Вы находитесь на странице: 1из 15

Renewable and Sustainable Energy Reviews 74 (2017) 387–401

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


journal homepage: www.elsevier.com/locate/rser

A review on recent developments and progress in the kinetics and MARK


deactivation of catalytic acetylation of glycerol—A byproduct of biodiesel

P.U. Okoye, A.Z. Abdullah, B.H. Hameed
School of Chemical Engineering, Engineering Campus, Universiti Sains Malaysia, 14300 Nibong Tebal, Penang, Malaysia

A R T I C L E I N F O A BS T RAC T

Keywords: The increased awareness and demand for green energy sources, such as biodiesel (BD), have impelled an
Acetylation increase in the production of its byproduct, glycerol, which is now considered an oversupplied commodity.
Biodiesel However, the excessive supply and physicochemical properties of glycerol render this compound as an attractive
Catalyst material for fine chemical synthesis, which can boost the BD market. Pathways for synthesis of fine chemicals
Deactivation
from glycerol have been proposed by a number of researchers by using different heterogeneous or homogeneous
Glycerol
Kinetics
catalysts, with the former being more environmentally benign. Glycerol acetylation pathway with acetic acid has
received increased research attention because of the vast potential applications of its products. Therefore,
studies were conducted to propose kinetic models for acetylation reaction and catalyst deactivation and
determine parameters influencing acetylation rate. This study also provides insights into assumptions of pseudo
first-order rate, PSSH, mass transfer, and diffusion limitation and discusses the future trends of catalytic
acetylation of glycerol.

1. Introduction compared with refined glycerol [16]. These challenges create a wide
cost margin between crude and refined glycerol, as reported by Quispe
Fossil fuels consumed by transport and energy industries are major et al. [17] in their study on global glycerol market and cost analysis. For
sources of CO2 and constitute over 60% of global pollution [1,2]. instance, in North Asia, refined glycerol is sold at USD 0.825–0.880/
Environmental degradation caused by CO2 released from human kg, whereas crude glycerol is sold at USD 0.044/kg; this cost difference
industrial activities into the atmosphere has raised awareness on the renders BD production via transesterification as unattractive.
use of renewable and eco-friendly sources of energy. Alternative energy Surplus glycerol commodity must be functionalized or upgrade to
sources, such as ethanol and biodiesel (BD), have been proposed to boost the BD market [18]. To avoid difficulties in glycerol refining,
avert or reduce the adverse impact of fossil fuels on the environment studies have focused on production of glycerol-free BD or its higher
[3–6]. BD has received increased research attention and commercia- molecular blend. Glycerol derivatives, such as monoglycerides or
lization either as pure BD (100%) or blended with conventional diesel glycerol carbonates, are produced when utilizing lipids and different
to reduce CO2 emissions by 52% compared with that from fossil diesel acyl acceptors, instead of methanol [19]. Acyl acceptors, such as ethyl
[7]. Triglycerides from biomass feedstocks, such as vegetable oil, acetate [20], methyl acetate [21], and dimethyl carbonate [14], produce
Jatropha seeds, algae, and animal fats, undergo transesterification three molecules of fatty acid methyl or ethyl esters and one molecule of
reaction in the presence of acid or basic catalysts to produce BD and glycerol carbonate or glycerol triacetate (triacetin). Glycerol triacetate
glycerol [8–14]. For every 100 kg of BD produced, 10 kg of glycerol is or glycerol carbonate blended with FAME exhibits physicochemical
synthesized as a by-product. Increased awareness of BD contribution in properties similar to those of biodiesel fuel and can be utilized in diesel
reducing CO2 emission has led to increased production and demand for engines without modification. Calero et al. [22] studied recent tech-
this type of diesel. According to the OECD report [15], the projected nologies of producing glycerol-free or glycerol-blended (Ecodiesel® and
production of BD in 2020 is estimated to be 42 billion liters, which Gliperol®) BD; they reported that 100% atom efficiency can be achieved
translates to 4.2 billion liters of glycerol produced in the market. In this because no glycerol is generated in situ, thereby avoiding purification
case, glycerol becomes a readily available commodity and combines steps and saving energy. However, glycerol-free BD production is
with contaminants, namely, methanol, soap, and water, which pose mostly performed on a laboratory scale because it is a new field that
difficulties in purification, to reduce the crude glycerol market value has yet to be fully commercialized.


Corresponding author.
E-mail address: chbassim@usm.my (B.H. Hameed).

http://dx.doi.org/10.1016/j.rser.2017.02.017
Received 5 November 2015; Received in revised form 8 December 2016; Accepted 3 February 2017
Available online 24 February 2017
1364-0321/ © 2017 Elsevier Ltd. All rights reserved.
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

Nomenclature Vi,1, Vi,2, Vi,3 Stoichiometric coefficient in the ith component at three-
step reactions (1−3)
CG, ΓG Concentration of glycerol N Number of experiments (3)
CA, ΓA Concentration of acetic acid CA0 Initial concentration of glycerol
2
CM , CW , CD, CT Concentrations of monoacetin (MAG), water, dia- CV Concentrations of vacant sites over the catalyst surface
cetin (DAG), and triacetin (TAG), respectively Cm Molar concentration of the total sites
rs Rate of surface reaction Cc Concentration of carbon on the catalyst surface (g/m2)
−dCG
dt
, −rG Rate of glycerol consumption A Fouling parameters that are likely to be a function of feed
−dCM
, YM , −rm,r1 Rate of MAG formation rate
dt
−dCD t Time
−rD, YD, , r2 Rate of DAG formation
dt C′ A , C′B Concentrations of reactant A and product B
−dCT
r3, −rT , YT , dt
Rate of TAG formation K ′ A , KB, kr Rates constants of consumption and formation of A and
−rAA Rate of acetic acid consumption B
KG, KA Equilibrium rate constants of glycerol and acetic acid, kr Reaction rate constant
respectively φA Fraction of the remaining active sites (activity function)
Kd Rate constant for desorption of MAG, DAG, and TAG K′d Sintering decay constant
KS Equilibrium rate constant for the surface reaction a (t ) Catalytic activity, time dependent
K1, K2, K3 Equilibrium rate constants for the forward reactions of Sa Catalyst active surface area
three consecutive steps Sa0 Initial catalyst active surface area
K ′2 , K ′4 , K ′6 Equilibrium rate constants for the reverse reaction of K Activated kinetic rate constant for sintering
three consecutive steps D Dispersion of metal
Γ Γ Γ
KEQ ,1, KEQ,2, KEQ,3 Thermodynamic equilibrium constants for three- Cp Concentration of poison in gas-phase feed mixture
step reactions, K =1–3 Cp0 Constant poison concentration
ΓEQ, i Composition of adsorbed phase for the ith component in n Rate decay order
the reaction

Glycerol is a highly functional material that requires defunctiona- Acetylation of glycerol has received increased attention because it
lization to synthesize more than a thousand fine chemicals. Pathways produces ester compounds, namely, monoacetin (MAG), diacetin
such as etherification [23], oligomerization [24], dehydration [25], (DAG), and triacetin (TAG), which can be applied in many processes,
dehydrogenation [26,27], glycerol oxidation [28,29], transesterification as summarized in Fig. 1 [49]. Studies on glycerol acetylation with acetic
[30], glycerol reforming [31–33], ketalization [34,35] and esterification acid (AcOH) showed that TAG is preferred to other esters, but its
[36–40] have been used to synthesize fine chemicals from glycerol. selectivity in the reaction is lower partly because the process is
Some of fine chemicals synthesized from glycerol are summarized in thermodynamically resistant; moreover, polar media are often formed
our previous study [41]. These chemicals can be directly applied as a in situ, and reaction parameters are poorly optimized [50]. Therefore,
food-grade material for poultry livestock, as a building block material knowledge of acetylation kinetics will facilitate understanding of the
substituted to petroleum-based propylene, and as a component for adsorption–desorption dynamics of glycerol acetylation for effective
manufacturing explosives, industrial foams and textiles and pharma- optimization of reaction parameters. Elucidating catalyst deactivation
ceutical products, such as cough syrup, toothpaste, and skin care kinetics is also vital for improving reactor operation and designing of
products. deactivation-resistant catalysts to obtain the optimum TAG yield.
Fine chemical synthesis using a catalyst can reduce energy require- Therefore, studies were conducted to propose kinetic models for
ments and speed up the reaction compared with high-temperature acetylation reaction and catalyst deactivation by poisoning from the
cracking of petroleum-based products without a catalyst. Based on their reaction by-product (water) on TAG yield and determine parameters
role in catalysis, catalysts have been categorized as heterogeneous, influencing acetylation rate. This study also provides insights into
homogeneous, and enzymes [42]. Homogeneous catalysts are used to assumptions of pseudo first-order rate, PSSH, mass transfer, and
catalyze glycerol upgrading processes but their applications are limited diffusion limitation and discusses the future trends of catalytic
because of their low recovery rate after the reaction, corrosive effects on acetylation of glycerol.
reactors and pipes (acid catalysts e.g., H2SO4), non-reusability, and
environmental challenges in terms of disposal [43,44]. As such, greener
and environmentally benign catalysts from microbial organisms, popu- 2. Heterogeneous catalysis for glycerol acetylation with
larly called enzymes, are used to synthesize fine chemicals from glycerol; acetic acid
enzymes exhibit a catalytic stability of up to 15 cycles of reuse, has no
corrosion effects on reactors, and can be easily separated from the Glycerol upgrading to obtain glycerol esters occurs through reaction
reaction media [45–47]. Enzymes can catalyze pathways involving crude with acylation agents, namely, AcOH or acetic anhydride (Ac2O) [57].
glycerol, but their use also presents serious limitations, including time- K-Montmorillonite and zeolite beta can catalyze glycerol acetylation
consuming, energy-intensive reactions (takes up to 24 h for the reaction to with Ac2O to yield 100% TAG under mild conditions (20 min at 60 °C).
reach equilibrium) and high cost, especially for large-scale production AcOH and Ac2O are also combined in a two-step acetylation process to
[48]. Furthermore, enzymes are easily deactivated by impurities in crude achieve 100% TAG yield [57]. Therefore, Ac2O, a simple carboxylic acid
glycerol or by-products of carboxylation during glycerol upgrading. To with two acyl groups bound to oxygen, is a thermodynamically
address the limitations of enzymes and homogeneous catalyst, current preferable alternative to AcOH in TAG production; the reaction
research focus shifts to the use of heterogeneous catalysts, which feature between Ac2O and glycerol proceeds exothermically and generates
reusability, easy separation from the reaction mixture, and tunability; negative Gibbs free energy, whereas glycerol acetylation with AcOH has
these catalyst can be modified through functionalization with active positive Gibbs free energy [50]. However, using Ac2O leads to release of
groups to obtain the product with high efficiency and less corrosive high heat of reaction, occurrence of violent reactions, and absence of
effects on reactors and pipes [13,43]. significant changes on thermodynamic parameters even at ambient

388
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

Fig. 1. Specific applications of glycerol acetylation with acetic acid and products [13,51–56].

temperature when the reaction temperature is varied [50]. Moreover, ratio, or a combination of these methods [49,57,60]. Acetylation with
Ac2O is a raw material for narcotic production and enlisted as contra- AcOH requires an oxygen electron lone pair of one hydroxyl group of
band in most countries, thereby prohibiting its use even for research glycerol to attack a carbonyl carbon of AcOH and consequently form
purposes. In this regard, the cost of Ac2O considerably increases, and MAG; thereafter, the formed MAG reacts with AcOH to form DAG and
its supply becomes relatively scarce; as such, Ac2O is impractical to use, TAG, releasing three molecules of water [58]. The schematic of the
in contrast to AcOH, which is readily available. Furthermore, synthesis proposed glycerol acetylation pathway with AcOH is presented in
of only TAG reduces glycerol acetyl esters analogues (MAG and DAG), Fig. 2. This acetylation reaction with AcOH differs from that of Ac2O,
that have vital industrial applications and can be obtained from low which proceeds in two possible pathways: Aac2 mechanism, which
cost glycerol [54,55]. The acetylation reaction of glycerol with AcOH or involves carbonyl oxygen lone pair protonation and nucleophilic attack
Ac2O is a three-step reaction, which includes MAG conversion to TAG in carbonyl to form intermediate quaternary carbon atom; and Aac1
and releases three molecules of water as a by-product [58,59]. These mechanism, in which protonation occurs at the oxygen atom of the
reactions are reversible and equilibrium controlled; hence, the reaction carbonyl group to form acylium ion [61].
is shifted to produce high amounts of TAG and DAG through several The rate at which the acetylation reaction proceeds depends on
approaches, such as continuous removal of in situ water molecules, various kinetic or thermodynamic parameters, which must be opti-
increase in reaction temperature, increase in AcOH/glycerol molar mized to achieve significant yield of higher esters, namely, DAG and

Fig. 2. Mechanism of glycerol acetylation with acetic acid [62].

389
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

TAG. These parameters include temperature, molar ratio, catalyst ⎛ moles of product formed ⎞
Selectivity(%) =⎜ ⎟X100
loading (i.e., number of available active sites), and speed of agitation, ⎝ moles of substrate consumed ⎠ (2)
which influence the reaction rate. The effects of these parameters are
studied, and their limits for specific reactions are determined. Several studies speculated that acetylation does not involve external
mass transfer limitation. This assumption can be validated using
appropriate kinetic and thermodynamic studies of the acetylation
process. As such, acetylation kinetics must be studied using process
3. Effects of influencing parameters on reaction rates data incorporated with influencing parameters to determine the effects
of these parameters on reaction rate and optimize them to obtain high
Reaction rates can be tailored to achieve the desired yield by yields of DAG and TAG. Table 1 presents catalysts used in various
altering reaction parameters, such as reducing the activation energy studies with their corresponding reaction conditions, selectivity of
barrier or promoting fast molecular collision to surmount the binding acetyl esters (MAG, DAG, and TAG), and glycerol conversion. The
energies of the reactants. For instance, temperatures ranging from combination of highly acidic condition, an AcOH/glycerol molar ratio
100 °C to 120 °C are ideal for glycerol acetylation with AcOH to attain of 6:1–9:1, and reaction temperatures between 80 and 115 °C results
efficient glycerol conversion and achieve the selectivity of MAG, DAG, in high DAG and TAG selectivity and glycerol conversion. Typically,
and TAG. In addition, glycerol/AcOH molar ratio is a key parameter to trial and error experiments are conducted to determine the effect of
obtain high ester yield [49,63,64]. Conversion of glycerol and selectiv- each reaction parameter on the reaction rate; in these experiments, all
ity of the reaction products can be calculated using Eqs. (1) and (2), parameters are kept constant while varying the parameter under
respectively. investigation. The results are then validated using kinetic rate laws or
other statistical modeling techniques.
⎛ initial mol − final mol ⎞
Conversion(%) = ⎜ ⎟ × 100
⎝ initial mol ⎠ (1)

Table 1
Heterogeneous catalyst applied in glycerol acetylation, reaction conditions, and selectivity.

Catalyst Acidity Reaction Parameters Selectivity (%) Ref.


(mmol/g)
Catalyst Loading Molar ratio Temp. Reaction time MAG DAG TAG Glycerol
(g) (AcOH: GL) (°C) (h)

Amberlyst-15 4.7d 5w% 6:1 105 10 0 12.3 83.9 100 [40]


Amberlyst-70 2.55d 5w% 6:1 105 10 0 7.5 87.6 100
STA/S11 0.274 5w% 6:1 105 10 1 55.5 35.8 100
STP/S11 0.1549 5w% 6:1 105 10 4.9 71.3 21.8 100
TPA3/MCM-41 0.627 0.15 6:1 100 6 25 60 15 87 [51]
TPA3/ZrO2 0.840 0.15 6:1 100 6 60 36 4 80
PMo3_NaUSY – 0.2 10.5:1 – 3 37 59 2 68 [54]
PW2_AC – 0.2 16:1 120 3 25 63 11 86 [56]
Amberlyst-15(dried) 4.7 2.65 9:1 110 5 7.8 47.7 44.5 97.1 [59]
Amberlyst-15(wet, moisture – 2.65 9:1 110 5 18.5 43.2 38.3 93.5
3.2%)
Blank (without catalyst) – – 9:1 110 4.20 84.7 13.8 1.5 73.5
SBAH-15(15)j 0.86 0.4 6:1 110 3 14 67 19 100 [63]
S-Mo(15) 0.69 0.4 6:1 110 3 38 53 9 74
AC-SA5 0.890 0.8 8:1 120 3 38 28 34 91 [64]
Amberlyst-15 4.9 0.64 6:1 80 8 21.1 63.8 15.1 100 [65]
Silica-alumina 3.2 0.64 6:1 80 8 88.5 11.2 0.3 71
HUSY 0.9 0.64 6:1 80 8 72.7 25.7 1.6 94
PrSO3H-SBA15 1.2 0.64 6:1 80 8 15.8 64.6 19.6 100
HPMo/SBA15 0.6 0.64 6:1 80 8 77.1 22.0 0.9 96
HPMo/Nb2O5 0.7 0.64 6:1 80 8 81.8 17.5 0.7 87
SCZ 4.1 0.64 6:1 80 8 84.7 14.8 0.5 81
SO3H-SBA15 0.8 0.64 6:1 80 8 11.1 61.9 27.0 100
SO3H-Cell 1.3 0.64 6:1 80 8 37.6 55.0 13.4 100
Amberlyst-36 5.4 0.25 1:8 105 10 70.3 4.5 – 95.6a [66]
Aamberlyst-15 4.7 0.25 1:8 105 10 70.3 2.5 – 95.3a
Dowex-2 4.8 0.25 1:8 105 10 80.8 5.1 – 95.2a
Dowex-4 – 0.25 1:8 105 10 71.6 4.2 – 94.8a
Dowex-8 – 0.25 1:8 105 10 72.9 4.7 – 94.7a
Amberlyst-15 0.0042 0.47b 3:1 110 0.5 31 54 13 97 [67]
K-10 0.0005 4.0b 3:1 110 0.5 44 49 5 96
Niobic Acid 0. 0003 6.3b 3:1 110 0.5 83 – – 30
HZSM-5 0.0012 1.6b 3:1 110 0.5 83 10 – 30
HUSY 0.0019 1.1b 3:1 110 0.5 79 14 – 14
Ar-SBA-15 1.15c 0.2 9:1 125 4 15 e 47 e 38 e 96e [68]
F-SBA-15 1.04c 0.2 9:1 125 4 14 e 50 e 36 e 90e
Pr-SBA-15 0.30c 0.2 9:1 125 4 17 e 44 e 39 e 80e

a=conversion of acetic acid.


b=catalytic loading to achieve 2 mmol of acid sites in each experiment.
c=meq/g (miliequilvalent/gram).
d=eq/kg (equivalent/kilogram).
e=the estimated values obtained from the bar plot.
j=modified with P123 and Brij S100 surfactant.

390
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

3.1. Effect of reaction temperature ing the rates of DAG and TAG formation [53,59,65].
Equilibrium molar concentrations are measured when a single
Glycerol esterification with AcOH to convert MAG into TAG is a reaction attains equilibrium. The concentrations of all molecules
highly endothermic reaction and requires large amount of heat to drive involved in the reaction are measured, and the ratio of product
the reaction. This phenomenon explains why MAG formation is concentration to reactant concentration is then calculated. AcOH/
dominant at low temperatures, whereas high concentrations of DAG glycerol molar ratio with excess AcOH affects the equilibrium time
and TAG are produced at high temperatures. Zhu et al. [49] determined and can delay the equilibrium product distribution of the ester
the effects of temperature on acetylation product distribution by using products; glycerol concentration exhibits major effects on ester dis-
HPA-supported zirconia catalyst. The results showed that an increase tribution and reaction time. Equilibrium constants are the same for a
in temperature from 60 °C to 120 °C increases glycerol conversion from given reaction (liquids and solids) and independent of initial concen-
54% to 100% and enhances selectivity of DAG and TAG compared with tration, which ideally determines the reaction rate; however, Zhou et al.
reduced selectivity of MAG. Testa et al. [69] and Patel and Singh [51] [59] reported that rate constants are independent of the initial
also reported a similar pattern in their studies using different solid concentrations, and that varying AcOH/glycerol molar ratios results
heterogeneous catalysts and reaction conditions. However, no signifi- in different K-values. This behavior is attributed to the nonlinear
cant changes were observed in the selectivity of DAG and TAG, and relationship or physical interactions that possibly occur at the mole-
glycerol conversion at temperature above 120 °C. This finding implied cular level. Melero et al. [68] observed through a 3D response surface
that equilibrium is rapidly attained because of fast reaction rates. methodology that glycerol/AcOH molar ratio is a predominant factor
Mufrodi and Sutijan [58] indicated that TAG selectivity decreases at that influences high TAG selectivity. The optimal molar ratio for DAG
reaction temperatures higher than 115 °C because of uncontrolled formation is 1:6; beyond which, selectivity declines because DAG is
evaporation of AcOH, which boils at 118 °C. In addition, the choice of converted into TAG.
catalyst is critical in temperature studies because heat can deactivate
thermally unstable catalysts. For instance, Amberlyst-15, a sulfonic 3.3. Effect of catalyst loading
acid resin, is thermally unstable when operated around 120 °C and
results in a sharp decrease in catalytic activity [57]. As such, the effect Although Arrhenius equation does not show the explicit relation-
of temperature on reaction rates can be determined by validating ship between catalyst and reaction rate, elementary chemistry shows
experimental runs of various temperature ranges (e.g., 60–120 °C) that chemical reactions proceed by forming charged ions at a high
with Arrhenius equation, as presented in Eq. (3). The equation shows energy level to overcome the activation barrier. High temperature input
the relationship between intrinsic temperature and equilibrium con- enhances molecular collision and bond dissociation and eventually
stants and thus can be used to obtain rate law parameters in constant surmounts the activation barrier, leading to the formation of a highly
volume batch reactors, creep rates, and other thermally induced energy-intensive reaction; as such, catalysts with active centers are
reactions. established to enhance the formation of the charged ions, thereby
reducing the activation energy barrier and speeding up reaction rates.
K = Ae−Ea / RT (3)
In glycerol acetylation with AcOH, the catalyst enhances the formation
−1 −1
where Ea (kJ/mol) is the activation energy, R (J mol K ) is the gas of acylium ion moiety, which undergoes electrophilic attack on the
constant or the Boltzmann constant obtained by multiplying with T hydroxyl group of glycerol. The active centers of the catalysts are
(°C), K is the reaction rate constant, and A(s−1 for the first-order usually anionic radicals of the so-called super acid groups (e.g., sulfuric
reaction) is the frequency factor, which varies with the rate order and or sulfonic acids, nobic acids, and phosphoric acids); the activity of
has the same unit as the rate constant. Based on Eq. (3), an increase in these radicals is directly related to their pore channel structure, nature
temperature or a decrease in activation energy (when catalyst is of the acid group, and acid strength [68].
applied) will result in increased rate constant, which is directly Functionalization using super acids reduces the specific surface
proportional to the rate of the reaction. RT is the average kinetic area of the catalyst support material and may likely block the catalyst
energy, and the exponential term, which is the ratio of Ea and RT , pore channel, thereby reducing catalytic activity; these acids also
implies that large ratios correspond to low rate constants and explains agglomerate and tend to be loosely bound on the catalyst surface
the negative sign. Hence, the rate constant K is the number of (depending on the synthesis technique), resulting in leaching of the
molecular collisions per second, the pre-exponential factor A is the active groups and undesired homogeneous catalysis [70]. The effects of
sum of all collisions, and e−Ea / RT is the probability that a collision will acidity on selectivity and conversion are discussed in Section 3.4.
result to a reaction. Although glycerol acetylation with AcOH can proceed without a catalyst
and generates more than 70% MAG yield, glycerol conversion is very
3.2. Effect of AcOH/glycerol molar ratio low and the desired products, namely, DAG and TAG, are not produced
[68]. Therefore, heterogeneous acid catalysts, which favor glycerol
The stoichiometric equation of glycerol/AcOH in the acetylation acetylation with AcOH, must be applied, considering that high glycerol
reaction with bimolecular consecutive steps suggests that three molar conversion ( > 90%) and higher acetyl esters are desirable.
AcOH is required to react with one molar glycerol to convert MAG into Catalysts such as AC-SA5 [64], sulfonic Amberlyst resins [59,66],
TAG. Previous studies showed that AcOH molar ratio above the HPA-functionalized zirconia [49], WO3 polypyrrole [71], and dodeca-
stoichiometric value is required to achieve high DAG and TAG molybdophosphoric acid-functionalized USY zeolite [54] have been
formation in the product distribution. Zhou et al. [59] investigated applied in acetylation reactions. Khayoon and Hameed [64] reported
the effects of glycerol/AcOH molar ratio (varying from 3:1–9:1, AcOH/ that an increase in catalyst loading from 0.2 to 0.8 g linearly increases
glycerol). The results showed that molar ratio as high as 9:1 could not the selectivity of DAG and TAG as a result of increased number of
result in 100% TAG selectivity because of thermodynamic limitations available active sites; however, saturation was attained at 0.8 g loading
caused by in situ aqua media, which hinder AcOH consumption (only without significant changes in selectivity recorded. Similar results were
74% of AcOH was consumed at the end of reaction). As such, a reactive reported by other studies [63,67,72].
distillation process should be used for continuously removing the water
by-product to positively shift the equilibrium toward fast and enhanced 3.4. Effect of surface acidity
TAG formation. Moreover, Dean–Stark apparatus can be used to
remove in situ water. Similarly, other studies demonstrated the Few studies have assessed the effect of catalyst surface acidity,
influence of glycerol/AcOH molar ratio on acetylation toward increas- surface strength, and type of acid site on the rate of glycerol acetylation

391
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

with AcOH. Existing studies indicate that acid density and acid for measuring if a heterogeneous catalytic reaction is limited by heat or
strength of heterogeneous solid catalysts are vital in tailoring acetyla- mass transfer by using the same catalyst dispersion; however, various
tion reaction to obtain high yields of DAG and TAG [73]. Surface loadings and comparison were performed under the same glycerol
acidity can be described by determining acid strength and number of conversion, where the observed similar turnover frequency indicates
acid centers. Typically, acid strength can be determined through that the reaction rates are not subjected to mass- or heat-transfer
neutralization titration for Bronsted acidity [74], temperature-pro- limitations [51,80]. Moreover, crushing the synthesized catalyst to fine
grammed desorption of ammonia (NH3-TPD) for total acidity, and powder (125–180 μm ) can enhance mass transport and eliminate
temperature-programmed desorption of acetonitrile (CH3CN-TPD) for intraparticle mass-transfer limitation [65]. Gelosa et al. [78] studied
Lewis acidity [75,76]. Inabe et al. [65] reported that acid strength the difference between transport time and reaction time by monitoring
affects the rate of acetylation because it increases intrinsic glycerol the concentration of the liquid phase during the reaction through gas
conversion turnover rate. However, comparison of different solid acids chromatography. When the characteristic reaction time is 1 h, the mass
showed that the proportional correlation between acid strength and transport process is about 10 min; this finding indicates that intra-
glycerol turnover rate is not valid; PrSO3H-SBA15 and Amberylst phase mass-transfer limitations can adversely affect the reaction rates.
catalysts with moderate acidity (4.9 and 1.2 mmol/g, respectively) To study the effect of mass-transfer limitations on bimolecular
exhibit superior selectivity of TAG. These findings are attributed to the reactions, models using the Wilke–Chang equation [78,81,82] or
spatial arrangement and configurations of surface acid moieties. The Scheibel relation [83] are used to determine the diffusivity of the
turnover rate (h−1) for glycerol conversion is the ratio of moles of reactants and products from the catalyst pores. In glycerol acetylation,
converted glycerol for each product species (MAG, DAG, and TAG) per the stoichiometric value of AcOH is in excess of glycerol; therefore, the
−1
unit catalyst amount and time (mol .gcat *h−1) to the acidity of the catalyst reaction is very likely to encounter external mass-transfer resistance,
−1
(mol .gcat ) [65]. which may occur from the bulk liquid phase to the catalyst surface, or
The preferred acid strength range is Bronsted acidity to obtain high an intraphase diffusion limitation, which can be determined using the
selectivity of combined DAG and TAG. Previous studies indicated that Glueckauf pore diffusion model [84]. Therefore, a similar case pre-
Bronsted acidity results in superior glycerol conversion and selectivity sented by Yadav et al. [85–87] can be adopted in the acetylation
of acetylation products compared with Lewis acids [62]. For instance, process for a solid–liquid reaction to evaluate external mass-transfer
the Bronsted acidity (0.405 mmol/g) of 20% (w/w) Cs-DTP/K-10 resistance, as presented in Eqs. (4)–(7).
catalyst can be ascribed to its higher DAG and TAG selectivity and
G + xAcOH ←⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯→ Products
glycerol conversion compared with UDCaT-4 and −5 catalysts with solid catalyst (4)
higher Lewis acidity (0.582 and 0.560, respectively) [62]. Konwar et al.
Overall, glycerol acetylation with AcOH is represented as a single-
[61] evaluated the effects of catalyst acidity for glycerol acetylation with
step reaction, which generates products such as MAG, DAG, and TAG.
acetic anhydride by using the same amount of active sites; they
where x is the stoichiometric ratio of AcOH (AcOH), which is ideally
reported that TAG selectivity increases with increasing surface acid
3 mol, and G is glycerol.
density, regardless of fixed catalyst amount. Furthermore, Khayoon
et al. [77] reported a higher selectivity for combined DAG and TAG RG = ∅SL − G aP*{[G0] − [GS ]} (5)
(34% and 55%, respectively) by using Y/SBA-3 catalyst compared with
Rate of mass transfer of glycerol G from bulk liquid to the catalyst
SBAH-15 and the sulfated activated carbon AC-SA5 under the same
surface.
conditions. The superior catalytic activity of Y/SBA-3 is ascribed to its
x
surface acidity, which decreases in the order of Y/SBA-3 > SBA-15 > RAA = ∅SL − ACOH ap{[AcOH0 ] − [AcOHs]x } (6)
AC-SA5 (1,342, 960, and 890 μmol/g, respectively). The derived rate
Rate of mass transfer of AcOH, RAA from bulk liquid to the catalyst
kinetic model (Table 2) incorporates the concentration of vacant sites
2 surface.
over the catalyst surface CV , which depends on surface acidity;
hence, reaction rate is influenced by catalyst acid strength. Moreover, =Rexp (7)
Khayoon and Hameed [63] used hybrid silica templated with molyb-
denum acid, SBAH15(15), and pure silica support in their study and Rate of reaction calculated from the experimental observations ,Rexp .
where [G0]and[GS ] are the glycerol concentrations in the bulk and
reported that catalytic activity depends on acid density as well as on
solid phases, respectively; [AcOH0 ]x and [AcOHs ]x are the AcOH concen-
surface acidity and type of support material. The hybrid SBAH15(15)
tration in the bulk and solid phases, respectively; ∅SL is the diffusion
with a surface acidity of 0.86 mmol/g exhibits superior catalytic activity
coefficient for glycerol and AcOH; and ap is the surface area per unit
compared with S–Mo (acidity=0.69 mmol/g).
volume.
Yadav and Krishnan [86] derived mathematical models on mechan-
isms controlling the rate of diffusion for different reaction scenarios,
3.5. Effect of agitation speed
where case 1,CG 0 / CAA0≠1 (molar ratio of glycerol/AcOH not equal to
unity), is applicable to the acetylation reaction. Therefore, the inverse
Acetylation of glycerol with AcOH and catalysts is a liquid–solid
of the observed rate must be higher than that of the glycerol and AcOH
slurry reaction that involves external mass transfer from the bulk liquid
rates to ascertain that external mass-transfer limitations are negligible:
phase to the catalyst, followed by intra- and inter-particle diffusion,
adsorption, reaction, and desorption of the products. These inter- and 1 1 1
≫ and
intraphase diffusion of the reactants encounters mass-transfer resis- Rexp ∅SL − Gap[G0] ∅SL − AcOHap [AcOH0 ]x (8)
tance, which results in slow or decreased yield. Various techniques can
be used to eliminate or determine if the acetylation reaction is mass Surface area per unit volume, aP , can be calculated from Eq. (9); the
transfer-controlled or -limited. A plausible method is by conducting mass transfer coefficients for glycerol and AcOH can be obtained from
experiments to determine interphase mass-transfer resistance by the Sherwood relation (Eq. (10)), assuming a low limiting value, which
varying the agitation speed, e.g., from 300 to 1200 rpm. The data is typically 2 [87].
obtained for each agitation speed are plotted for comparison, and 6w
ap =
overlapping of the plotted data indicates negligible mass-transfer ρp × dp (9)
limitations. Meanwhile, intraphase mass-transfer limitations can be
determined through blank experiment (nonreactive) using water, where w is the catalyst loading per unit liquid volume, ρp is the density
catalyst, and glycerol [78]. Koros and Nowak [79] proposed a technique of particle, and dp is the surface area of the particle.

392
Table 2
Kinetic models and rate mechanisms in glycerol acetylation with AcOH.

Catalyst Derived kinetic model Activation energy Rate order Rate Reaction Reaction Remarks Ref
(KJ/mole) mechanism regime conditions
P.U. Okoye et al.

H2SO4 −rm = k1CGCA − K ′2 CM CW − K3CM CA + K ′4 CDCW E1=3.94 1st order Power model – AcOH/Gly=3:1, The activation energy is similar [58]
−rD = k 3CM CA − K ′4 CDCW − K5CDCA + K ′6 CT CW E2=33.8012E3=51.7679 1.5 h, 100– to Cs-DTP/K−10 catalyst and
−rT = k5CDCA − K ′6 CT CW 120 °C, catalyst suggest that power model have
−rG = −k1CGCA − K ′2 CM CW loading= 2.5 w/w a trend in activation energy of
% three consecutive step reaction.
−rAA = −k1CGCA + K ′2 CM CW − K3CM CA + K ′4 CDCW − k5CDCA + K ′6 CT CW
However, the activation energy
values are not similar as
different reaction conditions
and catalysis was applied.

Amberlyst−15 CM K1K2 E1=57.26 Homogeneous LHHW Surface AcOH/Gly=3:1– Contrarily, the above [59]
YM = = [e−k 2t − e−k1t ]
CA0 K1 − K2 E2=31.87 1st order controlled 9:1, 6 h, 80– assertions, activation energy
CD K1K2 ⎡ e−k 2t − e−k 3t e−k1t − e−k 3t1 ⎤ E3=13.90 110 °C, catalyst increased in opposite direction
YD = = ⎢ K −K − ⎥ loading= 2.645 g in this study. Although no
CA0 K1 − K2 ⎣ 3 2 K3 − K1 ⎦
explanation of the possible
CT K1 K1K2 ⎡ e−k 2t − e−k 3t e−k1t − e−k 3t1 ⎤
YT = =1−e−k1t − [e−k 2t − e−k1t ] − ⎢ K −K − ⎥ causes was offered, however,
CA0 K1 − K2 K1 − K2 ⎣ 3 2 K3 − K1 ⎦
phase interactions between the
reactants and products, and
conditions of the reactions may
have likely favored TAG
formation and resulted in a
lower activation energy
compared to DAG and MAG.

393
Cs-DTP/K−10 −dCG
= K1CGCAW −K1CM CW W E1=45.35 1st order Power model Surface AcOH/Gly=9:1, Activation energy increases [62]
dt E2=57.28 controlled 4 h, 120–180 °C, from E3 > E2 > E1, implying
−dCM
= K1CGCAW −K1CM CW W −K2CM CAW +K2CDCW E3=80.5 catalyst loading= that increase in temperature
dt
−dCD 0.01 g/cm3 from 120 to 180 °C affects the
= K2CM CAW −K2CDCW W −K3CDCAW +K3CT CW W glycerol
dt rate. Also, the high activation
−dCT energy in E3 , indicates that
= K3CDCAW −K3CT CW W
dt
conversion from DAG to TAG is
most sensitive to temperature
changes and confirms the low
selectivity of TAG.
Y/SBA−3 2 21.54 PSSH LH Surface AcOH/Gly=4:1, Treated as one-step reaction, [77]
CV ⎛ Kd (CM + CD + CT ) ⎞
rs = ⎜K K C C − ⎟ controlled 2.5 h, 100– with an overall activation
Cm ⎝ G A G A KS ⎠
120 °C, catalyst energy which do not have any
loading= 0.4 g physical meaning as relates to
temperature or molar ratio.
2
However, CV term incorporates
mass transfer and intra-
particle limitations term used
to evaluate the reaction regime.

Amberlyst−15a Vi,1 ᴦ E1=62.8 1st order LH Mass AcOH/Gly=3.9, Also, the Amberlyst−15 [78]
r1 = K1ᴦ Gᴦ A[1−∏iN=1 (ᴦ EQ, i) /KEQ,1]
E2=17.57 transfer 0.5 h, 80 °C, activation energies are close in
Vi,2 ᴦ E3=18.41 controlled catalyst loading= values and trend to this study.
r2 = K2ᴦ Gᴦ A[1−∏iN=1 (ᴦ EQ, i) /KEQ,2]
N Vi,3 ᴦ
15 wt% TAG selectivity was favored in
r3 = K3ᴦ Gᴦ A[1−∏i =1 (ᴦ EQ, i) /KEQ,3] this case as reactive
chromatography was used to
remove in situ water formed in
the reaction media.
Renewable and Sustainable Energy Reviews 74 (2017) 387–401
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

shi = ∅SLi × id / Di LHHW is oversimplified from these assumptions, and its kinetic
p (10)
parameters derived from the fitted experimental data has no physical
where i is either AcOH or glycerol, D is the diffusivity value, sh is the meaning and cannot be used to theoretically distinguish MAG and DAG
Shearwood value, and ∅SLi is the diffusion coefficient for the ith compound. types (i.e., 1-monoacetin, 2-monoacetin, 1,3-diacetin, and 1,2-diacetin)
[94,95]. Therefore, density functional theory (DFT), molecular dy-
4. Kinetics of glycerol acetylation with acetic acid namics, and Monte Carlo (MC) simulations are adopted for modeling of
heterogeneous catalytic reactions to incorporate molecular data into
The kinetics of acetylation reaction with AcOH has received modeling of reactors [50,98]. DFT provides molecular information that
minimal research attention. Thus far, only a few kinetic studies on can be fed into MC simulations, where surface reaction rates can be
acetylation with heterogeneous catalysts were conducted computed as a function of fluid-phase partial pressure, adsorbate
[59,72,77,84,85,87–89]. Reactants are usually adsorbed in the catalyst structure, and temperature [98]. Furthermore, these methods can be
active centers, where protonation occurs. The rate of adsorption of applied to determine the most stable types of MAG and DAG. However,
these reactants is proportional to the fraction of the unoccupied active the evolved reaction rates are yet to be applied in models tractable to
sites on the catalyst, and the partial pressure of the component in the simulations of technical systems.
gas phase and species adsorption only occur at monolayer coverage. Most AcOH esterification reactions assume the surface reaction as
Acetylation reaction of glycerol and AcOH is a parallel consecutive the RDS [59,77,88,99], and that LH mechanism assumptions are often
three-step reaction from MAG to TAG; therefore, developing an based on pseudo-steady state hypothesis (PSSH) and imply the
intrinsic kinetic model of the process relies on observed experimental presence of relatively low-surface species activity; therefore, the
data. Kinetic analysis for acetylation process involves rate-determining reaction mechanism must be proposed based on small concentrations
steps (RDSs), which are expressed in terms of reactant and inter- of the active sites [100]. However, the active site concentrations of
mediate concentration; in RDS, the gas-phase concentrations of the heterogeneous catalysts are difficult to determine because they are
reactants and products are directly related to the concentrations of the mostly replaced by the total exposed surface area, which causes
intermediates, a phenomenon called adsorption isotherm. Changes in misleading results. Moreover, most heterogeneous catalysts, especially
concentration with respect to time, which are used to determine acetylation catalysts (zeolites, activated carbon, and functionalized
reaction rate constant and reaction order, are analyzed by applying zirconium), which require high Bronsted acidity and surface area, do
differential, regression, integral, or least square methods. These not fulfill the PSSH conditions (concentrations of the active sites must
analytical or statistical techniques provide insights into the dynamics be less than one tenth of the highest reactant concentration) [94].
of the underlying process. Therefore, if the PSSH criterion for LH kinetics is not satisfied, then the
Effect of reactant concentration on the reaction, diffusion limitations, experimental parameters can only be used in the region of the reported
and various mass transfer scenarios are evaluated using different experiment and its evolved kinetic parameters are reduced to only
mechanisms or approaches, such as simple power law [58,62,90,91], curve-fitting parameters; this phenomenon is clearly reported in Ref.
Eley–Rideal mechanism [92], neuclophilic substitution mechanism [89], [88], and additional details can be found in Ref. [100].
Langmuir–Hinshelwood (LH) mechanism or formalized Langmuir– Heterogeneous catalytic reactions are either kinetically, surface, or
Hinshelwood–Hougen–Watson (LHHW) model [48–52]; these mechan- intraparticle controlled. In particular, glycerol acetylation with AcOH
isms are usually adopted to describe the process kinetics for a given reaction regime is assumed to be surface controlled, and assumptions
reactor system. Pengyong and Zuhairi [89] derived and compared these on negligible intraparticle diffusion or mass transfer resistance, which
mechanisms (e.g., simple power law, neuclophilic substitution mechan- influences the reaction rates without validation, is likely to result in
ism, Eley–Rideal mechanism, and LH mechanism) to investigate the wrong kinetic representation of the system. For instance, in the kinetic
esterification reaction of glycerol with lauric acid. The derived models are studies of acetylation process by Zhou et al. [59], the resistance term of
based on different assumptions deemed theoretical for model simplifica- LHHW was ignored and AcOH was assumed to be completely
tion; as such, insights can be gained over the surface coverage of consumed because it is in excess during the reaction; hence, AcOH
intermediates, surface reactions, and reaction influencing parameter concentration is constant in the catalyst sites. Different K-values were
correlations [93]. The LH or LHHW is the most popular of these derived obtained at different initial AcOH/glycerol molar ratios, which can be
kinetic models for acetylation reaction [59,72,77,84,85,87–89]; this attributed to the physical-phase interaction of the reactants and
model optimally describes the process with high reliability and accuracy products and to the ignored mass transfer or intraparticle resistance
and involves equations that can be well-fitted to kinetic data [89]. In the effects. Therefore, the external mass transfer resistance in acetylation
acetylation reaction of glycerol with AcOH, assumptions are made to must be validated using the methods explained in a previous study
simplify the rate equation derivations, as summarized below [86]. Furthermore, kinetics study of acetylation reaction of glycerol
[59,72,77,84,85,87–89]: with AcOH has not considered intraparticle diffusion and but assumes
constant catalyst loading. Internal diffusion or Weber and Morris
1. Glycerol is assumed as a limiting reactant because AcOH is in excess model [101], which is vital in determining the adsorption rate for
and cannot be completely consumed during the reaction. liquid–solid systems, can be utilized to evaluate the influence of
2. Pseudo first-order reaction is proposed under isothermal condition. intraparticle diffusion resistance by varying catalyst loading at a fixed
3. The rate of forward reaction is faster than that of backward reaction, particle size and constant temperature with the Weisz–Prater criterion
thereby reducing the reaction rate constants to three for easy kinetic described in Ref. [102].
calculation. Acetylation from MAG to TAG is a reversible reaction and involves
4. The catalyst amount throughout the study is constant. six rate constants, three each for forward and backward reactions. The
5. The catalytic surface reaction is faster than that of the bulk liquid; rate equation derived from the three consecutive acetylation reaction
therefore, the reaction regime is surface controlled and not kineti- shown in Eqs. (11)–(13) is presented in Eqs. (14)–(19) [58].
cally controlled.
6. The adsorption and desorption of the reactants as well as the K1
G + AcOH ↔ MAG+H2O
products are very fast and at equilibrium. K2 (11)
7. The acetylation reaction is irreversible.
8. The adsorption of the main reactant is weak; thus, the resistance K3
term in the LHHW model can be ignored. MAG + AcOH ↔ DAG+H2O
9. No external mass-transfer limitation exists. K4 (12)

394
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

K5 deactivation, can be physical, chemical, or both and occurs in many


DAG + AcOH ↔ TAG+H2O categories, such as poisoning, coking or fouling, possible mechanical
K6 (13) failure because of attrition or erosion, thermal degradation, and phase
rG = −K1[G ][AcOH ] + K2[M ][W ] (14) transformation [104]. Catalyst deactivation is inevitable [105] but can
be slowed or avoided in some cases by removal or inhibiting the
rAA = −K1[G ][AcOH ] + K2[M ][W ] − K3[M ][AcOH ] + K4[D][W ] formation of identified deactivating agents. Catalyst deactivation in
glycerol acetylation with AcOH commonly occurs as a result of reduced
− K5[D][AcOH ] + K6[T ][W ] (15)
surface area caused by mechanical handling, attrition or functionaliza-
rM = K1[M ][AcOH ] − K2[M ][W ] − K3[M ][AcOH ] + K4[D][W ] (16) tion of the support, thermal degradation of Amberlyst resins, poison by
in situ formed water, leaching of active functional groups from the
rD = K3[M ][AcOH ] − K4[D][W ] − K5[D][AcOH ] − K6[T ][W ] (17) catalyst surface, and carbon deposition in the active centers of the
catalyst [49,57,69,71,73,106,107]. For instance, utilizing Amberlyst
rW = K1[G ][AcOH ] − K2[M ][W ] + K3[M ][AcOH ] − K4[D][W ] resins above the stated temperature limits can lead to decomposition of
+ K5[D][AcOH ] − K6[T ][W ] (18) the sulfonic functional groups, followed by loss of surface area and
reduced activity or complete loss of activity [108]. Furthermore, in situ
rT = K5[D][AcOH ] − K6[T ][W ] (19) water molecules, which are a reversible poison, are selectively adsorbed
where G is glycerol, AcOH is acetic acid, K1–K6 are rate constants of in the Amberlyst active sites, thereby reducing the number of active
forward and backward reactions, M is MAG, D is DAG, T is TAG, and W sites for adsorption and can lead to possible hydrolysis of the formed
is water. esters.
Although these equations can be solved to obtain the six rate Understanding the deactivation mechanism is vital for design of
constants for forward and reverse reactions, the formed assumptions deactivation-resistant catalysts, development of methods for regener-
stated that the rates of reverse reactions are less than the forward ating deactivated catalysts, and improvement of the operation of
reaction, thereby reducing the rate equation to three. Table 2 shows the industrial chemical reactors. The duration of catalyst life cycle for a
summary of derived kinetic models from different rate mechanisms given chemical process can be evaluated using reusability test on the
with remarks. In Ref. [58], a small reverse rate constant was obtained synthesized catalyst. Four cycles and above are usually used, and loss of
than the forward rate constant, which validates the assumption of activity is recorded. Moreover, leaching or homogeneity test using hot
neglecting the backward rates. In addition, conducted thermodynamics filtration method is conducted to determine if the functional groups in
studies on glycerol acetylation with AcOH at 105 °C indicated that the catalyst surface leach into the reaction media and constitute
when the Gibbs free energies of the three-step reaction are all positive, “undesired” homogeneous reactions. However, catalytic activity can
those of acetylation with acetic anhydride are negative; hence, glycerol be regenerated after each cycle by washing with solvents such as
acetylation with acetic anhydride is thermodynamically favored than methanol, water, and acetone. The catalyst is dried in the oven for
AcOH [50]. Moreover, the Gibbs energies of MAG and DAG reaction above 4 h and/or calcined thereafter. In-depth analysis on heteroge-
steps are relatively lower than those of TAG, explaining the usual slow neous catalyst deactivation and regeneration can be found in previous
selectivity or yield of TAG. Varying the reaction temperature to alter studies [104,105,109–112].
equilibrium rate constant and shift Gibbs free energy does not Thus far, no kinetics research has been conducted on acetylation
significantly change TAG selectivity. Gibbs free energy is energy change and catalyst deactivation, and only few studies describe heterogeneous
associated with substance formation at their most stable form (ele- deactivation mechanisms [104,112–116]. As such, developing suitable
ments of the substance) under standard conditions of 1 atm and 25 °C deactivation kinetics can provide insights into the process dynamics
[103]. Therefore, assuming that only glycerol is adsorbed on the and knowledge to optimize the design of deactivation-resistant cata-
catalyst surface (as AcOH is used in excess, its concentration is lysts and the operation of the catalytic process. Different mechanisms
assumed as constant during the reaction course) and considering the have been proposed by Butt and Petersen [114] to explain and model
external and internal mass-transfer limitations for the solid–liquid catalyst deactivation. Fogler [116] reported various empirical rate
reaction system, the rate at which glycerol is consumed is shown in Eq. decay orders and examples of such deactivation encountered in
(20). processes and provided with in-depth details of mathematical deriva-
tion of other deactivation mechanisms, namely, coking and sintering.
dXG
− = kmam(XG, b − XG, s ) The commonly encountered derived kinetic models for sintering,
dt (20) coking, and poisoning from the literature are presented in Table 3
where km, am , XG, b , XG, s are the mass transfer coefficient, the catalyst with corresponding remarks. Kinetics of catalyst deactivation varies
external surface area, the bulk concentrations, and the solid catalyst according to active-site blockage. The reported acetylation deactivation
phase, respectively. Assuming that the reaction rate is faster in the mechanisms exhibit different deactivation trends, ranging from sharp
catalyst surface than that in the bulk liquid phase and solving Eq. (20), decline in catalyst activity after the first cycle to gradual or stepwise
the linear equation form can be derived; this equation explains the decrease in catalyst activity up to four cycles [56,107]. Modeling
fractional conversion of glycerol and solves the mass transfer coeffi- acetylation rate of the reaction for different reactors by using hetero-
cient km from the slope of the plot (details of complete mathematical geneous catalyst should implicitly include deactivation terms for
derivation are presented in Ref. [77]). Eq. (21) represents glycerol efficient design of the reactors; the design should also incorporate
concentration as a function of the fractional conversion of glycerol θG adjustment of catalyst decay activity with time a(t ). Fogler [116]
and the initial glycerol concentration XG0 : divided reaction kinetics into separable and non-separable kinetics in
his study of chemical reactions and catalyst decay over time. Separable
XG = XG 0(1−θG ) (21)
kinetics is simpler to model because the rate law and activity can be
independently studied, whereas non-separable kinetics assumes non-
5. Kinetics of catalytic deactivation in glycerol acetylation ideal surface or modeling of several elementary reaction steps to
explain the deactivation mechanism. The activity of a deactivating
Loss of catalytic activity with time during the acetylation reaction of catalyst for considered separable case according to Fogler [116] can be
glycerol with AcOH is a major drawback that slows the industrial represented as a ratio of the kinetic rate of the reaction on the used
application of heterogeneous catalysts. This loss of activity, called catalyst at time t to that of the fresh catalyst at t=0 (Eq. (22)).

395
P.U. Okoye et al.

Table 3
Kinetic models for different deactivation mechanisms with remarks.

Deactivation Derived Kinetic model form Rate Mechanism Assumptions Remarks Ref.
Mechanism

Coking Cc=At n Empirical rate Equation Coke formation is independent of the hydrocarbon The validity of this assumption depends on the [125]
feed rate concentration of coke present on the catalyst surface.
Catalyst deactivated by coke can be regenerated by
calcination to burn off carbon.

Coking ⎛ C′ ⎞ ⎛ C′ ⎞ LHHW Single step reaction for A (reactant) to B (product). The second equation can be applied to consecutive [126,127]
krCtK ′A φA⎜C ′A − B ⎟ krCtK ′A φA⎜C ′A − B ⎟
⎝ K ⎠ ⎝ K ⎠ coke formation reaction is assumed to be parallel to step reactions and the both Equations can be solved
r= r=
1 + K ′A C ′A + KBC ′B 1 + K ′A C ′A + KBC ′B the main reaction path and the reaction rate is simultaneously accounting for both coke formation
Surface controlled. with respect to catalyst activity, temperature and
composition of feed mixture.

Sintering rd = K ′d a 2 2nd order rate expression – Current activity with respect to time is a function of [116]
Integral form of a(t ) is: with respect to current the decay rate constant and active surface species.
1 catalyst activity Because of the second order rate, expected nonlinear
a (t ) =

396
1 + k ′d t relationship is likely to predominate sintering
Amount of sintering with respect to active catalyst mechanism.
surface area is:
Sa 0
Sa =
1 + k ′d t
Sintering d (D / D 0 )
= − K (D /D0 − Deq /D0 )n Generalized power law – Ordinary power law expression are usually a functions [111,128]
dt expression, GPLE of time and cannot quantitatively calculate kinetic
parameters of deactivation, hence the use of GPLE.
Deq /D0term in GPLE, accounts for arbitrary limits
observed in time vs dispersion curve.
Poisoning-in feed df – Removal of poison from gas phase feed stream onto (1 − f ) is the fraction of unpoisoned site in the catalyst [116]
= Kd (1 − f )Cp
stream dt the catalyst sites is proportional to the fraction of and essentially is the activity of the catalyst with
unpoisoned active site and gas phase poison respect to time. This equation relates catalyst activity
concentration. with time and the concentration of poison in the gas
phase for petroleum feed contaminated by sulfur.
Poisoning- either by da n n Empirical decay rate law, For separable kinetics resulting in contacting a This can be applied to acetylation, by assuming [116]
− = rd = Kd′ Cpoa (t ) = Kda n
reactant or products dt with respect to n order of poison in the feed or product at constant constant poison (in situ water) concentration on the
decay. concentration Cp , with no spatial variation catalyst surface.
Renewable and Sustainable Energy Reviews 74 (2017) 387–401
P.U. Okoye et al.

Table 4
Heterogeneous catalyst and deactivation mechanism in acetylation.

Catalyst Reaction conditions No. of Mechanism of deactivation Report/Reasons for catalyst deactivation Ref.
reuse

HSiW/ZrO2, HPW/ZrO2 AcOH:Gly=10:1, 120 °C, t=4 h, C.L= 0.3 g 4 – Constant catalytic activity was observed with HSiW/ZrO2 catalyst for all the cycles [49]
while HPW/ZrO2 catalyst activity decreased slightly for cycle 1 and 2 then the glycerol
conversion remained constant. The stable performance of HSiW/ZrO2 catalyst was
ascribed to its hydrolytic stability in the in situ polar environment and strong bond
interaction between HSiW and ZrO2.

PW2_AC AcOH:Gly=1:16, 120 °C, t=3 h, C.L= 0.2 g 4 Slight leaching Leaching of the heteropoly acid used to functionalize activated carbon, AC and as [56]
such the activity of the catalyst decreased slightly but stabilized in the third cycle.

Amberlyst AcOH:Gly=0.3:0.1, 105 °C, t=4 h, 5 Possible thermal degradation and functional groups No significant deactivation recorded even when reaction time is extended to 12 h. [57]
C.L= 0.5 g; after 15 min add acetic volatilization and poisoning by in situ water. However, likely deactivation may occur via thermal degradation when reaction is
anhydride. performed at temperature limits of the catalyst (120 °C).

SBAH−15(15) AcOH:Gly= 6:1, 110 °C, t=3 h, C.L=0.38 4 Possible attrition resulting in reduced surface area Slight decrease of glycerol conversion and combined DAG and TAG was observed. [63]
and consequent acid sites decrease. Slight leaching. However, catalytic activity of the catalyst stabilized for the 3 and 4 cycle indicating
that active sites cannot be leached out. The stabilization in 3 and 4 cycle is attributed
to the restoration of the kegging structure on exposure of beta-MoO3 species to the in

397
situ byproduct water environment.

Amberlyst−36Dowex−2 AcOH:Gly=1:8, 105 °C, C.L= 0.25 g/40 ml 4 poisoning Recovered catalyst via centrifuge can be directly reused without any treatment and [66]
glycerol drying for four cycles without significant loss in activity, however, as the number of
water molecule occupy the resin active sites (selective adsorption because of their
high affinity to water than glycerol), pretreatment is desirable in order to prevent
slowed catalyst activity or hydrolysis of MAG, DAG, TAG.

SAS, SSBA, Amberlyst−15 AcOH:Gly= 3:1, 105 °C, t=1 h, C.L=50 mg 5 Leaching of active sites and poisoning SAS catalytic activity declined significantly due to leaching of the sulfonic [69]
functionalities from the catalyst. Amberlyst−15, was stable up to 3rd cycle then
declined at the end of the fifth cycle. SSBA catalytic activity was stable with no
significant decrease in activity and no leaching detected.

3%Y/SBA−3 AcOH:Gly= 4:1, 110 °C, t=2.5 h, C.L= 0.2 g 4 Poisoning (adsorbed reaction products) Catalytic activity decreased after 3 run from 100% to 80% for glycerol conversion and [77]
55–50% for TAG selectivity. The decrease was ascribed to mass transfer limitations
and reaction conditions.

SZ−1 AcOH:Gly=100g/l: 40 ml, 54.85 °C, 4 Leaching of active group Although the catalyst presents better mechanical resistance compared with other [65]
t=24 h, C.L=250 mg/40 ml of glycerol tested catalyst in their study, however, it deactivates linearly down during successive
cycle and was completely deactivated at the 4th cycle, due to leaching of sulfur groups
used in functionalizing the zirconia.

Note. C.L=catalyst loading, t=reaction time in hours (h) or minutes (min), Gly=glycerol.
Renewable and Sustainable Energy Reviews 74 (2017) 387–401
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

−rt
a (t ) =
−r(t =0) (22)

Szepe and Levenspiel [117] categorized the separability of reaction


rates as time-dependent kinetic rates and activity-dependent kinetics,
which is time-independent. Hence, Eq. (23), which is the ratio of the
number of available active sites on the fresh catalyst (non-deactivated
catalyst),N0 , to the number of active sties at time t of deactivation ,Nt ,
can be used to represent the fraction of the active sites on the catalyst
surface at time t.
Nt
γ=
N0 (23)

By combining Eqs. (22) and (23) for enhanced modeling, Butt and
Petersen [114] extended the LHHW mechanism to explain the
dynamics of systems with changing activity by using the dehydrogena-
tion reaction of methyl-cyclohexane to toluene with coke formation;
they considered surface reaction as the rate-limiting step and were able
to combine γand a(t ). Further details on the deactivation and its kinetics
for heterogeneous catalysts using various reactors are presented in
previous studies [98,111,113,118,119]. As shown in Table 4, some Fig. 3. Decline in catalytic activity for various types of poisoning with respect to fraction
catalysts during glycerol acetylation with AcOH deactivates either by of poisoned sites. (a) All types of poisoning, ∅ (Thiele modulus)≪2; (b) Uniform
Poisoning, ∅ ≫ 2; (c) core poisoning, ∅ = 3; (d) core poisoning, ∅ = 5; (e) core
poisoning, leaching of active sites, or thermal degradation. Hence,
poisoning, ∅ ≫ 5; (f) pore mouth poisoning, ∅ = 10; (g) pore mouth poisoning,
catalyst decay can be quantified as a decrease in the number of ∅ = 100 [114].
available active functional sites with respect to the initially available
active sites (fresh catalyst) because the activity of acid catalysts used in acetylation from the catalyst surface.
acetylation is related to the active functional groups on the catalyst where M, D, T, W, and S are MAG, DAG, TAG, water, and solid
surface. catalyst, respectively.
Poisoning mechanism can be determined from various models for
heterogeneous porous catalyst, considering the surface-controlled rM = kdXM − kdXM XV (25)
reaction assumed for determination of kinetic rates of glycerol acetyla-
rD = kdXD − kdXDXV (26)
tion with AcOH [104,113–116]. In situ water as poison deactivates
catalysts (e.g., zeolites) for aldol condensation reaction and has been rT = kdXT − kdXT XV (27)
kinetically modeled for enhanced reactor and catalyst design
[120,121]; glycerol acetylation with AcOH suffers similar challenges. rW = kdXW − kdXW XV (28)
Forzatti and Lietti [104] reported three limiting pore-poisoning
models, which are categorized as uniform, pore-mouth, and core where rM , rD, rT , rW are the desorption rates of MAG, DAG, and TAG, and
poisonings. The poison distribution on the catalyst surface for the water, respectively, from the catalyst. Kd is the desorption rate
three poisoning models was drawn from Wheeler's work [122]. For all constant; and XM , XD , XT , XW ,XV are the concentrations of desorbed
types of catalytic poisoning, decrease in catalytic activity with respect to MAG, DAG, TAG, water, and vacant site over the catalyst surface,
fraction of poisoned sites is presented in Fig. 3. In the presence of respectively. Therefore, the assumed surface-controlled rate law from
nonselective poison, pore-mouth and uniform poisonings are linearly Ref. [77] can be modified to consider time-dependent catalyst activity,
related to the amount of adsorbed poison under kinetic regime (Fig. 3 as presented in Eq. (29), for modeling the deactivation of glycerol
curve a). However, a different case exists for internal diffusional acetylation.
regime, where pore-mouth poisoning results in a sharp decline in XV2 n ⎛ k (X + XD + XT + XW ) ⎞
catalytic activity compared with that in the kinetic regime case (Fig. 3, rS = a (t )⎜kGkAAXGXAA − d M ⎟
Xm ⎝ ks ⎠ (29)
curves f and g). Under diffusion-controlled regime (only at ideal cases),
decrease in catalytic activity is slower than the chemical regime because where n is the order of catalyst decay, XW is the concentration of water
the reactants tend to utilize more surface areas than what should have formed as acetylation reaction byproduct, and a n(t ) is the fraction of
been used; this phenomenon results in less decline of rapid catalytic active sites available for adsorption. For separable kinetics rate law that
activity with the adsorbed amount (Fig. 3, curves b and c) [104]. These results from poisoning at constant poison concentration, assuming the
poisons can be selective or nonselective and result in eroding (through first-order rate decay to be at n=1, the activity of any given catalyst can
hydrolysis) or masking of active sites, respectively (Fig. 4). be represented as Eq. (30):
An appropriate heterogeneous catalyst design can enhance resis-
a = e−kdt (30)
tance to poisoning and is therefore needed because some poisons occur
on the outer surface of the catalyst. Using shape-selective catalyst with
non-uniform active surface distribution may theoretically favor the 6. Conclusion and future trends
desired reaction with respect to poisoning [123,124]. Furthermore, in
situ water that acts as poisons can be removed in real-time during Based on studies on glycerol acetylation with AcOH, proper
reaction through reactor modification by using the Dean stark appa- representation of reaction rates by using kinetic models facilitates
ratus. Therefore, considering poisoning from in situ water in glycerol understanding of the dynamic process and enhances robust reactor and
acetylation, consecutive steps derived in [77] (Eqs. (24)–(27)) include catalyst design. Heterogeneous catalyst can also be used to tailor the
a fourth equation, which is represented as Eq. (28): rate of acetylation reaction to the desired yield and in successive cycles.
(24) Reaction parameters such as temperature, catalyst loading, and molar
M . S + D. S + T . S + W . S → M + D + T + W +4S
ratio evidently influence the reaction rate of acetylation. In particular,
Eq. (24) is the general equation for product desorption of glycerol temperature and molar ratio of glycerol/AcOH influence the selectivity

398
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

Fig. 4. (a) Non-selective poisoning of active sites and (b) selective poisoning of active sites with time.

of higher acetyl esters (DAG and TAG). The LHHW kinetic model is emissions from fuel combustion; 2015.
[2] Charles C, Gerasimchuk I, Bridle R, Moerenhout T, Asmelash E, Laan T. Biofuels—
popularly and widely applied in acetylation process. However, this At what cost? A review of costs and benefits of EU biofuel policies. Manitoba,
model is oversimplified, resulting in lapses often ignored by research- Canada; 2013.
ers. Moreover, phase interactions or mass-transfer resistance (from [3] Serrano-Ruiz JC, Luque R, Sepúlveda-Escribano A. Transformations of biomass-
derived platform molecules: from high added-value chemicals to fuels via aqueous-
bulk liquid to catalyst surface) may likely exist because of excess AcOH, phase processing. Chem Soc Rev 2011;40:5266.
as evidenced by different rate constants obtained at different initial [4] Leung DYC, Wu X, Leung MKH. A review on biodiesel production using catalyzed
glycerol/AcOH molar ratios. Therefore, the assumption of no resistance transesterification. Appl Energy 2010;87:1083–95.
[5] Xiu S, Shahbazi A. Bio-oil production and upgrading research: a review. Renew
to mass transfer or intraparticle diffusion limitation must be validated. Sustain Energy Rev 2012;16:4406–14.
Most heterogeneous catalysts with high Bronsted acidity and surface [6] Pierpaolo C, Geoff M, Hiroyuki K, François C, Abbas G, Fulton L. Production costs
area are used in acetylation to validate the PSSH assumption. The of alternative transport fuels. France: Paris Cedex; 2013.
[7] Verbeek R, Smokers RTM, Kadijk G, Hensema A, Passier GLM, Rabe ELM et al.
validation process forms a basis for the LHHW model derivation and
Impact of biofuels on air pollutant emissions from road vehicles. Netherlands;
implies that the derived kinetic models are peculiar to the experiments 2008.
and conditions and cannot be applied outside the reaction zone. [8] Bart JCJ, Palmeri N, Cavallaro S. 3 - Oleochemical sources: basic science,
Finally, catalyst deactivation poses serious limitation to industrial processing and applications of oils. In: Bart JCJ, Palmeri N, Cavallaro S, editors.
Biodiesel sci technol. Woodhead Publishing; 2010. p. 62–113.
applications. Therefore, proper knowledge on mechanisms and/or [9] Bart JCJ, Palmeri N, Cavallaro S. 4 – Vegetable oil formulations for utilisation as
kinetics of deactivation is required to improve the designs of reactors biofuels. In: Bart JCJ, Palmeri N, Cavallaro S, editors. Biodiesel sci technol.
and deactivation-resistant catalysts. Shape-selective catalyst is recom- Woodhead Publishing; 2010. p. 114–29.
[10] Bart JCJ, Palmeri N, Cavallaro S. 7 – Transesterification processes for biodiesel
mended because of its resistance to certain molecular-shaped poisons production from oils and fats. In: Bart JCJ, Palmeri N, Cavallaro S, editors.
generated in situ. Biodiesel Sci Technol. Woodhead Publishing; 2010. p. 285–321.
[11] Bart JCJ, Palmeri N, Cavallaro S. 8 – Biodiesel catalysis. In: Bart JCJ, Palmeri N,
Cavallaro S, editors. Biodiesel Sci Technol. Woodhead Publishing; 2010. p.
Acknowledgement 322–85.
[12] Bart JCJ, Palmeri N, Cavallaro S. 10 – Biocatalytic production of biodiesel. In:
The first author acknowledges the Universiti Sains Malaysia, USM Bart JCJ, Palmeri N, Cavallaro S, editors. Biodiesel Sci Technol. Woodhead
Publishing; 2010. p. 434–61.
for their financial support under the USM fellowship scheme. The
[13] Bagheri S, Julkapli NM, Yehye WA. Catalytic conversion of biodiesel derived raw
authors acknowledge the financial support provided by the Ministry of glycerol to value added products. Renew Sustain Energy Rev 2015;41:113–27.
Education, Malaysia under the Transdisciplinary Research Grant [14] Ilham Z, Saka S. Two-step supercritical dimethyl carbonate method for biodiesel
production from Jatropha curcas oil. Bioresour Technol 2010;101:2735–40.
Scheme (TRGS) Phase2/2014 (203/PJKIMIA/6762002).
[15] Oecd, Fao. Biofuels. Agric Outlook 2011–2020; 2011. p. 77–93.
[16] Ardi MS, Aroua MK, Hashim NA. Progress, prospect and challenges in glycerol
References purification process: a review. Renew Sustain Energy Rev J 2015;42:1164–73.
[17] Quispe CAG, Coronado CJR, Carvalho JA. Jr., Glycerol: production, consumption,
prices, characterization and new trends in combustion. Renew Sustain Energy Rev
[1] International Energy Agency. Key trends in CO2 emissions excerpt from: CO2 2013;27:475–93.

399
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

[18] Kong PS, Kheireddine M, Mohd W, Wan A. Conversion of crude and pure glycerol [49] Zhu S, Zhu Y, Gao X, Mo T, Zhu Y, Li Y. Production of bioadditives from glycerol
into derivatives: a feasibility evaluation. Renew Sustain Energy Rev esterification over zirconia supported heteropolyacids. Bioresour Technol
2016;63:533–55. 2013;130:45–51.
[19] Luna C, Verdugo C, Sancho ED, Luna D, Calero J, Posadillo A, et al. Production of [50] Liao Xiaoyuan, Zhu Yulei, Wang Sheng-Guang, Hongmei Chen YL. Theoretical
a biodiesel-like biofuel without glycerol generation, by using Novozym 435, an elucidation of acetylating glycerol with acetic acid and acetic anhydride. Appl Catal
immobilized Candida antarctica lipase. Bioresour Bioprocess 2014;1:11. B Environ 2010;94:64–70.
[20] Kim Sok-Joong, Jung Sang-Min, Park Yong-Cheol, Park K. Lipase catalyzed [51] Patel A, Singh S. A green and sustainable approach for esterification of glycerol
transesterification of soybean oil using ethyl acetate, an alternative acyl acceptor. using 12-tungstophosphoric acid anchored to different supports: kinetics and
Biotechnol Bioprocess Eng 2007;12:441–5. effect of support. Fuel 2014;118:358–64.
[21] Tan KT, Lee KT, Mohamed AR. A glycerol-free process to produce biodiesel by [52] Rao PV, Rao BVA. Effect of adding Triacetin additive with Coconut oil methyl ester
supercritical methyl acetate technology: an optimization study via response (COME) in performance and emission characteristics of DI diesel engine; 2011. p.
surface methodology. Bioresour Technol 2010;101:965–9. 1.
[22] Calero J, Luna D, Sancho ED, Luna C, Bautista FM, Romero AA, et al. An overview [53] Mufrodi Z, Rochmadi R, Sutijan S, Budiman A. Synthesis acetylation of glycerol
on glycerol-free processes for the production of renewable liquid biofuels, using batch reactor and continuous reactive distillation column. Eng J
applicable in diesel engines. Renew Sustain Energy Rev 2015;42:1437–52. 2014;18:29–40.
[23] Clacens JM, Pouilloux Y, Barrault J. Selective etherification of glycerol to [54] Ferreira P, Fonseca IM, Ramos aM, Vital J, Castanheiro JE. Esterification of
polyglycerols over impregnated basic MCM-41 type mesoporous catalysts. Appl glycerol with acetic acid over dodecamolybdophosphoric acid encaged in USY
Catal A Gen 2002;227:181–90. zeolite. Catal Commun 2009;10:481–4.
[24] Martin A, Richter M. Oligomerization of glycerol – a critical review. Eur J Lipid Sci [55] Farinha J, Caiado M, Castanheiro JE. Valorisation of glycerol into biofuel additives
Technol 2011;113:100–17. over heterogeneous catalysts. A. Méndez-. Lisbon, Portugal: FORMATEX; 2013.
[25] Katryniok B, Paul S, Dumeignil F. Recent developments in the field of catalytic [56] Ferreira P, Fonseca IM, Ramos AM, Vital J, Castanheiro JE. Acetylation of glycerol
dehydration of glycerol to acrolein. ACS Catal 2013;3:1819–34. over heteropolyacids supported on activated carbon. Catal Commun
[26] Simakova O a, Davis RJ, Murzin DY. Biomass processing over gold catalysts. In: 2011;12:573–6.
Springer , editor. . SpringerBriefs Mol Sci, VIII. Springer; 2013. p. 11–33. [57] Liao X, Zhu Y, Wang SG, Li Y. Producing triacetylglycerol with glycerol by two
[27] Crotti C, Kaspar J, Farnetti E. Dehydrogenation of glycerol to dihydroxyacetone steps: esterification and acetylation. Fuel Process Technol 2009;90:988–93.
catalyzed by iridium complexes with P-N ligands. Green Chem [58] Mufrodi Z, Rochmadi S, Budiman A. Chemical kinetics for synthesis of triacetin
2010;12:1295–300. from biodiesel byproduct. Int J Chem 2012;4:101–7.
[28] Hejna A, Kosmela P, Formela K, Piszczyk Ł, Haponiuk JT. Potential applications [59] Zhou L, Nguyen TH, Adesina A a. The acetylation of glycerol over amberlyst-15:
of crude glycerol in polymer technology – current state and perspectives. Renew kinetic and product distribution. Fuel Process Technol 2012;104:310–8.
Sustain Energy Rev J 2016;66:449–75. [60] Zhou L, Al-Zaini E, Adesina A a. Catalytic characteristics and parameters
[29] Zhang C, Wang T, Liu X, Ding Y. Cu-promoted Pt/activated carbon catalyst for optimization of the glycerol acetylation over solid acid catalysts. Fuel
glycerol oxidation to lactic acid. J Mol Catal A Chem 2016;424:91–7. 2013;103:617–25.
[30] Malyaadri M, Jagadeeswaraiah K, Sai Prasad PS, Lingaiah N. Synthesis of glycerol [61] Konwar LJ, Mäki-Arvela P, Begum P, Kumar N, Thakur AJ, Mikkola J-P, et al.
carbonate by transesterification of glycerol with dimethyl carbonate over Mg/Al/ Shape selectivity and acidity effects in glycerol acetylation with acetic anhydride:
Zr catalysts. Appl Catal A Gen 2011;401:153–7. selective synthesis of triacetin over Y-zeolite and sulfonated mesoporous carbons.
[31] Silva JM, Soria MA, Madeira LM. Challenges and strategies for optimization of J Catal 2015;329:237–47.
glycerol steam reforming process. Renew Sustain Energy Rev 2015;42:1187–213. [62] Suraj Onkar Katole. Synthesis of green bioadditives through esterification of
[32] Aline C, José H, Andrade R, Alves F, Sequinel R, Rossato V, et al. Overview of glycerol. Mumbai, India: Institute of Chemical Technology (ICT); 2014.
glycerol reforming for hydrogen production. Renew Sustain Energy Rev [63] Khayoon MS, Hameed BH. Synthesis of hybrid SBA-15 functionalized with
2016;58:259–66. molybdophosphoric acid as efficient catalyst for glycerol esterification to fuel
[33] Remón J, Ruiz J, Oliva M, García L, Arauzo J. Effect of biodiesel-derived additives. Appl Catal A Gen 2012;433–434:152–61.
impurities (acetic acid, methanol and potassium hydroxide) on the aqueous phase [64] Khayoon MS, Hameed BH. Acetylation of glycerol to biofuel additives over
reforming of glycerol. Chem Eng J 2016;299:431–48. sulfated activated carbon catalyst. Bioresour Technol 2011;102:9229–35.
[34] Paul Ş, Pap T. Glycerol acetals and ketals as possible diesel additives. A review of [65] Kim I, Kim J, Lee D. A comparative study on catalytic properties of solid acid
their synthesis protocols. Renew Sustain Energy Rev J 2016;62:804–14. catalysts for glycerol acetylation at low temperatures. Appl Catal B Environ
[35] Nanda MR, Zhang Y, Yuan Z, Qin W, Ghaziaskar HS, Charles C. Catalytic 2014;148–149:295–303.
conversion of glycerol for sustainable production of solketal as a fuel additive: a [66] Dosuna-Rodríguez I, Gaigneaux EM. Glycerol acetylation catalysed by ion
review. Renew Sustain Energy Rev 2016;56:1022–31. exchange resins. Catal Today 2012;195:14–21.
[36] Ghaziaskar HS, Afsari S, Rezayat M, Rastegari H. Quaternary solubility of acetic [67] Gonçalves VLC, Pinto BP, Silva JC, Mota CJ a. Acetylation of glycerol catalyzed by
acid, diacetin and triacetin in supercritical carbon dioxide. J Supercrit Fluids different solid acids. Catal Today 2008;133–135:673–7.
2017;119:52–7. [68] Melero J a, van Grieken R, Morales G, Paniagua M. Acidic mesoporous silica for
[37] Murakami N, Oba M, Iwamoto M, Tashiro Y, Noguchi T, Bonkohara K, et al. L- the acetylation of glycerol: synthesis of bioadditives to petrol fuel. Energy Fuels
Lactic acid production from glycerol coupled with acetic acid metabolism by 2007;21:1782–91.
Enterococcus faecalis without carbon loss. J Biosci Bioeng 2016;121:89–95. [69] Testa ML, La Parola V, Liotta LF, Venezia AM. Screening of different solid acid
[38] San P, Kheireddine M, Mohd W, Wan A, Cognet P, Pérès Y. Enhanced microwave catalysts for glycerol acetylation. J Mol Catal A Chem 2013;367:69–76.
catalytic-esterification of industrial grade glycerol over Brønsted-based methane [70] Ghoreishi KB, Yarmo MA. Sol-gel sulfated silica as a catalyst for glycerol
sulfonic acid in production of biolubricant. Process Saf Environ Prot acetylation with acetic acid. J Sci Technol 2013;5:65–78.
2016;104:323–33. [71] Ghoreishi KB, Yarmo MA, Nordin NM, Samsudin MW. Enhanced catalyst activity
[39] Rosiene M, Arcanjo A, Silva IJ, Rodríguez-castellón E, Infantes-molina A, Vieira of WO3 using polypyrrole as support for acidic esterification of glycerol with acetic
RS. Conversion of glycerol into lactic acid using Pd or Pt supported on carbon as acid. J Chem 2013:2013.
catalyst. Catal Today 2017;279:317–26. [72] Bhorodwaj SK, Dutta DK. Activated clay supported heteropoly acid catalysts for
[40] Kale S, Armbruster U, Umbarkar S, Dongare M, Martin A. Esterification of esterification of acetic acid with butanol. Appl Clay Sci 2011;53:347–52.
glycerol with acetic acid for improved production of triacetin using toluene as an [73] Popova M, Szegedi Á, Ristić A, Tušar NN. Glycerol acetylation on mesoporous
entrainer. In: Proceedings of the 10th green chem conference; 2013. p. 70–71. KIL-2 supported sulphated zirconia catalysts. Catal Sci Technol
[41] Okoye PU, Hameed BH. Review on recent progress in catalytic carboxylation and 2014;4:3993–4000.
acetylation of glycerol as a byproduct of biodiesel production. Renew Sustain [74] Zhang C, Fu Z, Liu YC, Dai B, Zou Y, Gong X, et al. Ionic liquid-functionalized
Energy Rev 2016;53:558–74. biochar sulfonic acid as a biomimetic catalyst for hydrolysis of cellulose and
[42] Lam MK, Lee KT, Mohamed AR. Homogeneous, heterogeneous and enzymatic bamboo under microwave irradiation. Green Chem 2012;14:1928–34.
catalysis for transesterification of high free fatty acid oil (waste cooking oil) to [75] Penzien J, Müller TE, Lercher JA. Hydroamination of 6-aminohex-1-yne over zinc
biodiesel: a review. Biotechnol Adv 2010;28:500–18. based homogeneous and zeolite catalysts. Microporous Mesoporous Mater
[43] Zhou C-HC, Beltramini JN, Fan Y-X, Lu GQM. Chemoselective catalytic conver- 2001;48:285–91.
sion of glycerol as a biorenewable source to valuable commodity chemicals. Chem [76] Haw JF, Hall MB, Alvarado-swaisgoad AE, Mudsod EJ, Lin Z, Beck LW, et al.
Soc Rev 2008;37:527–49. Integrated NMR and Ab initio study of acetonitrile in zeolites: a reactive complex
[44] Cole-Hamilton DJ, Tooze RP. Homogeneous catalysis — advantages and pro- model of zeolite acidity. J Am Chem Soc 1994;116:7308–18.
blems. Catal. Sep. Recover. Recycl. Chem. Process Des., Springer; 2006, p. 1–8. [77] Khayoon MS, Triwahyono S, Hameed BH, Jalil AA. Improved production of fuel
[45] Tudorache M, Negoi A, Protesescu L, Parvulescu VI. Biocatalytic alternative for oxygenates via glycerol acetylation with acetic acid. Chem Eng J
bio-glycerol conversion with alkyl carbonates via a lipase-linked magnetic nano- 2014;243:473–84.
particles assisted process. Appl Catal B Environ 2014;145:120–5. [78] Gelosa D, Ramaioli M, Valente G, Morbidelli M. Chromatographic reactors:
[46] Waghmare GV, Vetal MD, Rathod VK. Ultrasound assisted enzyme catalyzed esterification of glycerol with acetic acid using acidic polymeric resins. Ind Eng
synthesis of glycerol carbonate from glycerol and dimethyl carbonate. Ultrason Chem Res 2003;42:6536–44.
Sonochem 2014;22:311–6. [79] Koros RM, Nowak EJ. A diagnostic test of the kinetic regime in a packed bed
[47] Kim SC, Kim YH, Lee H, Yoon DY, Song BK. Lipase-catalyzed synthesis of glycerol reactor. Chem Eng Sci 1967;22:470–5.
carbonate from renewable glycerol and dimethyl carbonate through transester- [80] Iglesia E, Michel B. Unimolecular and bimolecular formic acid decomposition on
ification. J Mol Catal B Enzym 2007;49:75–8. copper. J Phys Chem 1986;20:5272–4.
[48] Hruby SL, Shanks BH. Acid-base cooperativity in condensation reactions with [81] Wilke CR, Chang P. Correlation of diffusion coefficients in dilute solutions. AIChE
functionalized mesoporous silica catalysts. J Catal 2009;263:181–8. J 1955;1:264–70.

400
P.U. Okoye et al. Renewable and Sustainable Energy Reviews 74 (2017) 387–401

[82] Sitaraman R, Ibrahim SH, Kuloor NR. A generalized equation for diffusion in [107] Dosuna-Rodríguez I, Adriany C, Gaigneaux EM. Glycerol acetylation on sulphated
liquids. J Chem Eng Data 1963;8:198–201. zirconia in mild conditions. Catal Today 2011;167:56–63.
[83] Poling BE, Prausnitz JM, O’Connell JP. The properties of gases & liquids. vol. 1; [108] Singare PU. Thermal degradation studies of some strongly acidic Ca-tion
1988. http://dx.doi.org/10.1016/0894-1777(88)90021-0. exchange resins. Open J Phys Chem 2011;1:45–54.
[84] Lode F, Houmard M, Migliorini C, Mazzotti M, Morbidelli M. Continuous reactive [109] Okoye PU, Abdullah AZ, Hameed BH. Glycerol carbonate synthesis from glycerol
chromatography. Chem Eng Sci 2001;56:269–91. and dimethyl carbonate using trisodium phosphate. J Taiwan Inst Chem Eng
[85] Yadav GD, Ramesh P. Selectivity engineering in the 0- versus C-alkylation of p- 2016;0:1–8.
cresol with cyclohexene over sulfated zirconia. Can J Chem Eng 2000;78:917–27. [110] Okoye PU, Abdullah AZ, Hameed BH. Stabilized ladle furnace steel slag for
[86] Yadav GD, Krishnan MS. Solid acid catalysed acylation of 2-methoxy-naphthalene: glycerol carbonate synthesis via glycerol transesterification reaction with dimethyl
role of intraparticle diffusional resistance. Chem Eng Sci 1999;54:4189–97. carbonate. Energy Convers Manag 2016.
[87] Yadav GD, Joshi aV. A green route for the acylation of resorcinol withacetic acid. [111] Walke* Santosh, Kambale S. Catalyst deactivation and regeneration. Int J Sci Eng
Clean Technol Environ Policy 2002;4:157–64. Technol 2015;5:281–5.
[88] Akbay E, Altıokka MR. Kinetics of esterification of acetic acid with n- amyl alcohol [112] Moulijn JA, van Diepen AE, Kapteijn F. Catalyst deactivation: is it predictable?:
in the presence of Amberlyst-36. Appl Catal A: Gen 2011:13–9. what to do?. Appl Catal A Gen 2001;212:3–16.
[89] Hoo P, Abdullah AZ. Kinetics modeling and mechanism study for selective [113] Levenspiel O. Experimental search for a simple rate equation to describe
esterification of glycerol with lauric acid using 12-tungstophosphoric acid post- deactivating porous catalyst particles. J Catal 1972;25:265–72.
impregnated SBA-15. Ind Eng Chem Res 2015;54:7852–8. [114] Butt JB, Petersen EE. {CHAPTER} 10 – Deactivation in Fixed Beds. In: Petersen
[90] Baxter S, Zivanovic S, Weiss J. Molecular weight and degree of acetylation of high- JBBE, editor. Act Deactiv Poisoning Catal, Elsevier; 1988, p. 347–425.
intensity ultrasonicated chitosan. Food Hydrocoll 2005;19:821–30. [115] Jossens LW, Petersen EE. A novel reactor system that permits the direct
[91] Beard AD, James BB. Power-law kinetics of tracer washout from physiological determination of deactivation kinetics for a heterogeneous catalyst. J Catal
system. Ann Biomed Eng 1998;26:775–9. 1982;73:366–76.
[92] Venuto P. Organic reactions catalyzed by crystalline aluminosilicates III. [116] Fogler HS. 10.7 Catalyst deactivation. Catal Deactiv 2014:56–78.
Condensation reactions of carbonyl compounds. J Catal 1966;6:237–44. [117] Stephen Szépe, Levenspiel O, Laan T. Proc 4th Eur Symp Chem React Eng,
[93] Patrick G, Annie F. Aromatic benzoylation. In: Eric GD, editor. Catal FINE Chem Brussel, London: Pergamon Press; 1970.
Synth Microporous Mesoporous Solid Catal. vol 4, Lisbon, Portugal; 2006. p. 95– [118] Bartholomew CH. Sintering kinetics of supported metals: new perspectives from a
104. unifying GPLE treatment. Appl Catal A Gen 1993;107:1–57.
[94] Dezheng W. Experimental conditions for valid Langmuir-Hinshelwood kinetics. [119] Bartholomew CH. Mechanisms of catalyst deactivation. Appl Catal A Gen
Chin J Catal 2010;31:972–8. 2001;212:17–60.
[95] Kumar KV, Porkodi K, Rocha F. Langmuir–Hinshelwood kinetics – a theoretical [120] Wakabayashi F, Kondo JN, Domen K, Hirose C. FT-IR study of H218O adsorption
study. Catal Commun 2008;9:82–4. on H-ZSM-5: direct evidence for the hydrogen-bonded adsorption of water. J Phys
[98] Kunz L, Maier L, Tischer S, Olaf D. Modeling the rate of heterogeneous reactions. Chem 1996;100:1442–4.
Model Heterog Catal React From Mol Process to Tech Syst, Germany: Wiley-VCH, [121] Panov A, Fripiat J. Poisoning of aldol condensation reaction with H2O on acid
Weinheim 2011; 2011. p. 1–59. catalysts. Catal Lett 1999;57:25–32.
[99] Izci A, Bodur F. Liquid-phase esterification of acetic acid with isobutanol catalyzed [122] Wheeler A. Reaction rates and selectivity in catalyst pores. Adv Catal
by ion- exchange resins. React Funct Polym isobutanol catalyzed by ion-exchange 1951;3:249–327.
resins; 2015. [123] Karanth NG, Luss D. Selectivity enhancement in catalytic purification reactions by
[100] Heineken FG, Tsuchiya HM, Aris R. On the mathematical status of the pseudo- poisons. Chem Eng Sci 1975;30:695–8.
steady state hypothesis of biochemical kinetics. Math Biosci 1967;1:95–113. [124] Yang H, Chen H, Chen J, Omotoso O, Ring Z. Shape selective and hydrogen
[101] Tsibranska I, Hristova E. Comparison of different kinetic models for adsorption spillover approach in the design of sulfur-tolerant hydrogenation catalysts. J Catal
of heavy metals onto activated carbon from apricot stones. 43; 2011. p. 370–7. 2006;243:36–42.
[102] Fogler HS. Elements of chemical reaction engineering elements of chemical [125] Voorhies A. Carbon formation in catalytic cracking. Ind Eng Chem Res
reaction engineering, 3rd ed. New Delhi: Prentice-Hall International, Inc; 1995. 1945;37:318–22.
[103] Anderson DL. Thermodynamics and equations of state. Theory of the earth, 1998. [126] Froment GF, Bischoff KB. Kinetic data and product distributions from fixed bed
Boston: Blackwell Scientific Publications; 1989. p. 84–101. catalytic reactors subject to catalyst fouling. Chem Eng Sci 1962;17:105–14.
[104] Forzatti P. Catalyst deactivation. Catal Today 1999;52:165–81. [127] Froment GF, Bischoff KB. Non-steady state behaviour of fixed bed catalytic
[105] Argyle M, Bartholomew C. Heterogeneous catalyst deactivation and regeneration: reactors due to catalyst fouling. Chem Eng Sci 1961;16:189–201.
a review. Catalysts 2015;5:145–269. [128] Fuentes GA, Gamas ED. Towards a better understanding of sintering phenomena
[106] Farinha J, Caiado M, Castanheiro JE. Valorisation of glycerol into biofuel in catalysis. In: Bartholomew CH, Butt JB, editors. Catal Deactiv 1991
additives over heterogeneous catalysts; 2013. p. 422–9. Proceedings 5th Int Symp, vol. 68, Elsevier; 1991, p. 637–44.

401

Вам также может понравиться