Вы находитесь на странице: 1из 11

Ind. Eng. Chem. Res.

2005, 44, 5353-5363 5353

Synthesis of Biodiesel via Acid Catalysis


Edgar Lotero, Yijun Liu, Dora E. Lopez, Kaewta Suwannakarn, David A. Bruce, and
James G. Goodwin, Jr.*
Department of Chemical Engineering, Clemson University, Clemson, South Carolina 29634-0909

Biodiesel is synthesized via the transesterification of lipid feedstocks with low molecular weight
alcohols. Currently, alkaline bases are used to catalyze the reaction. These catalysts require
anhydrous conditions and feedstocks with low levels of free fatty acids (FFAs). Inexpensive
feedstocks containing high levels of FFAs cannot be directly used with the base catalysts currently
employed. Strong liquid acid catalysts are less sensitive to FFAs and can simultaneously conduct
esterification and transesterification. However, they are slower and necessitate higher reaction
temperatures. Nonetheless, acid-catalyzed processes could produce biodiesel from low-cost
feedstocks, lowering production costs. Better yet, if solid acid catalysts could replace liquid acids,
the corrosion and environmental problems associated with them could be avoided and product
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

purification protocols reduced, significantly simplifying biodiesel production and reducing cost.
This article reviews some of the research related to biodiesel production using acid catalysts,
including solid acids.
Downloaded via UNIV OF TOLEDO on June 30, 2018 at 06:33:44 (UTC).

Introduction Table 1. Values for the American Society for Testing and
Materials (ASTM) Standards of Maximum Allowed
Biodiesel is a nonpetroleum-based fuel that consists Quantities in Diesel and Biodiesel
of alkyl esters derived from either the transesterification property diesel biodiesel
of triglycerides (TGs) or the esterification of free fatty
acids (FFAs) with low molecular weight alcohols. The standard ASTM D975 ASTM D6751
composition HCa (C10-C21) FAMEb (C12-C22)
flow and combustion properties of biodiesel are similar Kin. viscosity (mm2/s) 1.9-4.1 1.9-6.0
to petroleum-based diesel and, thus, can be used either at 40 °C
as a substitute for diesel fuel or more commonly in fuel specific gravity (g/mL) 0.85 0.88
blends.1 As a point of comparison, pure biodiesel (B100) flash point (°C) 60-80 100-170
releases about 90% of the energy that normal diesel cloud point (°C) -15 to 5 -3 to 12
pour point (°C) -35 to -15 -15 to 16
does, and hence, its expected engine performance is
water (vol %) 0.05 0.05
nearly the same in terms of engine torque and horse- carbon (wt %) 87 77
power. Biodiesel, however, is made from renewable hydrogen (wt %) 13 12
resources, is biodegradable and nontoxic, and has a oxygen (wt %) 0 11
higher flash point than normal diesel. In addition, sulfur (wt %) 0.05 0.05
biodiesel increases lubricity (even at blends as low as cetane number 40-55 48-60
HFRRc (µm) 685 314
3% or less), which prolongs engine life and reduces the BOCLEd scuff (g) 3600 >7000
frequency of engine part replacement. Another signifi-
a Hydrocarbons. b Fatty acid methyl esters. c High-frequency
cant advantage of biodiesel is its low emission profile
and its oxygen content of 10-11%. Biodiesel is called reciprocating rig. d Ball-on-cylinder lubricity evaluator.
the environmentally friendly biofuel since it provides a Table 2. Average B100 and B20 Emissions (in %)
means to recycle carbon dioxide. In other words, bio- Compared to normal Diesel63
diesel does not contribute to global warming.
emission B100 B20
Table 1 shows a brief comparison of the ASTM
standards for diesel and biodiesel. As can be seen, carbon monoxide -48 -12
biodiesel exhibits characteristics that are comparable total unburned hydrocarbons -67 -20
particulate matter -47 -12
to traditional diesel fuel. Table 2 summarizes the typical nitrogen oxides +10 +2
emission profiles of biodiesel and one of its blends, B20, sulfates -100 -20
which consists of 20% biodiesel and 80% diesel, using air toxics -60 to -90 -12 to -20
petroleum-derived diesel emissions as the reference. The mutagenicity -80 to -90 -20.0
information in the table shows how biodiesel signifi-
cantly reduces emissions compared to diesel even when such as yellow and brown greases. Oils and fats belong
used as the minor component of a fuel blend. In to an ample family of chemicals called lipids. In general,
addition, the amount of sulfur in biodiesel is quite low, lipids are found in animals and plants. Typically, fats
which can significantly contribute to helping meet come from an animal source and oils from a plant
current sulfur emission standards in diesel vehicles. source. Fats and oils are primarily formed of TG
Commonly, biodiesel is prepared from TG sources molecules. A TG molecule is basically a triester of
such as vegetable oils, animal fats, and waste greases glycerol (a triol) and three fatty acids (long alkyl chain
carboxylic acids; see Table 3 for TG compositions of some
* To whom correspondence should be addressed. Phone: 864 common vegetable oils, animal fats, and greases). Mono-
656 2621. Fax (864) 656 0784. E-mail: james.goodwin@ and diglycerides can be obtained from TGs by substitut-
ces.clemson.edu. ing two and one fatty acid moieties with hydroxyl
10.1021/ie049157g CCC: $30.25 © 2005 American Chemical Society
Published on Web 01/25/2005
5354 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005

Table 3. Typical Fatty Acid Compositions of Vegetable Oils and Animal Fats
fatty acid composition, wt %
myristic palmitic palmitoleic stearic oleic linoleic linolenic sat.
fatty acid 14:00a 16:00 16:01 18:00 18:01 18:02 18:03 (%)
rapeseed oil64 3.5 0.9 64.4 22.3 8.2 4.4
virgin olive oil65 9.2 0.8 3.4 80.4 4.5 0.6 12.6
sunflower oil5 6.0 4.2 18.7 69.3 10.2
safflower oil5 5.2 2.2 76.3 16.2 7.4
soybean8 0.1 10.6 4.8 22.5 52.3 8.2 15.5
palm oil17 1.2 47.9 4.2 37 9.1 0.3 53.3
choice white grease66 23.3 3.5 11.0 47.1 11 1.0 37.8
poultry fat66 22.2 8.4 5.1 42.3 19.3 1.0 35.7
lard67 1.7 17.3 1.9 15.6 42.5 9.2 0.4 34.6
edible tallow67 4.8 28.4 14.8 44.6 2.7 52.0
yellow grease8 2.4 23.2 3.8 13.0 44.3 7.0 0.7 38.6
brown grease8 1.7 22.8 3.1 12.5 42.4 12.1 0.8 37.0
a 14:00, the alkyl chain contains 14 carbons and zero double bonds.

Figure 2. Formation of the active species in transesterification


reactions using a base catalyst.

using alkaline catalysts, such as sodium and potassium


methoxides and hydroxides. Industrially, NaOH and
KOH are preferred due to their wide availability and
low cost. Nonetheless, from a chemical standpoint the
active species with both types of catalysts are methoxide
ions. Methoxide ions form by dissociation of methoxide
salts in one case or when methanol reacts with hydroxyl
ions from added alkaline hydroxides in the second
situation (Figure 2). Once formed, the methoxide ions
are strong nucleophiles and attack the carbonyl moiety
in glyceride molecules to produce the alkyl esters.
Even though transesterification is feasible using
homogeneous base catalysts, the overall base-catalyzed
process suffers from serious limitations that translate
into high production costs for biodiesel. Strict feedstock
specifications are a main issue with this process. In
particular, the total FFA content associated with the
lipid feedstock must not exceed 0.5 wt %. Otherwise,
soap formation seriously hinders the production of fuel
grade biodiesel. Soap forms when the metal hydroxide
catalyst reacts with FFAs in the feedstock (Figure 3a).
Figure 1. Transesterification reactions of glycerides with metha- Soap production gives rise to the formation of gels,
nol. increases viscosity, and greatly increases product sepa-
ration cost.2 The alcohol and catalyst must also comply
groups, respectively. Lipid feedstocks used in biodiesel with rigorous specifications. The alcohol as well as the
production may contain a mixture of all these glyceride catalyst must be essentially anhydrous (total water
species plus some FFAs. The chemical transformation content must be 0.1-0.3 wt % or less).3 This is required
of lipid feedstocks to biodiesel involves the transesteri- since it is assumed that the presence of water in the
fication of glyceride species with alcohols to alkyl esters, feedstock promotes hydrolysis of the alkyl esters to
as shown in Figure 1. Among possible alcohols, metha- FFAs (Figure 3b) and, consequently, soap formation. To
nol is normally favored due to its low cost. The reaction conform to such demanding feedstock specifications
shown in Figure 1 uses methanol as a co-reagent for necessitates use of highly refined vegetable oils whose
transesterification. As can be seen, three consecutive price can account for 60-75% of the final cost of
reactions are required to complete the transesterifica- biodiesel.4 Other less expensive sources of TG feed-
tion of a TG molecule. All types of glyceride species stocks, such as yellow grease, can be used to counteract
participate in the reaction. If allowed to go to comple- the high price tag associated with biodiesel produced
tion, the net reaction produces 3 mol of alkyl esters and from refined oils, but this requires additional process
1 mol of glycerol for each mole of TG transformed. steps.
The transesterification reaction requires a catalyst in The demanding feedstock specifications for base-
order to obtain reasonable conversion rates. The nature catalyzed reactions have led researchers to seek cata-
of the catalyst is fundamental since it determines the lytic and processing alternatives that could ease this
compositional limits that the feedstock must conform difficulty and lower production costs. Methodologies
to. Furthermore, the reaction conditions and postsepa- based on acid-catalyzed reactions have the potential to
ration steps are predetermined by the nature of the achieve this since acid catalysts do not show measurable
catalyst used. Currently, most biodiesel is prepared susceptibility to FFAs. For this reason, and since there
Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5355

Figure 3. (a) Base catalyst reaction with FFAs to produce soap and water, both undesirable byproducts. (b) Water promotes the formation
of FFAs. These, then, can deactivate the catalyst and produce soap, as in (a).

are no previous reviews in the area of biodiesel synthesis ever, an alternative multistep process allows the use of
that primarily tackle the issue of acid-catalyzed reac- feedstocks having high FFA concentrations by first
tions, the development of acid-catalyzed methodologies carrying out the acid-catalyzed pre-esterification of the
is the focus of this paper. In an attempt to be complete, FFAs prior to the base-catalyzed TG transesterifica-
the second part of this paper also includes a review of tion.8,9 The combination of acid-catalyzed FFA pre-
work relating to heterogeneously acid-catalyzed reac- esterification followed by base-catalyzed TG transes-
tions due to the great potential that solid acids have in terification is commonly called the integrated process.
this field. Despite the added cost of production, the integrated
process is being increasingly applied to produce biodiesel
The Feedstock Issue from low-cost but high-FFA feedstocks with good re-
sults.
As previously mentioned, biodiesel can be prepared There are characteristics of biodiesel that are a direct
from a variety of sources including vegetable oils, animal consequence of the long alkyl chains associated with the
fats, and waste greases. Currently, refined vegetable oils alkyl esters that constitute the biofuel and strongly
are a major feedstock for biodiesel production. However, depend on the feedstocks used. For instance, depending
waste greases, such as yellow grease from used cooking on feedstock composition, biodiesel and its blends in cold
oils and animal fats, can also be employed because of climates can suffer significantly from high pour points,
their availability and low cost. It must be mentioned cloud points, and cold filter plugging points.10 In par-
that using waste greases to produce biodiesel allows for ticular, these problems are more apparent when the
a convenient route to recycle waste oils that otherwise sources of biodiesel feedstocks are animal fats. Here,
could only be sold as a low-value animal feed additive. the more saturated character of the fatty acid composi-
However, waste greases and animal fats are considered tion of animal fats (Table 3) is transferred to the alkyl
low-quality feedstocks compared to refined vegetable oils esters (biodiesel) obtained. In contrast, when biodiesel
in terms of FFA content. For instance, yellow grease, is prepared from vegetable oils, it mainly contains esters
obtained from rendered animal fats and restaurant of oleic and linoleic acids, which have unsaturated alkyl
waste oil, has FFA levels up to 15 wt %, and its price chains. In general, the unsaturation degree of the alkyl
varies from $0.09 to $0.20/lb.4 Another low-cost waste chains correlates well with the cold flow performance
grease, brown grease, obtained mainly from traps of biodiesel. This means that the more double bonds on
installed in commercial, municipal, or industrial sewage the alkyl chain, the lower the melting point of the esters
facilities, has a free fatty acid level greater than 15 wt obtained and, therefore, the colder it can get without
% and sells for $0.01-0.07/lb less than yellow grease. exhibiting undesirable high viscosity or solidifying.
Because of their price, use of waste grease feedstocks The problems outlined above could severely limit the
has been proposed as a way to lower biodiesel produc- use of biodiesel prepared from animal fats. However,
tion costs. In fact, some plants in the United States are the use of additives to improve the flow properties of
already producing biodiesel from yellow grease.5 biodiesel at low temperatures has been proposed as a
Predictably, low-cost feedstocks must undergo some counter measure for this problem.11 This topic will be
form of pretreatment before they can be used for expanded later in the heterogeneous-catalyzed glycerol
biodiesel production. Animal fats, for instance, often etherification section.
contain meat and bone particles, as well as other types
of organic matter. Commonly, the particulate matter in Homogeneous Acid-Catalyzed Reactions
these greases is removed using a cellulose filter. Other
pretreatment steps include water removal, steam distil- The liquid acid-catalyzed transesterification process
lation, and bleaching.6 Water removal via gravity does not enjoy the same popularity in commercial
separation in a horizontal tank is normally sufficient applications as its counterpart, the base-catalyzed
to reduce the concentration of water in the waste grease process. The fact that the homogeneous acid-catalyzed
to the acceptable level of 0.5 wt %.7 This drying step is reaction is about 4000 times slower than the homoge-
essential since yellow grease from used cooking oils may neous base-catalyzed reaction has been one of the main
contain up to 3 wt % water. In addition to water reasons.12 However, acid-catalyzed transesterifications
removal, the steam distillation process denatures and hold an important advantage with respect to base-
degrades residual proteins in the grease, and bleaching catalyzed ones: the performance of the acid catalyst is
removes spoiled proteins. not strongly affected by the presence of FFAs in the
Even though animal fats and waste greases constitute feedstock. In fact, acid catalysts can simultaneously
sources of inexpensive feedstocks for biodiesel, their catalyze both esterification and transesterification.
high concentrations of FFAs make them inappropriate Thus, a great-advantage with acid catalysts is that they
for the conventional direct base-catalyzed transesteri- can directly produce biodiesel from low-cost lipid feed-
fication route to biodiesel due to soap formation. How- stocks, generally associated with high FFA concentra-
5356 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005

Figure 4. Simplified BFD of the acid-catalyzed process. The diagram includes the following: feedstock pretreatment (1), catalyst
preparation (2), transesterification and esterification (3), alcohol recycle (4), acid catalyst removal (5), and biodiesel separation and
purification process (6).

tions (low-cost feedstocks, such as used cooking oil and


greases, commonly have FFAs levels of g6%). Recently,
it has been shown how acid-catalyzed production of
biodiesel can economically compete with base-catalyzed
processes using virgin oils, especially when the former
uses low-cost feedstocks.13,14 A simplified block flow
diagram (BFD) of a typical acid-catalyzed process is
shown in Figure 4 illustrating the most important steps
in biodiesel production.13
As already mentioned, the transesterification of TGs,
catalyzed by either bases or acids, consists of three
consecutive, reversible reactions. In the reaction se-
quence, TG is converted stepwise to diglyceride, mono-
glyceride, and finally glycerol accompanied with the
liberation of an ester at each step (Figure 1). The
transesterification chemical pathway shown in Figure
5, for an acid-catalyzed reaction, indicates how in the
catalyst-substrate interaction the key step is the pro-
tonation of the carbonyl oxygen. This in turn increases
the electrophilicity of the adjoining carbon atom, making
it more susceptible to nucleophilic attack. In contrast,
base catalysis takes on a more direct route, creating first Figure 5. Homogeneous acid-catalyzed reaction mechanism for
an alkoxide ion, which directly acts as a strong nucleo- the transesterification of triglycerides: (1) protonation of the
phile, giving rise to a different chemical pathway for carbonyl group by the acid catalyst; (2) nucleophilic attack of the
the reaction (Figure 6). This crucial difference, i.e., the alcohol, forming a tetrahedral intermediate; (3) proton migration
formation of a more electrophilic species (acid catalysis) and breakdown of the intermediate. The sequence is repeated
versus that of a stronger nucleophile (base catalysis), twice.
is ultimately responsible for the observed differences in
activity. the reaction was characterized by a mass-transfer-
To date, most studies have focused on base-catalyzed controlled regime that resulted from the low miscibility
transesterification of TGs. For acid-catalyzed systems, of the catalyst and reagents; i.e., the nonpolar oil phase
sulfuric acid has been the most investigated catalyst, was separated from the polar alcohol-acid phase. As
but other acids, such as HCl, BF3, H3PO4, and organic the reaction proceeded and ester products acted as
sulfonic acids, have also been used by different research- emulsifiers, a second-rate regime emerged, which was
ers.15 In a pioneering work, Freedman et al. examined kinetically controlled and characterized by a sudden
the transesterification kinetics of soybean oil with surge in product formation. Finally, the last regime was
butanol, using sulfuric acid as the catalyst.16 They found reached once equilibrium was approached near reaction
that the rate-limiting step varied over time, and three completion. In addition, these authors established that
regimes, in accordance with the observed reaction rate, a large molar ratio of alcohol-to-oil, 30:1, was needed to
could categorize the overall reaction process. Initially, have acceptable reaction rates. In this way, for the
Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5357

industry.18,19 As already mentioned, the low cost of


methanol makes it the first choice for the transesteri-
fication reaction. Ethanol, however, is derived from
agriculture products (renewable sources) and biologi-
cally less objectionable to the environment than metha-
nol; thus, ethanol is the ideal candidate for the synthesis
of a fully biogenerated fuel.18 Nonetheless, working with
higher molecular weight alcohols, such as butanol,
brings about interesting advantages. For instance,
butanol has better miscibility with the lipid feedstock
than smaller alcohols, contributing to a less pronounced
initial mass-transfer-controlled regime. Additionally, the
elevated boiling points of larger alcohols enable the
liquid reaction system to be operated at higher temper-
atures while maintaining moderate pressures. This is
an important issue with the acid-catalyzed transesteri-
fication since higher reaction temperatures are often
required to achieve faster reaction rates. However, only
a limited number of papers have actually addressed the
subject of the alcohol characteristics. Nye et al. com-
pared the series of linear alcohols from methanol to
butanol to find the most suitable for the transesterifi-
cation of waste frying oil.20 All reactions were carried
out using 0.1% sulfuric acid at reflux temperatures with
all other conditions kept constant. Butanol yielded the
highest reaction rates and conversion followed by 1-pro-
panol and then ethanol. Even though acid-catalyzed
Figure 6. Homogeneous base-catalyzed reaction mechanism for
the transesterification of TGs: (1) production of the active species, reactions carried out with methanol proved to be the
RO-; (2) nucleophilic attack of RO- to carbonyl group on TG, slowest, the reverse was true for the base-catalyzed
forming of a tetrahedral intermediate; (3) intermediate breakdown; process. These results suggest that initial reagent phase
(4) regeneration of the RO- active species. The sequence is miscibility was more critical in acid catalysis than in
repeated twice. base catalysis. Two factors could have affected phase
miscibility: the increased hydrophobicity associated
transesterification reaction (Figure 1), the forward with larger alcohols and the higher reaction tempera-
reactions followed pseudo-first-order kinetics, while the ture that was applied as the alcohol molecular weight
reverse reactions exhibited second-order kinetics. increased.
In a continuing effort to decrease biodiesel production
costs, researchers have attempted to ascertain what Thus, temperature plays an important role in the
optimum ratio of alcohol-to-TG should be used. Two acid-catalyzed synthesis of biodiesel.15,21 For instance,
factors must be considered during this optimization in the butanolysis of soybean oil catalyzed by 1 wt %
process. First, increasing the molar ratio of alcohol-to- H2SO4, five different temperatures from 77 to 117 °C
TG increases as well alcohol recovery and product were examined.16 Increasing temperature had a marked
separation costs. Second, the acid-catalyzed transes- effect on reaction rate with near complete conversion
terification achieves greater and faster conversions at of TGs requiring only 3 h at 117 °C, while comparable
high alcohol concentrations. To help identify an opti- conversions at 77 °C required 20 h. At higher temper-
mum alcohol-to-TG ratio, Canakci and Van Gerpen atures, the extent of phase separation decreases and
studied how the reagent molar ratio affected reaction rate constants increase, due to the higher temperature
rates and product yield in the transmethylation of as well as improved miscibility, leading to substantially
soybean oil by sulfuric acid.7 Five different molar ratios, shortened reaction times. The effect of temperature is
from 3.3:1 to 30:1, were studied. Their results indicated even more noticeable at higher temperatures and pres-
that ester formation increased with increasing the molar sures. In particular, at 240 °C and 70 bar, using 1.7 wt
ratio, reaching its highest value, 98.4%, at the highest % H2SO4, ester conversions greater than 90% could be
molar ratio used, 30:1. However, the benefits from obtained in only 15 min.22 Under such conditions, high-
higher alcohol-to-TG molar ratios became limited with FFA feedstocks (e.g., 44 wt % FFAs) could easily be
increasing ratio, ester formation increased sharply from transformed with continuous water removal. However,
77% at 3.3:1 to 87.8% at 6:1 and ultimately reaching a side reactions such as alcohol etherification could also
plateau value of 98.4% at 30:1. In a related study, be observed under such harsh conditions.
Crabbe et al. also determined the effect of molar ratio Reaction rates in acid-catalyzed processes may also
within the range of 3:1-23:1 and concluded that the be increased by the use of larger amounts of catalyst.
highest molar ratio required for complete transmeth- Typically, catalyst concentrations in the reaction mix-
ylation could be found between 35:1 and 45:1 by ture have ranged between 1 and 5 wt % in most
extrapolation.17 However, neither group determined an academic studies using sulfuric acid.16,21,23 Canakci and
optimum alcohol-to-TG molar ratio based on ester Van Gerpen used different amounts of sulfuric acid (1,
formation and separation efficiency. 3, and 5 wt %) in the transesterification of grease with
Alcohols used in acid-catalyzed transesterification methanol.7 In these studies, a rate enhancement was
have included methanol, ethanol, propanol, butanol, and observed with the increased amounts of catalyst, and
amyl alcohol. Methanol and ethanol are used most ester yield went from 72.7 to 95.0% as the catalyst
frequently in both laboratory research and the biodiesel concentration was increased from 1 to 5 wt %. The
5358 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005

dependence of reaction rate on catalyst concentration acid has for water, it is likely that the acid will interact
has been further verified by the same authors and other more strongly with water molecules than with alcohol
groups.9,17 It is known, however, that large quantities molecules. Thus, if water is present in the feedstock or
of acid catalyst can promote ether formation by alcohol produced during the reaction, the acid catalyst will
dehydration.24 A further complication of working with preferentially bind to water, leading to a reversible type
high acid catalyst concentrations becomes apparent of catalyst deactivation.
during the catalyst neutralization process, which pre- There are, therefore, two aspects that affect catalyst
cedes product separation. Since CaO addition during accessibility to the TG molecules. The first has been
neutralization is proportional to the concentration of previously discussed and has to do with macroscopic
acid needed in the reactor, high acid concentrations lead phase separation evidenced by the initial mass-transfer-
to increased CaO cost, greater waste formation, and controlled regime that can seriously hinder reaction
higher production cost. rate.16 In fact, when cosolvents, such as tetrahydrofuran
As previously mentioned, biodiesel production cost can (THF), are used to counteract the miscibility problem,
be reduced by using low-cost feedstocks with high FFA the methanolysis rate is known to dramatically in-
contents. However, the acid-catalyzed conversion of low- crease.27,28 The other less understood aspect that seems
cost feedstocks leads to the formation of significant to affect acid catalysis is the effect that water has on
quantities of water, which has a negative effect on catalyst activity. Water not only can bind to acid species
biodiesel production since water can hydrolyze the ester in solution (H+) more effectively than the alcohol (giving
products, producing (again) FFAs. Though ester hy- rise to a weaker acid) but also, with increasing water
drolysis can occur by either acid or base catalysis, acid concentrations, can give rise to water-rich clusters
catalysis is generally believed to be more tolerant of around protons.29 These extra molecules of water around
moisture and high FFA levels in the starting feedstock the catalyst could shield it from the hydrophobic TG
and, hence, more suitable for low-grade fats and molecules, inhibiting the reaction. Hydration of the acid
greases.15,19,21 In an effort to quantify the effect of water catalyst should have a more pronounced effect in
on sulfuric acid-catalyzed formation of biodiesel, Canak- transesterification than in esterification, given the
ci and Van Gerpen investigated the sensitivity of ester larger and more hydrophobic compounds involved in the
forming reactions to water and the FFAs levels present former (TGs vs FFAs). The polar carboxylic groups in
in soybean oil.7 Ester production was affected by as little FFAs and their capacity to easily interact through
as 0.1 wt % water concentration and was almost totally hydrogen bonds with water molecules facilitate the
inhibited when the water level reached 5 wt %. These FFAs-catalyst interaction and, therefore, esterification.
authors established that water content has to be kept As mentioned earlier, the acid-catalyzed pre-esteri-
under 0.5 wt % to achieve a 90% ester yield under their fication of FFAs is more commonly used to decrease FFA
reaction conditions. Recently, Kusdiana and Saka stud- levels in high-FFA feedstocks, before they can be used
ied the effect of water on methyl ester formation by the in the base-catalyzed transesterification. In fact, some
transesterification of rapeseed oil with methanol using acid catalysts, i.e., BF3 (a strong Lewis acid), have such
different catalysts.25 Reactions were conducted at the esterifying capacity that a method combining the alka-
same conditions except for the catalyst amount (1.5 wt line hydrolysis of feedstocks and the subsequent BF3-
% NaOH, 3 wt % H2SO4) and reaction time (1 h for the catalyzed esterification of FFAs has been proposed to
alkali, 48 h for the acid). These studies found that water prepare biodiesel in relatively short times.15 In practice,
concentration was more critical in acid catalysis than esterification is subjected to thermodynamic limitations,
in base catalysis. In the acid-catalyzed reaction, ester and water removal is desirable to obtain good conversion
conversion was reduced to 6% when 5 wt % water was rates and push the reaction to completion. Unfortu-
used in the starting reagent mixture. In contrast, the nately, calculations of average equilibrium constants in
alkali-catalyzed reaction was only slightly affected by these type of systems have not been reported given the
the presence of water, and ester conversion was ap- presence of a wide range of free fatty acids (C8-C24).
proximately 70% with an equivalent amount of water Canakci and Van Gerpen recently applied a two-step
in the reaction mixture, in agreement with the results sulfuric acid-catalyzed pre-esterification process of FFAs
of Canakci and Van Gerpen.7,9 Kusdiana and Saka also in used cooking oil to reduce FFA levels to below 1 wt
studied the esterification of FFAs.25 As expected, the %.9 The low FFA concentration was required for the
alkaline catalyst was completely saponified, but the subsequent alkaline-catalyzed transesterification. The
acid-catalyzed esterification showed excellent tolerance FFA pre-esterification step was preceded and followed
to water, maintaining an almost constant ester yield at by a relatively slow (∼24 h) water removal process that
the levels of water added. Unfortunately, neither Kus- employed large settling tanks as splitters. The use of
diana and Saka nor Canakci and Van Gerpen addressed gravity separation for water removal is lengthy and
the differences in the observed sensitivity to water for other methodologies have also been proposed for this
transesterification and esterification. task.13,30,31 However, in most cases gravity separation
In related work, Sridharan and Mathai noticed that in settling tanks is preferred due to economical reasons
the transesterification of small esters was retarded by arising from lower process utility costs.
the presence of spectator polar compounds.26 According The acid-catalyzed esterification follows a mechanistic
to their findings, the presence of polar compounds scheme similar to transesterification. Accordingly, in-
during acid-catalyzed alcoholysis reactions significantly stead of starting with a TG molecule, as in the trans-
reduced reaction rates. These authors attributed this esterification reaction (Figure 5), the starting molecule
retardant effect to the interference that polar com- is a FFA. Again, the key catalyst-FFA interaction is
pounds pose to the reaction by competing for hydrogen the protonation of the carboxylic moiety in the FFA. In
ions, hindering the availability of these ions for cataly- this case, the important factor influencing the reaction
sis. Water exerts similar effects in the transesterifica- rate at the molecular scale is the steric hindrance
tion of TGs. Considering the strong affinity that sulfuric inherent to both the carboxylic acid and the alcohol
Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5359

species.30 For example, in the esterification of palmitic


acid with either methanol or ethanol, ethanol was found
to be less efficient than methanol.9
Figure 7. Transesterification of β-ketoesters.
Heterogeneous Acid Catalysis
Homogeneous catalysts, although effective, lead to to be effective. In a related study, Waghoo et al. reported
serious contamination problems that make essential the on the transesterification of ethyl acetate with several
implementation of good separation and product purifi- alcohols over hydrous tin oxide to obtain larger esters.
cation protocols, which translate into higher production Linear and aromatic alcohols were tested in a temper-
costs. To be economically viable and to compete com- ature range of 170-210 °C. All reactions were com-
mercially with petroleum-based diesel fuel, processes for pletely selective for transesterification.34 In particular,
the synthesis of biodiesel need to involve continuous this catalyst presented an appreciable activity for reac-
processing in a flow system, have as few reaction steps tions involving n-butyl alcohol, n-octyl alcohol, and
as possible, limit the number of separation processes, benzyl alcohol.
and ideally use a robust heterogeneous (solid) catalyst. Amberlyst-15 has also been studied for transesteri-
The appropriate solid catalysts could be easily incorpo- fication reactions. However, mild reaction conditions are
rated into a packed bed continuous flow reactor, sim- necessary to avoid degradation of the catalyst. At a
plifying product separation and purification and reduc- relatively low temperature (60 °C), the conversion of
ing waste generation. In this section, three aspects of sunflower oil was reported to be only 0.7%, when
solid acid catalysis for biodiesel production will be carrying out the reaction at atmospheric pressure and
reviewed. The first section deals with the transesteri- a 6:1 methanol-to-oil initial molar ratio.35
fication by solid acid catalysts of esters with alcohols, On a related subject, Bronsted solid acids have also
the second topic relates to the solid acid-catalyzed been proposed for the transesterification of β-ketoesters
esterification of FFAs, while the third deals with solid to produce precursors for pheromones36 and other
acid-catalyzed etherification of glycerol. natural products (Figure 7).37 Among the catalysts used
1. Transesterification. Research on the direct trans- were Amberlyst-15,38 Envirocat EPZG,39 natural ka-
esterification of lipid feedstocks into biodiesel by solid olinite clay,39,40 B2O3/ZrO2,36 sulfated SnO2,41 and zeo-
acid catalysts has not been explored as one might expect lites.42 Even though β-ketoesters usually show a higher
given the amount and variety of catalytic materials now reactivity than simple esters for transesterification,
available. Actually, less than a handful of studies have under the right reaction conditions, a catalyst active for
explored the topic and without decisive conclusions the transesterification of β-ketoesters could be effective
about its potential. Even fundamental studies dealing in the transesterification of other types of esters as well.
with the reaction of model compounds of TGs on solid 2. Esterification. The esterification of carboxylic
acids are missing. Low expectations in terms of reaction acids by solid acid catalysts is important considering
rates and probable undesired side reactions have dis- that low-cost lipid feedstock contains high concentra-
couraged more significant research efforts into this tions of FFAs. Therefore, it is expected that a good solid
subject to date. acid catalyst must be able to carry out simultaneously
In one of the few published works dealing with the both esterification and transesterification. Since esteri-
transesterification of TG feedstocks using solid acid fication and transesterification share a common molec-
catalysts, Mittelbach et al. compared the activities of a ular pathway, evidence about catalyst reactivity for
series of layered aluminosilicates with sulfuric acid for esterification also provides evidence about transesteri-
the transesterification of rapeseed oil.32 These research- fication and vice versa. Thus, works cited in this section
ers used an initial molar ratio of 30:1 alcohol-to-oil and about esterification catalyzed by solid acids support the
5 wt % catalyst. Among the catalysts tested, sulfuric idea that heterogeneously acid-catalyzed transesterifi-
acid showed the highest activity. The solid catalysts cation of TGs should also be possible with these types
showed varied activities depending on reaction condi- of catalysts.
tions. The most active catalysts were activated by In addition, the reader should be aware that, although
sulfuric acid impregnation. For instance, activated the objectives of most works cited here have not been
montmorillonite KSF showed a 100% conversion after especially concerned with biodiesel synthesis, the fact
4 h of reaction at 220 °C and 52 bar. However, leaching that all carboxylic acids (in our specific case FFAs) share
of sulfate species compromised the reusability of this the same chemical functionality makes these results
clay. Thus, to maintain clay activity at constant values, relevant to biodiesel synthesis via esterification of FFAs
sulfuric acid re-impregnation had to be carried out after using solid acids.
each run. It is also likely that some degree of homoge- Currently, the esterification of fatty acids with alco-
neous catalysis was taking place due to sulfuric acid hols is commercially achieved using liquid catalysts,
leaching. such as sulfuric acid, hydrofluoric acid, and p-toluene-
Attempts to prepare other solid acid catalysts with sulfonic acid. The scientific literature contains a good
high activities for transesterification have been re- number of reports about the use of heterogeneous acid
ported, as well. In particular, Kaita et al. designed catalysts for esterification. For instance, esterification
aluminum phosphate catalysts with various metal-to- has been carried out using ion-exchange resins such as
phosphoric acid molar ratios (1:3-1:0.01) and used these Amberlyst-15 and Nafion with good results.43,44 In
materials for the transesterification of kernel oil with general, when using organic resins, the catalytic activity
methanol.33 According to the authors, durable and strongly depends on their swelling properties. Resin
thermostable catalysts were obtained with good reactiv- swelling capacity is fundamental since it controls sub-
ity and selectivity to methyl esters. However, the use strate accessibility to the acid sites and, therefore,
of these materials still needed high temperatures (200 affects its overall reactivity. Once swelled, the resin
°C) and high methanol-to-oil molar ratios (60:1) in order pores usually become macropores. This means that big
5360 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005

molecules with long hydrocarbon chains show no diffu- showed good activity at 110 °C (95% conversion of
sion limitations and can readily access the acid sites in 1-butanol). As expected, MCM-41-supported HPA was
the bulk.45 However, most ion-exchange resins are not more active than pure HPA. The enhanced activity could
stable at temperatures above 140 °C, which prohibits be ascribed to a high dispersion of the HPA on the
their application to reactions that require higher tem- MCM-41 internal surface, giving rise to a higher popu-
peratures. For this kind of application, inorganic acid lation of available acid sites than in pure HPA. How-
catalysts are generally more suitable. ever, the supported catalyst was considerably more
Among the different types of inorganic solids that hydrophilic than the original HPA. Water formation
have been used to produce esters, the most popular have caused HPA migration from the MCM-41 pores to the
been zeolites. Different characteristics make zeolites outer surface, facilitating the sintering of HPA species.
excellent catalysts for organic syntheses in general. For This was verified by measuring spent catalyst activities,
instance, zeolites can be synthesized with different which significantly decreased with catalyst reuse.
crystal structures, pore sizes, framework Si/Al ratios, Other composite catalysts, made using silica meso-
and proton-exchange levels. These characteristics per- porous materials modified with sulfonic groups, were
mit tailoring important catalytic properties, such as acid used in the pre-esterification of mixtures of FFAs and
strength. In zeolites, the acid strength can be adjusted soybean oil. Mbaraka et al. found that the activity of
such that it fits the reaction requirements.46 Too little these hybrid mesoporous silicas was highly dependent
acidity and the reaction may not proceed at a reasonable on their pore dimensions and probably on their hydro-
rate under the chosen conditions. Too high acidity and phobic character as well.50 Pore dimensions strongly
deactivation by coking may occur or undesirable byprod- affected reagent diffusion, especially problematic for
ucts might be formed requiring additional and expensive medium and small pore size materials. Acid strength
separation procedures. For esterification, however, op- also played an important role in these reactions. For
timum acidity has yet to be established. In addition, instance, catalysts prepared from a stronger acid pre-
zeolites provide the possibility to choose among different cursor containing benzenesulfonic acid groups were
pore structures and surface hydrophobicity, according more active than those containing only propylsulfonic
to the substrate’s size and polarity. For esterification, acid groups. Indeed, catalysts with a medium pore
the zeolite that is more suitable for a specific reaction diameter of 50 Å and benzenesulfonic acid groups
depends on the polarity and miscibility of the acid and showed activities comparable to sulfuric acid.
alcohol reagents. Nevertheless, when using zeolites, Recently, sulfated zirconia (SO4/ZrO2) has been shown
mass-transfer resistance becomes a critical issue due to have applicability for several acid-catalyzed reac-
to their microporous nature. In general, zeolite catalysis tions.51 For esterification, SO4/ZrO2 has shown some
of reactions using large molecules takes place on the promise as an active catalyst due to its high acid
external surface of zeolite crystals. Zeolites are active strength; however, SO4/ZrO2 deactivates due to sulfate
catalysts for the esterification of large carboxylic acids, leaching enhanced by hydrolysis.52 Sulfate groups can
but they catalyze the reaction rather slowly. Thus, only leach out as H2SO4 and HSO4-,46 which in turn can give
large-pore zeolites have been used with any success in rise to homogeneous acid catalysis, interfering with
fatty acid esterifications.47 But even in those successful measurements of the heterogeneous catalytic activity.
cases, the reaction always gives a variety of undesired To overcome the susceptibility of SO4/ZrO2 to water and
byproducts due to the high reaction temperature used. to improve its general characteristics, new preparations
In general, the catalytic activity of zeolites for the of SO4/ZrO2 have recently been proposed. For instance,
esterification of fatty acids increases with increasing Si/ Yadav and Murkute published a new route to preparing
Al ratio, indicating that reactivity is influenced by acid SO4/ZrO2 with higher sulfate loadings and resistant to
site strength as well as surface hydrophobicity. In sulfate leaching by hydrolysis.53 This catalyst was
conclusion, pore size, dimensionality of the channel prepared using a chlorosulfonic acid precursor dissolved
system (related to the diffusion of reagents and prod- in an organic solvent, instead of the conventional
ucts), and aluminum content of the zeolite framework sulfuric acid impregnation. The prepared SO4/ZrO2
strongly affect the zeolite’s catalytic activity for esteri- exhibited higher catalytic activity for esterification than
fication. the conventionally prepared SO4/ZrO2 and no sulfate
Related to zeolites, but with amorphous pore walls, leaching was observed. Additionally, the catalyst dem-
silica molecular sieves, such as MCM-41 mesoporous onstrated good retention of its activity for subsequent
materials, are generally not sufficiently acidic to cata- experiments.
lyze esterification reactions due to their pure silica Sulfated tin oxide (SO42-/SnO2), prepared from m-
structure. Introducing aluminum, zirconium, titanium, stannic acid, has shown activity superior to that of SO4/
or tin compounds into the silica matrix of these solids ZrO2 for the esterification of n-octanoic acid with
can significantly improve their acid properties. However, methanol at temperatures below 150 °C due to its
metal-doped materials behave more like weak acids and superior acid strength.54 However, a more widespread
can only be used for reactions that do not require a testing of SO42-/SnO2 has not been carried out due to
strong acid catalyst. For instance, the catalytic activity preparation inadequacies of the synthetic route initially
of Al-MCM-41 in the esterification of oleic acid with proposed to prepare this solid. New synthetic routes are
glycerol was found to be significantly lower than that making SO42-/SnO2 more accessible and additional
of zeolite beta with a similar Si/Al ratio.48 To enhance studies are expected with this promising material
the catalytic activity while keeping the benefits of large soon.55
pore diameters, strong acid species have been intro- The catalytic activities of hafnium and zirconium salts
duced to the pore interior of MCM-like solids. In have been investigated for the esterification of carboxy-
particular, MCM-41-supported heteropoly acids (HPAs) lic acids with primary and secondary alcohol in equimo-
have been used as catalysts in the gas-phase esterifi- lar ratios.56,57 In general, for esterification an equimolar
cation of acetic acid and 1-butanol.49 These catalysts ratio of reactants is preferred instead of excess alcohol.
Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5361

Equimolar ratios can reduce waste and simplify product To lower prices and to make biodiesel competitive
separation protocols, providing not only environmental with petroleum-based diesel, less-expensive feedstocks
but also economic benefits. In particular, hydrous such as waste greases have to be used in its production.
zirconium oxide showed good activity and selectivity for Indeed, the homogeneous acid-catalyzed process can
esterification. The low acid strength of this catalyst deal with high-FFA feedstocks, carrying out simulta-
helps to avoid undesirable side reactions, such as alcohol neously both esterification and transesterification, pro-
dehydration. In addition, the hydrated oxide was not vided that water is removed during processing. Other-
sensitive to water. Thus, esterification did not require wise, increasing water concentrations can seriously hurt
water-free conditions, an important characteristic given reaction conversions. An efficient way to remove water
that water is a reaction byproduct. is achieved by conducting feedstock transformation at
3. Glycerol Etherification. An issue some authors temperatures above 100 °C at moderate pressures and
have been recently exploring is the use of glycerol-based with continuous flow of an inert gas such as nitrogen.
additives to improve the flow properties of biodiesel at This way, with continuous alcohol feed, the in situ
low temperatures. For instance, glycerol ethers have separation of water by water/alcohol coevaporation helps
been prepared using an olefin such as isobutylene and to drive the reaction toward high conversions using
a solid acid catalyst like Amberlyst-15.58 Furthermore, relatively low-cost means.
Noureddini proposed a two-step process comprising the Solid acid catalysts have the capacity to replace strong
transesterification of soybean oil using a homogeneous liquid acids, eliminating corrosion problems and the
base catalyst and the separated etherification of glycerol environmental threat posed by them. However, research
byproduct with isobutylene or isoamylene using Am- dealing with the use of solid acid catalysts for biodiesel
berlyst-15.11 Afterward, the obtained mono-, di-, and synthesis has been limited due to pessimistic expecta-
tributyl ethers of glycerol were blended back with the tions about reaction rates and unfavorable side reac-
previously prepared biodiesel, resulting in a final fuel tions. In general, the use of solid catalysts to produce
with lower viscosity, cloud points, and pour points than biodiesel requires a better understanding of the factors
standard biodiesel, closely resembling a petroleum- that govern their reactivity. Thus far, it seems that an
based diesel. This approach not only makes good use of ideal solid acid catalyst should show some underlying
the reaction byproduct glycerol but potentially also could characteristics, such as an interconnected system of
increase the fuel yield by approximately 20 vol %. large pores, a moderate to high concentration of strong
However, a more appealing process would conduct the acid sites, and a hydrophobic surface. The former two
glycerol etherification in situ, without requiring glycerol requirements, the interconnected pore system of large
separation and subsequent back blending. The basic pores and a high population of strong acid sites, are very
problem with this is that using olefins and an alcohol, much evident. Large interconnected pores would mini-
such as methanol, in the presence of a strong acid mize diffusion problems of molecules having long alkyl
catalyst could result in methanol olefin etherification.59 chains, and strong acid sites are needed for the reaction
Another possible hurdle to in situ etherification of to proceed at an acceptable rate. The latter requirement,
glycerol could be the tendency of olefins to polymerize, having a hydrophobic surface, is essential to promote
giving rise to catalyst deactivation in the case of solid preferential adsorption of oily hydrophobic species on
acids and formation of gums when using liquid acids. the catalyst surface and to avoid possible deactivation
On the other hand, an olefin may not be needed to carry of catalytic sites by the strong adsorption of polar
out the etherification of glycerol. Ironically, the use of byproducts, such as glycerol and water. More research,
sulfuric acid as the catalyst at temperatures above 100 especially of a fundamental nature using model com-
°C has received some criticism due to the formation of pounds, is required in order to better delineate esteri-
byproducts such as ethers of glycerol and methanol. fication and transesterification reactions on solid acid
However, formation of such ethers could be taken as a catalysts. The results of such studies would aid in the
positive side reaction, rather than something that design of more active solid acid catalysts, certainly
should be avoided. Indeed, more detailed studies dealing impacting in a positive way biodiesel synthesis tech-
with the issue of glycerol ether additives to improve the nologies.
flow properties of biodiesel and their synthesis are
required. Acknowledgment
Conclusions
The authors acknowledge financial support for this
Despite all the promising characteristics related to work by the U. S. Department of Agriculture under
biodiesel, such as its low emission profile and the Grant 68-3A75-3-147 and by the Fats and Proteins
renewable character of its feedstocks, biodiesel is not Research Foundation.
yet the solution to any current fuel crisis. Several factors
play a role here. On one hand, current biodiesel produc-
Literature Cited
tion (about 2% of total diesel production in the United
States) is not sufficient to make an important impact (1) Kinast, J. A.; Tyson, K. S. Production of biodiesel from
on fuel markets. On the other hand, biodiesel production multiple feedstocks and properties of biodiesel and biodiesel/diesel
costs are still rather high compared to costs for produc- blends. Final report; NREL: Golden, CO, 2003.
ing petroleum-based diesel fuel (price for biodiesel not (2) Ma, F. R.; Hanna, M. A. Biodiesel production: a review.
counting tax breaks is about $2.02/gallon depending on Bioresour. Technol. 1999, 70, 1-15.
(3) Haas, M. J. The interplay between feedstock quality and
the feedstock used for its preparation. The price for
esterification technology in biodiesel production. Lipid Technol.
petroleum-based diesel is about $1.87/gallon).60,61 How- 2004, 16, 7-11.
ever, even if biodiesel is not the total solution to any (4) Talley, P. Biodiesel. Render 2004, (Sept).
energy crisis, it certainly is an important component of (5) Graboski, M. S.; McCormick, R. L. Combustion of fat and
a combined strategic approach to decrease our current vegetable oil derived fuels in diesel engines. Prog. Energy Combust.
dependence on fossil fuels.62 Sci. 1998, 24, 125-164.
5362 Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005

(6) Zappi, M.; Hernandez, R.; Sparks, D.; Horne, J.; Brough, (33) Kaita, J.; Mimura, T.; Fukuoda, N.; Hattori, Y. Catalysts
M.; Arora, S. M.; Motsenbocker, W. D. A review of the engineering for transesterification. U.S. Patent 6407269, June 18, 2002.
aspects of the biodiesel industry; Mississippi Biomass Council: (34) Waghoo, G.; Jayaram, R. V.; Joshi, M. V. Heterogeneous,
Jackson, MS, August 2003, 2003; p 71. catalytic conversions with hydrous SnO2. Synth. Commun. 1999,
(7) Canakci, M.; Van Gerpen, J. Biodiesel production via acid 29, 513-520.
catalysis. Trans. ASAE 1999, 42, 1203-1210. (35) Vicente, G.; Coteron, A.; Martinez, M.; Aracil, J. Applica-
(8) Canakci, M.; Van Gerpen, J. A pilot plant to produce tion of the factorial design of experiments and response surface
biodiesel from high free fatty acid feedstocks. Trans. ASAE 2003, methodology to optimize biodiesel production. Ind. Crops Products
46, 945-954. 1998, 8, 29-35.
(9) Canakci, M.; Van Gerpen, J. Biodiesel production from oils (36) Madje, B. R.; Patil, P. T.; Shindalkar, S. S.; Benjamin, S.
and fats with high free fatty acids. Trans. ASAE 2001, 44, 1429- B.; Shingare, M. S.; Dongare, M. K. Facile transesterification of
1436. beta-ketoesters under solvent-free condition using borate zirconia
(10) Tyson, K. S. Biodiesel, handling and use guidelines; solid acid catalyst. Catal. Commun. 2004, 5, 353-357.
NREL: Golden CO, September, 2001; p 22. (37) Balaji, B. S.; Chanda, B. M. Simple and high yielding
(11) Noureddini, H. System and process for producing biodiesel syntheses of beta-ketoesters catalysed by zeolites. Tetrahedron
fuel with reduced viscosity and a cloud point below thirty-two (32) 1998, 54, 13237-13252.
degrees fahrenheit. U.S. Patent 6174501, January 16, 2001. (38) Chavan, S. P.; Subbarao, Y. T.; Dantale, S. W.; Sivappa,
(12) Srivastava, A.; Prasad, R. Triglycerides-based diesel fuels. R. Transesterification of ketoesters using Amberlyst-15. Synth.
Renewable Sustainable Energy Rev. 2000, 4, 111-133. Commun. 2001, 31, 289-294.
(13) Zhang, Y.; Dube, M. A.; McLean, D. D.; Kates, M. Biodiesel (39) Bandgar, B. P.; Uppalla, L. S.; Sadavarte, V. S. Envirocat
production from waste cooking oil: 1. Process design and techno- EPZG and natural clay as efficient catalysts for transesterification
logical assessment. Bioresour. Technol. 2003, 89, 1-16. of beta-keto esters. Green Chem. 2001, 3, 39-41.
(14) Zhang, Y.; Dube, M. A.; McLean, D. D.; Kates, M. Biodiesel (40) Ponde, D. E.; Deshpande, V. H.; Bulbule, V. J.; Sudalai,
production from waste cooking oil: 2. Economic assessment and A.; Gajare, A. S. Selective catalytic transesterification, transthio-
sensitivity analysis. Bioresour. Technol. 2003, 90, 229-240. lesterification, and protection of carbonyl compounds over natural
(15) Liu, K. S. Preparation of Fatty-Acid Methyl Esters for Gas- kaolinitic clay. J. Org. Chem. 1998, 63, 1058-1063.
Chromatographic Analysis of Lipids in Biological-Materials. J. Am. (41) Chavan, S. P.; Zubaidha, P. K.; Dantale, S. W.; Kesha-
Oil Chem Soc. 1994, 71, 1179-1187. varaja, A.; Ramaswamy, A. V.; Ravindranathan, T. Use of solid
(16) Freedman, B.; Butterfield, R. O.; Pryde, E. H. Transes- superacid (Sulphated SnO2) as efficient catalyst in facile trans-
terification kinetics of soybean oil. J. Am. Oil Chem. Soc. 1986, esterification of ketoesters. Tetrahedron Lett. 1996, 37, 233-236.
63, 1375-1380. (42) Sasidharan, M.; Kumar, R. Transesterification over various
(17) Crabbe, E.; Nolasco-Hipolito, C.; Kobayashi, G.; Sonomoto, zeolites under liquid-phase conditions. J. Mol. Catal., A: Chem.
K.; Ishizaki, A. Biodiesel production from crude palm oil and 2004, 210, 93-98.
evaluation of butanol extraction and fuel properties. Process (43) Chen, X.; Xu, Z.; Okuhara, T. Liquid-phase esterification
Biochem. 2001, 37, 65-71. of acrylic acid with 1-butanol catalyzed by solid acid catalysts.
(18) Demirbas, A. Biodiesel fuels from vegetable oils via Appl. Catal., A 1999, 180, 261-269.
catalytic and noncatalytic supercritical alcohol transesterifications (44) Heidekum, A.; Harmer, M. A.; Hoelderich, W. F. Addition
and other methods: a survey. Energy Convers. Manage. 2003, 44, of carboxylic acids to cyclic olefins catalyzed by strong acidic ion-
2093-2109. exchange resins. J. Catal. 1999, 181, 217-222.
(19) Fukuda, H.; Kondo, A.; Noda, H. Biodiesel fuel production (45) Zhang, Z. Y.; Hidajat, K.; Ray, A. K. Determination of
by transesterification of oils. J. Biosci. Bioeng. 2001, 92, 405- adsorption and kinetic parameters for methyl tert-butyl ether
416. synthesis from tert-butyl alcohol and methanol. J. Catal. 2001,
(20) Nye, M. J.; Willianson, T. W.; Deshpande, S.; Schrader, J. 200, 209-221.
H.; Snively, W. H.; Yurkewich, T. P.; Frech, C. R. Conversion of (46) Corma, A.; Garcia, H. Organic reactions catalyzed over
used frying oil to diesel fuel by transesterification: preliminary solid acids. Catal. Today 1997, 38, 257-308.
test. J. Am. Oil Chem. Soc. 1983, 60, 1598-1601. (47) Corma, A.; Rodriguez, M.; Sanchez, N.; Aracil, J. Process
(21) Freedman, B.; Pryde, E. H.; T. L., M. Variables affecting for the selective production of monoesters of diols and triols using
the yields of fatty esters from transesterified vegetable oils. J. Am. zeolitic catalysts. WO9413617, 1994.
Oil Chem. Soc. 1984, 61, 1638-1643. (48) Perez-Pariente, J.; Diaz, I.; Mohino, F.; Sastre, E. Selective
(22) Khan, A. K. Research into biodiesel kinetics & catalyst synthesis of fatty monoglycerides by using functionalised meso-
development. University of Queensland, Brisbane, Queensland, porous catalysts. Appl. Catal., A 2003, 254, 173-188.
2002. (49) Verhoef, M. J.; Kooyman, P. J.; Peters, J. A.; van Bekkum,
(23) Formo, M. W. Ester reactions of fatty materials. J. Am. H. A study on the stability of MCM-41-supported heteropoly acids
Oil Chem Soc. 1954, 31, (11), 548-559. under liquid- and gas-phase esterification conditions. Microporous
(24) Keyes, D. B. Esterification processes and equipment. Ind. Mesoporous Mater. 1999, 27, 365-371.
Eng. Chem. 1932, 24 (10), 1096-1103. (50) Mbaraka, I. K.; Radu, D. R.; Lin, V. S. Y.; Shanks, B. H.
(25) Kusdiana, D.; Saka, S. Effects of water on biodiesel fuel Organosulfonic acid-functionalized mesoporous silicas for the
production by supercritical methanol treatment. Bioresour. Tech- esterification of fatty acid. J. Catal. 2003, 219, 329-336.
nol. 2004, 91, 289-295. (51) Yadav, G. D.; Nair, J. J. Sulfated zirconia and its modified
(26) Sridharan, R.; Mathai, I. M. Transesterification reactions. versions as promising catalysts for industrial processes. Mi-
Sci. Ind. Resour. 1974, 33, 178-187. croporous Mesoporous Mater. 1999, 33, 1-48.
(27) Boocock, D. G. B.; Konar, S. K.; Mao, V.; Sidi, H. Fast one- (52) Omota, F.; Dimian, A. C.; Bliek, A. Fatty acid esterification
phase oil-rich process for the preparation of vegetable oil methyl by reactive distillation: Part 2 - kinetics-based design for sul-
esters. Biomass Bioenerg. 1996, 11, 43-50. phated zirconia catalysts. Chem. Eng. Sci. 2003, 58, 3175-3185.
(28) Boocock, D. G. V. Single-phase process for production of (53) Yadav, G. D.; Murkute, A. D. Preparation of a novel
fatty acid methyl esters from mixtures of triglycerides and fatty catalyst UDCaT-5: enhancement in activity of acid-treated zir-
acids. U.S. Patent 6642399, November 4, 2003. conia-effect of treatment with chlorosulfonic acid vis-à-vis sulfuric
(29) Rived, F.; Canals, I.; Bosch, E.; Roses, M. Acidity in acid. J. Catal. 2004, 224, 218-223.
methanol-water. Anal. Chim. Acta 2001, 439 (2), 315-333. (54) Furuta, S.; Matsuhashi, H.; Arata, K. Catalytic action of
(30) Hoydonckx, H. E.; De Vos, D. E.; Chavan, S. A.; Jacobs, P. sulfated tin oxide for etherification and esterification in comparison
A. Esterification and transesterification of renewable chemicals. with sulfated zirconia. Appl. Catal., A 2004, 269, 187-191.
Top. Catal. 2004, 27, 83-96. (55) Matsuhashi, H.; Miyazaki, H.; Kawamura, Y.; Nakamura,
(31) Sonntag, V. O. V. Esterification and interesterification. J. H.; Arata, K. Preparation of a solid superacid of sulfated tin oxide
Am. Oil Chem Soc. 1979, 56, 751A∼754A. with acidity higher than that of sulfated zirconia and its applica-
(32) Mittelbach, M.; Silberholz, A.; Koncar, M. In Novel aspects tions to aldol condensation and benzoylation. Chem. Mater. 2001,
concerning acid-catalyzed alcoholysis of triglycerides, Oils-Fats- 13, 3038-3042.
Lipids 1995, Proceedings of the 21st World Congress of the (56) Wakasugi, K.; Misaki, T.; Yamada, K.; Tanabe, Y. Diphen-
International Society for Fats Research, The Hague, October 1995, ylammonium triflate (DPAT): efficient catalyst for esterification
1996; 1996; pp 497-499. of carboxylic acids and for transesterification of carboxylic esters
Ind. Eng. Chem. Res., Vol. 44, No. 14, 2005 5363

with nearly equimolar amounts of alcohols. Tetrahedron Lett. (64) Aranda, F.; Gomez-Alonso, S.; del Alamo, R. M. R.;
2000, 41, 5249-5252. Salvador, M. D.; Fregapane, G. Triglyceride, total and 2-position
(57) Ishihara, K.; Nakayama, M.; Ohara, S.; Yamamoto, H. fatty acid composition of Cornicabra virgin olive oil: Comparison
Direct ester condensation from a 1:1 mixture of carboxylic acids with other Spanish cultivars. Food Chem. 2004, 86, 485-492.
and alcohols catalyzed by hafnium(IV) or zirconium(IV) salts. (65) Engel, J. J.; Smith, J. W.; Unruh, J. A.; Goodband, R. D.;
Tetrahedron 2002, 58, 8179-8188. O’Quinn, P. R.; Tokach, M. D.; Nelssen, J. L. Effects of choice white
(58) Kesling, J. H. S.; Karas, L. J.; Liotta, J. F. J. Diesel Fuel grease or poultry fat on growth performance, carcass leanness,
U.S. Patent 5308365, May 3, 1994. and meat quality characteristics of growing-finishing pigs. J.
(59) Goodwin, J. G.; Natesakhawat, S.; Nikolopoulos, A. A.; Anim. Sci. 2001, 79, 1491-1501.
Kim, S. Y. Etherification on zeolites: MTBE synthesis. Catal. Rev.- (66) Paleari, M. A.; Moretti, V. M.; Bersani, C.; Beretta, G.;
Sci. Eng. 2002, 44, 287-320. Mentasti, T. Characterisation of a lard cured with spices and
(60) Goodwin, J. G.; Bruce, D. A.; Lotero, E.; Suwannakarn, aromatic herbs. Meat Sci. 2004, 67, 549-557.
K.; Lopez, D. E. Heterogeneous Catalyst Development for Biodiesel (67) Zheng, D. N.; Hanna, M. A. Preparation and properties of
Synthesis; Clemson University: Clemson, SC, October 2004; p 40. methyl esters of beef tallow. Bioresour. Technol. 1996, 57, 137-
(61) Energy Information Administration Gasoline and Diesel 142.
Fuel Update.
(62) Roadmap for biomass technologies in the United States; Received for review September 3, 2004
USDA, 2002. Revised manuscript received November 4, 2004
(63) Goering, C. E.; Schwab, A. W.; Dangherty, M. J.; Pryde, Accepted November 6, 2004
E. H.; Heakin, A. J. Fuel properties of eleven oils. Trans. ASAE
1982, 25, 1472-1483. IE049157G

Вам также может понравиться