Вы находитесь на странице: 1из 7

ARTICLE

pubs.acs.org/IECR

Silica-Supported Tin Oxides as Heterogeneous Acid Catalysts for


Transesterification of Soybean Oil with Methanol
Wenlei Xie,* Hongyan Wang, and Hui Li
School of Chemistry and Chemical Engineering, Henan University of Technology, Zhengzhou 450052, P. R. China

ABSTRACT: In this work, the SnO2/SiO2 materials with various Sn loadings ranging from 1 to 16 wt % were prepared and used as
heterogeneous acid catalysts for soybean oil transesterification to produce biodiesel. The catalyst with 8 wt % Sn loading and calcined
at 873 K exhibited the best catalytic activity, giving an oil conversion of 81.7%. The dispersed amorphous SnO2 species on the silica
surface are considered to be active sites for the transesterification reaction. The effect of reaction parameters such as catalyst amount,
reaction temperature, reaction time, and methanol-to-oil ratio was investigated to optimize the transesterification conditions. It is
shown that the free fatty acid (FFA) and water have no significant influence on the catalytic activity to the transesterification
reaction. In addition, the heterogeneous acid catalyst was found to have high activities toward the esterification of FFAs.
Furthermore, the catalyst could be recovered with a better reusability.

1. INTRODUCTION unsuitable for the industrial process based on alkaline catalysis.


Due to depletion of fossil fuels and environmental concerns, Therefore, homogeneous acid catalysts have the potential to replace
biodiesel fuels, consisting of methyl esters of long chain fatty alkali catalysts, since they show less sensitive to FFAs and able to
acids, have attracted much attention as a promising alternative conduct esterification and transesterification simultaneously.7,8
fuel with significantly lower exhaust emissions of particulate matter As a result, biodiesel produced by the transesterification of low-cost
and green-house gases such as carbon monoxide, carbon dioxide, waste oils using an acid catalyst could make it more cost-competitive
and sulfur oxides.1 Generally, biodiesel can be produced by with petroleum diesel as far as the economic benefit for biodiesel
the transesterification of vegetable oils with methanol, which is production was concerned. Among the homogeneous catalysts,
carried out in the presence of base or acid catalyst. In the con- sulfuric acid and hydrochloric acid are the most common acid
ventional biodiesel process, homogeneous base catalysts, such as catalysts used for the transesterification process. However, several
potassium hydroxide, sodium hydroxide, or alkoxides, are the drawbacks such as high reaction temperature, slow reaction rate,
most preferred choice of catalysts because the rate of transester- difficulty in separation, reactor corrosion, and the incapability for
ification reactions catalyzed by alkaline catalysts is much faster reuse have depreciated the utilization of the liquid acid catalyst
under moderate operating conditions than that of reactions cata- for biodiesel production.810
lyzed by the acid catalysts.2,3 However, these homogeneous In recent years, the recyclable solid acid catalysts have attrac-
catalysts suffer from various drawbacks, including sensitivity to ted much attention because of their potential to replace corrosive
moisture and free fatty acid (FFA) content and the need to deal homogeneous acid catalysts in industrial processes.2 Hetero-
with the waste from the neutralization processes, though they are geneous acid catalysts can be easily separated from the reaction
highly efficient and of low cost. The FFA present in the reactants mixture by centrifugation or filtration and do not require neu-
can react with the alkaline catalysts to form unwanted soap tralization processes, thus providing a more environmentally
byproduct, and these generated soaps could lead to the forma- benign process and reducing the cost of processing.4 In parti-
tion of emulsions, which inhibit the separation of biodiesel from cular, they have the ability to simultaneously catalyze the tran-
the reaction mixture and ultimately lowers the biodiesel yield. sesterification of triglycerides and the esterification of FFAs.11
Additionally, the removal of these base catalysts after the reaction Thus, the utilization of solid acid catalysts for the biodiesel
is technically difficult, and a large amount of wastewater is gen- preparation is expected to simplify significantly the manufactur-
erated in separating and cleaning the products. Therefore, the alkali- ing process when the low-cost oils are employed as feedstcocks
catalyzed procedure requires the use of higher purity vegetable and allows for good practices of catalyst recycling. There are
oils with less than 0.5 wt % FFA and anhydrous conditions.35 several reports about the use of heterogeneous acid catalysts to
At present, the major hurdle in the commercialization of bio- produce biodiesel, including zeolites,12 La/zeolite β,13 tung-
diesel production is the high cost of raw materials used, since the stated zirconia,14 sulfated zirconia,15,16 sulfated zirconia
high price of virgin oil can contribute potentially as much as 70% alumina (SZA),17 sulfated tin oxide,18 heteropoly acids,11,19 vanadyl
of total costs of biodiesel. For this reason, an effective way of phosphate,20 sulfonated polystyrene compounds,21 and carbon-
reducing the cost of raw material is to employ low-quality oils based solid acid catalyst.22
(such as waste cooking oils and nonedible oils), which are cheaply
available and can be regarded as attractive feedstocks for bio- Received: October 2, 2011
diesel production.6 However, the low-cost oils usually contain Accepted: November 27, 2011
large amounts of FFAs and water. As mentioned previously, the Revised: November 25, 2011
presence of high percentages of FFAs makes the low-cost oils Published: November 28, 2011

r 2011 American Chemical Society 225 dx.doi.org/10.1021/ie202262t | Ind. Eng. Chem. Res. 2012, 51, 225–231
Industrial & Engineering Chemistry Research ARTICLE

The silica used as a support presents significant advantageous different amounts of methanol, and the catalyst (5 wt % referred
features owing to its high specific surface area, large pore size, and to the initial oil weight) were incorporated together inside the
high thermal stability. In the present contribution, the SnO2/ reactor. To minimize mass transfer limitations, the reactants were
SiO2 catalysts were prepared and tested as a heterogeneous acid stirred at 350 rpm. The reactor temperature was raised to the
catalyst for the transesterification of soybean oil with methanol. required value, and after running the reaction for the desired
The catalytic performance of the catalyst in the transesterification duration, the reactor was allowed to cool to room temperature.
reaction was investigated regarding the conversion to methyl Afterward, the solid catalyst was separated from the product
esters, and more attention was paid to the effect of calcination mixture by filtration, and the excess methanol was recovered
temperatures and Sn loadings. Besides, the experimental variables under reduced pressure prior to the subsequent analyses. Experi-
such as the amount of catalyst, methanol/oil molar ratio, reaction ments were performed by varying reaction parameters such as
temperature, reaction time and the presence of FFAs and water methanol/oil ratio, reaction temperature, reaction time, and
were investigated to optimize the transesterification conditions. catalyst amount. The conversion of soybean oil to methyl esters
Moreover, the catalysts were characterized by various physico- was determined by measuring hydroxyl content on the transester-
chemical techniques such as powder X-ray diffractometry (XRD), ified soybean oil as previously described by us in the literature.23
thermogravimetric and differential thermogravimetric analysis 2.5. Esterification Reaction of FFAs. In the current work,
(TGDTG), and scanning electron microscopy (SEM). Further, oleic acid was used as a model free fatty acid and was deliberately
the reusability of the catalyst was checked and the catalytic acti- added into the soybean oil. The liquid-phase esterfication of oleic
vity of the catalyst toward the esterification reaction was also acid present in the feedstocks with methanol over the catalyst was
investigated in terms of the FFA conversion to methyl esters. carried out in a stainless steel high-pressure vessel using 5 wt % of
catalyst in the presence of 10 wt % FFA. The acid values of the
2. EXPERIMENTAL SECTION reaction mixtures were determined by titration of the aliquots,
taken out from the reaction mixture at different time intervals
2.1. Materials. Commercial edible-grade soybean oil was with an alkali solution. The conversion of FFAs in the reactant to
obtained from the local market. According to GC (Shimadzu methyl esters can be determined according to the following
DC-9A) analysis, the fatty acid profiles of the soybean oil used equation.
were as follows: palmitic acid, 12.3%; stearic acid, 5.8%; oleic
initial acid value  final acid value
acid, 26.5%; linoleic acid, 49.4%; and linolenic acid, 5.9%. The FFA conversion ð%Þ ¼
acid value was less than 0.1 mg KOH g1, and the average molar initial acid value
mass of triacylglycerols present in the soybean oil was 874 g ð1Þ
mol1, calculated from the saponification index (SV = 192.6 mg
KOH g1). All other employed materials were of analytical grade 3. RESULTS AND DISCUSSION
and were used as received without further purification.
2.2. Catalyst Preparation. The SnO2/SiO2 catalysts with 3.1. Screening of the Catalyst. A screening of different
various Sn loadings were prepared by an incipient wetness catalysts was performed under identical reaction conditions to
impregnation of powdered silica (SiO2) with an acetone solution identify the most promising solid acid catalysts for the transes-
containing an appropriate amount of dibutyltin dilaurate. Com- terification reaction. The best catalyst was selected on the basis of
mercial silica was used as the support. After impregnation, the maximum conversion to methyl esters. The experimental results
obtained sample was dried in an oven at 373 K, and finally, the are listed in Table 1. From this table it was seen that the selected
solid was calcined at designed temperature for 5 h under air support (Al2O3, ZrO2, and SiO2) showed low activities, with
atmosphere in a muffle furnace before use for the transesterifica- conversions below 30%. However, when the metal oxides were
tion reaction. The concentration of the precursor solution was loaded on the supports and activated at a high temperature, the
adjusted so as to yield the desired Sn loadings in the final cata- catalytic activity of the formed solid acid catalyst was improved
lysts. The tin content in the SnO2/SiO2 catalysts was varied from in the transesterification reaction. For example, over SnO2/
1 to 16 wt % based on the Sn loading. γ-Al2O3, a soybean oil conversion of 68.5% was obtained, and
2.3. Catalyst Characterization. Powder X-ray diffraction a conversion of 64.3% was observed with SnO2TiO2/SiO2 cata-
(XRD) measurements were conducted on a Rigaku D/MAX- lyst. Obviously, the Sn-based catalysts and the ZrO2-supported
3B powder X-ray diffractometer using a radiation source of catalysts showed higher catalytic activities. In particular, SnO2/
Cu Kα (λ = 0.154 nm) at 40 kV and 20 mA over a 2θ range of SiO2 solid exhibited superior catalytic activity, giving a conver-
20°70° at a scanning speed of 5° min1. The XRD diffraction sion of 81.6% after 5 h of reaction at a reaction temperature of
patterns thus obtained were compared with references from the 453 K, probably because of the presence of strong acid sites on
Power Diffraction File (PDF) database (JCPDS, International the surface of the catalyst (entry 14 in Table 1). It is expected that
Centre for Diffraction Data) to identify the phase present in the the incorporation of tin oxides to SiO2 leads to the formation of
samples. SnO2/SiO2 catalyst with an enhanced acidity in comparison with
TGDTG was carried out on a STA409PC thermal analyzer the silica. On the basis of the above results, the SnO2/SiO2 cata-
operating under a flow of air at a 20 K min1 heating rate up to lyst showed potential for use as a heterogeneous acid catalyst and
1273 K. The surface morphology of the catalysts was studied by therefore was chosen for the subsequent study.
scanning electron microscopy using a JEOL JSM-6700F 3.2. Catalyst Characterization. The XRD patterns for 8 wt %
instrument. SnO2/SiO2 calcined at different temperatures are illustrated in
2.4. Transesterification Procedures. All the transesterifica- Figure 1. For the samples calcined at temperatures below 873 K
tion reactions were carried out in a 250 mL stainless steel high- (curves ac in Figure 1), there was only a broad diffraction peak
pressure autoclave reactor fitted with a temperature controller centered at 2θ angles of 22°, due to the amorphous silica, to be
and mechanical stirrer. In a typical assay, 16.0 g of soybean oil, observed in the XRD patterns, and no diffraction peak corresponding
226 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231
Industrial & Engineering Chemistry Research ARTICLE

Table 1. Catalytic Activities of Different Catalystsa


entry catalysts Tp (K) conversion (%)

1 γ-Al2O3 873 17.8


2 ZrO2 873 10.5
3 SiO2 873 22.1
4 B2O3/γ-Al2O3 873 43.6
5 MoO3/γ-Al2O3 873 48.7
6 SnO2/γ-Al2O3 873 68.5
7 TiO2/γ-Al2O3 873 58.8
8 B2O3/ZrO2 873 61.7
9 MoO3/ZrO2 873 64.6
10 SnO2/ZrO2 873 58.6
Figure 2. XRD patterns for samples with different Sn loadings: (a) 2 wt %,
(b) 4 wt %, (c) 8 wt %, (d) 12 wt %, and (e) 16 wt %. The calcination
11 TiO2/ZrO2 873 63.3 temperature of supported catalysts was 873 K. (Δ) SiO2, (2) SnO2..
12 B2O3/SiO2 873 47.2
13 MoO3/SiO2 873 51.6
14 SnO2/SiO2 873 81.6 of the original one. This phenomenon shows that the SnO2
15 TiO2/SiO2 773 61.2
species is not changed during the transesterification processes.
Notably, the changes of XRD diffraction peaks with the calcina-
16 SnO2TiO2/SiO2 873 64.3
tion temperature for the catalyst are in agreement with the
17 SnO2MoO3/SiO2 873 62.1
a
changes in the catalytic activity toward the transesterification
Reaction conditions: methanol/oil molar ratio, 25:1; catalyst amount, reaction, as illustrated in the following section. In addition, all the
5 wt %; reaction temperature, 453 K; reaction time, 5 h.
catalysts maintained the amorphous structure of the support,
since their XRD patterns showed broad diffraction peaks at
2θ angles of 22°, similar to the amorphous silica. Accordingly, SnO2
amorphous species probably can be considered as catalytically
active sites for the transesterification reaction.
Figure 2 shows the XRD patterns of the catalyst with different
Sn loadings. As presented in Figure.2, when the Sn loading was
below 8 wt %, the major features of the XRD pattern basically
belonged to the amorphous silica support, and no characteristic
peak of crystalline SnO2 or another tin compound was detected
by the XRD techniques, indicating the good dispersion of tin
compounds on the silica.24,27 As the amount of Sn loading in-
creased to 12 wt %, the defined XRD peaks at 2θ angles of 26.7°,
33.9°, 37.9°, 71.8°, 54.8°, 61.8°, and 65.8° appeared on the XRD
patterns (curve e in Figure 2). These characteristic XRD peaks
Figure 1. XRD patterns for 8 wt % SnO2/SiO2 calcined at different were well-matched with the rutile cassiterte phase of SnO2
temperatures: (a) not calcined, (b) 573 K, (c) 873 K, (d) 1073 K, and (JCPDS 41-1445), and their intensities further grew with in-
(e) recovered catalyst; (Δ) SiO2, (2) SnO2.. creasing the Sn loading to 16 wt %, which suggests that a residual
crystalline phase of SnO2 is formed on the support at these
to SnO2 could be detected by XRD measurements, suggesting loadings.25 Besides, it is also shown that there is no crystallite
that tin compounds are highly dispersed on the silica at calcina- SnO2 species in the catalyst when the Sn loading amount is less
tion temperatures below 873 K.24 As expected, the incorporation than 8 wt %. Evidently, the XRD-undetectable phase of SnO2
of tin compounds into the silica and subsequent thermal treat- may have dispersed onto the support, which is important for the
ment could result in a remarkable increase in the number of acid generation of active sites in the SnO2/SiO2 catalysts. When the
sites and hence the catalytic activity. However, when the cal- Sn loading amount increases beyond a borderline, i.e., above its
cination temperature increased to 1073 K, the XRD pattern was dispersion capacity, the XRD SnO2 peaks could be detected,
completely different from the catalyst calcined at lower tempera- revealing that tin is dispersed in the form of an amorphous
tures. For the sample calcined at 1073 K, apart from the species at lower Sn loading and the SnO2 with crystal structure
diffraction peaks owing to the silica, some new diffraction peaks appears at higher Sn loading. Indeed, for those catalysts loaded
at 2θ angles of 26.7°, 33.9°, and 54.8°, which are the character- with tin oxides at more than the spontaneous dispersion capacity
istic peaks for the cassiterite SnO2 phase (curve d in Figure 1), on the silica, the residual bulk crystalline phase of SnO2 remains
emerged in the XRD pattern.25 Such a result indicates that the on the composite, which reduces the catalytic activity of the
local crystalline SnO2 is formed on the surface of the silica upon catalyst, since the crystalline SnO2 exhibits low activities for the
calcination at a temperature of 1073 K. Since the 1073 K-calcined transesterification reaction. In other words, the catalysts that con-
catalyst was found to have lower activities, the amorphous SnO2 tain only amorphous SnO2 are catalytically superior to the other
probably appeared to be catalytically better for the transester- catalysts that contain both amorphous and crystallite SnO2.
ification reaction than the SnO2 crystallites. This finding was also The thermal analysis of the catalyst was performed in an
reported in the literature.26 Besides, the XRD pattern of the re- attempt to gain information about the nature of the catalysts.
covered catalyst exhibited very similar diffraction peaks to those As shown in Figure 3, the weight loss was less than 1.5% upon
227 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231
Industrial & Engineering Chemistry Research ARTICLE

Figure 3. TGDTG curves of noncalcined 8 wt % SnO2/SiO2 sample.


Figure 5. Influence of calcination temperature on the conversion to
methyl esters. Reaction conditions: catalyst amount, 5 wt %; reaction
time, 5 h; reaction temperature, 453 K; methanol/oil molar ratio, 24:1.

Figure 4. SEM images of samples: (a) SiO2 and (b) 8 wt % SnO2/SiO2.


Figure 6. Influence of loading amount of Sn on the conversion to
methyl esters. Reaction conditions: catalyst amount, 5 wt %; reaction
heating to 423 K, and the total weight loss value was 21%. The time, 5 h; reaction temperature, 453 K; methanol/oil molar ratio, 24:1.
low-temperature weight loss below 423 K in the TG profile could
be assigned to the loss of residual solvent or water trapped in the
silica framework. Besides, The TG curve of the original sample calcination, the catalyst was not particularly active. After calcina-
showed a significant weight loss in a wide temperature between tion the catalyst showed considerable activities toward the tran-
453 and 873 K. This weight loss, accompanied by a large DTG sesterification reaction. With increasing the calcination tempera-
peak centered at 543 K, is mainly attributed to the progressively ture from 473 to 873 K, the conversion to methyl esters was
thermal decomposition of supported dibutyltin dilaurate, produ- gradually increased. At 873 K, the conversion of soybean oil was
cing tin oxides when the temperature is raised. After 873 K there attained a maximum value of 81.7% over the catalyst. The en-
was no change in the weight loss value in the TG curve. hancement of the catalytic activity for the solid catalyst is most
SEM analyses were used to elucidate the morphological char- likely caused by the different decomposition extents of the pre-
acteristics of the solid catalysts. The SEM photographs of the cursor compound upon increasing the calcination temperature
samples are presented in Figure 4. The morphology of the silica and subsequently the different amounts of the thus generated
and SnO2/SiO2 samples showed well-ordered particles with a active species. However, as the calcination temperature increased
size of about 212 μm and nearly polygonal and irregular in to 1073 K, it can be seen that the conversion was substantially
shape with defined edges. As shown in Figure.4, no significant reduced. Presumably, the decrease in the catalytic activity at higher
morphology change was observed, except small particles (maybe temperatures largely results from the formation of crystalline
SnO2 particles) on the SnO2/SiO2 catalyst that were not seen on SnO2, as was demonstrated by the XRD measurements. Therefore,
the silica. On the basis of the results, after loading of tin com- the optimal calcinatiion temperature is set at 873 K.
pounds, the silica retains its structure and the tin species are The conversions to methyl esters over the catalysts with dif-
distributed upon the silica. ferent Sn loadings are presented in Figure 6. As can be observed
3.3. Influence of Preparation Conditions on the Catalyst from this figure, when the Sn loading rose from 1 to 8 wt %, the
Performance. The transesterifications catalyzed by various SnO2/ conversion to methyl esters was gradually increased, and the best
SiO2 catalysts calcined at different temperatures were carried out conversion was reached at the Sn loading of 8 wt %. However,
using 5 wt % catalysts, and the results are illustrated in Figure 5. with the further enhancement in the Sn loading from 8 to 16 wt %,
It is shown that the calcination temperature is of importance in a diminishing trend in the conversion was noticed, probably due
determining the final activity of the catalysts. In the absence of to the formation of crystallites SnO2 that could cover the surface
228 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231
Industrial & Engineering Chemistry Research ARTICLE

Figure 7. Effect of methanol/oil molar ratio on the conversion to Figure 8. Effect of reaction time on the conversion to methyl esters.
methyl esters. Reaction conditions: catalyst amount, 5 wt %; reaction Reaction conditions: methanol/oil molar ratio, 24:1; reaction tempera-
time, 5 h; reaction temperature, 453 K. ture, 453 K; catalyst amount, 5 wt %.

active sites of the catalyst. As a result, the catalytic activity of the increasing methanol addition beyond the molar ratio of 24:1.
catalyst is largely dependent on the Sn loading, and the proper Sn From the results, the optimum molar ratio of methanol/oil for
loading is necessary for the catalyst to get high catalytic activity. the reaction is approximately 24:1. The excess methanol can be
Thus, it may then be suggested that the amorphous SnO2 species recovered from the reaction product by simple distillation, and
formed during the calcination processes could be considered to the byproduct of glycerol in the process can be gravitationally
be the mainly catalytically active phase for the reaction. Accord- separated by decantation without any complex process.
ingly, the SnO2/SiO2 catalyst with the Sn loading of 8 wt % is the The influence of catalyst loadings was also investigated by
most suitable catalyst for the transesterification reaction. using different catalyst loadings between 1 and 7 wt %. The tran-
3.4. Influence of Reaction Parameters on the Oil Conver- sesterification reactions were set at a methanol/oil molar ratio of
sion. The reaction temperature is one of the important variables 24:1, a temperature of 453 K, and a reaction time of 5 h. As the
affecting the transesterification reaction. Usually, higher reaction catalyst loading increased from 1 to 3 to 5 wt %, the correspond-
temperature can give higher reaction rate in the transesterifica- ing conversion to methyl esters was increased from 40.8 to 68.4
tion reaction, especially by using acid catalysts.2 The effect of to 81.7 and reached a plateau thereafter when the catalyst loading
reaction temperature was investigated with this catalyst by vary- was further increased beyond 5 wt %. The optimum catalyst
ing the reaction temperature in the range of 393493 K. The loading for obtaining high conversion to methyl esters was found
results achieved here showed that the soybean oil conversion was to be 5 wt %. As the catalyst loading increases, more catalytically
even less than 60%, though the reaction was run at 393 K. active sites are available to facilitate the reaction to occur, which
However, it could be noted that the conversion to methyl esters allows the conversion to methyl esters to increase. From the above
increased steadily from 56.5% to 70.8% to 81.7% upon increas- results, a catalyst loading of 5 wt % was chosen for subsequent
ing the reaction temperature from 393 to 423 to 453 K. The in- studies.
creased conversion with reaction temperature might be not only To examine the effect of reaction time on the conversion to
due to the effect of the increase in reaction rate by increasing methyl esters, experiments were performed by using different
temperature but also due to the improvement of the solubility of reaction times under the reaction conditions of 453 K and 24:1
methanol in soybean oil. However, the further increase in the ratio of methanol to oil. The reaction time was varied within a
reaction temperature gave an insignificant increase in conversion range from 1 to 9 h. From the results shown in Figure 8, it was
to methyl esters. Hence, it could be inferred that the optimum observed that when the reaction time extended to 5 h, the con-
operating temperature for the transesterification reaction is 453 K. version to methyl esters was increased and reached the maximum
In the reaction sequence, triglycerides are converted stepwise value of 81.7%, achieving the equilibrium of the reaction. The
to diglyceride, monoglyceride, and finally glycerol accompanied further increase in reaction time to 9 h resulted in the small
with the formation of methyl esters. Stoichiometrically, the tran- increase in the conversion to 83.6%. Therefore, the suitable reac-
sesterification reaction requires 3 mol of methanol for each mole tion time for the transesterification reaction is 5 h.
of triglyceride. However, the excess methanol is commonly em- The methodology based on acid-catalyzed transesterification
ployed for the transesterification reaction in order to shift the has raised a lot of interest because it allows for the preparation of
reaction equilibrium toward the desired product side.28 In gen- biodiesel by using low-cost oils as feedstocks. To study the influe-
eral, the higher methanol/oil molar ratio is required for the acid- nce of FFA and water on the activity of the catalysts, the sim-
catalyzed process to obtain a higher reaction rate, in comparison ulated waste oils with different amounts of FFA and water were
with the base-catalyzed process.4 The effect of methanol/oil prepared by deliberately adding oleic acid or water. Additional
molar ratio was investigated using this solid acid catalyst, and the experiments were performed with the simulated waste oils under
results are shown in Figure 7. With the increase in methanol/oil the optimized reaction conditions. When the FFA amount was
molar ratio, the conversion was gradually enhanced and a maxi- 1%, 3%, 5%, 7%, and 10% (based on refined soybean oil weight),
mum conversion after reaction for 5 h was achieved over the the conversions to methyl esters were 80.1%, 78.9%, 78.2%,
catalyst at the methanol/oil molar ratio of 24:1. However, there 77.1%, and 75.7%, respectively. Obviously, the conversion went
was no significant increase in the conversion to be observed with down slightly with the increase in the FFA amount within the
229 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231
Industrial & Engineering Chemistry Research ARTICLE

the silica, leading to the improvement of the catalyst stability.


However, when the catalyst was further used for a fifth cycle, the
conversion decreased remarkably to 54.3%. The loss of the cata-
lytic activity is probably owing to the original active sites being
blocked partially by adsorbed intermediate or product species,
which diminishes the amount of readily available catalytic sites.
In general, the acid-catalyzed transesterification usually in-
volves high reaction temperature, long reaction time, and low
conversion rate. In the present research, the SnO2/SiO2 catalyst
shows higher efficiency in biodiesel production when low-cost
feedstocks are employed. Over the catalyst, the soybean oil con-
version of 81.7% and the FFA conversion of 94.6% are obtained
under appropriate operating conditions. It is shown that the soy-
bean oil and FFA can be converted to biodiesel in a one-step
process catalyzed by the solid catalyst. Notably, the conversions
Figure 9. Effect of reaction time on the conversion of FFA to methyl achieved in this work are relatively low and are not of commercial
esters. Reaction conditions: methanol/oil molar ratio, 24:1; reaction
temperature, 453 K; catalyst amount, 5 wt %; reaction time, 5 h.
interest, and thus, for the industrial production of biodiesel, the
activity of the catalyst needs to be improved considerably. Alter-
natively, another stage of transesterification after glycerol separa-
range studied, and the conversion of soybean oil could retain tion would be expected to complete the reaction if the solid
75.7% even with 10% of FFA. It is reported that during the tran- catalyst is to be practically employed for the biodiesel production.
sesterification processes, the activity of solid catalysts is adversely
affected by the presence of water.4 By using the SnO2/SiO2 4. CONCLUSIONS
catalyst, the effect of the water amount on the catalytic activity
was studied. The conversions were 81.2%, 80.4%, 78.7%, 75.4%, The SnO2/SiO2 catalysts containing different amounts of Sn
72.1%, and 63.2%, respectively, when the water content was loading were prepared by impregnation with dibutyltin dilaurate
0.1%, 0.3%, 0.5%, 0.7%, 1%, and 1.5%. Clearly, the oil conversion and thermal activation at higher temperature for the production
declined with the increase in water content. Indeed, the water of biodiesel. The maximum activity is found for the catalyst with
content of less than 1.0% could lead to the conversion above 70% 8 wt % Sn loading and calcination at 873 K, which could catalyze
after 5 h of reaction. However, up to a water content higher than the transesterification reaction of soybean oil even containing a
1.0%, the conversion to methyl esters decreased rapidly. With large amount of FFA. The soybean oil conversion of 81.7% was
water content of 1.5%, the soybean oil conversion was low and obtained after 5 h of reaction, when a methanol/oil molar ratio of
only attained 63.2%. Therefore, the water content in oil had 24:1 and 5 wt % of catalyst loading were employed. The amor-
better be less than 1.0%, which is higher than the allowable water phous SnO2 on the silica support is shown to be catalytically
content of the base catalyst employed. active sites for the transesterification reaction. Besides, the cata-
In order to examine the catalytic activity of the catalyst toward lyst also exhibits high catalytic activities in the esterification
the esterification of FFAs, the esterification of oleic acid present reaction of FFA, with a FFA conversion of 94.6%. Further, the
in the oils with methanol over the catalysts was carried out in a SnO2/SiO2 catalyst could be easily recovered and reused without
stainless steel high-pressure vessel, using 5 wt % of catalyst at a serious loss in the catalytic activity for four cycles.
reaction temperature of 453 K. With the SnO2/SiO2 catalyst,
both the esterification and transesterification could be conducted ’ AUTHOR INFORMATION
in a single process. As illustrated in Figure 9, a FFA conversion of Corresponding Author
93.8% was attained after 3 h. When the reaction time increased to *Telephone: +86-371-67756302. Fax: +86-371-67756718. E-mail:
5 h, the conversion of FFA to methyl esters over the catalyst was xwenlei@163.com.
increased steadily and reached as high as 94.1%, followed by a
slight increase upon a further increase of reaction time. By draw-
ing on the results, the solid catalyst also exhibited catalytic acti- ’ ACKNOWLEDGMENT
vities toward the esterification of FFAs with methanol. This work was financially supported by research grants from
3.5. Reusability of the Catalyst. The reutilization of hetero- National Science Foundation of China (project No. 21076062)
geneous catalyst is an important aspect that makes it economic and Program for Science & Technology Innovation Talents in
and preferable over a homogeneous one. For the reusability Universities of Henan province in China (HASTIT).
investigation, the used catalysts were separated by filtration and
washed with petroleum ether and methanol. Before reuse in the
’ REFERENCES
second reaction cycle, the recovered catalysts were dried at 393 K
and subsequently used for new reactions under the optimum (1) Sharma, Y. C.; Singh, B.; Upadhyay, S. N. Advancements in
operating conditions. When the solid catalyst was used for one, development and characterization of biodiesel: A review. Fuel 2008, 87,
2355.
two, three, and four cycles, the conversion of soybean oil was
(2) Semwal, S.; Arora, A. K.; Badoni, R. P.; Tuli, D. K. Biodiesel
declined slightly from 80.2%, 78.3%, and 74.6% to 67.5%. Thus, production using heterogeneous catalysts. Bioresour. Technol. 2011, 102,
the catalyst retained a higher catalytic stability even after con- 2151.
tinuous use up to four cycles. It might be hypothesized that the (3) Leung, D. Y. C.; Wu, X.; Leung, M. K. H. A review on biodiesel
Sn is linked to the surface of the silica by a SiOSn chemical production using catalyzed transesterification. Appl. Energy 2010, 87,
bond arising due to the stronger interactions between SnO2 and 1083.

230 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231


Industrial & Engineering Chemistry Research ARTICLE

(4) Endalew, A. K.; Kiros, Y.; Zanzi, R. Inorganic heterogeneous (26) Park, P. W.; Kung, H. H.; Kim, D. W.; Kung, M. C. Character-
catalysts for biodiesel production from vegetable oils. Biomass Bioenergy ization of SnO2/Al2O3 lean NOx catalysts. J. Catal. 1999, 184, 440.
2011, 35, 3787. (27) Wang, S.; Ma, X.; Gong, J.; Yang, X.; Guo, H.; Xu, G. Transesterifica-
(5) Li, H.; Xie, W. Fatty acid methyl ester synthesis over Fe3+- tion of dimethyl oxalate with phenol under SnO2/SiO2 catalysts. Ind. Eng.
vanadyl phosphate catalysts. J. Am. Oil Chem. Soc. 2008, 85, 655. Chem. Res. 2004, 43, 4027.
(6) Gombotz, K.; Parette, R.; Austic, G.; Kannan, D.; Matson, J. V. (28) Li, Y.; Qiu, F.; Yang, D.; Li, X.; Sun, P. Preparation, character-
MnO and TiO solid catalysts with low-grade feedstocks for biodiesel ization and application of heterogeneous solid base catalyst for biodiesel
production. Fuel 2012, 92 (1), 9-15. production from soybean oil. Biomass Bioenergy 2011, 35, 2787.

(7) Oner, C.; Altun, S. Biodiesel production from inedible animal
tallow and an experimental investigation of its use as alternative fuel in a
direct injection diesel engine. Appl. Energ. 2009, 86, 2114.
(8) Zheng, S.; Kates, M.; Dube, M. A.; McLean, D. D. Synthesis of
biodiesel via homogeneous Lewis acid catalyst. Biomass Bioenergy 2006,
30, 267.
(9) Di Serio, M.; Tesser, R.; Dimiccoli, M.; Cammarota, F.; Nastasi,
M.; Santacesaria, E. Synthesis of biodiesel via homogeneous Lewis acid
catalyst. J. Mol. Catal. A 2005, 239, 111.
(10) Morales, G.; Bautista, L. F.; Melero, J. A.; Iglesias, J.; Sanchez-
Vazquez, R. Low-grade oils and fats: Effect of several impurities
on biodiesel production over sulfonic acid heterogeneous catalysts.
Bioresour. Technol. 2011, 102, 9571.
(11) Kulkarni, M. G.; Gopinath, R.; Meher, L. C.; Dalai, A. K. Solid
acid catalyzed biodiesel production by simultaneous esterification and
transesterification. Green Chem. 2006, 8, 1056.
(12) Ramos, M. J.; Casas, A.; Rodríguez, L.; Romero, R.; Perez, A.
Transesterification of sunflower oil over zeolites using different metal
loading: A case of leaching and agglomeration studies. Appl. Catal. A
2008, 346, 79.
(13) Shu, Q.; Yang, B.; Yuan, H.; Qing, S.; Zhu, G. Synthesis of
biodiesel from soybean oil and methanol catalyzed by zeolite beta
modified with La3+. Catal. Commun. 2007, 8, 2159.
(14) Jacobson, K.; Gopinath, R.; Meher, L. C.; Dalai, A. K. Solid acid
catalyzed biodiesel production from waste cooking oil. Appl. Catal. B
2008, 85, 86.
(15) Yee, K. F.; Lee, K. T.; Ceccato, R.; Abdullah, A. Z. Production of
biodiesel from Jatropha curcas L. oil catalyzed by SO42‑/ZrO2 catalyst:
Effect of interaction between process variables. Bioresour. Technol. 2011,
102, 4285.
(16) Garcia, C. M.; Teixeira, S.; Marciniuk, L. L.; Schuchardt, U.
Transesterification of soybean oil catalyzed by sulfated zirconia.
Bioresour. Technol. 2008, 99, 6608.
(17) Yee, K. F.; Wu, J. C. S.; Lee, K. T. A green catalyst for biodiesel
production from jatropha oil: Optimization study. Biomass Bioenergy
2011, 35, 1739.
(18) Furuta, S.; Matsuhashi, H.; Arata, K. Biodiesel fuel production
with solid superacid catalysis in fixed bed reactor under atmospheric
pressure. Catal. Commun. 2004, 5, 721.
(19) Juan, J. C.; Zhang, J.; Yarmo, M. A. 12-Tungstophosphoric acid
supported on MCM-41 for esterification of fatty acid under solvent-free
condition. J. Mol. Catal. A 2007, 267, 265.
(20) Di Serio, M.; Cozzolino, M.; Tesser, R.; Patrono, P.; Pinzari, F.;
Bonelli, B.; Santacesaria, E. Vanadyl phosphate catalysts in biodiesel
production. Appl. Catal. A 2007, 320, 1.
(21) Soldi, R. A.; Oliveira, A. R.S.; Ramos, L. P.; Cesar-Oliveira,
M. A. F. Soybean oil and beef tallow alcoholysis by acid heterogeneous
catalysis. Appl. Catal. A 2009, 361, 42.
(22) Rao, B. V. S. K.; Mouli, K. C.; Rambabu, N.; Dalai, A. K.; Prasad,
R. B. N. Carbon-based solid acid catalyst from de-oiled canola meal for
biodiesel production. Catal. Commun. 2011, 14, 20.
(23) Xie, W.; Li, H. Hydroxyl content and refractive index determi-
nations on transesterified soybean oil. J. Am. Oil Chem. Soc. 2006, 83, 869.
(24) Yang, X.; Ma, X.; Wang, S.; Gong, J. Transesterification of
dimethyl oxalate with phenol over TiO2/SiO2: Catalyst screening and
reaction optimization. AIChE J. 2008, 54, 3260.
(25) Burri, D. R.; Choi, K. M.; Han, D. S.; Sujandi; Jiang, N.; Burri, A.;
Park, S. E. Oxidative dehydrogenation of ethylbenzene to styrene with CO2
over SnO2ZrO2 mixed oxide nanocomposite catalysts. Catal. Today
2008, 131, 173.

231 dx.doi.org/10.1021/ie202262t |Ind. Eng. Chem. Res. 2012, 51, 225–231

Вам также может понравиться