Вы находитесь на странице: 1из 9

Fuel 89 (2010) 28442852

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Long term activity of modied ZnO nanoparticles for transesterication


Shuli Yan, Siddharth Mohan, Craig DiMaggio, Manhoe Kim, K.Y. Simon Ng, Steven O. Salley *
Department of Chemical Engineering and Materials Science, Wayne State University, Detroit, MI 48202, USA

a r t i c l e

i n f o

Article history:
Received 3 November 2009
Received in revised form 18 May 2010
Accepted 18 May 2010
Available online 31 May 2010
Keywords:
Biodiesel
Unrened and waste oils
Transesterication
Esterication
Catalyst life

a b s t r a c t
Biodiesel can be produced by the transesterication of natural oils with methanol using modied ZnO
nanoparticles as catalyst. Crude algae oil, corn oil from DDGs, crude palm oil, crude soybean oil, crude
coconut oil, waste cooking oil, food-grade soybean oil and food-grade soybean oil with 3% water and
5% FFA addition were converted into FAME within 3 h using this new catalyst. The ZnO nanoparticles
were reused 17 times without any activity loss in a batch stirred reactor and the average yield of FAME
was around 93.7%. ZnO nanoparticles were used continuously for 70 days in a x bed continuous reactor
and the average yield of FAME was around 92.3%. XRD, ICP, TEM and HRTEM were used to characterize
the long term used catalyst structure. Results show that this catalyst is a mixture of wurtzite ZnO nanoparticles and some amorphous materials and that the used catalysts have similar crystal structure to
fresh catalyst. ICP results show that this catalyst does not dissolve in biodiesel, methanol, oil and glycerinemethanol solutions. It has a stable crystal structure under the reaction conditions. The high catalytic
activity, long catalyst life and low leaching properties demonstrate these modied ZnO nanoparticles
have potential in a commercial biodiesel production process.
Published by Elsevier Ltd.

1. Introduction
Biodiesel, a renewable fuel with similar combustion properties
to petroleum diesel, is normally produced by the transesterication of highly rened oils with short-chain alcohols. Biodiesel can
signicantly decrease the exhaust emission of CO2, SOx and unburned hydrocarbons from motor vehicles [1,2]. It is environmentally benecial, and therefore is a promising alternative to fossil
diesel [3].
The transesterication reaction of triglycerides for the production of biodiesel is as follows:

Conventionally, transesterication is performed using strong base


or acid catalysts such as NaOH, NaOCH3 and H2SO4 [4,5]. These
traditional homogeneous catalysts have certain advantages includ* Corresponding author. Tel.: +1 313 577 5216; fax: +1 313 578 5636.
E-mail address: ssalley@eng.wayne.edu (S.O. Salley).
0016-2361/$ - see front matter Published by Elsevier Ltd.
doi:10.1016/j.fuel.2010.05.023

ing cost-effectiveness, high activity, and mild reaction conditions


(65150 C and 1530 Psi). However, these homogeneous catalysts
also have associated problems.
Firstly, usage of homogeneous catalysts is normally limited to
batch-mode processing followed by a catalyst separation step [5].
The catalysts are soluble in products and after reaction both fatty
acid methyl esters and glycerine phases are required to be neutralized by mineral acids or bases, washed with water, and dried by
vacuum. The post-reaction treatment process is long and complicated. Furthermore, a lot of waste washing water is generated
which is not environmentally benecial.

Homogeneous catalysts are also sensitive to free fatty acids


(FFAs) and water. FFAs react with basic catalysts (NaOH and
KOH) and form soaps, which complicates the glycerol separation,
and drastically reduces the methyl ester yield. Water in the feedstock leads to hydrolysis of oils and fatty acid methyl esters (FAME)
in the presence of strong basic or acidic catalysts. Thus, some

2845

S. Yan et al. / Fuel 89 (2010) 28442852

inexpensive oils, such as crude vegetable oils, waste cooking oil,


and rendered animal fats, which generally contain a high content
of FFA and water, cannot be directly utilized with homogeneous
catalysts. For conventional processes using homogenous catalysts,
FFA content in the feedstock must be lower than 0.50 (wt.%) [6]
and water content lower than 0.06 (wt.%) [7]. Thus the feedstock
cost for traditional biodiesel production processes is very high. It
has been reported that the cost of oil feedstock in 2006 accounts
up to a total of 80% of the biodiesel production cost [8,9], which
greatly inuenced the biodiesel price.
Since heterogeneous catalysts can be easily removed from
the reaction mixture and reused many times, the development
of a continuous reaction system based on a heterogeneous catalyst can greatly decrease the process and catalyst consumption
costs.
There have been many reports about highly active heterogeneous catalysts for biodiesel production. Kim et al. [10] prepared
a Na/NaOH/c-Al2O3 heterogeneous base catalyst and found it
showed almost the same activity under the optimized reaction
conditions compared to conventional homogeneous NaOH catalyst. Suppes et al. [11] carried out the transesterication of soybean oil with methanol in the presence of sodium species
loaded on NaX zeolite and ETS-10 zeolite. At 60 C, ETS-10 yielded
a conversion of 80.7%. Yan et al. [12] prepared supported CaO catalysts, achieving rapeseed oil conversion at 65 C as high as 92%
within 1 h. However, the catalyst durability is seldom reported,
which has a practical impact on the potential for catalyst
commercialization.
Of late, several researchers reported that some solid transesterication catalysts partially dissolved in the reaction mixtures
threatening the reusability of the catalyst. Granados et al. [13,14]
found that the transesterication mechanism of activated calcium
oxide catalysts is a combination of heterogeneous and homogeneous reaction. He observed that part of the transesterication
reaction takes place on basic sites at the surface of the catalyst,
while the rest is due to the dissolution of the activated CaO in
methanol that creates a homogeneous leached active phase. Later,
Martyanov and Sayari [15] investigated the catalytic activities of
sodium, magnesium and calcium methoxides and found that the
transesterication starts as a heterogeneous process, but then
the methoxide salts interact with products such as glycerol and
form an active homogeneous species. Kim et al. [16] prepared
ZnOAl2O3/ZSM-5 and SnOAl2O3/ZSM-5 catalysts and they
observed that the residue sodium on supported zeolite catalysts
provided the active site for the catalyst activity.
Many attempts have been made to develop a heterogeneous
catalyst with high tolerance to FFA and water in the raw materials.
Bournay and Hillion [17] claimed in a patent that a Lewis acid catalyst (zinc aluminate) for biodiesel production has a high tolerance
to FFA in oil. Insitut Francais Petrole (IFP) has promoted the Esterp-HTM process based on this kind of catalyst. Axens has commer-

cialized this process and Soproteol in Ste had it rst industrially


used in 2006 [18,19]. However, this catalyst is quite sensitive to
water, limiting water content in oils to 0.15 (wt.%). Thus, some
unrened oils cannot be used in this system. Furthermore, until
now there is no report about the long term catalyst activity and
catalyst structural stability.
In our previous work [20,21], we reported a series of ZnOLa2O3
catalysts which were highly active in biodiesel formation reactions
and highly tolerant to water and FFA. Yan et al. [20] reported that
ZnOLa2O3 catalysts can process soybean oil which has 3% water
and 5% oleic acid. The catalyst with 3:1 ratio of zinc to lanthanum
(Zn3La1) was found to simultaneously catalyze the oil transesterication and fatty acid esterication reactions, while minimizing
oil and biodiesel hydrolysis.
Our research goal is to develop a heterogeneous catalyst which
has the advantages of high catalytic activity, stable crystal structure under reaction conditions, long catalytic life, and high tolerance to water and FFA. In this study, we tested the catalytic
activity of Zn3La1 with multiple inexpensive oils in both batch
and continuous reactors as well as the catalyst durability and the
solubility of catalyst in reaction mixtures. We also investigated
the interaction between zinc and lanthanum and found how it
inuenced the catalyst life and crystal structures.

2. Materials and methods


2.1. Materials
Food-grade soybean oil was purchased from Costco warehouse
(Detroit, MI), crude coconut oil was from TWA Inc. (Selangor,
Malaysia), crude algae oil and corn oil from DDGs was obtained
from SRS (Dexter, MI), crude soybean oil was from BDI (Denton,
TX), crude palm oil was from Malaysia Palm Oil Board (Selangor,
Malaysia) and waste cooking oil was obtained from a local restaurant. The fatty acid compositions of these seven kinds of oil were
determined by GCMS (Table 1). Zinc nitrate hexahydrate (98%),
lanthanum nitrate hydrate (98%), and urea (99%), analysis grade,
were purchased from SigmaAldrich Company (St. Louis, MO).

2.2. Catalyst preparation and characterization


The modied ZnO nanoparticles were prepared by the urea
hydrolysis method as described in our previously published report
[20]. Solutions of 2 M Zn(NO3)2 and 1 M La(NO3)3 were prepared
with distilled water. Then, solutions with 3:1 ratios of Zn:La were
mixed with a 2 M urea solution. The mixture was boiled for 4 h,
and then dried at 150 C for 8 h, followed by step-rising calcination
at 250, 300, 350, 400 C, nally at 450 C for 8 h. The catalysts are
noted as Zn3La1.

Table 1
Fatty acid composition and TON of food-grade soybean oil, crude soybean oil, crude palm oil, waste cooking oil, crude corn oil from DDGs, crude algae oil and crude coconut oil.
Fatty acid
components

Food-grade soybean
oil (%)

Crude soybean
oil (%)

Crude palm oil


(%)

Waste cooking
oil (%)

Crude corn oil from


DDGs (%)

Crude algae oil


(%)

Crude coconut
oil (%)

C 12:0
C 14:0
C 16:0
C 16:1
C 18:0
C 18:1
C 18:2
C 18:3
Others
TAN

0
0
11.07
0.09
3.62
20.26
57.60
7.36
0
0.03

0
0.27
13.05
0.39
4.17
22.75
52.78
6.59
0
6.62

0
0.21
41.92
0.23
3.85
42.44
11.30
0.04
0
0.48

0
0
11.58
0.18
4.26
24.84
53.55
5.60
0
7.56

0
0
10.76
0
4.46
24.55
52.55
1.19
6.50
25.19

0
2.72
20.91
10.62
6.95
33.33
18.45
1.16
6.86
26.38

49.13
19.63
10.12
1.79
2.83
7.59
2.75
0.15
6.01
8.48

2846

S. Yan et al. / Fuel 89 (2010) 28442852

Powder X-ray diffraction (XRD) patterns were taken with a


Rigaku RU2000 rotating anode powder diffractometer (The
Woodlands, TX) equipped with Cu Ka radiation (40 kV, 200 mA).
Transmission Electron Microscopy (TEM) and high resolution
TEM (HRTEM) analyses were carried out using a JEOL 2010
(FasTEM, Japan) operating at 200 kV for microstructural, morphological and chemical quantication studies.
Inductively Coupled Plasma Optical Emission Spectrometry
(ICP-OES) was carried out with a PerkinElmer Optima 2100 DV
(Wellesley, MA). The dilutions of biodiesel, oil and FAME with kerosene were 0.1 g/g, and the dilutions of glycerine, methanol, and
glycerinemethanol solution with distilled water were 0.01 g/g.

2.3. Biodiesel reactions and product analysis


Batch catalytic reactions were carried out in a 500 mL stainless
steel stirred reactor (Parr 4575 HT/HP Reactor). The reaction mixture consisted of 126 g of oil, 180 g of methanol, and 3 g of catalyst.
Continuous catalytic reactions were carried out in a BTRS-JR Laboratory Reactor Systems (Parr) or a tubular continuous reactor
which has a similar structure to the BTRS-JR. The tube reactor
had dimensions of 20 mm i.d.  533 mm length in which 20 g of
Zn3La1 catalyst was packed. Reactants were premixed in a beaker
and then pumped into the top of the vertically oriented reactor
using an HPLc liquid pump (Chrom Tech Inc., Apple Valley, MN).
The ow rate was xed at 0.3 mL/min; with a residence time of
2 h; and a biodiesel production rate of 0.15 mL/min; the reaction
temperature was held at either 200 C or 140 C; reaction pressure
was 300 Psi; the molar ratio of methanol to oil was varied from
14:1 to 26:1.
The methanol was vaporized from the liquid product, and then
the remaining product was settled in a separating funnel. The
fatty acid methyl esters in the upper layer from the separating
funnel was characterized with a GCMS spectrometer (Clarus
500 MS System, PerkinElmer, Shelton, CT) equipped with a capillary column (Rtx-WAX Cat. No. 12426) (Bellefonte, PA). Methyl
arachidate (Nu-Chek Prep Inc., Elysian, MN) was used as an internal standard. The total acid number (TAN) in the FAME phase was
determined using a Brinkman/Metrohm 809 titrando (Westbury,
NY) according to ASTM D 664 (TAN is the amount of potassium

hydroxide in milligrams that is needed to neutralize the acids in


one gram of oil).
3. Results and discussion
3.1. Catalyst performance in a batch reactor
3.1.1. Oil exibility of Zn3La1 catalyst
Table 1 summarizes the composition of several oils for which
this catalyst was evaluated. Crude corn oil from DDGs contains
93% triglycerides and a total acid number (TAN) of 25.19 mg
KOH/g. Crude algae oil contains only 80% triglyceride, with a TAN
of 26.38 mg KOH/g. The TAN of crude coconut oil was 8.48 mg
KOH/g. It should be noted that many of these oils cannot be
directly used in the traditional biodiesel production processes
because of high TAN.
Fig. 1 shows the FAME yield versus time for the above oils when
using the Zn3La1 catalyst in a batch reactor. Nearly complete conversion is obtained within 3 h at 200 C and 500 Psi. The TAN of the
product of the algae oil reaction was 0.94 mg KOH/g while that of
the corn oil from DDGs product was 1.32 mg KOH/g. These results
imply that during the reaction process esterication of FFA is
simultaneously accomplished with the transesterication of
triglyceride.
3.1.2. Catalyst durability in the batch stirred reactor
The Zn3La1 catalyst was reused 17 times in the batch reactor for
the transesterication of rened soybean oil with methanol without any catalyst regeneration (Fig. 2a) and reused six times with
crude coconut oil and methanol (Fig. 2b). Fig. 1 shows that the
average yield of soybean methyl esters was maintained around
93.7% and coconut methyl esters around 91.3%. There was no drop
in the FAME yield during these catalyst durability tests.
3.2. Catalyst performance in a continuous reactor
3.2.1. Oil exibility of Zn3La1 in the continuous reactor
In our previous work [20], Zn3La1 was shown to be active for
esterication even when the reaction temperature was as low as
140 C. As crude corn oil contains a high content of FFA, the

100

Yield of FAME %

80

60

40

20

0
0

20

40

60

80

100

120

140

160

180

200

Tim e m in

soybean oil with 3ater and 5 % FFA


soybean oil
waste cooking oil
crude soybean oil
algae oil
crude palm oil
coconut oil
corn oil from DDGs
brown grease
H 2 SO 4
NaOH
Fig. 1. FAME yield of crude corn oil from DDGs, crude algae oil, crude coconut oil, crude palm oil, crude soybean oil, waste cooking oil, food-grade soybean oil and food-grade
soybean oil with 3% water and 5% oleic acid addition. Reaction conditions: 126 g of oil, 180 g of methanol, 3 g of catalyst, 200 C, 500 Psi, in the batch stir reactor. Note that all
of these oils were converted into FAME within 3 h.

2847

S. Yan et al. / Fuel 89 (2010) 28442852

100

a 100
90

80

Yield of FAME (%)

Yield of FAME %

80
70
60
50
40

60

40

30
20

20
10
0

0
0

10

12

14

16

18

Recycle times

Recycle times

Fig. 2. FAME yield in the batch reactor using the recycled Zn3La1 catalyst. (a) Yield of soybean methyl esters. Note the catalyst was reused for 17 times. The average yield of
soybean methyl esters is 93.7%. (b) Yield of coconut methyl eaters. Note the catalyst was reused for three times. The average yield of coconut methyl esters is 92.7%. Reaction
mixture is 126 g of oil, 180 g of methanol, 3 g of catalyst.

a 100

80

Yield of FAME %

80

Yield of FAME (%)

100

60

40

60

40

20

20

0
0

20

40

60

80

100

Time (hr)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Day

Fig. 3. FAME yield in the continuous reactor using crude corn oil from DDGs and crude coconut oil. (a) Yield of corn methyl esters. Note crude corn oil was run for 4 days.
Reaction conditions: 140 C, 300 Psi, 12:1 of mole ratio of methanol to oil, 20 g of catalyst, 0.3 mL/min of ow rate, 2.5 h of resident time. (b) Yield of coconut methyl eaters.
Note crude coconut oil was run for 1.5 days based on a 32 days used catalyst. Reaction conditions: 200 C, 300 Psi, 42:1 of mole ratio of methanol to oil, 8 g of 32-day-used
catalyst, 0.3 mL/min of ow rate, 2 h of resident time.

catalyst was utilized at 140 C, 300 Psi and 12:1 mol ratio of
methanol to oil for 4 days (Fig. 3a). The TAN of the methyl ester
product was 1.716 mg KOH/g and the average yield of corn methyl
esters was 72.5%, which is a little lower than the results in the
batch reactor. This can be explained by the low reaction temperature and low mole ratio of methanol to oil. Crude coconut oil was
reacted at 200 C, 450 Psi and 42:1 mol ratio of methanol to oil in
the continuous reactor for 1.5 days using the same catalyst which
had already been used for 32 days (Fig. 3b). The average yield of
coconut methyl esters was 85.4% with a TAN of the methyl ester
product of 1.23 mg KOH/g.
3.2.2. Catalyst durability in the continuous reactor
Fig. 4 shows the yield of FAME over 70 days in the Zn3La1 catalyst continuous reactor for rened soybean oil. As observed for
other long term heterogeneous catalysts [22], steady state opera-

tions were not obtained immediately. Fig. 4 shows that during


the initial 6 days there is an increasing trend in FAME yield, and
past the 7th day the yield of FAME is stable around 92.3%, continuing for 63 days. The delay in attaining steady, high yields may be
related to the time required to achieve good liquidsolid contact at
such low ow rates. To our knowledge, there is no transesterication catalyst with a reported longer catalyst life than Zn3La1. The
long term stability of FAME yield shows a potential for commercialization of this kind of ZnO-containing particles for biodiesel
production.
Mole ratio of methanol to oil has a signicant effect on the process cost of continuous biodiesel production. Therefore, raw materials with mole ratio of 13:1, 18:1, 26:1 were tested. Fig. 5ac
illustrate that the average yield of FAME for 13:1 is 40.5%, for
18:1 is 53.5%, and for 26:1 is 95.0%. Molar ratio of 26:1 is suggested
here. Effect of residence time was also tested (Fig. 6). In the same

2848

S. Yan et al. / Fuel 89 (2010) 28442852

3.2.3. Leaching from modied ZnO nanoparticles


Determination of the amount of metal (Zn or La) leached by
dissolution of the catalyst in the reaction medium is relevant for
the following two reasons. First, the presence of Zn2+ and La3+
in the biodiesel and glycerine products necessitates purication
of the products, especially the biodiesel. Second, leaching directly
affects the lifetime of the catalyst.
In this study, we prepared four different solutions: pure methanol, rened soybean oil, soybean biodiesel and 20% (wt.) glycerine
in methanol (glycerinemethanol), and measured leaching of
Zn3La1 in these solutions. One gram of catalyst was soaked in
500 mL of solution for 12 h with vigorous stirring at 500 rpm.
The mixture was then ltered, and the solid was washed with fresh
solution. Fig. 7 shows that Zn and La levels in the washing solution
were high in the initial period, which implies a partial leaching of
catalyst components. But Zn and La levels decreased quickly during
the washing process. After 200 mL of washing solution was used,
the Zn and La concentration were lower than 2 ppm. This low level
of Zn and La suggests that very little leaching of catalyst components occurs after the initial period.
The metal ion contents in the FAME and glycerine phases obtained during continuous reactor operation over 70 days are reported in Fig. 8. Fig. 8a shows that Zn and La levels in FAME
were relatively high in the rst 3 days, but decreased as operation
proceeded and stabilized around 6 and 2 ppm over the remainder
of the 67 days of operation. Fig. 8b shows the Zn and La levels in
the glycerine product. They decreased in the initial 7 days and

100

Yield of FAME (%)

80

60

40

20

0
0

10

20

30

40

50

60

70

Time (day)
Fig. 4. FAME yield of rened grade soybean oil in the continuous reactor. Note that
this catalyst has run for 70 days, and the average yield of FAME during stable stage
is 92.3%. Reaction conditions: 26:1 M ratio of methanol to oil, 20 g of catalyst,
0.3 mL/min, 200 C, 300 Psi.

reaction system, raw materials ow rate was increased to 0.6 mL/


min and residence time was decreased to 1 h. The results show
that FAME yields still maintain around 90.4%.

100

80

Yield of FAME %

Yield of FAME %

80

100

60

40

20

60

40

20

0
10

15

20

25

30

35

40

45

50

10

20

30

40

Time hr

50

60

70

80

Time hr

c 100

Yield (%)

80

60

40

20

0
0

10

12

14

16

18

20

Tim e (day)
Fig. 5. Effect of mole ratio of methanol to oil on the yield of FAME in the continuous reactor. 13:1 (a); 18:1 (b); and 26:1 (c). Note that the average yield of FAME increases
with mole ratio.

2849

S. Yan et al. / Fuel 89 (2010) 28442852

60

associated with the yield of FAME reaching the highest and steady
value of 92.3%. This phenomenon can be closely related with the
polarity of products. The polarization index of glycerine is much
higher than that of FAME, therefore, the solubility of metal ions
in glycerine is higher than in FAME and requiring more time to stabilize the catalyst in glycerine.

40

3.3. Catalyst structure and reaction mechanism

100

Yield (%)

80

20

0
1.0

1.5

2.0

2.5

3.0

3.5

4.0

Time (day)

Fig. 6. Yield of FAME basing on an increased ow rate of raw material (0.6 mL/min).

became stable around 8 and 4 ppm over the next 63 days. The low
level of Zn and La in FAME and glycerine products suggests that
leaching of catalyst components is negligible once the catalyst
bed attains a stable status.
The amount of metal leached stabilized more rapidly in the
FAME product than in the glycerol product, which took 7 days to
reach a stable value. The 7th day of the reaction period is also

Fig. 9 displays the XRD spectra of the fresh catalyst, catalyst


used 18 and 32 days in the continuous reactor and the catalyst
used 17 times in the batch stirred reactor. The used catalysts show
a similar crystal structure to the fresh Zn3La1 catalyst. This indicates that all of these catalysts are a mixture of hexagonal wurtzite
structure of zinc oxide and La2CO5. The mean grain size of ZnO in
the fresh catalyst is around 15.5 nm calculated by the DebyScherrer equation based on the reection peak of ZnO (1 0 1) while used
catalysts had a slightly reduced grain size. XRD results show that
the used catalysts have a similar distribution of average critical
size distribution. Fig. 9 also shows a slight shift of peaks of ZnO
(1 0 0) and ZnO (1 0 1). We calculated the crystal parameters of
ZnO, such as a, c, volume and density. Table 2 shows that the a
and c values slightly increased for the used catalyst. Fig. 9 also indicates that the peak intensity of the used catalyst slightly decreases
in comparison with fresh catalyst. This is possibly due to the

60

70

55

60

50
45

50

35
30

Zn/La content ppm

Zn/La content ppm

40

La content
Zn content

25
20
15

40

La content
Zn content

30

20

10

10

5
0

-5
0

100

200

300

400

100

200

300

400

Soybean biodiesel volumn (ml)

methanol volumn (ml)


90

18

80

16

La content
Zn content

60

14
12

Zn/La content ppm

Zn/La content ppm

70

50
40
30
20
10

10

La content
Zn content

8
6
4
2

0
0

100

200

300

soybean oil volumn (ml)

400

100

200

300

400

glycerine-methanol volumn (ml)

Fig. 7. Solubility of Zn3La1 in methanol (a), food-grade soybean oil (b), soybean biodiesel (c) and glycerinemethanol (d). Note that after washed by 200 mL of solutions Zn
and La contents are lower than 2 ppm.

2850

S. Yan et al. / Fuel 89 (2010) 28442852

b 1800
600

1600

500

Zn and La content (ppm)

Zn and La content (ppm)

400

La content in FAME phase


300

Zn content in FAME phase

200

100

1400
1200
1000
800
600

La content in glycerine phase


Zn content in glycerine phase

400
200
0

-200
0

10

20

30

40

50

60

70

80

10

20

30

40

50

60

70

80

Time (day)

Time (day)

Fig. 8. Zn and La contents in FAME (a) product and glycerine product (b). Note that after 3 days metal contents in FAME product reached a low level; after 7 days metal
contents in glycerine product reached a low level.

Fig. 9. XRD spectra of fresh Zn3La1, 18 days used Zn3La1 in the continuous reactor,
32 days used Zn3La1 in the continuous reactor, and the 17th times used Zn3La1in the
batch reactor. (1) ZnO and (2) La2CO5. Note that they have similar X-ray structure.

Table 2
Crystal size and crystal parameters of ZnO in fresh Zn3La1, 18 days used Zn3La1 in the
continuous reactor, 32 days used Zn3La1 in the continuous reactor and the 17th times
used Zn3La1 in the batch reactor.
Catalyst samples

Crystal size (nm)

a ()

c ()

V (3)

D (c)

Fresh Zn3La1
18 days used
32 days used
17th time reused

15.5
14.3
14.2
13.2

3.22
3.26
3.26
3.26

5.18
5.21
5.26
5.27

46.40
47.88
48.35
48.48

5.83
5.64
5.58
5.57

attached amorphous fat on the used catalyst surface, which will increase noise and bother the peak detection.
TEM images of fresh catalyst (Fig. 10a) show that Zn3La1 is a
mixture of amorphous material and hexagonal nanoparticles
around 2080 nm. Some of the nanoparticles indicate a single
crystalline textured orientation (Fig. 10b). This polyhedron is
enclosed by (0 0 0 1) (top and bottom surface), {1 0 1 0} (side surfaces), stepped {1 0 1 1} (inclined surface), and high index
planes with defect surface. The high resolution TEM image
(Fig. 10c) is recorded along [0 0 0 1] of the crystal. Fig. 10d indicates ZnO nanocrystals partially covered by small amorphous

materials. Fig. 10e is a TEM image of the 32 days used catalyst.


The size distribution of the used nanoparticles in Fig. 10d is closed
to the fresh catalyst in Fig. 10a.
ICP analysis shows that the molar ratio of Zn to La for the fresh
catalyst is 3.18 while the ratio is 3.29 for the catalyst used for
32 days. This suggests that some of the components in the catalyst
may have leached out during reaction, as indicated by the leaching
results above. Generally speaking, amorphous materials are not
stable and can easily be ushed out with reactants. Therefore it
is likely that the slight leaching in the initial period is caused by
the dissolution of amorphous materials, which are not considered
active centers for transesterication. Therefore we can assume that
only after amorphous materials are ushed out the catalyst shows
a high activity in transesterication.
Two kinds of surface could be supposed on the ZnO crystal, like
the polar faces of (0 0 0 1)-Zn, (0 0 0 1)-O, and the non-polar faces
of side (1 0 1 0) and stepped (5 0 5 1). The polar faces of
(0 0 0 1) consist of only Zn atoms and (0 0 0 1) consist of only O
atoms, while (1 0 1 0) and (5 0 5 1) are a mixture of Zn and O
atoms. The transesterication reaction rate on non-polar faces is
low [23,24]. Our previous study [20] has shown that O rich faces
have a high catalytic activity in the triglyceride transesterication
reaction and the Zn rich faces are active in fatty acid esterication
reaction. Thus having the {0 0 0 1} surface exposed to reactants becomes important.
In our previous work [20], we have found a strong interaction
between zinc and lanthanum species. Lanthanum partially replaces
zinc atoms in ZnO crystal. Thus, it weakens the bond of some of the
oxygen atoms in its neighborhood, and makes them more reactive
[25]. Therefore, the Zn3La1 catalyst shows a higher activity for
transesterication than ZnO. Since the crystal of wurtzite ZnO
doped with La is stable under reaction conditions, it has a long catalytic life in transesterication.
Fig. 11 illustrates a possible transesterication reaction mechanism for the catalytic process. For fresh catalysts some amorphous
fractions cover the (0 0 0 1) plane of ZnO. Therefore the catalyst
shows a lower activity in the initial period. After a washing process, amorphous material partially dissolves in reaction mixture
and active centers (atomic oxygen) are exposed. The bulk terminated ZnO (0 0 0 1)-O is composed of a topmost layer of nearlyclose packed oxygen anions, located at a short distance of 0.6
above a second layer of Zn cations and 2.6 above a second layer
of O anions [26,27]. Methanol is adsorbed on the oxygen and then
the oxygen anion forms. The nucleophilic attack of alcohol to the

S. Yan et al. / Fuel 89 (2010) 28442852

2851

Fig. 10. TEM images of fresh and long term used Zn3La1 catalysts. Fresh Zn3La1 (a), a ZnO nanoparticle in fresh Zn3La1 (b), HRTEM image of a ZnO nanoparticle in fresh Zn3La1
(c), ZnO nanoparticles sticked by small fractions (d) and the 32 days used Zn3La1 (e). Note that both fresh Zn3La1 and the used Zn3La1 are a mixture of nanoparticles and
amorphous material.

2852

S. Yan et al. / Fuel 89 (2010) 28442852

Fig. 11. Schematic representation of possible mechanism for transesterication of triglyceride with methanol.

esters produces a tetrahedral intermediate. Then the hydroxyl


group breaks and forms two kinds of esters. This transesterication
mechanism can be extended to di- and mono-glycerides.
4. Conclusion
The synthesis of FAME from unrened and waste oils was investigated using modied ZnO nanoparticles as catalysts. We have
found that at 200 C the catalyst is active in both transesterication
and esterication reactions, and can be directly used in inexpensive oil systems for biodiesel production, such as crude algae oil,
crude corn oil from DDGs, crude coconut oil, crude palm oil, crude
soybean oil, and waste cooking oil. There is a strong interaction between zinc and lanthanum species which is closely related with
long catalyst life and stable crystal structure under reaction conditions. At 200 C, this catalyst has continuously run for 70 days in a
x bed reactor and has been reused for 17 cycles in a batch stirred
reactor. Furthermore, leaching of catalyst components in reaction
mixtures and product streams is negligible. Hence this class of catalysts, which is relatively inexpensive because of low raw materials and manufacturing cost, signicantly simplies the oil
pretreatment process and product purication process, and greatly
decreases the feedstock cost and production cost of biodiesel.
Acknowledgements
Financial support from the Department of Energy (Grant
DEFG36-05GO85005) and Michigans 21st Century Job Fund is
gratefully acknowledged.

References
[1] Clark SJ, Wagner L, Schrock MD. J Am Chem Soc 1984;61:16328.
[2] Muniyappa PR, Brammer SC, Noureddini H. Bioresour Technol 1996;6:1924.
[3] Crabbe E, Nolasco HC, Kobayashi G, Sonomoto K. Process Biochem
2001;37:6571.
[4] Helwani Z, Othman MR, Aziz N, Kimc J, Fernando WJN. Appl Catal A: Gen
2009;363:110.
[5] Jothiramalingam R, Wang MK. Ind Eng Chem Res 2009;48:616272.
[6] Freedman B, Pryde EH, Mounts TL. J Am Oil Chem Soc 1984;61:163843.
[7] Ma F, Clements LD, Hanna MA. Trans ASAE 1998;41:12614.
[8] Canakci M. Bioresour Technol 2007;98:18390.
[9] Johnston M, Holloway T. Environ Sci Technol 2007;41:796773.
[10] Kim HJ, Kang BS, Kim MJ, Park YM, Kim DK, Lee JS, et al. Catal Today
2004;93:31520.
[11] Suppes GJ, Dasari MA, Doskocil EJ, Mankidy PJ, Goff MJ. Appl Catal A: Gen
2004;257:21323.
[12] Yan S, Lu H, Liang B. Energy Fuels 2008;22:64651.
[13] Granados ML, Poves MDZ, Alonso DM, Mariscal R, Galisteo FC, Moreno-Tost R,
et al. Appl Catal B: Environ 2007;73:31726.
[14] Granados ML, Alonso DM, Sdaba I, Mariscal R, Ocn P. Appl Catal B: Environ
2009;89:26572.
[15] Martyanov IN, Sayari A. Appl Catal A: Gen 2008;339:4552.
[16] Kim M, Yan S, Salley SO, Ng KYS. Catal Commun 2009;10:19139.
[17] Bournay L, Hillion G. European; 2003.
[18] <http://www.ifp.com/actualites/dossiers/les-biocarburants>.
[19] Bournay L, Casanave D, Delfortb B, Hillionb G, Chodorge JA. Catal Today
2005;106:1902.
[20] Yan S, Salley SO, Ng KYS. Appl Catal A: Gen 2009;353:20312.
[21] Yan S, Salley SO, Ng KYS. US; 2008.
[22] Spivey JJ. Catalysis. Royal Society of Chemistry; 2002.
[23] Cheng WH, Akhter S, Kung HH. J Catal 1983;82:34150.
[24] Akhter S, Lui K, Kung HH. J Catal 1984;85:43756.
[25] Lindsay R, Michelangeli E, Daniels BG, Ashworth TV, Limb AJ, Thornton G, et al.
J Am Chem Soc 2002;124:711722.
[26] Tasker PW. J Phys C: Solid State Phys 1979;12:497784.
[27] Lindsay R, Michelangeli E, Daniels BG, Ashworth TV, Limb AJ, Thornton G, et al.
J Am Chem Soc 2002;124:711722.

Вам также может понравиться