Sei sulla pagina 1di 7

Catalytic etherication of glycerol by tert-butyl alcohol to produce oxygenated

additives for diesel fuel


F. Frusteri
a,
*, F. Arena
b
, G. Bonura
a
, C. Cannilla
b
, L. Spadaro
a
, O. Di Blasi
a
a
CNR-ITAE, Istituto di Tecnologie Avanzate per lEnergia Nicola Giordano, Via S. Lucia 5, 98126 Messina, Italy
b
Dip. Chimica Industriale e Ingegneria dei Materiali, Universita` di Messina, Salita Sperone 31, 98166 Messina, Italy
1. Introduction
Recently, exhaust gases emitted by internal combustion
engines were considered primarily responsible for environmental
pollution and human diseases [15].
Biodiesel is currently used as a valuable fuel for diesel engines.
In spite of a slight power loss, exhausts contain less particulate
matter. Biodiesel is a mixture of methyl esters of fatty acids
(FAMEs) obtained by the transesterication reaction of vegetable
oils with methanol in presence of a basic catalyst [1,2]. Such a
catalytic process converts raw triglycerides into FAMEs, but
produces glycerol as side product.
If biodiesel is produced on a large scale, adequate technologies
capable of converting glycerol into added value chemicals are
necessary [3,4]. In particular, great attention has been already
devoted to the conversion of glycerol into oxygenated additives for
liquid fuels [3,57]. In this context, an industrially relevant route
for the conversion of glycerol into oxygenated chemicals involves
the etherication to tert-butyl ethers [5,710].
It is well known that the addition of oxygenated additives to
diesel fuels could represent a promising way enhancing the
combustion efciency in internal combustion engines with a
signicant reduction of pollutant emissions. Among several
oxygenated additives proposed to blend with diesel, the ethers of
glycerol could hold a prominent role [710]. In particular, tert-butyl
ethers of glycerol with a high content of di-ethers are considered
promising as oxygenated additives for diesel fuels (diesel, biodiesel
and their mixtures). However, mono-tert-butyl ethers of glycerol
(MBGEs) have a low solubility in diesel fuel; therefore, in order to
avoid an additional separation step, the etherication of glycerol
should address the formation of di- and tri-ethers [5,79,11].
The etherication of glycerol can be carried out using
heterogeneous acid catalysts like strong acid ion-exchange resins
[1218]; however, the utilization of large-pore zeolites has also
been widely investigated [1922]. Usually, low surface areas and
lack of thermal stability are the major drawbacks of sulfonic resins.
The incorporation of organosulfonic groups over mesostructured
silicas have generated effective solid acid catalysts with enhanced
catalytic properties as compared with conventional homogeneous
and heterogeneous acid catalysts [23]. Moreover, these type of
silica materials functionalized with organosulfonic acid groups
have been used previously for the conversion of biorenewable
molecules [2427], showing better catalytic performances than
that of the commercial sulfonated resins. Currently, these highly
surface materials characterized by interconnected mesopores and
high accessibility of acid sites represent the best systems for the
etherication reactions [28].
The synthesis of tert-butyl ethers (GTBEs) from isobutene and
glycerol on ion-exchange resins has been already investigated
extensively [810]. Isobutene (IB) is produced by catalytic cracking
and steam cracking fractions of petroleum rening and by
isobutane dehydrogenation [5,7]. What appears more attractive
Applied Catalysis A: General 367 (2009) 7783
A R T I C L E I N F O
Article history:
Received 5 May 2009
Received in revised form 23 July 2009
Accepted 24 July 2009
Available online 3 August 2009
Keywords:
Oxygenated compounds
Catalytic conversion
Glycerol ethers
Biodiesel
A B S T R A C T
The heterogeneous catalytic etherication of glycerol with tert-butyl alcohol was investigated in
presence of lab-made silica supported acid catalysts. As reference, two commercial acid ion-exchange
resins were also used. Experiments were carried out in batch mode at T
R
ranging from 303 to 363 K. An
increase in reaction temperature favors the formation of di-substituted ethers. The etherication
reaction proceeds according to a consecutive path and the surface reaction between adsorbed glycerol
and protonated tert-butanol (tertiary carbocation) can be considered as the rate determining step. Steric
hindrance phenomena and water hinder the formation of tri-substituted ether (TBGE). As expected,
water removal was necessary to allow the higher ethers formation. The specic activity (turnover
frequency, TOF) of A-15 catalyst is signicantly higher than that of the other studied acid systems, due to
the wide pore diameter that allows an easier accessibility of the reagent molecules.
2009 Elsevier B.V. All rights reserved.
* Corresponding author.
E-mail address: francesco.frusteri@itae.cnr.it (F. Frusteri).
Contents lists available at ScienceDirect
Applied Catalysis A: General
j our nal homepage: www. el sevi er . com/ l ocat e/ apcat a
0926-860X/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2009.07.037
is to produce GTBEs by the glycerol etherication reaction in a
solidliquid catalytic process, by using tert-butyl alcohol (TBA). In
fact, the use of TBA, as both reactant and solvent, instead of gaseous
isobutylene, allows to overcome the technological problems
arising from the need to use solvents able to dissolve glycerol
(i.e., dioxane, dimethyl sulfoxide) and typical drawbacks of a
complex three-phase system (mass transfer phenomena) [9,29].
This study focused on the etherication of glycerol with tert-
butyl alcohol over different solid acid systems. Attention was given
primarily to investigate the main limiting factors for large scale
process development.
2. Experimental
2.1. Catalysts and chemicals
Two solid acid supported catalysts were prepared by the
incipient wetness method using a silica carrier (S.A.
BET
,
250 m
2
g
1
) and two solutions containing 17 wt.% of Naon
1
ionomer (N-17) and 17 wt.% of tungstophosphoric heteropoly acid
(HPW-17), respectively. An aliquot of the HPW-17 sample was
mixed with a solution containing cesium to exchange a fraction of
H
+
protons with Cs
+
(Cs-HPW). In addition, two commercial acid
ion-exchange resins, Naon
1
on amorphous silica (SAC-13) and
Amberlyst
1
15 dry (A-15), were used as reference catalysts.
Glycerol anhydrous (purity 99.5%) and tert-butyl alcohol
(purity 99.7%), supplied by Fluka (Buchs, Switzerland), were used
as reactants. Standard compounds for GC analysis were supplied
on-demand by Aldrich.
2.2. Catalysts characterization
Surface area (SA
BET
) and pore volume (PV) were determined by
the nitrogen adsorption/desorption isotherms at 77 K using a Carlo
Erba (Sorptomatic Instrument) gas adsorption device. Before
analysis, all the samples were outgassed at 423 K under vacuum
for 2 h. The isotherms were elaborated according to the BET
method for surface area calculation, with the HorwarthKavazoe
(HK) and BarrettJoynerHalenda (BJH) methods used for micro-
pore and mesopore evaluation, respectively.
The active phase loading and thermal stability were evaluated
by thermo-gravimetric (TG) analyses in the range 293873 K using
a Netzsch STA409C analyzer, running in air atmosphere with a
heating rate of 10 K/min.
Acid sites were determined by potentiometric titrations with
the zero point charge (ZPC) method. About 0.1 g of each sample
was dispersed in an aqueous solution of NaNO
3
0.5 M under
stirring. Acidity, measured by an electrode Orion ROSS, was
calculated on the basis of the pH at which particles suspended in
solution had zero charge.
The list of catalysts used in this study is summarized in Table 1.
2.3. Catalytic testing
The etherication reaction between glycerol and tert-butyl
alcohol was carried out in liquid phase in a 100 cm
3
stainless steel
jacketed-batch reactor (Autoclave Engineers, Inc.) under a stirring
frequency of 1200 min
1
, in order to limit the inuence of external
mass transfer phenomena. Experiments were performed under
different reaction conditions: (i) under pressure; (ii) at reaction
temperatures ranging from 303 to 363 K; (iii) by operating at
different reaction times and (iv) at both different catalyst/glycerol
and alcohol/glycerol ratios.
The experimental procedure was the following: a well dened
amount of glycerol and dry catalyst were loaded into the reactor
and heated up to a prexed reaction temperature (in 10 min).
Before the addition of tert-butanol, the reactor was uxed with
nitrogen to remove the air; then, tert-butanol was injected into the
reactor by a syringe: this was taken as the starting point of the
reaction. At the end of the experiments, the reactor was cooled
down (at 298 K) by an ice-bath until the vapour pressure of the
mixture turned down to the atmospheric one, thus allowing all the
gas phase compounds to condense. After the opening of the
autoclave, we collected a completely liquid mixture, without any
solidication of TBA on the reactor walls. The liquid reaction
mixture was analyzed off-line by a gas chromatograph, HP 6890N,
provided with a capillary column HP Innowax (l, 30 m; i.d.,
0.53 mm; lm thickness, 1.0 mm) under the following oven
temperature program: from 40 to 220 8C (with a heating rate of
20 8C min
1
) and at 220 8C for 3 min. An automatic sampler Agilent
7683B Series was used (0.2 ml of samples were injected), each data
set being obtained, with an accuracy of 1%, from an average of
three independent measurements using the external standard
method (n-heptane, 8 wt.% in respect to the reaction mixture). The
samples were not collected at different reaction times for avoiding to
spill out a not representative sample of the reaction systemdue to the
different density of the mixture compounds. Water content was
calculated by considering the reaction stoichiometry and conrming
the result by a quantitative TCD analysis. In each experiment, carbon
balance was close to 98%.
3. Results and discussion
Physico-chemical properties of catalysts are summarized in
Table 2.
Catalysts are characterized by different surface area ranging
from 53 to 207 m
2
g
1
and porosity comprised between 0.07 and
0.80 cm
3
g
1
. In terms of SA and porosity, the data obtained with
SAC-13 sample are similar to that provided by the supplier.
Furthermore, all the catalytic systems show an average pore
diameter (APD) increasing with the porosity, apart from A-15 that
is characterized by a wide pore texture (300 A

), although it has the


lowest SA (53 m
2
g
1
). The acid capacity of A-15 sample was
signicantly higher than the other investigated samples.
In order to collect quantitative data approaching the equili-
brium composition and to obtain a reliable comparison of the
catalytic functionality at a prexed time (6 h), preliminary
experiments using different solid acid catalysts were carried out
at 0.1 MPa, 343 K, with a tert-butanol-to-glycerol molar ratio (R
A/G
)
equal to 4. The results shown in Fig. 1 demonstrate that a low
catalyst/glycerol ratio equivalent to 1.2 wt.% (much lower than
that so far reported in literature [710]) is adequate to guarantee
high glycerol conversion. Indeed, after 6 h of reaction, low glycerol
conversion levels were reached using SAC-13, N-17 and HPW-17
catalysts (815 mol.%), while by the Cs-HPW and A-15 samples,
characterized by higher acid capacity (Table 2), the reaction takes
place at higher rates, reaching at the end of the reaction a glycerol
conversion of 54 and 82 mol.%, respectively.
Considering that catalysts are characterized by different acid
capacity, the turnover frequency (TOF) of glycerol was reported as
a function of the number of acidic sites (see Fig. 2). TOF values were
determined fromthe initial reaction rate (glycerol conversion <5%)
Table 1
List of solid acid catalysts used in this study.
Code Active phase Carrier
HPW-17 Phosphotungstic acid SiO
2
(Cabosil LM50; S.A.
BET
=250m
2
g
1
)
Cs-HPW Phosphotungstic acid
exchanged with cesium
SiO
2
(Cabosil LM50; S.A.
BET
=250m
2
g
1
)
N-17 Naon
1
polymer SiO
2
(Cabosil LM50; S.A.
BET
=250m
2
g
1
)
SAC-13 Naon
1
polymer Amorphous silica
A-15 Amberlyst
1
-15
dry resin

F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 78


and normalized to the number of acid sites. As it can be seen, the
TOF appears to be comparable on all the systems, except for the A-
15 sample that exhibits a TOF value about 3 times higher. At rst
glance, this result could indicate that the glycerol etherication is a
structure sensitive reaction; however, what should be considered in
a reaction involving encumbered molecules is also the accessibility
of active sites. As reported in Table 2, it can be seen that A-15
catalyst is characterized by an average pore diameter (APD) much
higher (300 A

) in respect to the other catalysts (APD, 35155 A

);
then, the higher TOF observed for A-15 catalyst could not be due to
the higher specic activity of acidic sites but to the accessibility of
sites which is favored using macro-porous materials like Amber-
lyst. In conclusion, the ethericationof glycerol with TBAshould be
considered as an unstructured sensitive reaction.
On the basis of such results and considering that the aim of this
study was to explore the feasibility of the process based on the
conversion of glycerol with TBA, we have considered more
protable to evaluate the inuence of experimental parameters
using A-15 catalyst characterized by the highest acid capacity.
The inuence of reaction temperature on glycerol conversion
and products distribution is shown in Fig. 3. Glycerol conversion
linearly increased with reaction temperature growing from 10% at
303 K up to 85% at 343 K. It can be observed that, at high
temperatures (363 K), glycerol was not totally converted due to the
occurrence of de-etherication reactions, whichbecome important
as the reaction temperature increases [710,28].
In terms of products distribution, it is necessary to consider
that, according to literature evidences [710], the etherication
reaction proceeds according to a consecutive path and the
formation of ve different alkyl glycerol ethers could be expected:
3-tert-butoxy-1,2-propandiol (1-MBGE), 2-tert-butoxy-1,3-pro-
pandiol (2-MBGE), 1,3-di-tert-butoxy-2-propanol (1,3-DBGE),
1,2-di-tert-butoxy-3-propanol (1,2-DBGE) and tri-tert-butoxy-
propane (TBGE).
Taking into account the results reported in Fig. 3B, when glycerol
reacts with TBA, depending on the extent of the etherication
reaction, only the formation of four ethers was observed: two mono-
substitutedethers (1-MBGE and/or 2-MBGE) and two di-substituted
ethers (1,3-DBGE and/or 1,2-DBGE). Tri-substituted ether never
formed in the whole range of temperature investigated.
3-Butoxy-1,2-propandiol (1-MBGE) was the main mono-ether
formed and, depending on the reaction temperature, even to
different extent, its concentration was always much higher than
the correspective 2-tert-butoxy-1,3-propandiol (2-MBGE).
Indeed, since the reaction takes place between the tert-butyl
cation (tertiary carbocation) and glycerol [10,30], it is probable
that initially the electrophilic attack occurs on the primary carbon
of glycerol due to steric hindrance and electrostatic effects exerted
by OH groups of glycerol. As regards the formation of di-ethers,
1,3-di-tert-butoxy-2-propanol (1,3-DBGE) was the main com-
pound formed (its concentration linearly increased with reaction
temperature) while the concentration of 1,2-di-tert-butoxy-3-
propanol (1,2-DBGE) reached a maximum value of 7% at 343 K,
after which it remained almost constant. Then, the formation of
1,3-DBGE is favored for the same reasons noted above (steric
encumbrance and electrostatic effect).
Besides, it is also important to consider that, by dehydration of
TBA, isobutene could be formed. To eventually increase the
concentration of isobutene in liquid phase, experiments, under
pressure as a function of the reaction time, were carried out. The
results obtained by operating at 0.1 and 1.0 MPa are reported in
Fig. 4, in terms of glycerol conversion (A) and selectivity to GTBEs
(B).
It can be seen that, as the reaction proceeds, the glycerol
conversion is always higher by operating at 1.0 MPa. The most
Table 2
Physico-chemical properties of the studied samples.
Catalyst Loading (wt.%) of active phase S.A.
BET
(m
2
g
1
) P.V. (cm
3
g
1
) APD (A

) Acidity (mmol
H+
g
1
)
HPW-17 15 200 0.62 124 0.74
Cs-HPW 0.5 (Cs) 207 0.80 155 0.81
N-17 17 81 0.07 35 0.31
SAC-13 13 189 0.52 110 0.15
A-15
a
53 0.40 300 4.70
a
Data supplied by Rohm and Haas.
Fig. 1. Conversion of glycerol on solid acid catalysts: T
R
= 343 K; P
R
= 0.1 MPa; R
A/
G
= 4.0; cat = 1.2 wt.%/glycerol; reaction time = 6 h.
Fig. 2. TOF of glycerol as a function of the catalyst acidity: T
R
= 343 K; P
R
= 0.1 MPa;
R
A/G
= 4.0; reaction time: 2 h.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 79
evident pressure effect was observed after 2 h of reaction: the
glycerol conversion values increases from 53% (at 0.1 MPa) to 72%
(at 1.0 MPa).
Furthermore, after 6 h, no differences were seen in terms of
glycerol conversion, since the reacting system reaches the
equilibrium. In terms of selectivity to ethers, apart from the
reaction pressure, a progressive decreasing in MBGEs concentra-
tion can be observed as a function of the reaction time. In any case,
no TBGE forms due to steric hindrance.
Therefore, considering that the etherication reaction occurs at
higher rate under pressure, the reaction mechanismshould involve
the electrophilic attack of a tertiary carbocation, formed through
the protonation of both TBA and isobutene, to the adsorbed
glycerol, to formethers. Indeed, the tert-butyl cation concentration
in liquid phase is ruled out by a dynamic equilibrium existing
between TBA and isobutene. Naturally, the concentration of
isobutene in liquid phase is higher by operating under pressure.
To assess how the products distribution and yield to higher
ethers could be affected by the amount of catalyst used, several
experiments were carried out with different catalyst/glycerol
weight ratios. Results reported in Table 3 clearly point out that
without catalysts the reaction does not take place, while a
signicant rise in the glycerol ether formation can be observed as
the amount of catalyst increases. In particular, the highest
cumulative yield (28%) to higher ethers of glycerol (DBGEs + TBGE)
was obtained with an amount of catalyst equivalent to 7.5 wt.%
with respect to the glycerol weight.
Moreover, it is noteworthy that increasing the catalyst weight,
the conversion of glycerol becomes strongly limited by the
growing formation of water (see Fig. 5), which negatively affects
the etherication equilibrium. It is also important to consider that,
under the reaction conditions, a certain amount of isobutene (<5%)
forms due to tert-butanol dehydration. However, we did not
observe the formation of oligomers that normally form when
Fig. 3. Glycerol conversion and products distribution vs. reaction temperature: A-15 catalyst; P
R
= 0.1 MPa; R
A/G
= 4.0; cat = 1.2 wt.%/glycerol; reaction time = 6 h.
Fig. 4. Glycerol conversion (A) and selectivity to ethers (B) at different reaction pressure as a function of time: A-15 = 1.2 wt.%/glycerol; T
R
= 343 K; rpm = 1200 min
1
.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 80
isobutylene is used as the reactant [710,28], because the
formation of water should account for hindering the extent of
oligomerisation.
To evaluate how the reaction time affects the evolution of
reaction and catalyst activity, several experiments were carried out
at 1, 2, 6, 24 and 30 h, using a catalyst-to-glycerol weight ratio of
7.5%. The results obtained are shown in Fig. 6. After 2 h of reaction,
glycerol was almost totally converted and the 1-MBGE was the
main product.
As the reaction proceeded, the concentration of MBGE
decreased and the corresponding concentration of di-ether (DBGE)
increased. This transformation occurred with a low reaction rate;
in fact, after 30 h, MBGEs selectivity was still 50%.
Experimental data obtained by operating at different tert-
butanol/glycerol molar ratios (R
A/G
) are reported in Table 4. It can
be seen that the increase of R
A/G
does not affect the reaction in
terms of products distribution.
As such result was not expected, reaction order with respect to
glycerol (GLY) and tert-butanol (TBA) was determined considering
initial glycerol conversion values (<5%), under the following
reaction conditions: (a) glycerol volume was changed from 5 to
10 ml, while TBA volume was maintained in a large excess; (b) TBA
volume was changed from 5 to 10 ml, while glycerol volume was
maintained in a large excess. In these conditions, the concentration
of the exceeding reactant can be considered almost invariant.
On the basis of the results obtained, logarithmic relationships
were established between the reaction rate and molar concentra-
tion of reactants, thus enabling the discovery of the following rate
law:
rate kGLY
0:3
TBA
1:7
(1)
From such an equation, the kinetic constant at each temperature
was calculated and results obtained are summarized in Table 5.
These data are thermodynamically consistent since k increased
with temperature as expected for an endothermic reaction. The
apparent activation energy was close to 70 3 kJ mol
1
, which well
matches the values so far reported in similar kinetic studies [31].
According to the rate Eq. (1), considering that the reaction order
referred to TBA is much higher than one, it can be conrmed that
the etherication reaction occurs with a molecular mechanism
Table 3
Inuence of amount of catalyst on product distribution and yield.
Catalyst (wt.%/glycerol) Product distribution (wt.%) Yield (%)
IB TBA MBGEs DBGEs TBGE GLY H
2
O DBGEs +TBGEs
0.0 0.0 76.0 0.0 0.0 0.0 23.4 0.1
0.3 2.8 52.4 18.7 7.1 0.0 13.4 4.0 10.7
1.2 4.9 28.6 36.6 13.7 0.0 6.4 7.5 16.7
7.5 4.2 16.6 41.4 25.0 0.4 2.6 8.7 28.2
A-15: T
R
=343K; P
R
=0.1MPa; reaction time =6h; rpm=1200min
1
.
Fig. 5. Glycerol conversion and water formation as a function of amount of catalyst
used: A-15 catalyst; P
R
= 0.1 MPa; R
A/G
= 4.0; reaction time = 6 h.
Fig. 6. Glycerol conversion and ethers yield as a function of reaction time: A-15
catalyst; P
R
= 0.1 MPa; R
A/G
= 4.0; catalyst = 7.5 wt.%/glycerol.
Table 4
Effect of tert-butanol-glycerol molar ratio (R
A/G
).
R
A/G
Conv. (%) Selectivity (%) Yield (%)
GLY 1-MBGE 2-MBGE 1,3-DBGE 1,2-DBGE TBGE DBGE+TBGE
2.0 93.1 67.8 1.6 22.4 7.6 0.6 28.5
4.0 93.6 68.0 1.8 20.8 8.9 0.4 28.2
5.0 95.3 70.3 2.0 18.7 9.1 0.0 26.5
A-15: 7.5wt.%/glycerol; T
R
=343K; P
R
=0.1MPa; reaction time =6h; rpm=1200min
1
.
Table 5
Rate constants at different temperature.
Catalyst k10
2
(M
1
min
1
)
303K 323K 343K 363K
A-15 0.5 2.2 13.8 45.0
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 81
wherein TBA is quickly protonated on the acid sites forming a
tertiary carbocation able to react with the glycerol strongly
adsorbed (partial order 0.3) on the catalyst surface.
In order to shift the reaction equilibrium and favor the
consecutive path of glycerol etherication, some experiments
were carried out to better assess how water affects catalytic
activity. On this account, an experiment was stopped at a reaction
time of 6 h and, after dehydration of the reaction mixture by
zeolites, the run was continued for further 6 hours, thus obtaining
a net increase in the yield of DBGEs from 28.5% (before water
removal) to 41.5%. This result clearly demonstrates that the
difculty in obtaining higher ethers is due to the presence of water,
as previously inferred. Indeed, it is noteworthy that water also
competes with tert-butanol and glycerol on the active site
adsorption. Really, the higher water acidity in relation to tert-
butanol results in a lowering of catalyst activity due to the
formation of solvated sites, as reported in the literature [3236].
Since the A-15 resin could swell during reaction due to the
inclusion of water in its polymeric structure, we decided to stop the
run after 24 h of reaction, substituting the used catalyst with a
same amount of fresh catalyst and continuing the experiment for
25 more hours. Results obtained are summarized in Table 6.
It is possible to observe that the replacement of the catalyst only
slightly favored the formation of higher ethers; therefore, the
difculty in obtaining high yields to higher ethers is not
attributable to catalyst deactivation but to the water formation.
However, just to conrm that catalyst does not change during the
reaction, an experiment employing the used catalyst, after its
ltration and regeneration by a drying procedure, was performed.
As Fig. 7 clearly shows, the results obtained with both fresh and
used catalysts are similar, conrming that the A-15 catalyst is
stable and even if it swells during the reaction, its catalytic
properties do not change signicantly.
Further attempts to elucidate how experimental procedures
could affect catalyst performance and to investigate if, during the
reactor heating, experimental conditions could change due to the
occurrence of thermal and catalytic reactions, different runs were
carried out under the following conditions: (i) after the mixing of
glycerol with the catalyst, the reactor was heated up to reaction
temperature and then alcohol was added; (ii) after the mixing of
alcohol with the catalyst, the reactor was heated up to reaction
temperature and then glycerol was added; (iii) after the heating of
the reactor, alcohol, glycerol and the catalyst were added at same
time.
Even if the alkylation agent is usually injected into the reactor
after mixing of glycerol and catalyst [810,28,37], in all the
investigated cases no differences were observed, either in terms of
glycerol conversion or products distribution, thus conrming that
the formation of ethers is strictly dependent upon the surface
reaction between adsorbed glycerol and tert-butyl cation, which
can be considered as the rate limiting step of the etherication
reaction.
4. Conclusions
The main ndings of this study can be summarized as follows:
the etherication of glycerol with tert-butyl alcohol effectively
takes place on solid acids catalysts, hence providing a promising
way to transform glycerol into value added products to be used
as oxygenated additives blended with diesel fuels;
the accessibility of acid sites plays a fundamental role in
promoting catalyst activity and the systems with large pores are
more indicate to perform reaction at high rate;
a low catalyst/glycerol ratio equivalent to 1.2 wt.%, much lower
than that so far reported in literature, is adequate to guarantee
high glycerol conversion;
the reaction pressure signicantly affects the reaction kinetic,
the tert-butyl cation concentration in liquid phase being ruled
out by a dynamic equilibrium existing between TBA and
isobutene;
no oligomers form during reaction since, under the reaction
conditions, the formation of water limits the extent of
oligomerisation reaction;
water formed during the reaction inhibits the glycerol ether-
ication and its removal from the reaction medium is necessary
for the formation of di- and tri-ethers. On this account, the design
of a water-separating reaction medium could increase the
formation of higher ethers.
References
[1] G. Knothe, J. Van Gerpen, J. Krahl (Eds.), The Biodiesel Handbook, AOCS Press,
Champaign, IL, 2005.
[2] M. Mittelbach, C. Remschmidt, BiodieselThe Comprehensive Handbook, M.
Mittelbach, Graz, Austria, 2004.
[3] S. Fernando, S. Adhikari, K. Kota, R. Bandi, Fuel 86 (2007) 28062809.
[4] J.M. Marchetti, V.U. Miguel, A.F. Errazu, Renew. Sust. Energy Rev. 11 (2007) 1300
1311.
[5] F. Ancillotti, V. Fattore, Fuel Process. Technol. 57 (1998) 163.
[6] P. Satge de Caro, Z. Mouloungui, G. Vaitilingom, J.Ch. Berge, Fuel 80 (2001) 565
574.
[7] R.S. Karinen, A.O.I. Krause, Appl. Catal. A: Gen. 306 (2006) 128133.
[8] K. Klepa cova , D. Mravec, E. Ha jekova , M. Bajus, Petrol. Coal 45 (2003) 5457.
[9] K. Klepa cova , D. Mravec, M. Bajus, Appl. Catal. A: Gen. 294 (2005) 141147.
[10] K. Klepa cova , D. Mravec, A. Kaszonyi, M. Bajus, Appl. Catal. A: Gen. 328 (2007)
113.
[11] J. Deutsch, A. Martin, H. Lieske, J. Catal. 245 (2007) 428435.
Table 6
Reaction advancement after catalyst replacing.
Reaction time (h) Conv. (%) Yield (%) Products distribution (wt.%)
GLY MBGEs DBGEs TBGE IB TBA MBGEs DBGEs TBGE GLY H
2
O
24.00 95.9 54.0 41.0 1.0 3.0 12.3 35.1 37.7 1.1 1.7 8.5
50.00
a
93.6 47.5 44.2 1.8 2.9 13.3 31.6 39.8 2.0 2.5 7.3
A-15: 7.5wt.%/glycerol; T
R
=343K; P
R
=0.1MPa; rpm=1200min
1
.
a
After 24h, the used catalyst was unloaded from the reactor and replaced by fresh catalyst.
Fig. 7. Products distribution recorded by using fresh and re-used A-15 catalysts.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 82
[12] A. Juha, A. Linnekoski, O.I. Krause, L.K. Struckmann, Appl. Catal. A: Gen. 170 (1998)
117126.
[13] J.F. Knifton, J.C. Edwards, Appl. Catal. A: Gen. 183 (1999) 113.
[14] Y. Pouilloux, S. Abro, C. Vanhove, J. Barrault, J. Mol. Catal. A: Chem. 149 (1999)
243254.
[15] B.-L. Yang, S.-B. Yang, R.-Q. Yao, React. Funct. Polym. 44 (2000) 167175.
[16] R. Alca ntara, L. Canoira, C. Fernandez-Martin, M.J. Franco, J.I. Martinez-Silva, A.
Navarro, React. Funct. Polym. 43 (2000) 97104.
[17] C. Park, M.A. Keane, J. Mol. Catal. A: Chem. 166 (2001) 303322.
[18] J.M. Adams, D.E. Clement, S.H. Graham, J. Chem. Res. (1981) S254.
[19] A.A. Chin, S.S.F. Wong, Zeolites 17 (56) (1996) 524.
[20] I. Hoek, et al. Appl. Catal. A: Gen. 266 (2004) 109116.
[21] K.J. Won, K.D. Jung, H.J. Uk, K. Min, K.J. Man, Y.J. Eui, Catal. Today 87 (14) (2003)
195203.
[22] P.S.E. Dai, J.F. Knifton, Zeolites 18 (56) (1997) 417418.
[23] J.A. Melero, R. van Grieken, G. Morales, Chem. Rev. 106 (2006) 3790.
[24] I.K. Mbaraka, D.R. Radu, V.S.-Y. Lin, B.H. Shanks, J. Catal. 219 (2003) 329.
[25] I.K. Mbaraka, B.H. Shanks, J. Catal. 229 (2005) 365.
[26] P.L. Dhepe, M. Ohashi, S. Inagaki, M. Ichikawa, A. Fukuoka, Catal. Lett. 102 (2005)
163.
[27] J.A. Bootsma, B.H. Shanks, Appl. Catal. A: Gen. 327 (2007) 44.
[28] J.A. Melero, G. Vicente, G. Morales, M. Paniagua, J.M. Moreno, R. Rolda n, A.
Ezquerro, C. Pe rez, Appl. Catal. A: Gen. 346 (2008) 4451.
[29] P.M. Somkiewicz, Appl. Catal. A: Gen. 313 (2006) 7485.
[30] R.A. Van Santen, P.V.N.M. Van Leeuwen, J.A. Moulijn, B.A. Averill, Stud. Surf. Sci.
Catal. 123 (1999).
[31] F. Cunill, M. Iborra, C. Fite , J. Tejero, J.F. Izquierdo, Ind. Eng. Chem. Res. 39 (2000)
12351241.
[32] F. Ancillotti, F. Mauri, E. Pescarollo, J. Catal. 46 (1977) 4957.
[33] F. Ancillotti, F. Mauri, E. Pescarollo, L. Romagnoni, J. Mol. Catal. 4 (1978) 3748.
[34] A. Gicquel, B. Torck, J. Catal. 83 (1983) 918.
[35] P. Rys, W.J. Steinegger, J. Am. Chem. Soc. 101 (1979) 48014806.
[36] W.J. Casey, D.J. Pietrzyk, Anal. Chem. 45 (1973) 14041407.
[37] D.E. Lo pez, J.G. Goodwin Jr., D.A. Bruce, J. Catal. 245 (2007) 381391.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 7783 83

Potrebbero piacerti anche