdiana and Saka, 2004a), because hydrolysis of triacylglycerides caused by the presence of water, if it occurs, leads to the formation of free fatty acids, which are then esterified into methyl esters. Therefore, a small amount of water in the vegetable oil or animal fat is not a major concern in supercritical biodiesel production.

Short chain, primary alcohols are the logical choice for transesterifying vegetable oils and animal fats (Warabi et al., 2004a). However, other solvents can also be used in supercritical processing for biodiesel production, such as dimethyl carbonate (Ilham and Saka, 2010), methyl acetate and other carboxylate esters (Saka and Isayama, 2009; Goembira et al., 2012; Niza et al., 2012).

Recognizing the potential advantage of triacylglyceride hydrolysis in supercritical methanol to free fatty acids and subsequent esterification to methyl esters, a two-step process has also been studied by researchers. In this two-step process, vegetable oil was hydrolyzed in sub-critical water for 20 min at 270°C into free fatty acids and glycerin byproduct. After the mixture of glycerin and excess water was separated out, the free fatty acids were fed into a supercritical methanol reactor and esterified into methyl esters at a comparable operating temperature and for a comparable reaction time (Kusdiana and Saka, 2004b; Minami and Saka, 2006). Using acetic acid to replace water for triacylglyceride hydrolysis under subcritical conditions was also studied by the same group and claimed to have better effectiveness (Saka et al., 2010).

14.4.4 Purification by adsorbents and resins

Magnesium silicate is widely used to remove free fatty acids and other polar compounds from used cooking oils to extend their life. It can also be used to purify biodiesel by adsorbing free glycerin, soaps, methanol as well as monoglycerides and sterol glucosides. Approximately 1% magnesium silicate powder is added to the biodiesel at 60–65°C. However, the exact treatment level depends on the amount of contaminants to be removed. The mixture is agitated for 20 minutes and then the adsorbent is removed by filtration. Typically, the residual methanol is removed from the crude biodiesel before the adsorbent is added so the active sites on the adsorbent particles are not overwhelmed by the alcohol (Berrios and Skelton, 2008).

An alternative process uses ion exchange resins to remove free glycerin and soaps. The resins usually consist of small (∼0.5 mm) styrene beads with treated surfaces. The beads are placed in fixed beds and the biodiesel is pumped through the beads. When soap molecules in the biodiesel contact the resin beads, hydrogen ions from the beads are exchanged with the sodium (or potassium) ions from the soap (Wall et al., 2011). Thus, removal of the soap causes an increase in the acid value of the biodiesel. Soap levels above about 2500 ppm may cause the resulting fuel to exceed the ASTM specified acid value. Glycerin is removed by adsorption to the resin bead surface, a connection that can be overcome by washing the beads with methanol, this allows the resin beads to regenerate. It is also claimed that magnesium silicate can also be utilized in fixed beds with regeneration by methanol wash.

14.5 Traditional and emerging feedstocks

14.5.1 Traditional feedstocks

Currently, soybean oil is the dominant feedstock in the US while canola and corn oils, animal fats, and yellow greases, are the supplemental feedstocks. Soybean oil is a co-product of the soybean meal industry. Conventional soybean seeds contain about 18–20% oil. Despite the low oil content, soybean oil produced in the US is still available in large quantities due to the vast amount of soybean production. For example, 2.97 billion bushels or approximately 80 million metric tons of soybeans were produced in 2012 (USDA, 2012a), which is approximately equivalent to 17.6 × 106 m3 (or approx. 4.6 billion gallons) of oil. In 2012, 1.12 million metric tons (2.48 billion pounds) of canola seed was produced, which is equivalent to approximately 146 million gallons of oil (USDA, 2012b). However, most of the traditional feedstocks have established markets and cannot easily migrate to the biodiesel industry.

The ‘double-zero’ rapeseed or canola oil is the major feedstock for biodiesel in Europe. Rapeseed adapts well to a wide variety of climate and soil conditions and is cultivated widely around the world. The Canadian modified cultivar of rapeseed or canola is very low in the undesirable erucic acid and glucosinolates, and is now the main cultivar planted for its oil in Europe and North America. Total biodiesel production by the EU-27 was 9.57 million tons (approx. 2.31 billion gallons) in 2010 (European Biodiesel Board, 2012). Although there were imported feedstocks such as palm oils, this production was mainly from rapeseed/canola oils.

It is predicted that rapeseed/canola production in 2012 will be 20 million metric tons in the EU-27 and 14.5–15.5 million metric tons in Canada (Mielke, 2012). If 45% oil content (CGC, 2012) is assumed, the canola oil produced by the EU-27 and Canada in 2012 will be approximately 9.0 million tons (or 2.6 billion gallons) and 6.75 million tons (or 1.95 billion gallons), respectively. It is expected that the biodiesel production by the EU-27 will increase in 2012. Therefore, in order to keep growing, biodiesel producers in Europe must explore other feedstock opportunities while competing with the food market for canola oil.

Vegetable oils are composed mainly of triacylglycerides, which are the glycerin esters of different fatty acids. When processed for food use, these oils contain very low free fatty acids, and are very low in sulfur and other heterogeneous chemicals. This makes them high quality feedstocks for biodiesel production, especially when using homogeneous alkali-catalyzed conversion processes. When using food-grade vegetable oil as the feedstock, soap formation is typically not a concern in biodiesel production.

Different oils and fats are characterized by the fatty acid chains that are present in the triacylglycerides. The nomenclature for identifying fatty acids is commonly CX:Y, where X is the number of carbon atoms in the fatty acid chain and Y is the number of double bonds. The fatty acid profiles for common vegetable oils and animal fats are shown in Table 14.2.

Table 14.2

Fatty acid compositions of common seed oils and animal fats

Oils/fats Fatty acid profiles
C12:0 C14:0 C16:0 C18:0 C18:1 C18:2 C18:3 C20:0 C22:1
Coconut 45–53 16–21 7–10 2–4 5–10 1–2.5    
Corn  1–2 8–16 1–3 20–45 34–65 1–2   
Cottonseed  0–2 20–25 1–2 23–35 40–50    
Palm  0.5–2 39–48 3–6 36–44 9–12    
Rapeseed, low erucic or canola   1–3 2–3 50–60 15–25 8–12   
Rapeseed, high erucic   1–3 0–1 10–15 12–15 8–12 7–10 45–60
Soybean, high linoleic   6–10 2–5 20–30 50–60 5–11   
Soybean, high oleic   2–3 2–3 80–85 3–4 3–5   
Lard  1–2 25–30 10–20 40–50 6–12 0–1   
Tallow  3–6 22–32 10–25 35–45 1–3    

Image

Chemically, animal fats are composed of the same compounds as those in seed oils but with fatty acid profiles that contain larger fractions of saturated fatty acids (C16:0 and C18:0). Animal fats are the byproducts of the livestock and poultry industries, and are relatively inexpensive feedstocks. Although large quantities of animal fats are produced annually, these fats have established markets as food and food additives (such as butter, lard, and shortenings), animal feed additives, and industrial products such as fatty acids, soap, paints, and lubricants.

Greases are generally used vegetable oils and/or animal fats. Yellow grease, or waste vegetable oil (WVO), consists mostly of used cooking oils, and brown grease is typically recovered from grease traps of restaurants and food processing plants. Greases usually contain large amounts of free fatty acids and other polymerized and/or hydrolyzed compounds after repeated uses at elevated temperatures. Due to the manner of its collection, greases are commonly high in water content which complicates the conversion process. Brown grease is even more problematic due to its contamination by cleaning agents. Therefore, greases are low quality feedstocks for biodiesel production, and special processing and/or pretreatment are required to avoid problems caused by impurities in order to have a quality biodiesel product.

14.5.2 Emerging feedstocks

Emerging feedstocks, including camelina, jatropha, pennycress, and microalgae, have also been researched and utilized for biodiesel production. Camelina, a member of the Brassica family, has many species but the most important for biodiesel is Camelina sativa (Vollmann et al., 1996). It has recently attracted the attention of farmers in the northern US (McVay and Lamb, 2008). Camelina adapts well to marginal lands having dry and cool climate conditions with a short growing season that, with adequate rainfall, allows for double cropping. The yield of camelina in North America is about 890–1,350 kg/ha (800–1,200 lb/acre) and the oil content of camelina seeds varies from 30 to 40% (McVay and Lamb, 2008). Camelina oil consists of up to 88% unsaturated fatty acids (Putnam et al., 1993), which contribute to the desirable cold flow properties of the biodiesel produced from it. However, camelina seeds are very small (only up to 2 g per 1,000 seeds) and the plant is very strong when ripe, which makes harvesting camelina very difficult.

Jatropha is a perennial shrub that grows in various harsh climates and poor soil conditions. The jatropha plant and its fruits and seeds contain poisonous toxins so it is considered to be a non-food crop (Dias et al., 2012). Jatropha seeds contain up to 35% non-edible oil (Kumar and Sharma, 2008) and its oil yield can be as high as 1,900 kg/ha (1,690 lb/acre), making it a good candidate for biodiesel production. Jatropha fruits ripen year round and harvesting is very labor-intensive. Unless automated harvesting machines are invented, jatropha fruit harvesting is unlikely to be cost-effective. Jatropha is being actively explored by countries that have limited agricultural land but are still interested in seeking oil sources for biofuels such as India, China, and other developing countries (Kumar et al., 2012; Yang et al., 2012; Mofijur et al., 2012). However, the suitability of jatropha as a large-scale domesticated crop has been questioned and discussions are underway to find sustainable ways to produce it (Contran et al., 2013; Kumar et al., 2012).

Pennycress is a collective name applied to a group of species belonging to the Brassicaceae family. Field pennycress, Thlaspi arvense, is a variety that has been identified as having industrial potential, especially for use as a feedstock for biodiesel. The seeds of pennycress contain up to 36% oil and are high in erucic and linoleic acids (Moser et al., 2009b). Pennycress seeds are small. Seeds harvested in Peoria, Illinois, have average dimensions of 1.87 × 1.35 × 0.74 mm and 1000 seeds weighed only 0.97 g at 9.5% moisture (Evangelista et al., 2012). However, the seed yield can approach 1,420 kg/ha (1,265 pounds/acre). Field pennycress is basically considered to be an annual weed, growing in open spaces and roadsides (Mitich, 1996). This species is tough and its seeds can have a long life, lasting approximately 20–30 years (Koundinya and Hansen, 2012). Evaluation of biodiesel produced from pennycress oil has reported desirable properties due to the high contents of erucic (32.8 wt%) and linoleic (22.4 wt%) acids. The biodiesel produced from pennycress oil has a cetane number close to 60 and exhibits favorable properties at low temperatures (Moser et al., 2009a). To be established as a feasible feedstock for biodiesel production, more studies are needed on pennycress to establish its agronomic, harvesting, and processing requirements as well as its environmental and social impacts.

One of the most promising feedstocks for biodiesel production is believed to be microalgae oil (Wu et al., 2012; Ahmad et al., 2011; Stephens et al., 2010; Chisti, 2008). As a potential source for biofuel production, microalgae have been exploited and researched extensively (e.g., Sheehan et al., 1998). Microalgae can be cultivated in open ponds or in bioreactors (Chen et al., 2011). If strains of microalgae contain 30% lipids, the productivity of algal oil can be as high as 58,700 L/ha (6,270 gal/acre); as a comparison to traditional oil seeds, canola productivity is only 1,190 L/ha (127 gal/acre) (Chisti, 2007). Practical production of microalgal lipids for biodiesel production still faces many challenges. Among them, cost effectiveness is the most critical (Stephens et al., 2010; Harun et al., 2011). Difficulties exist not only in producing oil efficiently from high lipid strains but also in cost-effective cultivation and harvesting systems. Of particular significance is the high energy requirement for the de-watering and drying processes. Currently, research and development on producing biodiesel from microalgal oils is very active, and operations at pilot and demonstration scales are reported, but no reports are available yet on commercial microalgal biodiesel production. The fatty acid profiles of these emerging plant oils are provided in Table 14.3.

Table 14.3

Fatty acid compositions of oils from emerging feedstocks

Oils Fatty acid profiles
C14:0 C16:0 C18:0 C18:1 C18:2 C18:3 C20:0 C20:1 C20:5 C22:1
Algal oil 12–15 10–20   4–19  1–2  5–8   35–48  
Camelina   7 ~ 8 2 ~ 3 12–16 15–25 30–40  12–15  2–3
Jatropha  11–16 6–15 34–45 30–50  3–5    
Pennycressa 0.1 3.1 0.5 12.6 22.4 11.8 0.3 8.6  32.8

Image

aMoser et al., 2009a.

14.6 Feedstock quality issues

14.6.1 Feedstocks with high content of free fatty acids

Most biodiesel is currently produced using high quality soybean oil catalyzed by homogeneous alkali catalysts. However, other feedstocks such as waste vegetable oils and/or microalgal lipids can be high in FFAs requiring a different approach. The FFA content ranges between 2 and 7% for used vegetable oils, 5 and 30% for animals fats, and 0 to 100% for some low quality trap grease (Van Gerpen et al., 2004). If a feedstock contains high levels of FFA, typically 5% or higher, alkaline catalysts are not suitable for use in triacylglyceride transesterification. This is because the alkaline catalysts, such as sodium methoxide, will react with the FFA directly to produce fatty acid salts or ‘soaps’. Soap formation not only consumes the catalysts but also causes the reacting mixture to emulsify. This can lead to difficulties in processing and to a low quality biodiesel product. Early research by Freedman et al. (1984) demonstrated the significant FFA effect on the transesterification process catalyzed by alkaline catalysts. At the same operating conditions, the product yield of methyl esters was reduced from 93–98% to 67–86% due to the presence of a significant amount (approx. 6.7%) of free fatty acids in the oil. Soap generally partitions into the byproduct glycerin and is considered waste if not recovered with additional processing. Generally speaking, if a feedstock contains no more than 1–2% FFA in alkaline-catalyzed transesterification, the FFA can be ignored; if 2–5% FFA, additional alkaline catalyst is needed to account for the consumption of catalyst due to soap formation and special attention needs to be paid to the procedures. An acid-catalyzed pretreatment, or distillation to remove the FFAs, is required if the oil or fat contains more than 5% FFAs.

It has been shown that strong acid-catalyzed transesterification is too slow and requires much higher alcohol levels; however, acid catalysts are very effective in catalyzing the esterification of FFA to methyl esters (Canakci and Van Gerpen, 1999). Therefore, a two-step process was proposed and is now considered to be the standard process for converting high FFA feedstocks, e.g., waste vegetable oils, into biodiesel (Canakci and Van Gerpen, 2001; Van Gerpen et al., 2004). In this two-step process, a strong acid is used to first convert the FFA into alkyl esters through an esterification process, and then the triacylglycerides are converted to alkyl esters via alkaline-catalyzed transesterification. The byproduct of the FFA esterification reaction is water. It has to be removed before the mixture proceeds to the transesterification process, which follows the same steps as the transesterification of high quality oil described previously.

14.6.2 Impurities that affect product quality

Moisture

Moisture in biodiesel is problematic, especially during long-term storage. High moisture in biodiesel may cause the fuel to deteriorate. Hydrolytic, oxidative, and other chemical reactions may occur, especially when minerals are present, as may be the case when using metal storage tanks (Waynick, 2005). Water promotes hydrolysis of the methyl esters, which leads to an increase in acidity; high acidity accelerates the decomposition of biodiesel when oxygen is available. This destructive loop effect will eventually lead to rancidity of the fuel. Another negative consequence of high moisture content in biodiesel is the potential for microbial growth. Some airborne microorganisms can grow on biodiesel and produce biomass. This biomass will quickly plug fuel filters and cause operational problems (Zhang et al., 2011).

Biodiesel contains oxygen in the form of carboxyl groups and tends to be polar, and thus will be hygroscopic. Thorough investigation into biodiesel moisture absorption and retention has shown that biodiesel can hold 1,000–1,700 ppm of moisture at 4–35°C. This is much higher than that in petroleum diesel, in which moisture retention of 40–114 ppm was observed within the same temperature range (He et al., 2007a).

The high moisture content in biodiesel can be attributed to multiple causes. High moisture is expected if water washing is used to purify the fuel. Despite careful drying after the water wash, moisture can still exist in biodiesel due to its hygroscopic nature. Vacuum drying can reduce the moisture in biodiesel to a level of 200–300 ppm. Biodiesel can also absorb moisture from the air during long-term storage without a nitrogen blanket. Therefore, adequate drying, if a water wash is used during processing, and careful sealing of the biodiesel storage containers, including the use of nitrogen blanketing, are necessary to prevent high moisture content in the biodiesel.

Sterol glucosides

Sterol glucosides are compounds of plant sterols, also known as phytosterols, which are present in all plants as important structural components to stabilize the phospholipid bilayers in plant membranes (Piironen et al., 2000). The major plant sterols in higher plants are β-sitosterol and its glycoside β-sitosterolin. Acylated sterol glucosides are the major form and present in all parts of vegetables including fruit, tuber, root, stem, leaf, and cereals (Sugawara and Miyazawa, 1999).

Due to the small quantity in vegetable oils, measurement of plant sterols is difficult. Biodiesel producers have observed that insoluble particles are seen in biodiesel after the fuel has been stored for a few days. According to Lee et al. (2007), these fine insoluble particles consist of the non-acylated form of sterol glucosides. These particles then act as nuclei where other impurities, such as saturated monoacylglycerides, crystallize and consequently plug the fuel filter. The non-acylated form appears to be produced from the acylated form, which is soluble, during transesterification.

To resolve this problem, many producers perform a large version of ‘cold soak filtration’, one of the standard tests in the ASTM Standard D6751 (ASTM, 2011), during bulk production of the fuel. The biodiesel is cooled down to 3–5°C overnight to allow sterol glucosides to precipitate and be removed by filtration. This process is simple and effective, but is very costly. There are some adsorbents that suppliers claim will remove sterol glucosides but third-party validation is not available.

Phosphorus

High levels of phosphorous in biodiesel will cause several negative consequences. It will poison the vehicle’s catalytic converter and decrease its efficiency (NREL, 2009). This will negatively affect the exhaust emissions. Therefore, phosphorus is strictly regulated. The biodiesel specification for maximum phosphorous content is 0.001% or 10 ppm by both ASTM 6751 and EN 14214.

Crude vegetable oils always contain small amounts of phospholipids, commonly referred to as gums. Degumming is typically performed to remove the impurity before an oil is used for biodiesel production. Acidic solutions of phosphoric or sulfuric acid are used to hydrate the gums to an insoluble form that can be separated from the oil. However, residual phosphorous may still be present in degummed oils at the level of a few hundreds of ppm.

High levels of phosphorous in the biodiesel are mainly caused by inadequate pretreatment of the feedstock. A study by Van Gerpen and Dvorak (2002) showed that phosphorous has considerable effect on biodiesel yield. If the oil contains 50 ppm or more of phosphorus, the biodiesel yield was reduced by 3–5%. The phosphorous is removed in soap form, which stays at the interface of the biodiesel and crude glycerin layers and thus makes the product separation difficult. The phosphorous is not carried into the biodiesel during processing, which is good news to biodiesel producers. This advantageous effect is attributed to the miscibility of the saponified phospholipids with polar compounds like methanol and glycerin. Reduction of methanol in the biodiesel layer helps bring the soapy materials into the interface or glycerin layer (Mendow et al., 2011). Due to the quality of feedstock used, i.e., mainly soybean oil, and rigorous process control, biodiesel produced in the US generally has phosphorus levels of about 1 ppm or less (NREL, 2009).

14.7 Advantages and limitations of biodiesel

Biodiesel is made from renewable plant and animal sources such as soybean oil, animal fats, and waste vegetable oils. Biodiesel burns in diesel engines with the same efficiency as petroleum-based diesel fuel. Its high flash point makes it safer to use and store. Biodiesel generally has a higher cetane number than petroleum diesel fuel, in the range of 45–55 for soybean biodiesel and 49–62 for rapeseed biodiesel (Mittelbach and Remschmidt, 2005). Environmentally, biodiesel is non-toxic and biodegradable. It contains very low levels of sulfur and nitrogen, and emits much fewer pollutants, particularly smoke, than petroleum diesel. It can be blended with petroleum diesel in any ratio, utilize existing distribution infrastructure, and requires no engine modifications. Biodiesel possesses excellent lubricity. Even low level blending of biodiesel, e.g., 2%, with petroleum diesel fuel will improve the lubricity of the blend to a satisfactory level and thus extend the engine’s life (Van Gerpen et al., 2006).

However, biodiesel also has limitations. Biodiesel contains less energy than petroleum diesel, approximately 8% less per unit volume or 12% less per unit mass. Biodiesel generally has higher cloud and pour points, which make it less favorable for use in low temperature environments. Biodiesel has good biodegradability but it is also less stable chemically, thermally, and oxidatively, compared with petroleum diesel (Jain and Sharma, 2011). Biodiesel is a stronger solvent than petroleum diesel. It may not be compatible with some fuel lines and/or gasket materials in engines and it may cause degradation or other compatibility issues. Biodiesel may also be incompatible with some metals and cause corrosion problems (Diaz-Ballote et al., 2009; Hu et al., 2012; Singh et al., 2012). Another drawback of biodiesel is the possibility of elevated nitrogen oxide emissions, however this is controversial. The discussions in the following two sections focus mainly on biodiesel’s feedstock availability, cold flow properties, and oxidative stability.

14.7.1 Feedstock availability

Despite its advantages, biodiesel production is constrained by the availability of feedstock. Biodiesel production in the US was approximately 1.05 billion gallons in 2011 (US EPA, 2012a), and approximately 830 million gallons in the first nine months of 2012 (US EPA, 2012b). The National Biodiesel Board expects the total biodiesel production to be up to 2 billion gallons (or 7.3 × 106 m3) in the next few years (Jobe, 2012). The estimate requires a considerable increase in feedstock supply, and most likely will require major new sources beyond soybean oil.

As the world’s largest biodiesel producers, the EU-27 and North America are facing a challenge of limited feedstock supply. Biodiesel production in the US will grow as the Renewable Fuel Standard mandates expand and world demand for diesel fuel increases. Biodiesel production in the US will be limited to approximately 2 billion gallons/year because of feedstock availability, as mentioned above. Before economically viable microalgal and jatropha oils are available for biodiesel production, biodiesel producers will find no easy solution to the issue of feedstock shortage. Feedstock suppliers will balance their profitability between selling to existing markets, especially the food market, and to the biodiesel industry.

Another limitation preventing the biodiesel industry from expanding is the cost relative to petroleum diesel fuel. It is generally agreed that biodiesel is more expensive than petroleum diesel in the market, and the major cause of this is the feedstock cost. An estimate for a plant of 17,400 tons/year (5 million gallons/year) has revealed that oil is the dominant contributor, at 80.6%, of the total production cost (Van Gerpen et al., 2006). This observation is consistent with an early study by Haas et al. (2006) on a 37,850 m3/year (or 10 million gallon/year) plant, in which the feedstock (soybean oil) cost was 88% of the total cost. Although the byproduct, glycerin, can be recovered and marketed to offset some of the cost, the dominance of the oil cost will not change much.

Varying production costs are expected when different business models are adopted; however, the production cost is still dependent on the cost of the feedstock. It is logical, therefore, to use low cost feedstocks, such as animal fats and waste vegetable oils or greases. In today’s market, low cost feedstock usually means low quality feedstock. Additional cost is needed to upgrade and handle the low quality feedstocks.

14.7.2 Cold flow properties and oxidative stability

Biodiesel is prone to the issue of poor low temperature operability due to its content of long-chain, saturated fatty acids. The cold flow properties of biodiesel, as characterized by its cloud point (CP), pour point (PP), and cold filter plugging point (CFPP), are less satisfactory than petroleum diesel. CP is the temperature at which a fuel starts to show observable crystals upon cooling under defined conditions (ASTM D2500, 2005). PP is the lowest temperature at which a fuel can maintain its flowability or ability to be pumped. The CFPP for biodiesel is defined as the highest temperature at which a given volume of fuel fails to pass through a standardized filtration device under specified testing conditions.

All cold flow properties are related to the melting points (m.p.) of the fuel constituents and their solubility in the fuel. A high m.p. component will crystallize and precipitate out when its concentration in the fuel is beyond its solubility. The components of biodiesel generally have higher melting temperatures than those of petroleum diesel, especially the long-chain, saturated fatty acid esters. The longer the fatty acid chain, the higher the melting point of the component. When unsaturated bonds are present in the alkyl chain, the m.p. of the component decreases considerably. For example, the m.p. of methyl oleate (C18:1) is −20°C, while the m.p. of linoleate (C18:2) is −35°C. Another factor affecting the cold flow properties is the solubility of a component in the fuel. A high m.p. component will not crystallize unless its concentration is higher than the quantity the fuel can dissolve. The final CP or PP of a biodiesel is the collective outcome of the properties of the individual methyl esters and their liquid–solid equilibria in the fuel. The CP and PP of biodiesel largely depend on the fatty acid profiles of the feedstock, as shown in Table 14.4. Traceable impurities, such as plant sterol glucosides, can also affect the biodiesel cold flow properties. Plant sterol glucosides have higher melting points and limited solubility in biodiesel. Once in biodiesel, they may crystallize and serve as the nuclei for other high m.p. components to agglomerate.

Table 14.4

Examples of cloud points of biodiesel from different feedstocks (Imahara et al., 2006)

No. Oil/fat Methyl ester composition (wt%) Cloud point
C16:0 C18:0 C18:1 C18:2 C18:3 Others (K) (°C)
1 Beef tallow 23.9 17.5 43.9 2.3 0.1 12.3 286 13
2 Palm 39.5 4.1 43.2 10.6 0.2 2.4 283 10
3 Sunflower 6.1 4.2 24 63.5 0.4 1.8 274 1
4 Soybean 10.7 3.2 25 53.3 5.4 2.5 272 − 1
5 Linseed 6.7 3.7 21.7 15.8 52.1 0 268 − 5
6 Olive 10.7 2.6 78.7 5.8 0.7 1.5 268 − 5
7 Safflower 6.4 2.2 13.9 76 0.2 1.3 267 − 6
8 Rapeseed 4.3 1.9 61.5 20.6 8.3 3.1 267 − 6

Image

Source: Imahara et al., 2006.

Biodiesel is relatively unstable compared to petroleum diesel. This is mainly due to the chemical characteristics of the biodiesel constituents. The fatty acid chains contain variable numbers of unsaturated carbon to carbon (C–C) bonds, depending on the feedstock, with the carboxyl groups linked with alkyl groups at one end. These unsaturated C–C double bonds are non-conjugated in structure and are thermochemically vulnerable to oxidation. The mechanism is that alkyl radicals are formed first at the positions of unsaturation following an attack by oxidants; then hydroperoxides are formed, followed by a series of degradation and polymerization reactions. The presence of free mineral acids catalyzes the formation of alkyl free radicals (Shahidi, 2005; Frankel, 2005). The relative rate of oxidation depends on the level of unsaturation. The more double-bonds an oil or a fat has, the more it is prone to oxidation. The relative vulnerability of the oleic (C18:1), linoleic (C18:2), and linolenic (C18:3) chains is 1:12:26 after 30 days of storage (Chapman et al., 2009). Other literature reports even higher oxidation rates of 1:40:80 for the same fatty acids (Witting, 1965). This difference is largely attributed to the environment to which the oil or fat is exposed. A large number of chemicals can be produced from such oxidation reactions, including aliphatic alcohols, aldehydes, short-chain organic acids, and polymers (Loury, 1972; Neff et al., 1993; Andersson and Lingnert, 1998; Bondioli et al., 2002). The adverse consequences of such chemicals in biodiesel are reduced flash point, rancidity, increased acidity, increased viscosity, and accelerated fuel degradation. Biodiesel oxidative instability may be accelerated due to the impurities present in the fuel. Minerals such as rust from the containers serve as catalysts for decomposition reactions. Incompletely reacted products, indicated by the presence of excess monoglycerides and free glycerin, likely provide a reactive environment. The potential for high moisture retention in biodiesel (15–25 times more than that in petroleum diesel; He et al., 2007a) also provides favorable conditions for biodiesel oxidation.

To ensure its quality in long-term storage, biodiesel needs to strictly meet the ASTM D6751 or EN21414 specifications. Rusty tanks need to be thoroughly cleaned before being filled with biodiesel. Direct sunlight and high temperatures should be avoided. Nitrogen blanketing should be used to reduce air contact and moisture absorption in the biodiesel. For the long-term storage of biodiesel, a biocide application is recommended to prevent biological contamination.

14.8 Conclusion and future trends

Current production of both biodiesel and renewable diesel are driven by the requirements of the Renewable Fuel Standard (RFS) which is mandated by the Energy Independence and Security Act of 2007 (EIA, 2012). The RFS required that petroleum refiners in the US utilize 1 billion gallons of biomass-based diesel fuel in 2011. These fuels are now designated as Advanced Biofuels, with mandated levels that will ramp up to even higher levels in future years. Although alternative diesel requirements have been met every year, the level of lignocellulosic ethanol required by the RFS has not met expected levels and this may require adjustment of the target production levels. The uncertainty regarding the government’s continued commitment to the RFS is one of the greatest challenges in financing alternative fuel products in the US.

14.9 Sources of further information and advice

Books and proceedings

1. Knothe G, Van Gerpen J, eds. The Biodiesel Handbook. 2nd edn AOCS Publishing 2010; ISBN-10:1893997626, ISBN-13: 978–1893997622.

2. Van Gerpen J, Pruszko R, Clements D, Shanks B, eds. Building a Successful Biodiesel Business: Technology Considerations, Developing the Business, Analytical Methodologies. 2nd edn Biodiesel Basics 2006; ISBN-10: 097863490X, ISBN-13: 978–0978634902.

3. Mario Marchetti J, Fang Z, eds. Biodiesel: Blends, Properties and Applications (Energy Science, Engineering and Technology). Nova Science 2011; ISBN-10: 1613246609, ISBN-13: 978–1613246603.

4. Mario Marchetti J, ed. Biodiesel: Blends, Properties and Applications (Energy Science, Engineering and Technology). Nova Science 2010; ISBN-10: 1616689633, ISBN-13: 978-1616689636.

5. Bart J, Cavallaro S, Palmeri N, eds. Biodiesel Science and Technology: From Soil to Oil. Woodhead Publishing 2010; ISBN-10: 1845695917, ISBN-13: 978–1845695910.

6. Luque R, Antonio Melero J, eds. Advances in Biodiesel Production: Processes and Technologies. Woodhead Publishing 2012; ISBN-10: 0857091174, ISBN-13: 978–0857091178.

Online documents

1. NREL/TP-540-43672 Biodiesel Handling and Use Guide. 4th edn National Renewable Energy Laboratory, US Department of Energy http://www.nrel.gov/vehiclesandfuels/npbf/feature_guidelines.html; 2009.

2. Technical publications on biodiesel. National Renewable Energy Laboratory, US Department of Energy. Available at: http://www.nrel.gov/vehiclesandfuels/npbf/pubs_biodiesel.html.

3. Renewable and Alternative Fuels. US Environmental Protection Agency (EPA). Available at: http://www.epa.gov/otaq/fuels/alternative-renewablefuels/index.htm.

4. Online documents and information. National Biodiesel Board. Available at: http://www.biodiesel.org/.

5. Studies and Reports. European Biodiesel Board. Available at: http://www.ebb-eu.org/studies.php.

6. Online documents and information. IEA Bioenergy Task 39 – Commercializing Liquid Biofuels. Available at: http://www.task39.org/.

14.10 References

1. Ahmad A, Mat Yasin N, Derek C, Lim J. Microalgae as a sustainable energy source for biodiesel production: a review. Renewable and Sustainable Energy Reviews. 2011;15(1):584–593. doi 10.1016/j.rser.2010.09.018.

2. Andersson K, Lingnert H. Influence of oxygen and copper concentration on lipid oxidation in rapeseed oil. JAOCS. 1998;75(8):1041–1046.

3. ASTM Standard D2500-05. Standard Test Method for Cloud Point of Petroleum Products West Conshohocken, PA: ASTM International; 2005. doi 10.1520/D2500-05.

4. ASTM (American Society for Testing and Materials). D 6751 Standard Specification for Biodiesel Fuel (B100) Blend Stock for Distillate Fuels West Conshohocken, PA: ASTM International; 2011. doi 10.1520/D6751-11B.

5. Benjapornkulaphong S, Ngamcharussrivichai C, Bunyakiat K. Al2O3-supported alkali earth metal oxides for transesterification of palm kernel oil and coconut oil. Chem Eng J. 2009;145:468–474. doi 10.1016/j.cej.2008.04.036.

6. Berrios M, Skelton RL. Comparison of purification methods for biodiesel. Chem Eng J. 144:459–465. doi 10.1016/j.cej.2008.07.019.

7. Bondioli P, Gasparoli A, Bella L, Tagliabue S. Evaluation of biodiesel storage stability using reference methods. Eur J Lipid Sci Technol. 104. doi 10.1002/1438-9312(200212)104:12<777::AID-EJLT777>3.0.CO;2-#.

8. Borges M, Diaz L. Recent developments on heterogeneous catalysts for biodiesel production by oil esterification and transesterification reactions: a review. Renewable and Sustainable Energy Reviews. 2012; 16. doi 10.1016/j. rser.2012.01.071.

9. Borges M, Brito A, Hernandez A, Diaz L. Alkali metal exchanged zeolite as heterogeneous catalyst for biodiesel production from sunflower oil and waste oil: studies in a batch/continuous slurry reactor system. International Journal of Chemical Reactor Engineering. 2011;9:1–20.

10. Canakci M, Van Gerpen J. Biodiesel production via acid catalysis. Trans ASAE. 1999;42(5):1203–1210.

11. Canakci M, Van Gerpen J. Biodiesel production from oils and fats with high free fatty acids. Trans ASAE. 2001;44(6):1429–1436.

12. Carmo Jr AL, de Souza C, da Costa E, Longo J Zamian, da Rocha Filho G. Production of biodiesel by esterification of palmitic acid over mesoporous aluminosilicate Al-MCM-41. Fuel. 2009;88:461–468. doi 10.1016/j.fuel.2008.10.007.

13. CGC (Canadian Grain Commission). Oil content – quality of western Canadian canola 2011. Available at http://www.grainscanada.gc.ca/canola/harvest-recolte/2011/hqc11-qrc11-06-eng.htm; (accessed Nov. 20, 2012).

14. Chapman TM, Kim HJ, Min DB. Prooxidant activity of oxidized α-tocopherol in vegetable oils. J of Food Sci. 74(7):C536–C542. doi 10.1111/j.1750-3841.2009.01262.x.

15. Chen C, Yeh K, Aisyah R, Lee D, Chang J. Cultivation, photobioreactor design and harvesting of microalgae for biodiesel production: a critical review. Bioresource Technol. 2011;102(1):SI71–S181. doi 10.1016/j.biortech.2010.06.159.

16. Chen C, Huang C, Tran D, Chang J. Biodiesel synthesis via heterogeneous catalysis using modified strontium oxides as the catalysts. Bioresour Technol. 2012;113:8–13. doi 10.1016/j.biortech.2011.12.142.

17. Chisti Y. Biodiesel from microalgae. Biotechnol Adv. 25:294–306. doi 10.1016/j.biotechadv.2007.02.001.

18. Chisti Y. Biodiesel from microalgae beats bioethanol. Trends in Biotechnol. 2008;26(3):126–131. doi 10.1016/j.tibtech.2007.12.002.

19. Chouhan A, Sarma A. Modern heterogeneous catalysts for biodiesel production: a comprehensive review. Renewable and Sustainable Energy Reviews. 15(9):4378–4399. doi 10.1016/j.rser.2011.07.112.

20. Cintas P, Mantegna S, Gaudino E, Cravotto G. A new pilot flow reactor for high-intensity ultrasound irradiation: application to the synthesis of biodiesel. Ultrasonics Sonochemistry. 2010;17(6):985–989. doi 10.1016/j.ultsonch.2009.12.003.

21. Contran N, Chessa L, Lubino M, Bellavitea D, Roggero P, Enne G. State-of-the-art of the Jatropha curcas productive chain: from sowing to biodiesel and by-products. Industrial Crops and Products. 2013;42:202–215. doi 10.1016/j. indcrop.2012.05.037.

22. Dias L, Missio R, Dias D. Antiquity, botany, origin and domestication of Jatropha curcas (Euphorbiaceae), a plant species with potential for biodiesel production. Genet Mol Res. 2012;11(3):2719–2728.

23. Diaz-Ballote L, Lopez-Sansores J, Maldonado-Lopez L, Garfias-Mesias L. Corrosion behavior of aluminum exposed to a biodiesel. Electrochemistry Communications. 2009;11(1):41–44. doi 10.1016/j.elecom.2008.10.027.

24. Di Serio M, Tesser R, Casale L, D'Angelo A, Trifuoggi M, Santacesaria E. Heterogeneous catalysis in biodiesel production: the influence of leaching. Topics in Catalysis. 2010;53(11–12):811–819. doi 10.1007/s11244-010-9467-y.

25. Donnis B, Egeberg RG, Blom P, Knudsen KG. Hydroprocessing of bio-oils and oxygenates to hydrocarbons: understanding the reaction routes. Topics in Catalysis. 2009;52:229–240. doi 10.1007/s11244-008-9159-z.

26. EIA. Energy Information Agency. Available at: www.eia.gov; 2012; (accessed August 30, 2012).

27. European Biodiesel Board. Statistics – 2010 production by country. Available at: http://www.ebb-eu.org/stats.php; (accessed Nov. 26, 2012).

28. Evangelista R, Isbell T, Cermak S. Extraction of pennycress (Thlaspi arvense L.) seed oil by full pressing. Industrial Crops and Products. 2012;37(1):76–81. doi 10.1016/j.indcrop.2011.12.003.

29. Faria E, Ramalho H, Marques J, Suarez P, Prado A. Tetramethylguanidine covalently bonded onto silica gel surface as an efficient and reusable catalyst for transesterification of vegetable oil. Appl Catal A: Gen. 2008;338:72–78. doi 10.1016/j. apcata.2007.12.021.

30. Flint E, Suslick K. The temperature of cavitation. Science. 1991;253(5026):1397–1399. doi 10.1126/science.253.5026.1397.

31. Frankel EN. Lipid Oxidation Bridgwater, UK: The Oily Press; 2005.

32. Freedman B, Pryde E, Mounts T. Variables affecting the yields of fatty esters from transesterified vegetable oils. JAOCS. 1984;61:1638–1643. doi 10.1007/BF02541649.

33. Furimsky E. Catalytic hydrodeoxygenation. Appl Catal A: Gen. 2000;199:147–190. doi 10.1016/S0926-860X(99)00555-4.

34. Furuta S, Matsuhashi H, Arata K. Biodiesel fuel production with solid superacid catalysis in fixed bed reactor under atmospheric pressure. Catal Commum. 5(12):721–723. doi 10.1016/j.catcom.2004.09.001.

35. Garcia C, Teixeira S, Marciniuk L, Schuchardt U. Transesterification of soybean oil catalyzed by sulfated zirconia. Bioresour Technol. 2008;99:6608–6613. doi 10.1016/j.biortech.2007.09.092.

36. Goembira F, Matsuura K, Saka S. Biodiesel production from rapeseed oil by various supercritical carboxylate esters. Fuel. 97:373–378. doi 10.1016/j. fuel.2012.02.051.

37. Gomes J, Puna J, Goncalves L, Bordado J. Study on the use of MgAl hydrotalcites as solid heterogeneous catalysts for biodiesel production. Energy. 36(12):6770–6778. doi 10.1016/j.energy.2011.10.024.

38. Granados M, Poves M, Alonso D, et al. Biodiesel from sunflower oil by using activated calcium oxide. Appl Catal B: Environ. 2007;73:317–326. doi 10.1016/j.apcatb.2006.12.017.

39. Haas M, McAloon A, Yee W, Foglia T. A process model to estimate biodiesel production costs. Bioresour Technol. 2006;97:671–678. doi 10.1016/j. biortech.2005.03.039.

40. Hanh H, Nguyen D, Okitsu K, Nishimura R, Maeda Y. Biodiesel production through transesterification of triolein with various alcohols in an ultrasonic field. Renewable Energy. 2009;34(3):766–768. doi 10.1016/j.renene.2008. 04.007.

41. Harun R, Davidson M, Doyle M, Gopiraja R, Danquaha M, Forde G. Technoeconomic analysis of an integrated microalgae photobioreactor, biodiesel and biogas production facility. Biomass Bioenergy. 2011;35:741–747. doi 10.1016/j. biombioe.2010.10.007.

42. He B, Van Gerpen JH. Application of ultrasonication in transesterification processes for biodiesel production. Biofuels. 2012a;3(4):479–488. doi 10.4155/bfs.12.35.

43. He B, van Gerpen JH. Analyzing biodiesel for contaminants and moisture retention. Biofuels. 2012b;3(3):351–360. doi 10.4155/bfs.12.19.

44. He B, Thompson J, Routt D, Van Gerpen J. Moisture absorption in biodiesel and its petro-diesel blends. Appl Eng Agri. 2007a;23(1):71–76.

45. He H, Wang T, Zhu S. Continuous production of biodiesel fuel from vegetable oil using supercritical methanol process. Fuel. 2007b;86:442–447. doi 10.1016/j.fuel.2006.07.035.

46. Hsin T, Chen S, Guo E, Tsai C, Pruski M, Lin V. Calcium containing silicate mixed oxide-based heterogeneous catalysts for biodiesel production. Topics in Catalysis. 2010;53(11–12):746–754. doi 10.1007/s11244-010-9462-3.

47. Hu E, Xu Y, Hu X, Pan L, Jiang S. Corrosion behaviors of metals in biodiesel from rapeseed oil and methanol. Renewable Energy. 2012;37(1):371–378. doi 10.1016/j.renene.2011.07.010.

48. Huber GW, Chheda JN, Barrett CJ, Dumesic JA. Production of liquid alkanes by aqueous-phase processing of biomass-derived carbohydrates. Science. 2005;308:1446–1450. doi 10.1126/science.1111166.

49. Ilham Z, Saka S. Two-step supercritical dimethyl carbonate method for biodiesel production from Jatropha curcas oil. Bioresour Technol. 2010;101:2735–2740. doi 10.1016/j.biortech.2009.10.053.

50. Imahara H, Minami E, Saka S. Thermodynamic study on cloud point of biodiesel with its fatty acid composition. Fuel. 85(12–13):1666–1670. doi 10.1016/j.fuel.2006.03.003.

51. Imahara H, Minami E, Hari S, Saka S. Thermal stability of biodiesel in supercritical methanol. Fuel. 2008;8:1–6. doi 10.1016/j.fuel.2007.04.003.

52. Jain S, Sharma M. Thermal stability of biodiesel and its blends: a review. Renewable and Sustainable Energy Reviews. 2011;15(1):438–448. doi 10.1016/ j.rser.2010.08.022.

53. Ji J, Wang J, Li Y, Yu Y, Xu Z. Preparation of biodiesel with the help of ultrasonic and hydrodynamic cavitation. Ultrasonics. 44:E411–E414. doi 10.1016/j.ultras.2006.05.020.

54. Jitputti J, Kitiyanan B, Rangsunvigit P, Bunyakiat K, Attanatho L, Jenvanitpanjakul P. Transesterification of crude palm kernel oil and crude coconut oil by different solid catalysts. Chem Eng J. 2006;116:61–66. doi 10.1016/j. cej.2005.09.025.

55. Jobe J. Opening address to the Annual Technical Workshop of the National Biodiesel Board, Kansas City, MO. Oct. 29:2012.

56. Kalva A, Sivasankar T, Moholkar V. Physical mechanism of ultrasound-assisted synthesis of biodiesel. Ind & Eng Chem Res. 2009;48(1):534–544. doi 10.1021/ie800269g.

57. Kawashima A, Matsubara K, Honda K. Development of heterogeneous base catalysts for biodiesel production. Bioresour Technol. 2008;99(9):3439–3443. doi 10.1016/j.biortech.2007.08.009.

58. Kawashima A, Matsubara K, Honda K. Acceleration of catalytic activity of calcium oxide for biodiesel production. Bioresour Technol. 2009;100:696–700. doi 10.1016/j.biortech.2008.06.049.

59. Kiss F, Jovanovic M, Boskovic G. Economic and ecological aspects of biodiesel production over homogeneous and heterogeneous catalysts. Fuel Processing Technology. 91(10):1316–1320. doi 10.1016/j.fuproc.2010. 05.001.

60. Knothe G. Biodiesel and renewable diesel: a comparison. Progress in Energy and Combustion Science. 2010;36:364–373. doi 10.1016/j.pecs.2009.11.004.

61. Kondamudi N, Mohapatra S, Misra M. Quintinite as a bifunctional heterogeneous catalyst for biodiesel synthesis. Appli Cat A: Gen. 2011;393(1–2):36–43. doi 10.1016/j.apcata.2010.11.025.

62. Koundinya V, Hansen R. Pennycress Agricultural Marketing Resource Center, Iowa State University. Available at http://www.agmrc.org/commodities__products/grains__oilseeds/pennycress/; (accessed Nov. 21, 2012).

63. Kouzu M, Kasuno T, Tajika M, Yamanaka S, Hidaka J. Active phase of calcium oxide used as solid base catalyst for transesterification of soybean oil with refluxing methanol. Appl Catal A: Gen. 2008;334:357–365. doi 10.1016/j.apcata.2007.10.023.

64. Kumar A, Sharma S. An evaluation of multipurpose oil seed crop for industrial uses (Jatropha curcas L.): a review. Industrial Crops and Products. 28:1–10. doi 10.1016/j.indcrop.2008.01.001.

65. Kumar D, Kumar G, Poonam J, Singh C. Fast, easy ethanolysis of coconut oil for biodiesel production assisted by ultrasonication. Ultrasonics Sonochemistry. 2010a;17(3):555–559. doi 10.1016/j.ultsonch.2009.10.018.

66. Kumar D, Poonam J, Singh C. Ultrasonic-assisted transesterification of Jatropha curcus oil using solid catalyst, Na/SiO2. Ultrasonics Sonochemistry. 2010b;17(5):839–844. doi 10.1016/j.ultsonch.2010.03.001.

67. Kumar S, Chaube A, Jain S. Sustainability issues for promotion of Jatropha biodiesel in Indian scenario: a review. Renewable and Sustainable Energy Reviews. 2012;16(2):1089–1098. doi 10.1016/j.rser.2011.11.014.

68. Kusdiana D, Saka S. Kinetics of transesterification in rapeseed oil to biodiesel fuel as treated in supercritical methanol. Fuel. 2001a;80(5):693–698. doi 10.1016/S0016-2361(00)00140-X.

69. Kusdiana D, Saka S. Methyl esterification of free fatty acids of rapeseed oil as treated in supercritical methanol. J Chem Eng Japan. 2001b;34(3):383–387. doi 10.1252/jcej.34.383.

70. Kusdiana D, Saka S. Effects of water on biodiesel fuel production by supercritical methanol treatment. Bioresour Technol. 2004a;91(3):289–295. doi 10.1016/S0960-8524(03)00201-3.

71. Kusdiana D, Saka S. Two-step preparation for catalyst-free biodiesel fuel production – hydrolysis and methyl esterification. Appl Biochem Biotechnol. 2004b;113:781–791.

72. Lee I, Pfalzgraf LM, Poppe GB, Powers E, Haines T. The role of sterol glucosides on filter plugging. April 6 In: Biodiesel Magazine. http://www.biodieselmagazine.com/articles/1566/the-role-of-sterol-glucosides-on-filter-plugging/; 2007.

73. Li Y, Qiu F, Yang D, Li X, Sun P. Preparation, characterization and application of heterogeneous solid base catalyst for biodiesel production from soybean oil. Biomass & Bioenergy. 35(7):2787–2795. doi 10.1016/j. biombioe.2011.03.009.

74. Liang X, Gao S, Wu H, Yang J. Highly efficient procedure for the synthesis of biodiesel from soybean oil. Fuel Process Technol. 2009;90:701–704. doi 10.1016/j.fuproc.2008.12.012.

75. Liu X, He H, Wang Y, Zhu S. Transesterification of soybean oil to biodiesel using SrO as a solid base catalyst. Catal Commun. 2007;8:1107–1111. doi 10.1016/j.catcom.2006.10.026.

76. Liu X, He H, Wang Y, Zhu S. Transesterification of soybean oil to biodiesel using CaO as a solid base catalyst. Fuel. 2008;87:216–221. doi 10.1016/j. fuel.2007.04.013.

77. Loury M. Possible mechanisms of autoxidative rancidity. Lipids. 1972;7:671–675. doi 10.1007/BF02533075.

78. Mason T. Large scale sonochemical processing: aspiration and actuality. Ultrasonics Sonochemistry. 2000;7(4):145–149. doi 10.1016/S1350-4177(99)00041-3.

79. McVay K, Lamb P. Camelina production in Montana. Available at http://msuextension.org/publications/AgandNaturalResources/MT200701AG.pdf; 2008; (accessed Nov. 20, 2012).

80. Mendow G, Monella F, Pisarello M, Querini C. Biodiesel production from non-degummed vegetable oils: phosphorus balance throughout the process. Fuel Processing Technol. 2011;92(5):864–870. doi 10.1016/j.fuproc.2010.11.029.

81. Mielke T. World supplies of rapeseed and canola likely to remain tight in the 2012/13 season. Inform. 2012; 24(3):136–138.

82. Minami E, Saka S. Kinetics of hydrolysis and methyl esterification for biodiesel production in two-step supercritical methanol process. Fuel. 2006; 85:2479–2483. doi 10.1016/j.fuel.2006.04.017.

83. Mitich L. Field pennycress (Thlaspi arvense L.): the stinkweed. Weed Technology. 1996;10(3):675–678.

84. Mittelbach M, Remschmidt C. Biodiesel, the Comprehensive Handbook Vienna: Boersedruck; 2005.

85. Mofijur M, Masjuki H, Kalam M, et al. Prospects of biodiesel from Jatropha in Malaysia. Renewable and Sustainable Energy Reviews. 16(7):5007–5020. doi 10.1016/j.rser.2012.05.010.

86. Monnier H, Wilhelm A, Delmas H. Influence of ultrasound on mixing on the molecular scale for water and viscous liquids. Ultrasonics Sonochemistry. 1999a;6(1–2):67–74. doi 10.1016/S1350-4177(98)00034-0.

87. Monnier H, Wilhelm A, Delmas H. The influence of ultrasound on micromixing in a semi-batch reactor. Chem Eng Sci. 1999b;54(13–14):2953–2961. doi 10.1016/S0009-2509(98)00335-2.

88. Mootabadi H, Salamatinia B, Bhatia S, Abdullah A. Ultrasonic-assisted biodiesel production process from palm oil using alkaline earth metal oxides as the heterogeneous catalysts. Fuel. 2010;89(8):1818–1825. doi 10.1016/j.fuel.2009.12.023.

89. Moser B, Shah S, Winkler-Moser J, Vaughn S, Evangelista R. Composition and physical properties of cress (Lepidium sativum L.) and field pennycress (Thlaspi arvense L.) oils. Industrial Crops and Products. 2009a;30(2):199–205. doi 10.1016/j.indcrop.2009.03.007.

90. Moser B, Knothe G, Vaughn S, Isbell T. Production and evaluation of biodiesel from field pennycress (Thlaspi arvense L.) oil. Energies and Fuels. 2009b;23(8):4149–4155. doi 10.1021/ef900337g.

91. Nakagaki S, Bail A, dos Santos V, et al. Use of anhydrous sodium molybdate as an efficient heterogeneous catalyst for soybean oil methanolysis. Appl Catal A: Gen. 351(2):267–274. doi 10.1016/j. apcata.2008.09.026.

92. Neff WE, Mounts TL, Rinsch WM, Konishi H. Photooxidation of soybean oils as affected by triacylglycerol composition and structures. JAOCS. 1993;70(2):163–168. doi 10.1007/BF02542620.

93. Niu L, Gao L, Xiao G, Fu B. Study on biodiesel from cotton seed oil by using heterogeneous super acid catalyst SO42 -/ZrO2. Asia-Pacific Journal of Chemical Engineering. 2012;7:S222–S228. doi 10.1002/apj.532.

94. Niza N, Tan K, Lee K, Ahmad Z. Biodiesel production by non-catalytic supercritical methyl acetate: thermal stability study. Appl Energy. 101(special Issue):198–202. doi 10.1016/j.apenergs.2012.03.033.

95. NREL. NREL/TP-540–43672 Biodiesel Handling and Use Guide Golden, CO: National Renewable Energy Laboratory, US DOE; 2009.

96. Park YM, Lee JY, Chung SH, et al. Esterification of used vegetable oils using the heterogeneous WO3/ZrO2 catalyst for production of biodiesel. Bioresour Technol. 2010;101:S59–S61. doi 10.1016/j.biortech.2009.04.025.

97. Peng B, Shu Q, Wang J, Wang G, Wang D, Han M. Biodiesel production from waste oil feedstock by solid acid catalysis. Process Saf Environ Prot. 86(B6):441–447. doi 10.1016/j.psep.2008.05.003.

98. Piironen V, Lindsay DG, Miettinen TA, Toivo J, Lampi AM. Plant sterols: biosynthesis, biological function and their importance to human nutrition. J Sci Food Agri. 2000;80:939–966. doi 10.1002/(SICI)1097-0010 (20000515)80:7<939::AID-JSFA644>3.3.CO;2-3.

99. Putnam D, Budin J, Field L, Breene W. Camelina: a promising low-input oilseed. In: Janick J, Simon JE, eds. New Crops. New York: Wiley; 1993;314–322.

100. Saka S, Isayama Y. A new process for catalyst-free production of biodiesel using supercritical methyl acetate. Fuel. 2009;88:1307–1313. doi 10.1016/j.fuel.2008. 12.028.

101. Saka S, Kusdiana D. Transesterification of rapeseed oils in supercritical methanol to biodiesel fuels. Vol. 1 In: Overend R, Chornet E, eds. Proceedings of the Fourth Biomass Conference of the Americas. Pergamon: Qxford; 1999;747. Oakland, CA.

102. Saka S, Kusdiana D. Biodiesel fuel from rapeseed oil as prepared in supercritical methanol. Fuel. 2001a;80:225–231. doi 10.1016/S0016-2361(00)00083-1.

103. Saka S, Kusdiana D. Kinetics of transesterification in rapeseed oil to biodiesel fuel as treated in supercritical methanol. Fuel. 2001b;80:693–698. doi 10.1016/S0016-2361(00)00140-X.

104. Saka S, Kusdiana D, Minami E. Non-catalytic biodiesel fuel production with supercritical methanol technologies. J Scientif Ind Res. 2006;65(5):420–425.

105. 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. doi 10.1016/j.fuel.2009.10.018.

106. Sakai T, Kawashima A, Koshikawa T. Economic assessment of batch biodiesel production processes using homogeneous and heterogeneous alkali catalysts. Bioresour Technol. 2009; 100(13):3268–3276. doi 10.1016/j.biortech.2009.02.010.

107. Samart C, Sreetongkittikul P, Sookman C. Heterogeneous catalysis of transesterification of soybean oil using KI/mesoporous silica. Fuel Process Technol. 90:922–925. doi 10.1016/j.fuproc.2009.03.017.

108. Sebos I, Matsoukas A, Apostolopoulos V, Papayannakos N. Catalytic hydroprocessing of cottonseed oil in petroleum diesel mixtures for production of renewable diesel. Fuel. 2009;88:145–149. doi 10.1016/j.fuel.2008.07.032.

109. Semwal S, Arora A, Badoni R, Tuli D. Biodiesel production using heterogeneous catalysts. Bioresour Technol. 2011;102(3):2151–2161. doi 10.1016/j. biortech.2010.10.080.

110. Bailey’s Industrial Oil and Fat Products. Vol. 1. Shahidi F, ed. 6th edn New York: John Wiley & Sons; 2005; Section 1.7 Lipid Oxidation: Theoretical Aspects.

111. Sheehan J, Dunahay T, Benemann J, Roessler P. Report NREL/TP-580-24190 A Look Back at the US Department of Energy’s Aquatic Species Program–Biodiesel from Algae Golden, CO: National Renewable Energy Laboratory, US DOE; 1998.

112. Shol A. Industrial applications of ultrasound. In: Suslick K, ed. Ultrasound: Its Chemical, Physical, and Biological Effects. New York: VCH Publishers; 1988;97–122.

113. Shu Q, Yang B, Yuan H, Qing S, Zhu G. Synthesis of biodiesel from. Catal Commun. 2007;8:2159–2165. doi 10.1016/j.catcom.2007.04.028.

114. Singh A, Fernando S, Hernandez R. Base-catalyzed fast transesterification of soybean oil using ultrasonication. Energy and Fuels. 2007;21(2):1161–1164. doi 10.1021/ef060507g.

115. Singh B, Korstad J, Sharma Y. A critical review on corrosion of compression ignition (CI) engine parts by biodiesel and biodiesel blends and its inhibition. Renewable and Sustainable Energy Reviews. 2012;16(5):3401–3408. doi 10.1016/j.rser.2012.02.042.

116. Stavarache C, Vinatoru M, Nishimura R, Maeda Y. Fatty acids methyl esters from vegetable oil by means of ultrasonic energy. Ultrasonics Sonochemistry. 2005;12(5):367–372. doi 10.1016/j.ultsonch.2004.04.001.

117. Stephens E, Ross I, King Z, et al. An economic and technical evaluation of microalgal biofuels. Nature Biotechnol. 2010;28:126–128. doi 10.1038/nbt0210-126.

118. Sugawara T, Miyazawa T. Separation and determination of glycolipids from edible plant sources by high-performance liquid chromatography and evaporative light-scattering detection. Lipids. 1999;34(11):1231–1238. doi 10.1007/s11745-999-0476-3.

119. Teixeira L, Assis J, Mendonca D, et al. Comparison between conventional and ultrasonic preparation of beef tallow biodiesel. Fuel Processing Technology. 2009;90(9):1164–1166. doi 10.1016/j. fuproc.2009.05.008.

120. Thompson L, Doraiswamy L. Sonochemistry: science and engineering. Ind Eng Chem Res. 1999;38(4):1215–1249. doi 10.1021/ie9804172.

121. USDA. US Soybeans Production National Agricultural Statistics Service. Available at http://www.nass.usda.gov/; 2012a; (accessed Nov. 20, 2012).

122. USDA. US National Statistics for Canola National Agricultural Statistics Service. Available at http://www.nass.usda.gov/; 2012b; (accessed Nov. 20, 2012).

123. US DOE. United States Department of Energy. US billion ton update – biomass supply for a bioenergy and bioproducts industry. Available at http://www1.eere.energy.gov/biomass/pdfs/billion_ton_update.pdf; 2011; (accessed August 30, 2012).

124. US EPA. 2012 RIN Generation and Renewable Fuel Volume Production. Available at http://www.epa.gov/otaq/fuels/rfsdata/2011emts.htm; 2012a; (accessed Nov. 26, 2012).

125. US EPA. 2011 RIN Generation and Renewable Fuel Volume Production. Available at http://www.epa.gov/otaq/fuels/rfsdata/2012emts.htm; 2012b; (accessed Nov. 26, 2012).

126. Van Gerpen J. Biodiesel processing and production. Fuel Process Technol. 2005;86(10):1097–1107. doi 10.1016/j.fuproc.2004.11.005.

127. Van Gerpen J, Dvorak B. The effect of phosphorus level on the total glycerol and reaction yield of biodiesel. In: Bioenergy 2002, the 10th Biennial Bioenergy Conference. Boise, ID. Sept. 22–26. 2002.

128. Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G. Project report: NREL/SR-510-36244. Biodiesel Production Technology Golden, CO: National Renewable Energy Laboratory, US Department of Energy; 2004.

129. Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G. Building a Successful Biodiesel Business 2nd edn Dubuque, IA: Biodiesel Basics; 2006.

130. Vollmann J, Damboeck A, Eckl A, Schrems H, Ruckenbauer P. Improvement of Camelina sativa, an underexploited oilseed. In: Janick J, ed. Progress in New Crops. Alexandria, VA: ASHS Press; 1996;357–362.

131. Wall J, Thompson J, Van Gerpen J. Soap and glycerin removal from biodiesel using waterless processes. Trans ASABE. 2011;54(2):535–541.

132. Wang J, Chen K, Huang S, Chen C. Application of Li2SiO3 as a heterogeneous catalyst in the production of biodiesel from soybean oil. Chinese Chem Lett. 2011;22(11):1363–1366. doi 10.1016/j.cclet.2011.05.041.

133. Wang L, Yang J. Transesterification of soybean oil with nano-MgO or not in supercritical and subcritical methanol. Fuel. 2007;86:328–333. doi 10.1016/j. fuel.2006.07.022.

134. Warabi Y, Kusdiana D, Saka S. Biodiesel fuel from vegetable oil by various supercritical alcohols. Appl Biochem Biotechnol. 2004a;115:793–801.

135. Warabi Y, Kusdiana D, Saka S. Reactivity of triglycerides and fatty acids of rapeseed oil in supercritical alcohols. Bioresour Technol. 2004b;91(3):283–287. doi 10.1016/S0960-8524(03)00202-5.

136. Waynick JA. Technical literature review, TP-540–39096 Characterization of biodiesel oxidation and oxidation products Golden, CO: National Renewable Energy Laboratory, US DOE; 2005.

137. Wen G, Yan Z, Smith M, Zhang P, Wen B. Kalsilite based heterogeneous catalyst for biodiesel production. Fuel. 2010;89(8):2163–2165. doi 10.1016/j.fuel.2010.02.016.

138. Witting L. Lipid peroxidation in vivo. JAOCS. 1965;42:908–913. doi 10.1007/BF02632443.

139. Wu P, Yang Y, Colucci J, Grulke A. Effect of ultrasonication on droplet size in biodiesel mixtures. JAOCS. 2007;84(9):877–884. doi 10.1007/s11746-007-1114-9.

140. Wu X, Ruan R, Du Z, Liu Y. Current status and prospects of biodiesel production from microalgae. Energies. 2012;5(8):2667–2682. doi 10.3390/en508266.

141. Xie W, Liu Y, Chun H. Biodiesel preparation from soybean oil by using a heterogeneous CaxMg2 -xO2 catalyst. Catal Lett. 2012;142:352–359. doi 10.1007/s10562-012-0776-6.

142. Yamaguchi T, Benmore C, Soper A. The structure of subcritical and supercritical methanol by neutron diffraction, empirical potential structure refinement, and spherical harmonic analysis. J Chem Phys. 2000;112:8976–8987. doi 10.1063/1.481530.

143. Yang C, Fang Z, Li B, Long Y. Review and prospects of Jatropha biodiesel industry in China. Renewable and Sustainable Energy Reviews. 2012;16(4):2178–2190. doi 10.1016/j.rser.2012.01.043.

144. Yang Z, Xie W. Soybean oil transesterification over zinc oxide modified with alkali earth metals. Fuel Process Technol. 2007;88:631–638. doi 10.1016/j. fuproc.2007.02.006.

145. Yu D, Tian L, Wu H, et al. Ultrasonic irradiation with vibration for biodiesel production from soybean oil by Novozym 435. Process Biochem. 2010;45(4):519–525. doi 10.1016/j.procbio.2009.11.012.

146. Zhang T, Chao Y, Liu N, et al. Case study of biodiesel-diesel blends as a fuel in marine environment. Adv in, Chem Eng Sci 2011;(1):65–71.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset