Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Renewable Energy 114 (2017) 794e804

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Thermo-kinetic and diffusion studies of glycerol dehydration to


acrolein using HSiW-g-Al2O3 supported ZrO2 solid acid catalyst
Amin Talebian-Kiakalaieh, Nor Aishah Saidina Amin*
Chemical Reaction Engineering Group (CREG), Faculty of Chemical Engineering, Universiti Teknologi Malaysia (UTM), 81310 Skudai, Johor, Malaysia

a r t i c l e i n f o a b s t r a c t

Article history: The thermo-kinetic study of gas-phase glycerol dehydration reaction using a supported g-Al2O3 nano-
Received 26 February 2017 particle based solid catalyst (SiW20-Al/Zr10) has been investigated. The kinetic model was established
Received in revised form based on the reaction mechanism, taking into account two parallel reactions of glycerol degradation into
31 May 2017
acrolein or acetol. Reaction rate constants and activation energies for all the products in the glycerol
Accepted 23 July 2017
dehydration reaction were determined at various reaction temperatures (280e340  C). The first-order
Available online 25 July 2017
kinetic model and the experimental data fitted-well. Also, based on thermodynamic analysis the
values of DHz, DSz, and DGz for all the endothermic reactions were determined by Eyring equation.
Keywords:
Glycerol
Finally, the absence of internal and external diffusions were confirmed by Weisz-Prate Criterion (CWP <1)
Acrolein and Mear's Criterion (CM < 0.15), respectively for pellet diameters less than dP < 5 mm. The results from
Kinetic this study are useful for future reactor modeling and simulation work.
Heterogeneous catalysis © 2017 Elsevier Ltd. All rights reserved.
Diffusion
Thermodynamics

1. Introduction procedure involving an active component (HSiW) supported on g-


Al2O3 nanoparticles (NPs) and finally grafting the supported sample
Glycerol is the main by-product of the biodiesel process, ac- with ZrO2 is proposed. Indeed, it is envisaged the g-Al2O3 NPs could
counting for almost 10 wt% of the total products [1]. There is a tune the acidity and provides a supported sample with enlarged
surplus of glycerol supply due to increasing biodiesel production specific surface area and pore diameter compared to the bulk HSiW.
rate in the last decade [2]. Catalytic conversion of glycerol to other Finally, ZrO2 is added to reduce the effect of catalyst deactivation
value-added chemicals has attracted much attention recently [3,4]. and improved the life-long stability of catalyst for gas phase
Consequently, large numbers of studies have focused on the design dehydration of glycerol to acrolein.
and development of effective heterogeneous catalyst for dehydra- From the economical point of view, the cost of an acid catalyst
tion of glycerol to acrolein [5e7]. Acrolein is one of the most (e.g. SiW20-Al-Zr10) is much cheaper than a multi-component metal
important intermediates for production of acrylic acid, methionine, sample (e.g. bismuth molybdate (Bi2Mo3O12)) currently is being
1,3-propanediol and lactic acid [8,9]. The highest ever reported used in the petroleum-based process of acrolein production.
acrolein selectivity was obtained by application of supported het- Indeed, the retail price (Sigma-Aldrich) of the active compound
eropoly acid catalyst in gas phase [10]. HPAs are environmentally Bi2Mo3O12 sample costs (USD 5.4/g) over twice the price of HSiW
friendly and reusable in various chemical reactions. However, small (USD 2.5/g). If we assume that equal amount of catalysts are used in
surface area (1e10 m2/g) and low thermal stability are the main both petroleum- and glycerol-based processes then the costs of
drawbacks of HPAs. HPAs are usually supported over carriers such supported acid catalyst will be approximately half of the multi-
as Al2O3, ZrO2, and SiO2 to overcome these limitations. In addition, component sample. Definitely, other factors such as density and
strong acidity and high Bronsted acidic sites are the main reasons long-life stability have significant influence on the cost of catalyst.
for increasing coke deposition on the catalyst surface, deactivating The SiW20-Al/Zr10 catalyst is recently synthesized in our
the catalyst rapidly. Therefore, a novel catalyst preparation research group. It is registered 87.6% acrolein selectivity at 97.1%
glycerol conversion [11]. The latter sample exhibits long-life sta-
bility with >75% acrolein selectivity and >80% glycerol conversion
even after 40 h of reaction time. Thus, SiW20-Al/Zr10 is a suitable
* Corresponding author.
E-mail addresses: amin.talebian63@gmail.com (A. Talebian-Kiakalaieh), catalyst to be investigated for further studies including kinetic,
noraishah@cheme.utm.my (N.A.S. Amin). mass transfer, modeling and simulation. Various characterization

http://dx.doi.org/10.1016/j.renene.2017.07.096
0960-1481/© 2017 Elsevier Ltd. All rights reserved.
A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804 795

studies confirmed that catalyst activity is dominantly affected by nanoparticles was prepared via the incipient-wetness impregna-
the acidity and textural characteristics of the prepared catalyst [11]. tion method. Aqueous HSiW solution was added drop-wise to the
Only a limited number of recent kinetic studies are performed in g-Al2O3supportinitially. The suspension was rigorously stirred for
a continuous system and at atmospheric pressure [12,13]. All the 12 h followed by drying at 110  C for 18 h. The HSiW-Al2O3 sup-
previous kinetic studies on glycerol dehydration to acrolein were ported catalyst was denoted as SiW20-Al. The final catalyst was
performed at supercritical water (SCW) conditions without catalyst prepared via the impregnation of 10 wt% ZrO2 on the dried SiW20-
or in presence of some simple catalyst [14e16], which has limited Al precursor in a similar aqueous impregnation procedure. Finally,
potential for industrialization due to the high production costs and the sample was dried in the oven at 120  C for 18 h and referred to
inherent technical challenges (e.g. high pressure and temperature). as SiW20-Al/Zr10.
Indeed, kinetic modeling is vital in unraveling the reaction mech-
anism, analyzing the catalyst characteristics effects on reaction rate, 2.3. Catalyst characterization
and even designing a catalyst with long-life stability and high ac-
tivity [17]. The SiW20-Al/Zr10 catalyst was characterized by field-emission
Catalyst deactivation as a result of coke deposition on catalyst scanning electron microscopy and energy dispersive X-ray tech-
surface is reported as the main obstacle for successful industrial niques, elemental analyzer, and temperature-programmed
application of the majority of synthesized catalysts in green desorption. Also, the pore size and surface area were determined
chemistry [18,19]. Also, in a heterogeneous catalytic system, effec- by Nitrogen adsorptiondesorption and Brunauer, Emmett, and
tive utilization of catalyst surface depends on the diffusion limita- Teller (BET) methods, respectively. The details of different charac-
tions inside the pore structure in the pellet [20]. It is well terization results have been discussed elsewhere [11]. Table 1
established that diffusion has significant effect on the rate of re- summarizes some of the physico-chemical properties of SiW20-
action and product formation due to different phases [21]. Thus, Al/Zr10 for the fresh and spent catalysts.
internal and external mass transfer studies by application of Weisz
Prater criterion (CWP) and Mear's criterion (CM) are essential to 2.4. Catalytic reaction
confirm the absence of diffusion inside the pores and the full uti-
lization of catalyst surface, respectively. The gas-phase dehydration of glycerol was conducted at atmo-
Therefore, the main objective of the paper is to present a spheric pressure in a vertical packed-bed quartz reactor (30 cm
comprehensive kinetic study to determine the kinetic parameters length, 11 mm i.d.) using a 0.5 g catalyst (dp ¼ 1e5 mm) sandwiched
(e.g. reaction rate constants, activation energies, and frequency between plugs of glass wool. Fig. 1 illustrates the schematic dia-
factors) for all the products in glycerol dehydration reaction using gram of gas phase glycerol dehydration to acrolein. Prior to reac-
the new synthesized solid acid catalyst (SiW20-Al/Zr10). Thermo- tion, the catalyst was pretreated at reaction temperature (300  C)
dynamic analysis is also performed to determine DHz, DSz, and DGz under nitrogen (N2) flow (1200 mL/h) for 1 h. Liquid aqueous
for all reactions and products in the glycerol dehydration reaction glycerol (10 wt %) was fed by a syringe pump at a 2 mL/h flow rate.
by using the Eyring equation. Finally, the existence of internal and The liquid was then vaporized in a pre-heater, mixed with inert
external diffusions over various feed flow rates (mg), pellet diam- carrier gas, and run to the reactor. The gas hourly space velocity
eter (dP), and catalytic bed volume (Vcat) are investigated using (GHSV) of the inert carrier gas was 1200 h1. The reaction products
Wiesz-Prater criterion (CWP) and Mear's criterion (CM) to evaluate and unconverted glycerol were condensed and collected for anal-
the catalyst efficiency. ysis after 3 h of reaction. n-butanol was added to the condensed
products as the internal standard. The final solution was analyzed
2. Experimental by a gas chromatography (GC) instrument equipped with capillary
column (DB-Wax: 30 m  0.53 mm  0.25 mm) and flame ionization
2.1. Materials detector (FID).The glycerol conversion, acrolein selectivity, and
yield are defined in Eqs (1)e(3):
Glycerol (>99%), silicotungstic acid (HSiW, 99.9%), aluminum
oxide (g-Al2O3, 99.9%) nanoparticle, and zirconium oxide (ZrO2, M Gl; in feed  M Gl; in outlet
ConversionGl ð%Þ ¼  100% (1)
>99%) were purchased from Sigma-Aldrich (Malaysia) as were M Gl; in feed
other chemicals including hydroxyacetone, acetone, propanal, and
etc (HPLC grade 99.8%). Acrolein (98%) at reagent grade was
MC in product
supplied by Scientific Trends (M) Sdn. Bhd. All the chemicals were
SelectivityN ð%Þ ¼
used as received without further purification. MC in Gl; feed  M C in Gl; outlet
 100%
2.2. Catalyst preparation
(2)
First, a catalyst with 20 wt% HSiW loading on g-Al2O3

Table 1
Schematic diagram of reactor setup.

Catalyst SBETa (m2/g) VPb (cm3/g) DPc (nm) Coke content (wt%) Loading (wt%)

HSiW g-Al2O3 ZrO2

Fresh SiW20-Al/Zr10 96.9 0.4 19.1 e 20 70 10


Spent SiW20-Al/Zr10 93.0 0.4 18.2 1.3 20 70 10
a
Specific surface area.
b
Pore volume.
c
Pore diameter.
796 A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804

If the adsorption and desorption steps are very fast during the
above reaction mechanisms, then the concentration of the adsor-
bed species can be determined by assuming the adsorption and
desorption steps are at equilibrium. Therefore, the concentrations
of all adsorbed species are determined by Eqs. (9e11):

k1
rAds ¼ k1 :CA :CS  k1 :CAS ¼ 00CAS ¼ :C :C 0CAS
k1 A S
¼ K1 :CA :CS (9)

k3
rdes ¼ k3 :CBS  k3 :CB :CS ¼ 00CBS ¼ :C :C 0CBS
k3 B S
1
¼ :C :C (10)
K3 B S

sffiffiffiffiffiffiffiffi
2 k4
rdes ¼ k4 :CCS  k4 :CC2 :CS2 ¼ 00CCS ¼ :C :C 0CCS
k4 C S
1
¼ pffiffiffiffiffiffi:CC :CS
K4
(11)
The new equation for the reaction rate is generated in Eq. (12) by
substituting the adsorbed species of A, CAS in Eq. (8):

Fig. 1. Schematic diagram of reactor setup.


rA ¼ k2 :K1 :CA :CS3 (12)
The value of Cs can be determined from the site balance ac-
cording to Eq. (13)e(16):
YieldN ð%Þ ¼ ConversionGl ð%Þ  SelectivityN ð%Þ (3)
Ct;s ¼ CAS þ CBS þ CCS þ CS (13)
where MGl is moles of glycerol in feed or outlet streams, MC, is
moles of acrolein, and N is moles of each product. In this work, the
products are identified as acrolein, acetol, acetone, acetaldehyde, CB :CS CC :CS
0Ct;s ¼ K1 :CA :CS þ þ pffiffiffiffiffiffi þ CS (14)
and minor products. All the experimental runs were repeated at K3 K4
least 3 times and the average values with ±5% variance have been
reported as the final results.” !
CB CC
0Ct;s ¼ CS K1 :CA þ þ pffiffiffiffiffi
ffiþ1 (15)
2.5. Kinetic study K3 K4

The Langmuir-Hinshelwood-Hougen-Watson (LHHW) model is Ct;s


applied for finding the reaction rate equation. The LHHW mecha- 0CS ¼ (16)
K1 :CA þ KCB3 þ pCffiffiffiffi
K
C
þ1
nism includes adsorption of the glycerol (A), the reaction between 4

adsorbed reactants (glycerol) on the catalyst active site (S) and


where, Ct,s is the total available active sites.
finally desorption of acrolein (B) and water (C) from the catalyst
Therefore, the rate of reaction (Eq. (17)) is determined by
surface. Hence, the following steps are taken to delineate the rate
substituting Cs from Eq. (16) into Eq. (12):
equations:
3
k2 :K1 :CA :Ct;s
k1
A þ S ⇔ A:S (4) rA ¼ h i3 (17)
k1 K1 :CA þ KCB3 þ pCffiffiffiffi
K
C
þ1
4

k2
AS þ 2S ⇔ B:S þ 2C:S ðRate determining stepÞ (5) However, at initial condition (t ¼ 0, -rA¼ -rA0; therefore CB¼Cc ¼ 0).
k2 Thus, the reaction rate is obtained as Eq. (18):

k3 3
k2 :K1 :CA :Ct;s
B:S ⇔ B þ S (6) rA ¼ (18)
k3 ½K1 :CA þ 13

k4
The reaction rate can be expressed as Eq. (19) by assuming
2C:S ⇔ 2C þ 2S (7) K1CA«1 in gas phase reactions:
k4

If the surface reaction is the rate-controlling step during dehy- rA ¼ kSR :CA (19)
dration reaction, the rate of reaction is given by Eq. (8):
where, kSR ¼ apparent reaction rate constant ¼ k2 :K1 :Ct;s .
rA ¼ k2 :CAS :CS2 (8) The LHHW model confirmed that the glycerol dehydration re-
action rate is a function of glycerol (A) concentration only. A similar
A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804 797

procedure for calculation of reaction rate is also reported elsewhere


[13,22]. rA0 :rb :R:n
〈0:15 (24)
Unfortunately, the detailed calculation procedure for reaction kc :CAB
rate constants (ki) is not commonly discussed in the majority of
kinetic papers. Thus, a detail step by step procedure is described as where, n is reaction order, R is catalyst pellet radius (m), rb is bulk
follows: first, we should find various differential equations based density of catalyst bed (kg/m3) (rb¼(1- ɸ).rc), rc is solid density of
on different reaction rates presented in the glycerol dehydration catalyst (kg/m3), CAB is bulk reactant concentration (mol/dm3), and
reaction. Next, the obtained differential equations should be solved kc is mass transfer coefficient (m/s). However, to calculate the
and the analytical expression of various products concentrations external mass transfer coefficient (kc), we first need to find the
expressed. Finally, the ordinary differential equations (ODE's) Reynolds number (Re), Schmidt number (Sc), and Sherwood num-
should be solved as a function of contact time (t) to determine the ber (Sh), based on the following equations:
reaction rate constants. There are various approaches for calcula- y0
tion of ODE's such as Excel and MATLAB [23], but Polymath has U¼ (25)
Ac
been selected in this study.
U:dp
Re0 ¼ (26)
2.6. Diffusion study ð1  fÞ:y

2.6.1. Internal diffusion y


Sc ¼ (27)
The Weisz-Prater Criterion (CWP) uses measured values of the DAB
rate of reaction to determine if internal diffusion is limiting the
1= 1
reaction. In fact, a lack of significant intraphase diffusion effects (i.e.
Sh0 ¼ ðRe0 Þ :ðScÞ
=
2 3
(28)
h ¼ 0.95) on an irreversible, isothermal, first-order reaction in a
spherical catalyst pellet can be assessed by the CWP [24]. The in-  
1  f DAB
ternal effectiveness factor for a first order reaction with a spherical kc ¼ :Sh0 (29)
catalyst pellet is calculated by equation (20) [25,26]: f dp

rA0 ðobsÞ:rC :R2 where, y0 is the reactant flow rate, Ac is the cross sectional area of
CWP ¼ hf21 ¼ 3ðf1 Cothf1  1Þ ¼ (20) reactor, dp is the pellet diameter, ɸ is bed porosity, y is kinematic
Deff :CAS
viscosity and DAB is gas phase diffusivity.
Meanwhile, the Thiele modules (ɸ21) for a first-order reaction can
be calculated by equation (21) based on the reaction rate constant 2.6.3. Overall effectiveness factor
(k) with dimension (1=time or s1) [27]: The overall effectiveness factor is calculated by equation (30) for
a first order reaction [25,26]:
sffiffiffiffiffiffiffiffi
f1 ¼ R:
k1 h
Deff
(21) U¼ h:k001 :Sa :rb
(30)
1þ kc :ac
The gas diffusivity (DAB) in this study is calculated by equation
To find the overall effectiveness factor (U) we need to first find
(22):
the unknown parameters such as external mass transfer coefficient
T (kc), as discussed in section 2.6.2, external area per unit reactor
DAB ¼ 1:173  1016 ðf:MB Þ: (22) volume (ac), and bulk density of bed (rb ¼ rc(1-ɸ)).
mB :VA0:6
6ð1  fÞ
where MB is the molecular weight of solvent (water), mB is the ac ¼ (31)
dp
viscosity of solvent (water) in gas phase at 300  C and VA is the
solute molar volume [28]. The calculation method and values for
each material are mentioned elsewhere [29]. The value of ɸ, the
associated parameter, is 2.6 for water [29]. 3. Results and discussion
The effective diffusivity is calculated by equation (23):
ε 3.1. Reaction route
Deff ¼ :DAB (23)
t
Glycerol dehydration to acrolein has been widely studied at
where, t is tortuosity and ε is pellet porosity. various (ambient and even supercritical) conditions in gas or liquid
After finding all the unknowns (k1 [s1], Deff, ɸ21 and R) the CWP phases. Consequently, various products have been reported based
can be calculated by equation (20). Finally, CWP « 1indicates there is on application of different catalysts and reaction conditions. The
no internal diffusion limitation, but for CWP[1 internal diffusion major products are 3-hydroxepropanal (3-HPA), acetol, acrolein,
severely limits the reaction. Similar procedure for calculation of acetaldehyde, 1,2-propanediol, acetone, acetic acid, allyl alcohol,
CWP has been reported elsewhere [26]. and propanal [31e34]. Identification of the intermediate steps and
by-products formation is necessary since the mechanisms, reaction
rates, and activation energies related to the majority of glycerol
2.6.2. External diffusion dehydration reaction by-products are still unknown. Fig. 2a illus-
Mear's criterion (CM), like CWP, uses the measured rate of reac- trates a reaction mechanism proposed for glycerol dehydration to
tion to determine if mass transfer from the bulk gas phase to the acrolein using supported solid acid (SiW20-Al/Zr10) catalyst in this
catalyst surface can be neglected [30]. Indeed, CM < 0.15 reveals that study.
the external diffusion can be neglected as indicated by Eqn (24). The proposed mechanism reveals that acrolein is obtained after
798 A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804

a) b)

Fig. 2. (a) Reaction mechanism of glycerol dehydration to acrolein over solid acid catalyst, (b) Simplified reaction routes for calculation of kinetic parameters.

two glycerol dehydration steps. 3-HPA and acetol, which are the As mentioned earlier, two of the intermediate products (3-HPA
products of the first dehydration step, are produced by two distinct and 1,2-propanediol) were not detected in this study. We could
and independent pathways. 3-HPA is produced as a result of conclude that all the 3-HPA and 1,2-propanediol transformed to
removal of the central alcohol function (secondary hydroxyl group) acrolein and acetone through R1 and R5, respectively (Fig. 2b). The
in the glycerol molecule, while acetol formation is due to removal of proposed reaction mechanism exhibits various reaction path ways
the glycerol terminal alcohol (one of the eOH) groups for acetaldehyde production. Thus, to simplify the reaction mech-
[12,18,35e37]. anism we assumed that all the acetaldehyde and minor by-
The 3-HPA is sufficiently reactive and readily converted into products formed from acrolein by R3 and R4, respectively. The
acrolein by second dehydration step. In fact, 3-HPA is one of the final reaction routes for calculation of kinetic parameters are
intermediate products which is not detected in this study due to its summarized in Fig. 2b.
high reactivity to acrolein, while acrolein is a less reactive product.
However, the hydrogenation of acrolein can form propanal and allyl
alcohol [35]. Allyl alcohol can also be obtained directly from glyc- 3.2. Kinetic modeling
erol [36]. The experimental results confers a wide range of prod-
ucts, including formaldehyde, 1,3-dioxan-5-ol, but the Based on the LHHW model (Section 2.5) and previous studies
concentration of each compound was minute. Thus, propanal, allyl [12,38] we can assume first order reaction for all the products. The
alcohol, formaldehyde, 1, 3-dioxan-5-ol and etc are grouped as a dehydration steps (R1 and R2) are assumed to be pseudo-first order
minor by-products [12]. Acetaldehyde is formed from acrolein reaction [12,38]. The water molecules release during the dehydra-
through intermediate 3-HPA [38]. Acetaldehyde can also be formed tion steps (R1 and R2). However, we assumed that total amount of
by acetol C-C bond cleavage [12]. Acetol is highly reactive and it is water is constant due to the large proportion of water (90 wt%) and
hydrogenated to acetone with acid catalyst through the 1,2- little amount of glycerol (10 wt%) in the feed solution. Thus, the
propanediol as intermediate, which is not detected in this study reaction rates of Ri can be described as follows:
[18,35,39]. Hydrogen, which is involved in the hydrogenation pro-
cess, originated from the acid catalyst [39], Lewis acidic sites of ri ¼ ki : Cr ði ¼ 1  5Þ (32)
catalyst [10], and coke species [33].
where, ki is pseudo-first order kinetic rate constant and Cr is the
A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804 799

concentration of each reactant. Consequently, based on various of (39e44) by the non-linear least square regression method in
reaction rates, the following differential equations (33e38) were Polymath 6.10 software. Finally, all the differential equations were
obtained: used to fit the experimental data.
Fig. 3 exhibits the experimental data and calculated curves of
dCGl glycerol conversion and different product yields versus W/F at 280,
¼ k1 CGl  k2 CGl ¼ kG CGl (33)
dt 300, 320, and 340  C. The experimental data are explained well by
the proposed reaction routes. Fig. 3aed clearly illustrate that
dCAC increasing temperature from 280 to 340  C significantly surged the
¼ k1 CGl  k3 CAC  k4 CAC ¼ k1 CGl  kA CAC (34)
dt pace of conversion increment to reach the maximum level (99%) at
lower W/F. Also, acrolein production amount is reduced by
dCAct increasing the reaction temperature, particularly at higher W/F
¼ k2 CGl  k5 CAct (35) ratios of 2e3 (103 kg-cat.s/m3). Moreover, the production rate of
dt
by-products such as acetone, acetaldehyde, and minor products is
enhanced significantly at higher temperatures. Acetol production
dCAd
¼ k3 CAC (36) reached a peak at 320  C reaction temperature and W/F ratios of
dt
0.5e1 (103 kg-cat.s/m3) and then reduced at higher temperatures
and W/F ratios (340  C and W/F ¼ 3 (103 kg-cat.s/m3)). The main
dCMp
¼ k4 CAC (37) reason is more acetol has been transformed to acetone. Indeed,
dt Fig. 3ced clearly exhibit acetone yield increases from approxi-
mately 0.06% at 320  C to around 0.07% at 340  C, which is in good
dCA agreement with previous result [12]. Indeed, further degradation of
¼ k5 CAct (38)
dt acrolein is minimal through the catalyst bed whereas acetol tends
to be transformed into acetone in this study. In addition, the yield of
where, kG ¼ k1 þ k2 and kA ¼ k3 þk4. minor products also increased steadily with temperature. The re-
The above differential equations were solved and the analytical sults are consistent with reports that higher temperature reduced
expressions of concentrations for various products are expressed in acrolein selectivity and enhanced by-products yield [12,18,40]. The
Eqs (39e44): carbon balance (CB) for each experimental run at constant W/
F ¼ 2.1 [103 kg-cat.s/m3] is reported in Table 2.
CGl ¼ CG0 :expð  kG :tÞ (39)
The obtained reaction rate constants (k1 to k5, kG, and kA),
activation energies and frequency factors are listed in Table 3. The
k1 CG0 last step in kinetic study is determination of activation energy (Ea)
CAC ¼ ðexpðkG tÞ  expðkA tÞÞ (40)
kA :kG and frequency factor (A). Based on the Arrhenius equation (Eq (46)),
ln(k) versus inverse of temperature(1/T) is plotted in Fig. 4.
k2 CG0 Meanwhile, the activation energies and the frequency factors are
CAct ¼ ðexpðkG tÞ  expðk5 tÞÞ (41) calculated from the slope and intercept, respectively.
k5  kG

 
k3 k1 CG0 kA ð1  expðkG tÞÞ  kG ð1  expðkA tÞÞ Ea
CAd ¼ ln k ¼ ln A  (46)
kA  kG kA :kG RT
(42) Table 3 and Fig. 4 reveal that by increasing the reaction tem-
perature from 280 to 340  C, the rate constant for each reaction
 
k4 k1 CG0 kA ð1  expðkG tÞÞ  kG ð1  expðkA tÞÞ increase significantly. The rate constant of glycerol degradation to
CMp ¼
kA  kG kA :kG acrolein (k1) is larger than glycerol to acetol (k2) at the any of the
investigated temperatures (Fig. 4). Thus, a higher acrolein yield is
(43)
obtained compared to acetol. The latter findings are in good
  agreement with our descriptions in section 3.1. In fact, the presence
k5 k2 CG0 k5 ð1  expðkG tÞÞ  kG ð1  expðk5 tÞÞ of Bronsted acid sites attributes to reaction with the secondary
CA ¼
k5  kG kG :k5 hydroxyl groups of glycerol and produce acrolein instead of acetol.
(44) The reaction rate constants for acetaldehyde (k3) and minor
products (k4) or acrolein decomposition (kA) were significantly
The above ordinary differential equations (ODEs) are functions
    smaller than the reaction rate constant of acrolein formation from
of the contact time (t) where t ¼ W=F Catalyst weight
Feed flowrate
¼ kgcat:s
m3
. glycerol (k1) at the tested reaction temperatures (280e340  C) as
evident in Fig. 4 [12,38]. This indicates that just a small amount of
CGl, CAC, CAct, CAd, CMp, CA, and CG0 are the concentrations of glycerol,
acrolein is transformed to acetaldehyde and other minor products.
acrolein, acetol, acetaldehyde, minor by-products, acetone, and
The latter results are also in good agreement with the proposed
initial concentration of glycerol, respectively. Also, k1 to k5, kG, and
reaction mechanism. Acrolein is a less reactive product and hy-
kA are the reaction rate constants of R1 to R5, glycerol dehydration
drogenation of acrolein formed minor products (e.g. propanal and
to acrolein and acetol, and acrolein decomposition, respectively.
allyl alcohol). In addition, the reaction rate constant of minor
The calculations of reaction rate constants are divided into three
products (k4) is higher than acetaldehyde (k3). The activation en-
steps. First, the glycerol dehydration to acrolein and acetol rate
ergies reveals that acrolein formation requires less activation en-
constant (kG) was determined from the conversion equation (Eq.
ergy (Ea ¼ 46.0 kJ/mol) compared to glycerol decomposition to
(45)) as a function of contact time (t ¼ W/F):
acetol (Ea ¼ 53.3 kJ/mol). Indeed, acetol formation requires the
highest activation energy among various products and high tem-
XGl ¼ 1  expð  ðk1 þ k2 ÞtÞ ¼ 1  expðkG :tÞ (45)
perature favors acetol formation [41].
Second, k1 to k5 and kA were obtained by solving the equations Table 4 compares the major kinetic studies for glycerol
800 A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804

a) b)
Conversion, Yield [-]

Conversion, Yield [-]


W/F [103kg-cat.s/m3]
W/F [103kg-cat.s/m3]
c) d)
Conversion, Yield [-]

Conversion, Yield [-]

W/F [103kg-cat.s/m3] W/F [103kg-cat.s/m3]

Fig. 3. Glycerol conversions and product yields at (a) 280  C, (b) 300  C, (c) 320  C, and (d) 340  C, Experimental results: ( ) Glycerol conversion, ( ) AC, ( ) Act, ( ) Ad, ( ) A,
and ( ) Mp. Fitting results: ( ) Glycerol conversion, ( ) AC, ( ) Act, ( ) Acd, ( ) A, and ( ) Mp.

Table 2 continuous catalytic process. The first four rows in Table 4 present
Performance of SiW20-Al/Zr10 catalyst.a some studies at SCW conditions with or without catalyst presence.
For instance, Watanabe et al. [14] and Ott et al. [15] obtained 146
T ( C) X (%) YAC (%) YAct (%) YAd (%) YMp (%) YA (%) CB (%)
and 140 kJ/mol activation energies for glycerol dehydration to
280 86.0 68.0 9.2 2.3 6.0 3.0 88.5 acrolein reaction. Akizuki et al. [38] pioneered a comprehensive
300 93.0 78.0 7.5 2.6 7.0 1.2 96.3
320 96.0 76.5 9.5 3.0 8.0 3.0 97.0
kinetic study to determine all the reaction rate constants for various
340 98.5 75.0 7.2 3.5 7.5 5.5 98.7 products in glycerol dehydration reaction. Unfortunately, they did
a not report the (Ea) and (A) since their experimental studies was just
Reaction condition: Feed: 2 ml/h, 10 wt% aqueous glycerol solution, Carrier gas:
N2 with 20 ml/min flow, W/F: 2.1 [103kg-cat.s/m3]. performed at one reaction temperature only. Instead, their main
concern was to evaluate WO3 loading on TiO2 sample to determine
the best catalyst for glycerol dehydration to acrolein. In another
dehydration to acrolein reaction during the last decade. The value development, Park et al. [12] was the earliest to report the kinetic
of activation energy ¼ 46.0 kJ/mol obtained in this study is not only studies in a continuous process. They performed a comprehensive
far lower than the ones reported at supercritical water (SCW) kinetic study by application of two commercial solid catalysts. The
condition, but it also corroborated with a few recent studies in a obtained activation energies were 40.4 and 44.6 kJ/mol for

Table 3
Apparent kinetic and thermodynamic parameters for the dehydration of glycerol over SiW20-Al/Zr10 catalyst calculated by the best fitted curve.
 
rate constants Temperature 
(C) Ea ðkJ=molÞ A m3 DHz (kJ/mol) DSz (J/mol.K) DGz(kJ/mol)
  kg  cat:s

m3 280 300 320 340
kg  cat:s

kG 1.051  103 1.439  103 2.116  103 2.801  103 46.9 27.7 41.3 166.7 91.03
k1 9.240  104 1.306  103 1.855  103 2.441  103 46.0 20.7 40.6 169.6 91.20
k2 1.270  104 1.326  104 2.607  104 3.603  104 53.3 12.1 47.6 174.5 99.65
kA 6.869  105 7.673  105 7.718  105 7.832  105 5.7 2.5  104 3.5 258.4 80.58
k3 2.092  105 2.300  105 2.310  105 2.345  105 5.0 6.3  105 2.8 269.7 83.25
k4 4.777  105 5.373  105 5.408  105 5.487  105 6.1 1.8  104 3.9 260.6 81.64
k5 1.042  104 1.432  104 1.863  104 2.888  104 46.6 2.6 43.3 183.4 98.01
A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804 801

1/T (1/K) synthesized catalyst (SiW20-Al/Zr10) exhibited quite comparable


0.0016 0.00165 0.0017 0.00175 0.0018 0.00185 results (46.0 kJ/mol) with the commercial ones [12].
-5
kG
3.3. Thermodynamic analysis
-5.5 k1
k2 There are actually very little studies on thermodynamic analysis
-6 related to glycerol dehydration to acrolein reaction [13,42]. Thus,
kA determination of thermodynamic parameters (DHz, DSz, and DGz,
k3 double dagger symbol represents the activated complex [41,43])
-6.5
has significant effect on improving the understanding of the
k4 dehydration reaction process. Indeed, thermodynamic studies can
-7 provide important information about the feasibility of reaction
k5
environment and activation parameters [43].
Ln (k)

-7.5 The Eyring equation (Eq. (47)) was applied to determine the
thermodynamic parameters based on the transition state theory
-8  
k DН ± 1 k DS±
ln ¼ : þ ln B þ (47)
T R T h R
-8.5
where, k is reaction rate constant (1/s), T is temperature (K), DHz is
-9 enthalpy of activation (kJ/mol), R is gas constant (8.314 J/mol.K), kB
is Boltzmann constant (1.381  1023 J/K), h is Planck's constant
(6.626  1034 J s) [44], and DSz is entropy of activation (J/mol.K).
-9.5
The activation enthalpy (DHz) and the activation entropy (DSz) can
be calculated from the slope and intercept of the ln (k/T) versus 1/T
-10 plot, respectively.
The values of DHz and DSz for all the reactions (R1 to R5), re-
-10.5 ported in Table 3, are based on the fitted results (slops and in-
tercepts) in Fig. 5. The positive values of DHz infers the glycerol
dehydration to acrolein reaction is endothermic. The calculated DHz
-11
values by Eyring equation are much closer to the Ea values from the
Fig. 4. Arrhenius plot of the reaction rate constants for the estimation of activation Arrhenius equation (Eq. (46)). Similar to the Ea, formation of desired
energies (Ea) and frequency factors (A). products such as acrolein (R1) confers lower DHz ¼ 40.6 kJ/mol
compared to the acetol (R2) formation with DHz ¼ 47.6 kJ/mol. In
fact, the lower enthalpy of activation (DHz) value indicated easier
commercial HZSM-5 and ASPN-40 samples, respectively. They also formation of a product (e.g. acrolein), while higher DHz value
reported all the reaction rate constants for different products (3- indicated complexity in the product formation (e.g. acetol) which is
HPA, acrolein, acetaldehyde, acetol, and lump). Our new absolutely in agreement with our experimental results.

Table 4
Summary of kinetic studies on glycerol dehydration to acrolein.

Reaction condition Ea (kJ/mol)a A (s1) Description Ref

T: 300e400  C 146 e Just overall glycerol dehydration to acrolein reaction [16]


P: 34.5 MPa evaluated
Cat: H2SO4
SCW condition
T: 300e360  C 140 1.3  108 Just overall glycerol dehydration to acrolein reaction [17]
P: 25 MPa evaluated
Cat: Zinc-Sulfate
SCW condition
T: 200e400  C 39.6 e Just overall glycerol dehydration to acrolein reaction [18]
P: 30 MPa evaluated
No catalyst
SCW condition
T: 400  C e e A comprehensive study to find reaction rate constants for all [31]
P: 33 MPa the glycerol dehydration products.
Cat: TiO2& WO3/TiO2 No (Ea) and (A) were reported due for all the experimental
SCW condition runs performed at one temperature.
T: 250e300  C 40.4 for e A comprehensive study to find reaction rate constants for all [14]
P: Atmospheric HZSM-5 the glycerol dehydration products.
Cat: HZSM-5 & ASPN-40 44.6 for
ASPN-40
T: 280e340  C 46.0 20.70b A comprehensive study to find reaction rate constants for all This study
P: atmospheric the glycerol dehydration products.
Cat: Nano g-Al2O3 based (SiW20-Al/Zr10)
a
Activation energy related to the glycerol to acrolein step.
b
Unit [m3/kg-cat. S].
802 A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804

1/T (1/K) 3.4. Diffusion studies


0.0016 0.00165 0.0017 0.00175 0.0018 0.00185
-4 3.4.1. Internal diffusion
ln (kG/T) According to BET results the pellet porosity, ε ¼ 0.7, and using
Eqn (23) the tortuosity (t) is estimated to be about t ¼ 1.56 [29].
ln (k1/T) Therefore, the effective diffusivity (Deff) is equal to
y = -4968.x + 3.632 ln (k2/T) Deff ¼ 4.12  108 m2/s. In order to evaluate the existence of in-
-5 ternal diffusion, all the unknowns should be calculated at various
ln (kA/T)
catalyst pellet diameter. Consequently, the CWP values (Table 5)
ln (k3/T) indicate no internal diffusion for pellet diameter dp < 5 mm, but
y = -4881.x + 3.359
ln (k4/T) internal diffusion limits the reaction rate if dp > 5 mm. Fig. 6a and b
-6 ln (k5/T) show the interaction between different pellet diameter (dp) and
Thiele modules (ɸ1) with the effectiveness factors (h), respectively.
ln (k/T)

In addition micrographs of pellet diameter and spherical shapes,


based on FESEM, are depicted in Fig. 6c.
-7 y = -5729.x + 2.778

3.4.2. External diffusion


As mentioned earlier, to calculate the CM, we first need to
y = -5207.x + 1.697
determine the mass transfer coefficient (kc). The procedure begins
-8 with calculation of superficial velocity (U) which is obtained by Eq.
y = -423.5x - 7.319 (25). The reactant flow rate (n0) is 3:33  107 m3/s and also the
cross sectional area (Ac) of the reactor with ID ¼ 1.1 cm is equal to
y = -473.9x - 7.581
Ac ¼ 9:5  105 m2. Finally, the superficial velocity (U) is equal to
-9 3:51  105 m/s. The other unknown factor, used for calculation of
Schmidt number (Sc), is kinematic viscosity (y) equals to
y = -330.9x - 8.673

-10

Fig. 5. Eyring plot of the reaction rate constants and temperatures for estimation of
DHz and DSz

DSz shows the extent of desire between transition and ground


state of a reaction. In fact, lower DSz value reveals that the reaction
requires less energy and can perform easily [43]. In fact, the second
law of thermodynamic indicated that ease of reaction is associated
with the lower DSz value. For instance, glycerol dehydration to
acrolein (R1) possess lower DSz ¼ 169.6 J/mol.K value compared
to the glycerol dehydration to acetol (R2) with DSz ¼ 174.5 J/
mol.K. The other by products in R3 to R5 reactions exhibited very
large (even 2 times larger) DSz values compared to the glycerol
dehydration to acrolein (R1) reaction. The standard Gibbs free en-
ergy change (DGz) can be found from the standard values of
enthalpy and entropy according to Eq. (48). The DGz values are
reported in Table 3.

DGz ¼ DHz  D T$DSz (48)


The Gibbs free energy of activation revealed the feasibility and
ease of reaction, as the lower the free energy the more the feasi-
bility and vice-versa [43]. Indeed, the lower DGz value for the for-
mation of acrolein (91.20 kJ/mol) confirms this reaction is more Fig. 6. (a) Dependency of effectiveness factor (h) versus Thiele modules (ɸ1), (b) Effect
feasible compared to the formation of acetol with DGz equal to of different catalyst pellet diameters (dp) on obtained effectiveness factor (h), and (c)
99.65 kJ/mol. various catalyst pellet diameters (dp) based on FESEM estimation.

Table 5
Results of internal and external diffusions based on Wiesz-Prater Criterion (CWP), Mear's Criterion (CM) and overall effectiveness factor amount at various pellet diameters (dp).

No dP (mm) ɸ1 h CWP Re’ Sc Sh’ kc (m/s) CM ac (m2/m3) U


1 0.5 0.046 1 0.002 4.465  105 583.38 0.056 0.029 0.009 8841600 1
2 2 0.185 1 0.034 1.79  104 583.38 0.112 0.014 0.07 2210400 1
3 5 0.460 1 0.212 4.465  104 583.38 0.180 9.22  103 0.28 884160 0.9
4 20 1.853 0.83 2.850 1.79  103 583.38 0.354 4.53  103 2.30 221040 0.54
A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804 803

m
5:335  105 m2/s based on equation of y ¼ r steam@300 C . The calcu- to acrylic acid over Mo/V and W/V oxide catalysts, Chem. Eng. J. 244 (2014)
steam@300 C 168e177.
lated values of Re’, Sc, Sh’, and kc at various pellet diameters and [8] R. Estevez, S. Lopez-Pedrajas, F. Blanco-Bonilla, D. Luna, F.M. Bautista, Pro-
based on Eqs (26-29) are summarized in Table 5. Results indicate duction of acrolein from glycerol in liquid phase on heterogeneous catalysts,
Chem. Eng. J. 282 (2015) 179e186.
that through Mear's method the existence of external mass transfer [9] T. Ma, Z. Yun, W. Xu, L. Chen, L. Li, J. Ding, R. Shao, Pd-H3PW12O40/Zr-MCM-41:
limitations is confirmed for catalysts with pellet diameter, an efficient catalyst for the sustainable dehydration of glycerol to acrolein,
dp > 5 mm as CM is greater than 0.15 (CM > 0.15). Chem. Eng. J. 294 (2016) 343e352.
[10] A. Alhanash, E.F. Kozhevnikova, I.V. Kozhevnikov, Gas-phase dehydration of
glycerol to acroleincatalysed by caesium heteropolysalt, Appl. Catal. A Gen.
3.4.3. Overall effectiveness factor 378 (2010) 11.
Table 1 clearly reports the SiW20-Al/Zr10 sample exhibits just [11] A. Talebian-Kiakalaieh, N.A.S. Amin, Coke-tolerance SiW20-Al/Zr10 catalyst for
glycerol dehydration to acrolein, Chin. J. Catal. (2017) in press.
1.3 wt% coke content on the catalyst surface after a 3 h reaction.
[12] Y.S. Park, T.Y. Yun, K.R. Kim, J. Lee Baek, J. Yi, Kinetics of the dehydration of
Indeed, a catalyst tolerate about 6e10 wt% coke content with glycerol over acid catalysts with an investigation of deactivation mechanism
limited amount of activity loss, particularly when coke is deposited by coke, Appl. Catal. B Environ. 176 (2015) 1.
[13] A. Talebian-Kiakalaieh, N.A.S. Amin, Kinetic modeling, thermodynamic, and
on inactive sites [45,46]. Thus, the effect of coke content on the
mass-transfer studies of gas-phase glycerol dehydration to acrolein over
effectiveness factor can be neglected. supported silicotungstic acid catalyst, Ind. Eng. Chem. Res. 54 (2015) 8113.
First, the external area per unit reactor volume (ac) is deter- [14] M. Watanabe, T. Iida, Y. Aizawa, T.M. Aida, H. Inomata, Acrolein synthesis from
mined based on Eq (31) in order to calculate the overall effective- glycerol in hot-compressed water, Bioresour. Technol. 98 (2007) 1285.
[15] L. Ott, M. Bicker, H. Vogel, Catalytic dehydration of glycerol in sub- and su-
ness factor (U). Finally, the overall effectiveness factor, (U) (Eqn percritical water: a new chemical process for acrolein production, Green.
(30)) for various h, kc, and ac are calculated and the results tabu- Chem. 8 (2006) 214.
lated in the last column of Table 5. Only catalyst with pellet [16] L. Qadariyah, Mahfud Sumarno, S. Machudah, S. Wahyudiono, M. Sasaki,
M. Goto, Degradation of glycerol using hydrothermal process, Bioresour.
diameter of dp < 5 mm have no external diffusion due to U > 0.9. In Technol. 102 (2011) 9267.
fact, dp < 5 mm led to full utilization of catalyst surface due to the [17] J.V. Belleghem, D. Constales, J.W. Thybaut, G.B. Marin, Numerical methods and
absence of mass transfer limitations (U > 0.9) inside the pore Fischer-Tropsch complex reaction network generation for steady state iso-
topic transient kinetic analysis, Com. Chem. Eng. (2016), http://dx.doi.org/
structure in the pellets. 10.1016/j.compchemeng.2016.07.003.
[18] A. Talebian-Kiakalaieh, N.A.S. Amin, Z.Y. Zakaria, Gas phase selective conver-
4. Conclusion sion of glycerol to acrolein over supported silicotungstic acid catalyst, J. Ind.
Eng. Chem. 34 (2016) 300.
[19] H. Atia, U. Armbruster, A. Martin, Dehydration of glycerol in gas phase using
The thermo-kinetic study of gas-phase glycerol dehydration heteropolyacid catalysts as active compounds, J. Catal. 258 (2008) 71.
reaction using supportedg-Al2O3 nanoparticle based solid catalyst [20] S.M. Baek, J.H. Kang, K. Lee, J.H. Nam, A numerical study of the effectiveness
factors of nickel catalyst pellets used in steam methane reforming for resi-
(SiW20-Al/Zr10) has been investigated. The kinetic study reveals all
dential fuel cell applications, Int. J. Hydrog. Energy 39 (2014) 9180.
the rate constants increase with temperature, and the activation [21] R. Klaewkla, M. Arend, W.F. Hoelderich, A review of mass transfer controlling
energies of glycerol dehydration to acrolein and acetol are 46.0 and the reaction rate in heterogeneous catalytic systems, mass transfer e
53.3 kJ/mol, respectively. The thermodynamic parameters (DHz, advanced aspects, in: Hironori Nakajima (Ed.), ISBN: 978-953-307-636-2,
InTech, 2011, http://dx.doi.org/10.5772/22962. Available from: http://www.
DSz, and DGz) are determined for all the reactions (R1 to R5) by intechopen.com/books/masstransfer-advanced-aspects/a-review-of-mass-
applying the Eyring equation. The obtained reaction rate constants, transfer-controlling-the-reactionrate- in-heterogeneous-catalytic-systems.
activation energies, enthalpies and entropies are essential for [22] R.V. Sharma, P. Kumar, A.K. Dalai, Selective hydrogenolysis of glycerol to
propylene glycol by using Cu:Zn:Cr: Zr mixed metal oxides catalyst, Appl.
reactor modeling and simulation studies in future. Finally, the Catal. A Gen. 477 (2014) 147.
diffusion analysis confers the absence of internal and external dif- [23] R. Sharma, P.N. Sheth, A.M. Gujrathi, Kinetic modeling and simulation: py-
fusions limitations over pellet diameter dp < 5 mm. rolysis of Jatropha residue de-oiled cake, Renew. Energy 86 (2016) 554e562.
[24] P.B. Weisz, C.D. Prater, Interpretation of measurements in experimental
catalysis, Adv. Catal. 6 (1954) 143.
Acknowledgments [25] H.S. Fogler, Elements of Chemical Reaction engineering.4thedition, Pearson
International Edition, Upper Saddle River, New Jersey, US, 2009.
[26] A. Talebian-Kiakalaieh, N.A.S. Amin, Theoretical and experimental evaluation
The authors would like to express their sincere gratitude to the of mass transfer limitation in gas phase dehydration of glycerol to acrolein
Ministry of Science, Technology and Innovation (MOSTI), Malaysia over supported HSiW catalyst, J. Taiwan. Inst. Chem. Eng. 59 (2016) 11.
for supporting the project under project no. 03-01-06-SF0963. Also, [27] L.K. Chan, A.F. Sarofim, J.M. Beer, Combust. Flame 52 (1983) 37.
[28] L.E. Bas, The Molecular Volumes of Liquid Chemical Compounds, David McKay
the authors are grateful to Dr. N.A.S. Ramli and Dr. M. Akizuki for Co., Inc., New York, 1915.
their assistance during the calculation of ODE's using Polymath and [29] C.J. Geankoplis, Transport Processes and Separation Process Principles, 4thE-
Excel. dition, Pearson Education International, Upper Saddle River, New Jersey. US,
2009.
[30] D.E. Mears, Ind. Eng. Chem. Process Des. Dev. 10 (1971) 541.
References [31] F. Wang, J. Dubois, W. Ueda, Catalytic dehydration of glycerol over vanadium
phosphate oxides in the presence of molecular oxygen, J. Catal. 268 (2009)
[1] A.B. Leoneti, V. Aragao-Leoneti, S.V.W.B. de Oliveira, Glycerol as a by-product 260.
of biodiesel production in Brazil: alternatives for the use of unrefined glycerol, [32] E. Tsukuda, S. Sato, R. Takahashi, T. Sodesawa, Production of acrolein from
Renew. Energy 45 (2012) 138e145. glycerol over silica supported heteropolyacids, Catal. Commun. 8 (2007) 1349.
[2] M. Hunsom, P. Saila, Electrochemical conversion of enriched crude glycerol: [33] A. Corma, G.W. Huber, L. Sauvanaud, P. O'Connor, Biomass to chemicals:
effect of operating parameters, Renew. Energy 74 (2015) 227e236. catalytic conversion of glycerol/water mixtures into acrolein, reaction
[3] M.N.N. Shahirah, J. Gimbun, A. Ideris, M.R. Khan, C.K. Cheng, Catalytic pyrolysis network, J. Catal. 257 (2008) 163.
of glycerol into syngas over ceria-promoted Ni/a-Al2O3 catalyst, Renew. En- [34] J. Deleplanque, J.L. Dubios, J.F. Devaux, W. Udea, Production of acrolein and
ergy 107 (2017) 223e234. acrylic acid through dehydration and oxydehydration of glycerol with mixed
[4] J.A. Sullivan, S. Burnham, The use of alkaline earth oxides as pH modifiers for oxide catalysts, Catal. Today 157 (2010) 351.
selective glycerol oxidation over supported Au catalysts, Renew. Energy 78 [35] M.H. Haider, N.F. Dummer, D. Zhang, P. Miedziak, T.E. Davies, S.H. Taylor,
(2015) 89e92. D.J. Willock, D.W. Knight, D. Chadwick, G.J. Hutchings, Rubidium- and
[5] A.S. de Oliveiraa, S.J.S. Vasconcelosa, J.R. de Sousab, F.F. de Sousacd, J.M. Filhoc, caesium-doped silicotungstic acid catalysts supported on alumina for the
A.C. Oliveira, Catalytic conversion of glycerol to acrolein over modified mo- catalytic dehydration of glycerol to acrolein, J. Catal. 286 (2012) 206.
lecular sieves: activity and deactivation studies, Chem. Eng. J. 168 (2011) [36] B. Katryniok, S. Paul, M. Capron, C. Lancelot, V. Bellere-Beca, P. Rey,
765e774. F. Dumeignil, A long-life catalyst for glycerol dehydration to acrolein, Green.
[6] L. Shen, H. Yin, A. Wang, Y. Feng, Y. Shen, Z. Wu, T. Jiang, Liquid phase Chem. 12 (2010) 1922.
dehydration of glycerol to acrolein catalyzed by silicotungstic, phospho- [37] I. Martinuzzi, Y. Azizi, J. Devaux, S. Tretjak, O. Zaharaa, J.P. Leclerc, Reaction
tungstic, and phosphomolybdic acids, Chem. Eng. J. 180 (2012) 277e283. mechanism for glycerol dehydration in the gas phase over a solid acid catalyst
[7] L. Shen, H. Yin, A. Wang, X. Lu, C. Zhang, Gas phase oxidehydration of glycerol determined with on-line gas chromatography, Chem. Eng. Sci. 116 (2014) 118.
804 A. Talebian-Kiakalaieh, N.A.S. Amin / Renewable Energy 114 (2017) 794e804

[38] M. Akizuki, Y. Oshima, Kinetics of glycerol dehydration with WO3/TiO2 in based acrolein production, Chem. Sustain. Chem. 5 (2012) 1162.
supercritical water, Ind. Eng. Chem. Res. 51 (2012) 12253. [43] K.C. Badgujar, B.M. Bhanage, Thermo-chemical energy assessment for pro-
[39] S. Thanasilp, J.W. Schwank, V. Meeyoo, S. Pengpanich, M. Hunsom, Prepara- duction of energy-rich fuel additive compounds by using levulinic acid and
tion of supported POM catalysts for liquid phase oxydehydration of glycerol to immobilized lipase, Fuel Proc. Technol. 138 (2015) 139.
acrylic acid, J. Mol. Catal. A Chem. 380 (2013) 49. [44] X. Yuan, T. He, H. Cao, Q. Yuan, Cattle manure pyrolysis process: kinetic and
[40] A. Talebian-Kiakalaieh, N.A.S. Amin, Supported silicotungstic acid on zirconia thermodynamic analysis with isoconversional methods, Renew. Energy 107
catalyst for gas phase dehydration of glycerol to acrolein, Catal. Today 256 (2017) 489e496.
(2015) 315. [45] P.G. Menon, Coke on catalysts. harmful, harmless, invisible and benefical
[41] N.A.S. Ramli, N.A.S. Amin, Kinetic study of glucose conversion to levulinic acid types, J. Mol. Catal. 59 (1990) 207e220.
over Fe/HY zeolite catalyst, Chem. Eng. J. 283 (2016) 150. [46] Y. Yang, H. Wang, The role of H2 in n-butane isomerization over Al-promoted
[42] L. Liu, X.P. Ye, J.J. Bozell, A comparative review of petroleum-based and bio- sulfated zirconia catalyst, J. Mol. Catal. A Chem. 304 (2009) 65e70.

You might also like