Sie sind auf Seite 1von 7

Applied Catalysis A: General 367 (2009) 77–83

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Catalytic etherification 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 l’Energia ‘‘Nicola Giordano’’, Via S. Lucia 5, 98126 Messina, Italy
b
Dip. Chimica Industriale e Ingegneria dei Materiali, Università di Messina, Salita Sperone 31, 98166 Messina, Italy

A R T I C L E I N F O A B S T R A C T

Article history: The heterogeneous catalytic etherification of glycerol with tert-butyl alcohol was investigated in
Received 5 May 2009 presence of lab-made silica supported acid catalysts. As reference, two commercial acid ion-exchange
Received in revised form 23 July 2009 resins were also used. Experiments were carried out in batch mode at TR ranging from 303 to 363 K. An
Accepted 24 July 2009
increase in reaction temperature favors the formation of di-substituted ethers. The etherification
Available online 3 August 2009
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
Keywords:
hindrance phenomena and water hinder the formation of tri-substituted ether (TBGE). As expected,
Oxygenated compounds
water removal was necessary to allow the higher ethers formation. The specific activity (turnover
Catalytic conversion
Glycerol ethers frequency, TOF) of A-15 catalyst is significantly higher than that of the other studied acid systems, due to
Biodiesel the wide pore diameter that allows an easier accessibility of the reagent molecules.
ß 2009 Elsevier B.V. All rights reserved.

1. Introduction glycerol could hold a prominent role [7–10]. In particular, tert-butyl


ethers of glycerol with a high content of di-ethers are considered
Recently, exhaust gases emitted by internal combustion promising as oxygenated additives for diesel fuels (diesel, biodiesel
engines were considered primarily responsible for environmental and their mixtures). However, mono-tert-butyl ethers of glycerol
pollution and human diseases [1–5]. (MBGEs) have a low solubility in diesel fuel; therefore, in order to
Biodiesel is currently used as a valuable fuel for diesel engines. avoid an additional separation step, the etherification of glycerol
In spite of a slight power loss, exhausts contain less particulate should address the formation of di- and tri-ethers [5,7–9,11].
matter. Biodiesel is a mixture of methyl esters of fatty acids The etherification of glycerol can be carried out using
(FAMEs) obtained by the transesterification reaction of vegetable heterogeneous acid catalysts like strong acid ion-exchange resins
oils with methanol in presence of a basic catalyst [1,2]. Such a [12–18]; however, the utilization of large-pore zeolites has also
catalytic process converts raw triglycerides into FAMEs, but been widely investigated [19–22]. Usually, low surface areas and
produces glycerol as side product. lack of thermal stability are the major drawbacks of sulfonic resins.
If biodiesel is produced on a large scale, adequate technologies The incorporation of organosulfonic groups over mesostructured
capable of converting glycerol into added value chemicals are silicas have generated effective solid acid catalysts with enhanced
necessary [3,4]. In particular, great attention has been already catalytic properties as compared with conventional homogeneous
devoted to the conversion of glycerol into oxygenated additives for and heterogeneous acid catalysts [23]. Moreover, these type of
liquid fuels [3,5–7]. In this context, an industrially relevant route silica materials functionalized with organosulfonic acid groups
for the conversion of glycerol into oxygenated chemicals involves have been used previously for the conversion of biorenewable
the etherification to tert-butyl ethers [5,7–10]. molecules [24–27], showing better catalytic performances than
It is well known that the addition of oxygenated additives to that of the commercial sulfonated resins. Currently, these highly
diesel fuels could represent a promising way enhancing the surface materials characterized by interconnected mesopores and
combustion efficiency in internal combustion engines with a high accessibility of acid sites represent the best systems for the
significant reduction of pollutant emissions. Among several etherification reactions [28].
oxygenated additives proposed to blend with diesel, the ethers of The synthesis of tert-butyl ethers (GTBEs) from isobutene and
glycerol on ion-exchange resins has been already investigated
extensively [8–10]. Isobutene (IB) is produced by catalytic cracking
* Corresponding author. and steam cracking fractions of petroleum refining and by
E-mail address: francesco.frusteri@itae.cnr.it (F. Frusteri). isobutane dehydrogenation [5,7]. What appears more attractive

0926-860X/$ – see front matter ß 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2009.07.037
78 F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83

is to produce GTBEs by the glycerol etherification reaction in a ‘‘jacketed-batch reactor’’ (Autoclave Engineers, Inc.) under a stirring
solid–liquid catalytic process, by using tert-butyl alcohol (TBA). In frequency of 1200 min1, in order to limit the influence of external
fact, the use of TBA, as both reactant and solvent, instead of gaseous mass transfer phenomena. Experiments were performed under
isobutylene, allows to overcome the technological problems different reaction conditions: (i) under pressure; (ii) at reaction
arising from the need to use solvents able to dissolve glycerol temperatures ranging from 303 to 363 K; (iii) by operating at
(i.e., dioxane, dimethyl sulfoxide) and typical drawbacks of a different reaction times and (iv) at both different catalyst/glycerol
complex three-phase system (mass transfer phenomena) [9,29]. and alcohol/glycerol ratios.
This study focused on the etherification of glycerol with tert- The experimental procedure was the following: a well defined
butyl alcohol over different solid acid systems. Attention was given amount of glycerol and dry catalyst were loaded into the reactor
primarily to investigate the main limiting factors for large scale and heated up to a prefixed reaction temperature (in 10 min).
process development. Before the addition of tert-butanol, the reactor was fluxed with
nitrogen to remove the air; then, tert-butanol was injected into the
2. Experimental reactor by a syringe: this was taken as the starting point of the
reaction. At the end of the experiments, the reactor was cooled
2.1. Catalysts and chemicals 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
Two solid acid supported catalysts were prepared by the gas phase compounds to condense. After the opening of the
incipient wetness method using a silica carrier (S.A.BET, autoclave, we collected a completely liquid mixture, without any
250 m2 g1) and two solutions containing 17 wt.% of Nafion1 solidification of TBA on the reactor walls. The liquid reaction
ionomer (N-17) and 17 wt.% of tungstophosphoric heteropoly acid mixture was analyzed off-line by a gas chromatograph, HP 6890N,
(HPW-17), respectively. An aliquot of the HPW-17 sample was provided with a capillary column HP Innowax (l, 30 m; i.d.,
mixed with a solution containing cesium to exchange a fraction of 0.53 mm; film thickness, 1.0 mm) under the following oven
H+ protons with Cs+ (Cs-HPW). In addition, two commercial acid temperature program: from 40 to 220 8C (with a heating rate of
ion-exchange resins, Nafion1 on amorphous silica (SAC-13) and 20 8C min1) and at 220 8C for 3 min. An automatic sampler Agilent
Amberlyst1 15 dry (A-15), were used as reference catalysts. 7683B Series was used (0.2 ml of samples were injected), each data
Glycerol anhydrous (purity 99.5%) and tert-butyl alcohol set being obtained, with an accuracy of 1%, from an average of
(purity 99.7%), supplied by Fluka (Buchs, Switzerland), were used three independent measurements using the external standard
as reactants. Standard compounds for GC analysis were supplied method (n-heptane, 8 wt.% in respect to the reaction mixture). The
on-demand by Aldrich. samples were not collected at different reaction times for avoiding to
spill out a not representative sample of the reaction system due to the
2.2. Catalysts characterization different density of the mixture compounds. Water content was
calculated by considering the reaction stoichiometry and confirming
Surface area (SABET) and pore volume (PV) were determined by the result by a quantitative TCD analysis. In each experiment, carbon
the nitrogen adsorption/desorption isotherms at 77 K using a Carlo balance was close to 98%.
Erba (Sorptomatic Instrument) gas adsorption device. Before
analysis, all the samples were outgassed at 423 K under vacuum 3. Results and discussion
for 2 h. The isotherms were elaborated according to the BET
method for surface area calculation, with the Horwarth–Kavazoe Physico-chemical properties of catalysts are summarized in
(HK) and Barrett–Joyner–Halenda (BJH) methods used for micro- Table 2.
pore and mesopore evaluation, respectively. Catalysts are characterized by different surface area ranging
The active phase loading and thermal stability were evaluated from 53 to 207 m2 g1 and porosity comprised between 0.07 and
by thermo-gravimetric (TG) analyses in the range 293–873 K using 0.80 cm3 g1. In terms of SA and porosity, the data obtained with
a Netzsch STA409C analyzer, running in air atmosphere with a SAC-13 sample are similar to that provided by the supplier.
heating rate of 10 K/min. Furthermore, all the catalytic systems show an average pore
Acid sites were determined by potentiometric titrations with diameter (APD) increasing with the porosity, apart from A-15 that
the zero point charge (ZPC) method. About 0.1 g of each sample is characterized by a wide pore texture (300 Å), although it has the
was dispersed in an aqueous solution of NaNO3 0.5 M under lowest SA (53 m2 g1). The acid capacity of A-15 sample was
stirring. Acidity, measured by an electrode Orion ROSS, was significantly higher than the other investigated samples.
calculated on the basis of the pH at which particles suspended in In order to collect quantitative data approaching the equili-
solution had zero charge. brium composition and to obtain a reliable comparison of the
The list of catalysts used in this study is summarized in Table 1. catalytic functionality at a prefixed time (6 h), preliminary
experiments using different solid acid catalysts were carried out
2.3. Catalytic testing at 0.1 MPa, 343 K, with a tert-butanol-to-glycerol molar ratio (RA/G)
equal to 4. The results shown in Fig. 1 demonstrate that a low
The etherification reaction between glycerol and tert-butyl catalyst/glycerol ratio equivalent to 1.2 wt.% (much lower than
alcohol was carried out in liquid phase in a 100 cm3 stainless steel that so far reported in literature [7–10]) is adequate to guarantee
high glycerol conversion. Indeed, after 6 h of reaction, low glycerol
Table 1
List of solid acid catalysts used in this study. conversion levels were reached using SAC-13, N-17 and HPW-17
catalysts (8–15 mol.%), while by the Cs-HPW and A-15 samples,
Code Active phase Carrier characterized by higher acid capacity (Table 2), the reaction takes
HPW-17 Phosphotungstic acid SiO2 (Cabosil LM50; S.A.BET = 250 m2 g1) place at higher rates, reaching at the end of the reaction a glycerol
Cs-HPW Phosphotungstic acid SiO2 (Cabosil LM50; S.A.BET = 250 m2 g1) conversion of 54 and 82 mol.%, respectively.
exchanged with cesium
Considering that catalysts are characterized by different acid
N-17 Nafion1 polymer SiO2 (Cabosil LM50; S.A.BET = 250 m2 g1)
SAC-13 Nafion1 polymer Amorphous silica capacity, the turnover frequency (TOF) of glycerol was reported as
A-15 Amberlyst1-15 – a function of the number of acidic sites (see Fig. 2). TOF values were
dry resin determined from the initial reaction rate (glycerol conversion <5%)
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83 79

Table 2
Physico-chemical properties of the studied samples.

Catalyst Loading (wt.%) of active phase S.A.BET (m2 g1) P.V. (cm3 g1) APD (Å) Acidity (mmolH+ g1)

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-15a – 53 0.40 300 4.70
a
Data supplied by Rohm and Haas.

and normalized to the number of acid sites. As it can be seen, the formation of five different alkyl glycerol ethers could be expected:
TOF appears to be comparable on all the systems, except for the A- 3-tert-butoxy-1,2-propandiol (1-MBGE), 2-tert-butoxy-1,3-pro-
15 sample that exhibits a TOF value about 3 times higher. At first pandiol (2-MBGE), 1,3-di-tert-butoxy-2-propanol (1,3-DBGE),
glance, this result could indicate that the glycerol etherification is a 1,2-di-tert-butoxy-3-propanol (1,2-DBGE) and tri-tert-butoxy-
structure sensitive reaction; however, what should be considered in propane (TBGE).
a reaction involving encumbered molecules is also the accessibility Taking into account the results reported in Fig. 3B, when glycerol
of active sites. As reported in Table 2, it can be seen that A-15 reacts with TBA, depending on the extent of the etherification
catalyst is characterized by an average pore diameter (APD) much reaction, only the formation of four ethers was observed: two mono-
higher (300 Å) in respect to the other catalysts (APD, 35–155 Å); substituted ethers (1-MBGE and/or 2-MBGE) and two di-substituted
then, the higher TOF observed for A-15 catalyst could not be due to ethers (1,3-DBGE and/or 1,2-DBGE). Tri-substituted ether never
the higher specific activity of acidic sites but to the accessibility of formed in the whole range of temperature investigated.
sites which is favored using macro-porous materials like Amber- 3-Butoxy-1,2-propandiol (1-MBGE) was the main mono-ether
lyst. In conclusion, the etherification of glycerol with TBA should be formed and, depending on the reaction temperature, even to
considered as an unstructured sensitive reaction. different extent, its concentration was always much higher than
On the basis of such results and considering that the aim of this the correspective 2-tert-butoxy-1,3-propandiol (2-MBGE).
study was to explore the feasibility of the process based on the Indeed, since the reaction takes place between the tert-butyl
conversion of glycerol with TBA, we have considered more cation (tertiary carbocation) and glycerol [10,30], it is probable
profitable to evaluate the influence of experimental parameters that initially the electrophilic attack occurs on the primary carbon
using A-15 catalyst characterized by the highest acid capacity. of glycerol due to steric hindrance and electrostatic effects exerted
The influence of reaction temperature on glycerol conversion by –OH groups of glycerol. As regards the formation of di-ethers,
and products distribution is shown in Fig. 3. Glycerol conversion 1,3-di-tert-butoxy-2-propanol (1,3-DBGE) was the main com-
linearly increased with reaction temperature growing from 10% at pound formed (its concentration linearly increased with reaction
303 K up to 85% at 343 K. It can be observed that, at high temperature) while the concentration of 1,2-di-tert-butoxy-3-
temperatures (363 K), glycerol was not totally converted due to the propanol (1,2-DBGE) reached a maximum value of 7% at 343 K,
occurrence of de-etherification reactions, which become important after which it remained almost constant. Then, the formation of
as the reaction temperature increases [7–10,28]. 1,3-DBGE is favored for the same reasons noted above (steric
In terms of products distribution, it is necessary to consider encumbrance and electrostatic effect).
that, according to literature evidences [7–10], the etherification Besides, it is also important to consider that, by dehydration of
reaction proceeds according to a consecutive path and the 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

Fig. 1. Conversion of glycerol on solid acid catalysts: TR = 343 K; PR = 0.1 MPa; RA/ Fig. 2. TOF of glycerol as a function of the catalyst acidity: TR = 343 K; PR = 0.1 MPa;
G = 4.0; cat = 1.2 wt.%/glycerol; reaction time = 6 h. RA/G = 4.0; reaction time: 2 h.
80 F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83

Fig. 3. Glycerol conversion and products distribution vs. reaction temperature: A-15 catalyst; PR = 0.1 MPa; RA/G = 4.0; cat = 1.2 wt.%/glycerol; reaction time = 6 h.

evident pressure effect was observed after 2 h of reaction: the To assess how the products distribution and yield to higher
glycerol conversion values increases from 53% (at 0.1 MPa) to 72% ethers could be affected by the amount of catalyst used, several
(at 1.0 MPa). experiments were carried out with different catalyst/glycerol
Furthermore, after 6 h, no differences were seen in terms of weight ratios. Results reported in Table 3 clearly point out that
glycerol conversion, since the reacting system reaches the without catalysts the reaction does not take place, while a
equilibrium. In terms of selectivity to ethers, apart from the significant rise in the glycerol ether formation can be observed as
reaction pressure, a progressive decreasing in MBGEs concentra- the amount of catalyst increases. In particular, the highest
tion can be observed as a function of the reaction time. In any case, cumulative yield (28%) to higher ethers of glycerol (DBGEs + TBGE)
no TBGE forms due to steric hindrance. was obtained with an amount of catalyst equivalent to 7.5 wt.%
Therefore, considering that the etherification reaction occurs at with respect to the glycerol weight.
higher rate under pressure, the reaction mechanism should involve Moreover, it is noteworthy that increasing the catalyst weight,
the electrophilic attack of a tertiary carbocation, formed through the conversion of glycerol becomes strongly limited by the
the protonation of both TBA and isobutene, to the adsorbed growing formation of water (see Fig. 5), which negatively affects
glycerol, to form ethers. Indeed, the tert-butyl cation concentration the etherification equilibrium. It is also important to consider that,
in liquid phase is ruled out by a dynamic equilibrium existing under the reaction conditions, a certain amount of isobutene (<5%)
between TBA and isobutene. Naturally, the concentration of forms due to tert-butanol dehydration. However, we did not
isobutene in liquid phase is higher by operating under pressure. observe the formation of oligomers that normally form when

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; TR = 343 K; rpm = 1200 min1.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83 81

Table 3
Influence of amount of catalyst on product distribution and yield.

Catalyst (wt.%/glycerol) Product distribution (wt.%) Yield (%)

IB TBA MBGEs DBGEs TBGE GLY H2O 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: TR = 343 K; PR = 0.1 MPa; reaction time = 6 h; rpm = 1200 min1.

Fig. 6. Glycerol conversion and ethers yield as a function of reaction time: A-15
catalyst; PR = 0.1 MPa; RA/G = 4.0; catalyst = 7.5 wt.%/glycerol.

Fig. 5. Glycerol conversion and water formation as a function of amount of catalyst volume was changed from 5 to 10 ml, while glycerol volume was
used: A-15 catalyst; PR = 0.1 MPa; RA/G = 4.0; reaction time = 6 h. 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
isobutylene is used as the reactant [7–10,28], because the were established between the reaction rate and molar concentra-
formation of water should account for hindering the extent of tion of reactants, thus enabling the discovery of the following rate
oligomerisation. law:
To evaluate how the reaction time affects the evolution of
reaction and catalyst activity, several experiments were carried out rate ¼ k½GLY0:3 ½TBA1:7 (1)
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, From such an equation, the kinetic constant at each temperature
glycerol was almost totally converted and the 1-MBGE was the was calculated and results obtained are summarized in Table 5.
main product. These data are thermodynamically consistent since ‘‘k’’ increased
As the reaction proceeded, the concentration of MBGE with temperature as expected for an endothermic reaction. The
decreased and the corresponding concentration of di-ether (DBGE) apparent activation energy was close to 70  3 kJ mol1, which well
increased. This transformation occurred with a low reaction rate; matches the values so far reported in similar kinetic studies [31].
in fact, after 30 h, MBGEs selectivity was still 50%. According to the rate Eq. (1), considering that the reaction order
Experimental data obtained by operating at different tert- referred to TBA is much higher than one, it can be confirmed that
butanol/glycerol molar ratios (RA/G) are reported in Table 4. It can the etherification reaction occurs with a molecular mechanism
be seen that the increase of RA/G does not affect the reaction in
terms of products distribution. Table 5
Rate constants at different temperature.
As such result was not expected, reaction order with respect to
glycerol (GLY) and tert-butanol (TBA) was determined considering Catalyst k  102 (M1 min1)
initial glycerol conversion values (<5%), under the following 303 K 323 K 343 K 363 K
reaction conditions: (a) glycerol volume was changed from 5 to
A-15 0.5 2.2 13.8 45.0
10 ml, while TBA volume was maintained in a large excess; (b) TBA

Table 4
Effect of tert-butanol-glycerol molar ratio (RA/G).

RA/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: 7.5 wt.%/glycerol; TR = 343 K; PR = 0.1 MPa; reaction time = 6 h; rpm = 1200 min1.
82 F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83

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 H2O

24.00 95.9 54.0 41.0 1.0 3.0 12.3 35.1 37.7 1.1 1.7 8.5
50.00a 93.6 47.5 44.2 1.8 2.9 13.3 31.6 39.8 2.0 2.5 7.3

A-15: 7.5 wt.%/glycerol; TR = 343 K; PR = 0.1 MPa; rpm = 1200 min1.


a
After 24 h, the ‘‘used’’ catalyst was unloaded from the reactor and replaced by ‘‘fresh’’ catalyst.

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 [8–10,28,37], in all the
investigated cases no differences were observed, either in terms of
glycerol conversion or products distribution, thus confirming 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 etherification
reaction.

Fig. 7. Products distribution recorded by using ‘‘fresh’’ and ‘‘re-used’’ A-15 catalysts. 4. Conclusions

The main findings of this study can be summarized as follows:


wherein TBA is quickly protonated on the acid sites forming a
tertiary carbocation able to react with the glycerol strongly  the etherification of glycerol with tert-butyl alcohol effectively
adsorbed (partial order 0.3) on the catalyst surface. takes place on solid acids catalysts, hence providing a promising
In order to shift the reaction equilibrium and favor the way to transform glycerol into value added products to be used
consecutive path of glycerol etherification, some experiments as oxygenated additives blended with diesel fuels;
were carried out to better assess how water affects catalytic  the accessibility of acid sites plays a fundamental role in
activity. On this account, an experiment was stopped at a reaction promoting catalyst activity and the systems with large pores are
time of 6 h and, after dehydration of the reaction mixture by more indicate to perform reaction at high rate;
zeolites, the run was continued for further 6 hours, thus obtaining  a low catalyst/glycerol ratio equivalent to 1.2 wt.%, much lower
a net increase in the yield of DBGEs from 28.5% (before water than that so far reported in literature, is adequate to guarantee
removal) to 41.5%. This result clearly demonstrates that the high glycerol conversion;
difficulty in obtaining higher ethers is due to the presence of water,  the reaction pressure significantly affects the reaction kinetic,
as previously inferred. Indeed, it is noteworthy that water also the tert-butyl cation concentration in liquid phase being ruled
competes with tert-butanol and glycerol on the active site out by a dynamic equilibrium existing between TBA and
adsorption. Really, the higher water acidity in relation to tert- isobutene;
butanol results in a lowering of catalyst activity due to the  no oligomers form during reaction since, under the reaction
formation of solvated sites, as reported in the literature [32–36]. conditions, the formation of water limits the extent of
Since the A-15 resin could swell during reaction due to the oligomerisation reaction;
inclusion of water in its polymeric structure, we decided to stop the  water formed during the reaction inhibits the glycerol ether-
run after 24 h of reaction, substituting the ‘‘used’’ catalyst with a ification and its removal from the reaction medium is necessary
same amount of ‘‘fresh’’ catalyst and continuing the experiment for for the formation of di- and tri-ethers. On this account, the design
25 more hours. Results obtained are summarized in Table 6. of a ‘‘water-separating’’ reaction medium could increase the
It is possible to observe that the replacement of the catalyst only formation of higher ethers.
slightly favored the formation of higher ethers; therefore, the
difficulty in obtaining high yields to higher ethers is not References
attributable to catalyst deactivation but to the water formation.
[1] G. Knothe, J. Van Gerpen, J. Krahl (Eds.), The Biodiesel Handbook, AOCS Press,
However, just to confirm that catalyst does not change during the Champaign, IL, 2005.
reaction, an experiment employing the ‘‘used’’ catalyst, after its [2] M. Mittelbach, C. Remschmidt, Biodiesel—The Comprehensive Handbook, M.
filtration and regeneration by a drying procedure, was performed. Mittelbach, Graz, Austria, 2004.
[3] S. Fernando, S. Adhikari, K. Kota, R. Bandi, Fuel 86 (2007) 2806–2809.
As Fig. 7 clearly shows, the results obtained with both ‘‘fresh’’ and
[4] J.M. Marchetti, V.U. Miguel, A.F. Errazu, Renew. Sust. Energy Rev. 11 (2007) 1300–
‘‘used’’ catalysts are similar, confirming that the A-15 catalyst is 1311.
stable and even if it swells during the reaction, its catalytic [5] F. Ancillotti, V. Fattore, Fuel Process. Technol. 57 (1998) 163.
[6] P. Satgé de Caro, Z. Mouloungui, G. Vaitilingom, J.Ch. Berge, Fuel 80 (2001) 565–
properties do not change significantly.
574.
Further attempts to elucidate how experimental procedures [7] R.S. Karinen, A.O.I. Krause, Appl. Catal. A: Gen. 306 (2006) 128–133.
could affect catalyst performance and to investigate if, during the [8] K. Klepáčová, D. Mravec, E. Hájeková, M. Bajus, Petrol. Coal 45 (2003) 54–57.
reactor heating, experimental conditions could change due to the [9] K. Klepáčová, D. Mravec, M. Bajus, Appl. Catal. A: Gen. 294 (2005) 141–147.
[10] K. Klepáčová, D. Mravec, A. Kaszonyi, M. Bajus, Appl. Catal. A: Gen. 328 (2007)
occurrence of thermal and catalytic reactions, different runs were 1–13.
carried out under the following conditions: (i) after the mixing of [11] J. Deutsch, A. Martin, H. Lieske, J. Catal. 245 (2007) 428–435.
F. Frusteri et al. / Applied Catalysis A: General 367 (2009) 77–83 83

[12] A. Juha, A. Linnekoski, O.I. Krause, L.K. Struckmann, Appl. Catal. A: Gen. 170 (1998) [25] I.K. Mbaraka, B.H. Shanks, J. Catal. 229 (2005) 365.
117–126. [26] P.L. Dhepe, M. Ohashi, S. Inagaki, M. Ichikawa, A. Fukuoka, Catal. Lett. 102 (2005)
[13] J.F. Knifton, J.C. Edwards, Appl. Catal. A: Gen. 183 (1999) 1–13. 163.
[14] Y. Pouilloux, S. Abro, C. Vanhove, J. Barrault, J. Mol. Catal. A: Chem. 149 (1999) [27] J.A. Bootsma, B.H. Shanks, Appl. Catal. A: Gen. 327 (2007) 44.
243–254. [28] J.A. Melero, G. Vicente, G. Morales, M. Paniagua, J.M. Moreno, R. Roldán, A.
[15] B.-L. Yang, S.-B. Yang, R.-Q. Yao, React. Funct. Polym. 44 (2000) 167–175. Ezquerro, C. Pérez, Appl. Catal. A: Gen. 346 (2008) 44–51.
[16] R. Alcántara, L. Canoira, C. Fernandez-Martin, M.J. Franco, J.I. Martinez-Silva, A. [29] P.M. Słomkiewicz, Appl. Catal. A: Gen. 313 (2006) 74–85.
Navarro, React. Funct. Polym. 43 (2000) 97–104. [30] R.A. Van Santen, P.V.N.M. Van Leeuwen, J.A. Moulijn, B.A. Averill, Stud. Surf. Sci.
[17] C. Park, M.A. Keane, J. Mol. Catal. A: Chem. 166 (2001) 303–322. Catal. 123 (1999).
[18] J.M. Adams, D.E. Clement, S.H. Graham, J. Chem. Res. (1981) S254. [31] F. Cunill, M. Iborra, C. Fité, J. Tejero, J.F. Izquierdo, Ind. Eng. Chem. Res. 39 (2000)
[19] A.A. Chin, S.S.F. Wong, Zeolites 17 (5–6) (1996) 524. 1235–1241.
[20] I. Hoek, et al. Appl. Catal. A: Gen. 266 (2004) 109–116. [32] F. Ancillotti, F. Mauri, E. Pescarollo, J. Catal. 46 (1977) 49–57.
[21] K.J. Won, K.D. Jung, H.J. Uk, K. Min, K.J. Man, Y.J. Eui, Catal. Today 87 (1–4) (2003) [33] F. Ancillotti, F. Mauri, E. Pescarollo, L. Romagnoni, J. Mol. Catal. 4 (1978) 37–48.
195–203. [34] A. Gicquel, B. Torck, J. Catal. 83 (1983) 9–18.
[22] P.S.E. Dai, J.F. Knifton, Zeolites 18 (5–6) (1997) 417–418. [35] P. Rys, W.J. Steinegger, J. Am. Chem. Soc. 101 (1979) 4801–4806.
[23] J.A. Melero, R. van Grieken, G. Morales, Chem. Rev. 106 (2006) 3790. [36] W.J. Casey, D.J. Pietrzyk, Anal. Chem. 45 (1973) 1404–1407.
[24] I.K. Mbaraka, D.R. Radu, V.S.-Y. Lin, B.H. Shanks, J. Catal. 219 (2003) 329. [37] D.E. López, J.G. Goodwin Jr., D.A. Bruce, J. Catal. 245 (2007) 381–391.

Das könnte Ihnen auch gefallen