Sie sind auf Seite 1von 13

Available online at www.sciencedirect.

com

Fuel 87 (2008) 933–945


www.fuelfirst.com

Catalytic deoxygenation of unsaturated renewable feedstocks


for production of diesel fuel hydrocarbons
M. Snåre, I. Kubičková, P. Mäki-Arvela, D. Chichova, K. Eränen, D.Yu. Murzin *

Laboratory of Industrial Chemistry, Process Chemistry Centre, Åbo Akademi University, Biskopsgatan 8 FIN-20500 Åbo-Turku, Finland

Received 25 October 2006; received in revised form 4 June 2007; accepted 5 June 2007
Available online 16 July 2007

Abstract

The liquid-phase deoxygenation reaction of unsaturated renewables has been investigated in a semi-batch reactor. The reactants
examined were the monounsaturated fatty acid, oleic acid, the diunsaturated fatty acid, linoleic acid and the monounsaturated fatty acid
ester, methyl oleate. The reactions were carried out over a Pd/C catalyst under constant pressure and temperature in the following
domain, 15–27 bar and 300–360 C, respectively. The influence of carrier gas was additionally investigated. The impact as solvent (mesit-
ylene) was studied as well and reaction pathways were proposed. Furthermore, continuous deoxygenation experiments were conducted,
facilitating understanding of the catalyst stability and catalyst deactivation. The deoxygenation catalyst was characterized by physisorp-
tion, temperature programmed desorption (TPD), thermogravimetric analysis (TGA), X-ray photoelectron spectroscopy (XPS) and
scanning electron microscopy (SEM).
 2007 Elsevier Ltd. All rights reserved.

Keywords: Deoxygenation; Renewables; Diesel

1. Introduction 100,000 kg oil/km2, but oil yields of 10–20 times are


reported [3]. Moreover, non-edible and high oil yielding
There are several reasons for the growing interest of uti- crops that can withstand desert like areas and grow with
lizing renewable feedstocks as alternative fuels for the a minimal amounts of water are being developed [4–7].
future, such as the rapid decrease of fossil fuel reservoirs Natural oils and fats are complex mixtures of triglycer-
and currently existing major environmental concerns. ides, which contain three fatty acids and a glycerol moiety
Natural oils and fats are renewable feeds that could be [8]. As it is demonstrated in Table 1 [9–11], natural oils and
used for energy utilization. Approximately 100 million tons fats comprise of saturated and unsaturated fatty acids with
of oils and fats were produced worldwide in 2003 [1], from a carbon numbers of C4 and C24, typically C16 and C18.
animal and vegetable feedstocks. At that time the produc- The high carbon numbers, hence high heating values, of
tion of biodiesel (defined as alkyl fatty acid esters) was, the fatty acid alkyl chains in the natural oils and fats make
nevertheless, relatively low, 1.4 million tons [2]. The annual them promising potential candidates as diesel fuel if selec-
production of oils and fats, however, could be largely tively deoxygenated. The selective deoxygenation can be
increased, without touching the sensitive ethical dilemma done by direct removal of the carboxyl group via carbon
of using farmland for energy production. Recently, novel dioxide and/or carbon monoxide release [12] (decarboxyl-
solutions have been made in order to increase availability ation and/or decarbonylation). It has recently been demon-
and oil yields. Typically, an oil-bearing crop can yield ca. strated that saturated feeds over heterogeneous catalysts in
the liquid-phase tend to deoxygenate [13–15], when the
*
Corresponding author. Tel.: +358 2 215 4985; fax: +358 2 215 4479. reactants are brought together with, i.e. Pd/C catalyst
E-mail address: dmurzin@abo.fi (D.Yu. Murzin). under elevated temperatures and pressures.

0016-2361/$ - see front matter  2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2007.06.006
934 M. Snåre et al. / Fuel 87 (2008) 933–945

Table 1
Content of fatty acids (wt%) in typical natural oils and fats [9–11]
Natural fats and oils Fatty acids
Saturated Unsaturated
C4:0 C6:0 C8:0 C10:0 C12:0 C14:0 C16:0 C18:0 C20:0 C22:0 C24:0 C16:1 C18:1 C20:1 C22.1 C18:2 C18:3
a
Castor 1.1 3.1 4.9 1.3
Coconut 7.1 54.1 17.4 6.1 1.6 5.1 1.3
Linseed 5.1 2.5 0.3 18.9 18.1 55.1
Maize 3.9 11.2 1.8 25.4 60.3 1.1
Olive 7.3 11.0 2.2 67.0 0.3 8.9 0.6
Palm 2.5 40.8 3.6 45.2 7.9
Palm kernel 0.2 4.0 4.3 50.4 17.3 7.9 2.3 11.9 2.1
Peanut 0.1 0.7 0.4 13.7 2.3 1.3 3.0 1.2 0.1 47.8 29.2
Rapeseed 0.6 0.1 5.1 2.1 0.2 0.2 57.9 1.0 0.2 24.7 7.9
Sesame 13.1 3.9 52.8 30.2
Soybean 0.1 0.3 10.9 3.2 0.1 0.1 0.3 24.0 54.5 6.8
a
Castor oil contains ricinoleic, 89.6%, which is a fatty acid with a hydroxyl group CX:Y X and Y defines number of carbons and number of double bonds,
respectively, in the fatty acid molecule.

Nonetheless, a great portion of natural oils and fats con- metric analysis (TGA), X-ray photoelectron spectroscopy
tain unsaturated compounds (Table 1) and must therefore (XPS), temperature programmed desorption of hydrogen
be taken into consideration in the deoxygenation biodiesel (TPD-H2) and scanning electron microscopy (SEM). The
technology. analytical procedure of nitrogen physisorption technique
Production of hydrocarbons by thermolysis of vegetable can be found elsewhere [12].
oils was performed over metal oxide catalysts [16]. The pre-
sented results showed that during the thermolysis severe 2.1.1. X-ray photoelectron spectroscopy (XPS)
cracking occurred. Analogously production of synthetic X-ray photoelectron spectroscopy (XPS) measurements
crude petroleum was carried out over an alumina catalyst were utilized to determine the oxidation states of hydrogen
[17]. Although, several other studies of hydrocarbon pro- pretreated and non-pretreated palladium catalyst. The XPS
duction from natural oils and fats have been reported experiments were carried out in a Perkin-Elmer 5400 X-ray
[18–21], no selective deoxygenation of triglyceride based spectrometer with a base pressure below 9.3 · 107 Pa. Mg
feeds for the production of diesel fuel hydrocarbons, Ka X-rays (1253.6 eV) was used as a primary excitation at
according to our knowledge, has been demonstrated. 200 W power. The hemispherical energy analyzer, with
Studies related to hydrocarbon production from natural instrumental resolution not lower than 0.53, was used with
oils and fats are mainly performed over homogeneous cat- 90 angle between X-ray source and analyzer.
alysts [22,23]. However, gas phase selective heterogeneous
catalytic decarboxylation of carboxylic acids, such as octa- 2.1.2. Scanning electron microscopy–energy dispersive X-ray
noic, benzoic, salicylic acid, etc., has been demonstrated Analysis (SEM–EDX)
over palladium and nickel catalysts [24]. Furthermore, The morphology and the chemical composition of the
selective manufacture of unsaturated linear hydrocarbons catalyst surface were analyzed by scanning electron micros-
from saturated fatty acid and fatty acid esters has been copy. The LEO 1530 scanning system was equipped with a
reported over a nickel type catalyst [25]. According to the ThermoNoran Vantage analyzer using an acceleration
previous results obtained from a thorough catalyst screen- voltage of 13 keV and a take-off angle of 34.5153.
ing study in the deoxygenation of stearic acid [12], palla-
dium on carbon is the catalyst of choice, since it 2.1.3. Thermogravimetric analysis (TGA)
displayed superior deoxygenation properties. Analysis of coke deposits on spent catalysts was per-
The aim of the present work is to investigate the possi- formed by a thermogravimetric technique. The spent cata-
bility to deoxygenate unsaturated renewables yielding die- lyst (5 mg) was burnt in synthetic air with a Mettler
sel-like hydrocarbons over palladium catalysts, which Toledo TGA/SDTA 851e instrument. The temperature
were previously demonstrated to be efficient for saturated ramp of 10 C/min was applied and the catalyst weight loss
fatty acids and esters. as a function of temperature was registered.

2. Experimental 2.1.4. Temperature programmed desorption of hydrogen


(TPD-H2)
2.1. Catalyst characterization Temperature programmed desorption of hydrogen on
the palladium catalyst was performed with the Micromeri-
The 5 wt% Pd on charcoal catalyst, provided by Aldrich, tics Autochem 2910 coupled to GC–MS (Balzers Instru-
was characterized by nitrogen physisorption, thermogravi- ments, Omnistar). The catalyst was reduced with flowing
M. Snåre et al. / Fuel 87 (2008) 933–945 935

hydrogen prior to TPD with flowing hydrogen with follow- The catalyst reduction was followed by direct heating and
ing temperature program: 5 C/min–200 C (2 h), after pressurizing from reduction temperature and pressure
which the adsorbed hydrogen was flushed with helium for (mentioned previously) to desired reaction temperature
30 min. TPD was carried out with a temperature program: and pressure, respectively. The reaction mixture was con-
10 C/min–1000 C and a carrier gas flow rate 50 ml/min. tinuously fed through the catalyst bed with the volumetric
flow of 0.05 ml/min.
2.2. Experimental setup and procedure
2.3. Product analysis
The catalytic experiments were carried out in a 300 ml
semi-batch reactor coupled to a condenser and a heating 2.3.1. Liquid phase analysis
jacket. The reaction temperature was kept constant during Liquid phase samples were withdrawn from the reactor
reaction at temperatures between 300 and 360 C. The vessel via a sampling valve during the experiments. Sam-
reaction pressure was adjusted in order to keep the reaction ples were silylated with N,O-bis(trimethyl)-trifloroaceta-
in liquid phase, generally between 1.5 and 4.2 MPa. Stir- mide, BSTFA (Acros Organics, 98+%) in order to
ring speed was maintained suitably high, 1100 rpm, to pre- analyze by GC. Generally 100 wt% excess of BSTFA was
vent external mass transfer limitations. The flow of carrier added to the sample. After addition of silylation agent,
gas and the reaction pressure were controlled by a flow the samples were kept in an oven at 60 C for 30 min.
(Brooks 58505S) and a pressure controller (Brooks 5866), The samples were analyzed ex situ with a gas chromato-
respectively. graph (GC, HP 6890, Series). The GC was equipped with
A comparison of batch and continuous operations were a non-polar colonn (DB-5) with dimensions of
performed. The continuous deoxygenation of unsaturated 60 m · 0.32 mm · 0.5 lm and with a FI- detector. One
renewables was carried out in a fixed bed tubular reactor microliter of sample was injected into the GC with a split
(reactor dimension: length, 175 mm; internal diameter, ratio 50:1 and the carrier gas (helium) flow rate was
4.4 mm) over the same commercial Pd catalyst. The reac- 137 ml/min. The injector and detector temperature were
tion was carried out in an upward two-phase system and 265 C and 290 C, respectively. The following chromato-
thus the pressurizing media (inert gas) was introduced after graphic temperature program was used for analysis:
the reactor. 130 C (constant 5 min)–169 C (1 C/min, constant
5 min)–246 C (5 C/min)–250 C (1 C/min, constant
2.2.1. Reaction chemicals 3 min)–300 C (5 C/min, constant 45 min). The chromato-
The commercial 5 wt% Pd supported on carbon graphic pressure program was as well adjusted to achieve
(20.568-0) was supplied by Aldrich. The model compounds satisfactory separation of the desired product and its iso-
used were, oleic acid (>58% purity, Riedel-de Haën), lino- mers. The initial pressure at 172.4 kPa was kept for
leic acid (60–74% purity, Fluka) and methyl oleate (>70% 30 min, thereafter the pressure was ramped with
purity, Aldrich). Detailed composition of technical grade 34.5 kPa/min until reaching the ending pressure 221 kPa.
feedstock as well as the influence of feed impurities on catal- A number of chemical standards were purchased enabling
ysis will be discussed below. The reaction and the reduction product identification and calibration. Product identifica-
gases, argon, hydrogen (5 vol%)–argon mixture and tion was validated with a gas chromatograph –mass spec-
hydrogen were provided by AGA. The reaction solvent, trometer (GC–MS).
mesitylene (C9H12), was supplied by Fluka (>98% purity).
3. Results and discussion
2.2.2. Reaction procedure
The Pd/C catalyst (1 g) was reduced in situ by hydrogen at 3.1. Catalyst characterization
constant reduction temperature and pressure, 200 C and
2 bars, respectively. The catalyst reduction was followed by 3.1.1. Temperature programmed desorption of hydrogen
introduction of 63 g of the solvent to the reactor via a bub- Desorption of hydrogen from the Pd/C catalyst exhibits
bling unit preventing reoxidizing the catalyst. After solvent a maximum at 200 C. This maximum amount of hydrogen
introduction the reactor was opened and 23 g (25 wt%) is desorbed from the catalytically active metal. Further-
of feed was placed into it. The reactor was then flushed more, the hydrogen desorbed above 800 C is originating
thoroughly with an inert gas. Before the reaction started from dehydrogenation of the carbon support [26] and it
25 ml/min of inert gas was passed through the reactor until is indicating gasification of the support, i.e. aromatic com-
reaching the reaction pressure. Afterwards the reactor mix- pounds and hydrogen are formed. Other gaseous products,
ture was heated with a temperature rate of 15 C/min to such as CO2 and H2O, have desorption maximum at 250
desired reaction temperature. At this point stirring and reac- and 220 C, respectively, while CO has a maximum at
tion time started. After the reaction the catalyst particles were 700 C (Fig. 1). Based on the literature [26–28] CO2 origi-
filtered with 200 ml of acetone for further characterization. nates from lactone, anhydride and/or carboxyl groups on
In the continuous system, the 5% Pd/C powder catalyst the catalyst support, while formation of CO is thought
was placed between a layer of quartz sand and quartz wool. to arise from the decomposition of quinonic, phenolic,
936 M. Snåre et al. / Fuel 87 (2008) 933–945

1.00E-12 2.00E-12 Furthermore, the carbon support seems to originate from


woody raw materials, since the particles are rather irregular
8.00E-13 1.60E-12 and they exhibit a channel-shaped structure [29]. Palladium
H2 agglomerates of around 2 lm are observed (Fig. 2).
6.00E-13 1.20E-12 Predictably, the catalyst particles did not notably change
CO in their shape after pretreatment with hydrogen. Moreover,
a.u.

no sintering on micrometer scale was detected after pre-


4.00E-13 8.00E-13
treatment, hence, the visual metal particles maintained
H 2O
their original size (Fig. 3). Nevertheless, sintering in princi-
2.00E-13 4.00E-13
ple can occur, on the nanoscale, since hydrogen pretreated
CO2
CH4 palladium particles supported on an activated carbon sim-
0.00E+00 0.00E+00 ilar to those used in the present study have been reported to
0 200 400 600 800 1000
Temperature (°C)
agglomerate [30].
As expected, the palladium on carbon catalyst surface
Fig. 1. Temperature programmed desorption of hydrogen and other was mainly composed of C, O and Pd (84.2%, 8.5% and
gaseous emissions. 6.2%). The trace compounds were Na, S, Cl, Si and Fe
(Table 2).
carbonyl and/or ether groups. The carboxyl group is pro-
posed to decompose in the temperature range of 100– 3.1.3. X-ray photoelectron spectroscopy (XPS)
400 C and lactone at higher temperature. Water probably The non-pretreated and the hydrogen pretreated palla-
evolves from moisture and/or the removal of oxygen, orig- dium catalysts were examined by XPS analysis. The metal-
inating from catalyst O2-passivation (typical technique for lic Pd was observed in Pd 3d5/2 with the binding energy
commercial catalysts) and/or the carbon support, by (BE) of 335.2–335.4 eV, which is close to the values
hydrogen. reported in [31,32]. As anticipated, the relative value of
metallic palladium increased with the hydrogen pretreat-
3.1.2. Scanning electron microscopy–energy dispersive X-ray ment procedure, the non-pretreated catalyst contained
analysis (SEM–EDX) 14%, while the pretreated catalyst had 59% of metallic state
The metal supported carbon particles are in the region palladium. The palladium oxide in Pd 3d exhibited a bind-
of 20 lm and very porous based on SEM analysis, which ing energy of 336.0–336.3 eV in the Pd/C catalyst [33].
is in accordance with previously stated textural properties. PdCl2 exhibited binding energies of 336.7–337.2 eV, which

Fig. 2. SEM images of non-pretreated Pd/C catalyst at different magnifications.

Fig. 3. SEM images of a pretreated catalyst at different magnifications.


M. Snåre et al. / Fuel 87 (2008) 933–945 937

Table 2 the average diameter of metal crystallites is 6.2 nm. How-


Surface composition of 5% palladium on carbon measured by SEM–EDX ever, an investigation of dispersion of palladium at
Element Fraction (wt%) Atom (%) 350 C reduction temperature showed that the dispersion
C 84.26 91.82 decreased to 6% [30]. Mean catalyst particle size is 15 lm
O 8.48 6.94 (Table 4) to avoid interference of internal mass transfer.
Al 0.01 0.01 A decrease of surface area was observed for all examined
S 0.26 0.1
Cl 0.13 0.05
reactants. The most severe surface area decrease of spent
Fe 0.12 0.03 catalyst, indicating coke formation and catalyst deactiva-
Na 0.36 0.2 tion, was detected in the linoleic acid experiment and for
Si 0.12 0.06 deoxygenation of oleic acid at higher temperatures. The
Mg 0.02 0.01 surface area and the pore size distribution of fresh and used
Pd 6.15 0.76
Ca 0.04 0.01
catalysts are illustrated in Table 5.
P 0.05 0.02
3.1.5. Thermogravimetric analysis (TGA)
Table 3 The spent catalysts were studied by TGA, where the cat-
State of palladium compounds supported on carbon for non-pretreated alyst was burnt in synthetic air and the weight loss was reg-
and pretreated catalyst measured by XPS istered. The catalyst used in deoxygenation of linoleic acid
State of Pd/C under Ar atmosphere demonstrated higher relative amount
palladium of weight loss in the temperature range of 200–800 C, than
Non-pretreated Pretreated
Relative area BE (eV) Relative area BE (eV) the catalyst used for oleic acid deoxygenation under
(%)a (%)a Ar + H2 atmosphere (Fig. 4). Furthermore, the catalyst
Pd(0)/C 14.5 335.3–335.4 56.0 335.3–335.4 used in deoxygenation of methyl oleate under pure hydro-
PdO (d+)/C 9.2 335.9 22.3 336.0–336.3 gen atmosphere illustrated the lowest relative amount of
PdCI2 (ll)/C 54.8 336.9–337.0 10.9 336.7–337.2 weight loss as a function of temperature, thus, indicating
337.4–337.9 337.7–338.2 that lighter coke molecules are formed in inert atmosphere
Other 21.6 339.1–339.2 10.9 339.4–339.8
than under hydrogen rich conditions. Possible explanation
a
Relative to the Pd 3d peak. is that CO, which is formed during deoxygenation, is deac-
tivating the catalyst in inert atmosphere, while under
are in accordance with the literature [32]. The total chlorine hydrogen atmosphere CO is being reduced by hydrogen
content was relatively high also according to EDXA anal- to CH4.
ysis in the non-pretreated fresh Pd/C, whereas the chlorine
content in the pretreated catalyst was much lower. Other 3.2. Catalytic deoxygenation of unsaturated renewables
binding energies observed were 337.4–338.2 and 339.1–
339 eV (Table 3). The most common unsaturated fatty acids in vegetable
oils, e.g. oleic acid and linoleic acid as well as oleic acid
3.1.4. Textural properties of the deoxygenation catalyst methyl ester were investigated (Table 1 and Fig. 5) and
The commercial palladium on carbon is a high surface methyl oleate.
area catalyst with specific pore volume of 0.43 ml/g. The The experiments were conducted under constant tem-
metal dispersion is 18% after pretreatment at 200 C, thus perature (300–360 C) and pressure (15–27 bar) with tech-

Table 4
Textural properties of fresh 5% Pd/C catalyst [14]
Metal loading Palladium metal Particle size Surface area Specific pore volume
(wt%) Surface area (m2/gmet) Dispersion (%) Mean (lm) >50 lm (%) (m2/gcal) (ml/gcat)

5 80.2 18.0 15 96.6 1214 0.431

Table 5
Physisorption results of spent 5% Pd/C catalysts used under different feeds
Reactant Reaction conditions N2-physisorption
Temperature (C) Pressure (bar) Surface area Micropore volume
2
Used (m /gcat) Decrease (%) Used (ml/gcat) Decrease (%)
Oleic acid 300 15 (Ar) 492 59.5 0.19 55.9
Oleic acid 330 21 (Ar + H2) 250 79.4 0.09 79.1
Linoleic acid 300 15 (Ar + H2) 367 69.8 0.19 55.9
Methyl-oleate 300 15 (H2) 591 51.3 0.21 51.3
938 M. Snåre et al. / Fuel 87 (2008) 933–945

100 Table 6
Oleic acid (330 C, Ar+H2)
Composition of technical grade oleic acid, linoleic acid and methyl oleate
based on GC analysis
Linoleic acid (300 C, Ar+H2)
Weight loss of combustibles (wt%)

80
Methyl oleate (300 C, H2) Fatty acid Structure Technical grade feedstocks
Oleic Linoleic Methyl
acid acid oleatea
60
Palmetic acid C16:0 4.3 8.4 4.7
Palmitoleic C16:1(cis-9) 4 0 3.8
acid
40 Stearic acid C18:0 1.1 4.8 1.7
Oleic acid C18:1(cis-9) 63.7 15.1 59.8
Elaidic acid C18:1(trans-9) 11.2 0.9 6.8
20 Vaccenic acid C18:1(cis and 0.7 0.1 0.6
trans-11)
Linoleic acid C18:2(cis–cis-9-12) 5.2 67.6 12.1
Other 9.8 3.1 10.5
0
150 250 350 450 550 650
impurities
a
Temperature (°C) Fatty acids in the methyl ester molecule.

Fig. 4. Weight loss in TGA for catalysts used in the deoxygenation of


unsaturated feeds.
ation reaction was additionally accomplished via hydroge-
nation of double bonds and further through deoxygenation
of saturated acid (Fig. 6).
O
O Under inert atmosphere the source of hydrogen for the
CH3-(CH2)7- -(CH2)7- C CH3-(CH2)4- -(CH2)7- C hydrogenation process was provided through a simulta-
Linoleic acid OH
Oleic acid OH neous, hydrogen re-distribution reaction, e.g. dehydroge-
O nation of the feed to unsaturated by-products such as
CH3-(CH2)7- -(CH2)7- C diunsaturated acids, triunsaturated acids as well as aromat-
Methyl-oleate O-CH3 ics and additionally through preadsorbed H2 on the surface
(Figs. 6–8). Furthermore, positional and geometrical isom-
Fig. 5. Chemical structure of the unsaturated model compounds, oleic
acid, linoleic acid and methyl oleate.
erization of oleic acid, e.g. to elaidic and vaccenic acid,
seems to be the intermediate reactions prior to the hydro-
genation step, which is in accordance with literature
nical grade reactants. Notably, when using technical grade [34,35]. Thereafter, the deoxygenation reaction is taking
chemicals one has to take into consideration the analytical place, the main deoxygenation product is n-heptadecane,
complexity. The three different feeds varies substantially in but minor amounts of unsaturated positional isomers
purity and the impurities can be cumbersome to separate (1-, 3-, and 8-heptadecene) and C17 aromatics are detected.
from the main compound in the gas chromatograph In addition, some formation of heavier products was
(GC) leading to a summation of isomers in order to achieve observed as well as the traces of lighter products formed
reprocessable data. The initial concentrations of the model by cracking. The heavier molecules could not be properly
compounds and the main impurities according to GC anal- identified, however, according to GC–MS analysis the
ysis are summarized in Table 6. molecular masses were above 400.
The catalytic performance was evaluated based on initial A summary of the experiments conducted with unsatu-
rate of feed consumption after 1 min, conversion of feed rated renewables is provided in Table 7.
after 6 h and selectivity towards product formation. Con-
version of feed is defined as
3.2.1.1. Effect of solvent and pretreatment under inert
C 0;RC18:Y  C t;RC18:Y atmosphere in oleic acid deoxygenation. The effect of cata-
Xt ¼  100% ðIÞ
C 0;RC18:Y lyst pretreatment by hydrogen and the solvent nature were
investigated under inert atmosphere at 300 C. The results
where C 0;RC18:Y is initial concentration of all unsaturated showed that the pretreated catalyst, accelerated initially the
fatty acids or esters. In oleic acid experiments this includes isomerisation and hydrogenation rates compared to the
vaccenic acid and elaidic acid, while in linoleic acid exper- non-pretreated catalyst. The higher rates are presumably
iments, linoleic isomers are included. Similarly methyl ole- due to differences in palladium oxidation states, since the
ate isomers are incorporated in the original feed. At the hydrogen pretreated catalyst had a significantly higher
arbitrary time, t, the feed concentration is C t;RC18:Y . amount of metallic palladium than the non-pretreated
one as from XPS results. The solvent free reaction exhib-
3.2.1. Deoxygenation of monounsaturated fatty acid ited the same degree of conversion even though the concen-
Preliminary results indicated that direct deoxygenation tration was fourfold (0.83 mol/l and 3.15 mol/l) compared
of oleic acid to olefins occured. However, the deoxygen- to the experiment conducted in mesitylene. Since the
M. Snåre et al. / Fuel 87 (2008) 933–945 939

Heavy byproducts (dimers, aromatics) Light byproducts (cracking)

Polyunsaturated acids
Other C18:X acids
+H2/-H2
O
CH3-(CH2)4- -(CH2)7- C
+H2/- Linoleic acid OH +H2/-
H2
H2 O
O
CH3-(CH2)7- CH3-(CH2)5- -(CH2)9- C
-(CH2)7- C
OH cis-Vaccenic acid OH
Oleic acid

CH3-(CH2)7- O CH3-(CH2)5- O
+H2
-(CH2)7- C +H2 -(CH2)9- C
Elaidic acid OH +H2 +H2 trans-Vaccenic acid OH
O
CH3-(CH2)16- C -CO2
OH
Stearic acid
Other C17 hydrocarbons
-CO2

CH3-(CH2)15-CH3
Heptadecane

Fig. 6. Reaction scheme of oleic acid deoxygenation under inert atmosphere.

9
70 Linoleic acid
8
Oleic acid
60 7 n-heptadecane
Concentration (mol%)

Elaidic acid
Other C17-products
Vaccenic acid 6
Lighter
Concentration (mol%)

50
5 Heavier
Other C18:x -acids
40 4
Stearic acid

3
30
2
20
1

0
10 0 60 120 180 240 300 360
Reaction time (min)
0
0 60 120 180 240 300 360 Fig. 8. Typical final product concentration profiles in the catalytic
Reaction time (min) deoxygenation of oleic acid over 5% Pd/C. Reaction conditions: coleic acid =
3.15 mol/l (solvent free), mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼
Fig. 7. Typical reactant and intermediate concentration profiles in the 25 ml=min ðArÞ.
catalytic deoxygenation of oleic acid over 5% Pd/C. Reaction conditions:
coleic acid = 3.15 mol/l (solvent free), mcatalyst = 1 g, T = 300 C,
p = 15 bar, V_ carrier gas ¼ 25 ml=min ðArÞ.
products are not present in solvent free systems, whereas
in mesitylene a relatively high amount of heavies are
solvent, mesitylene, is highly unsaturated it should not act detected. Additionally, the very concentrated solvent free
as an hydrogen donor, but on the contrary as a hydrogen mixture might be affected by mass transfer limitations
acceptor [36]. Subsequently, the deficit of hydrogen in the (Table 6 (runs I–III) and Fig. 9).
solvated system retards the hydrogenation step, while the
solvent free condition enhances the reaction (Figs. 7–9b). 3.2.1.2. Effect of reaction atmosphere in oleic acid deoxy-
The presence of hydrogen is, of course, necessary in the genation. Separate experiments were conducted under
hydrogenation step, however, hydrogen is, as well, known hydrogen atmosphere in order to suppress the dehydroge-
to enhance the isomerisation rate [35,37]. When comparing nation step. Consequently, oleic acid was deoxygenated
selectivity towards formation of heavier products in mesit- under an Ar + H2 atmosphere. As expected, over Pd/C in
ylene and solvent free system, it is noticed that heavier the presence of hydrogen the conversion of unsaturated
940
Table 7
Summary of semi-batch deoxygenation results after 6 h reaction time
Run Reactant Reaction conditions Conversion Selectivityd
c
Temperature Pressure XMain feed XTotal feed SME-SA SSA Sn-C17 S RC17 SLight SHeavy
(C) (bar) (%) (%) X = 35% t=6h X = 35% t=6h X = 35% t=6h X = 35% t=6h X = 35% t=6h X = 35% t=6h
a
I Oleic acid 300 15 (Ar) 79 49 – – 79 77 4 6 9 9 7 7 1 0

M. Snåre et al. / Fuel 87 (2008) 933–945


IIb Oleic acid 300 15 (Ar) 93 45 – – 63 64 5 5 18 17 13 12 1 2
III Oleic acid 300 15 (Ar) 74 50 – – 61 60 4 6 20 21 12 9 3 4
IV Oleic acid 300 15 98 78 – – 57 63 4 14 20 13 7 9 12 2
(Ar + H2)
V Oleic acid 330 21 99 92 – – 54 25 8 22 23 14 13 24 3 14
(Ar + H2)
VI Oleic acid 360 27 99 91 – – 54 36 5 26 26 17 12 16 2 4
(Ar + H2)
VII Linoleic 300 15 100 34 – – 64 64 3 3 11 11 13 13 9 9
acid (Ar + H2)
VIII Methyl- 300 15 99 84 87 87 2 2 0 3 9 5 1 2 1 1
oleate (Ar + H2)
IX Methyl- 300 15 (H2) 100 96 – 71 – 0 – 26 – 2 – 1 – 0
oleate
X Methvl- 300 15 (Ar) 91 44 80 81 3 2 0 0 7 7 8 7 3 3
oleate
Reaction conditions: coleic acid in mesitylene = 0.83 mol/1 (solvent free = 3.15 mol/1), mcatalyst = l g, V_ carrier gas ¼ 25 ml=min ðArÞ.
a
No solvent, 100% oleic acid.
b
No catalyst pretreatment.
c
Conversion of total feed is defined as conversion of all identified unsaturated acids or esters.
d
Selectivity at 35% of conversion of total 1eed (linoleic acid at 34%) and after 6 h of reaction , Et-SA = methyl stearate, SA = stearic acid, n-C17 = n-heptadecane, RC17 ¼ other C17 products.
M. Snåre et al. / Fuel 87 (2008) 933–945 941

70
a 70 Linoleic acid Linoleic acid
Oleic acid Oleic acid
Elaidic acid 60 Elaidic acid
60
Vaccenic acid Vaccenic acid
Other unsaturated acids Other unsaturated acids

Concentration (mol%)
Concentration (mol%)

50 50 Stearic acid
Stearic acid
n-heptadecane
n-heptadecane
40 Other C17 products
Other C17 products 40

30
30

20
20

10
10

0
0 60 120 180 240 300 360 0
Reaction time (min) 0 60 120 180 240 300 360
Reaction time (min)
b 70 Linoleic acid
Oleic acid Fig. 10. Deoxygenation of oleic acid over Pd/C. Reaction conditions:
60
Elaidic acid coleic acid = 0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼
Vaccenic acid 25 ml=min ðAr þ H2 -mixtureÞ.
Other unsaturated acids
Concentration (mol%)

50 Stearic acid
n-heptadecane
40 Other C17 products Conversion of unsaturated C18-acids (%)
100
30
80
20

60
10

0 40 360°C & 27 bar


0 60 120 180 240 300 360
Reaction time (min) 330°C & 21 bar
20
300°C & 15 bar
Fig. 9. Effect of 5% Pd/C catalyst reduction in the deoxygenation of oleic
acid. Concentration profile versus reaction time for a) non pretreated and 0
b) hydrogen pretreated palladium catalyst. Reaction conditions: coleic acid = 0 60 120 180 240 300 360
0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼ Reaction time (min)
25 ml=min ðArÞ.
Fig. 11. The effect of temperature and pressure on conversion of
unsaturated C18-acids under H2:Ar (5:95 vol%) atmosphere. Reaction con-
ditions: coleic acid = 0.83 mol/l, mcatalyst = 1 g, V_ carrier gas ¼ 25 ml=min
acids increased compared to experiments under inert atmo- ðAr þ H2 -mixtureÞ.
sphere. Furthermore, the formation of unwanted polyun-
saturated acids was substantially diminished, whereas the
formation of saturated hydrocarbons increased (Table 6 The consumption rate of unsaturated acids is more
(runs III and IV) and Fig. 10). retarded after 1 h for the reaction at higher temperatures,
implying stronger catalyst deactivation at elevated temper-
3.2.1.3. Effect of reaction temperature and pressure in oleic atures. Most likely, the catalyst deactivation originates
acid deoxygenation. The effect of temperature in oleic acid from the solvent (as depicted earlier) or from the reactant
deoxygenation was studied in the temperature range of causing coke which covers the catalyst surface. The
300–360 C. The reaction pressure was varied in accor- assumption is strengthened by the surface area measure-
dance with the vapour pressure of the reaction mixture. ments of the spent catalysts (see Table 5). The specific sur-
Reaction rate, as expected, increased with increasing tem- face areas of the spent catalysts used at high reaction
perature. At 300 C the conversion of unsaturated acids temperatures exhibited the lowest surface areas. If the cat-
(lumped) was 78% after 6 h, while the conversion was alytically active metal particle is located inside the blocked
approximately 93% at 330 and 360 C (Fig. 11). The visible pore the catalytic activity is decreased at the same time.
‘‘conversion plateau’’ appearing in the beginning of the The selectivity patterns as a function of conversion
reaction is attributed to the internal isomerisation of unsat- seems to be dependent on temperatures, hence the forma-
urated acids. The isomerisation step seems to be sufficiently tion of products is affected by reaction temperature
slower at lower temperatures. The apparent activation (Fig. 12). At 300 C, the selectivity to stearic acid increases,
energy, based on the Arrhenius equation and first order initially, up to ca. 40% of conversion, thereafter formation
kinetics after 45 min, is 59 kJ/mol. of n-heptadecane is enhanced, indicating a consecutive
942 M. Snåre et al. / Fuel 87 (2008) 933–945

Stearic acid n-heptadecane reported in the deoxygenation of saturated fatty acid esters
60
[13] was not detected. One of the possible explanations is
fast deoxygenation of stearic acid compared to methyl stea-
50
rate. At the same time comparison of the reaction rates
40
indicate that more probable is the direct deoxygenation
Selectivity (%)

of methyl stearate (methyl oleate). Similarly, the deoxygen-


30
300°C & 15 bar ation of ethyl stearate was additionally suggested to pro-
330°C & 21 bar
ceed via direct formation of n-heptadecane through CO
360°C & 27 bar
20 and ethanol release [38].

10
3.2.2.1. Effect of reaction atmosphere in methyl oleate
0
deoxygenation. The effect of hydrogen atmosphere in the
0 20 40 60 80 100 deoxygenation reaction of methyl oleate was investigated
Conversion of unsaturated C18-acids (%) by using pure hydrogen, a mixture of Ar + H2 and inert
Fig. 12. The effect of temperature and pressure on product selectivities gas, Ar. The results demonstrated that hydrogen has a con-
versus conversion of unsaturated C18-acids under H2:Ar (5:95 vol%) siderable effect on the deoxygenation reaction. Methyl ole-
atmosphere. Reaction conditions: coleic acid = 0.83 mol/l, mcatalyst = 1 g, ate was immediately completely converted to methyl
V_ carrier gas ¼ 25 ml=min ðAr þ H2 -mixtureÞ. stearate under neat hydrogen atmosphere, while the hydro-
genation rate significantly decreased with lower partial
reaction. However, formation of heavier products is pre- pressures of hydrogen (Figs. 14 and 15).
dominant in the beginning of the reaction. Speculatively, Under Ar and Ar + H2 (5%) atmosphere the formation
the unidentified heavier product could be intermediate acid of C17 products was noticed at lower conversions indicat-
isomers, which could explain the observed phenomenon. ing that methyl oleate is reacting directly to olefinic C17
Furthermore, an increase in temperature slightly improved products, which is in accordance with previous reports.
the rate of the consecutive deoxygenation reaction. The hydrogenation and isomerisation commenced at
higher conversions. The formation of n-heptadecane was,
however, limited to the reactions taking place under pure
3.2.2. Deoxygenation of monounsaturated fatty acid methyl hydrogen atmosphere, where the hydrogenation and isom-
ester erisation steps were instantaneous, whereas in hydrogen
The deoxygenation of the oleic acid derivative, methyl deficit conditions, under Ar and Ar + (5%)H2 atmosphere,
oleate, was studied at 300 C over Pd/C. The deoxygen- the formation of deoxygenation products were barely
ation of methyl oleate proceeded via a similar pathway as detectable (Fig. 15).
previously stated in oleic acid deoxygenation (Fig. 13), Since the hydrogenation step was almost instantaneous
i.e. the main reaction routes were isomerization and hydro- under neat hydrogen atmosphere, the deoxygenation step
genation. Methyl stearate was an intermediate product could be investigated by plotting selectivity towards C17-
formed via hydrogenation of methyl oleate and subse- products versus conversion of the intermediate, methyl
quently deoxygenation to n-heptadecane. However, the stearate. Results showed high selectivity towards the
formation of the fatty acid intermediate, previously desired deoxygenation products (Fig. 16), with above

Linoleic methyl ester


100
Oleic methyl ester
70 Elaidic methyl ester
Vaccenic methyl ester
unsaturated metylesters (%)

60 Other unsaturated methyl ester 80


Stearic methyl ester
Concentration (mol%)

n-heptadecane
50 Other C17 products
Conversion of

60
40

30 40

20 Ar
20
10
Ar-5%H
Ar + 5%2
H2
Ar
0 0
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Reaction time (min) Reaction time (min)

Fig. 13. Deoxygenation of methyl oleate over Pd/C. Reaction conditions: Fig. 14. The effect of reaction atmosphere on conversion of unsaturated
cmethyl oleate = 0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼ methyl-C18 esters over 5% Pd/C. Reaction conditions: coleic acid = 0.83 mol/l,
25 ml=min ðAr þ H2 -mixtureÞ. mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼ 25 ml=min.
M. Snåre et al. / Fuel 87 (2008) 933–945 943

= methyl stearate, = n-heptadecane and = other C17 products 80 Linoleic acid


100 Oleic acid
Elaidic acid
70
80 Vaccenic acid
Other unsaturated acids
Selectivity (%)

Ar 60

Concentration (mol %)
60 Stearic acid
Ar-5%H2 n-heptadecane
Fig. 16 50 Other C17 products
40 H2
40
20
30
0
0 20 40 60 80 100 20
Conversion of unsaturated methyl esters (%)
10
Fig. 15. The effect of reaction atmosphere on product selectivities as a
function of unsaturated methyl-C18-ester over 5% Pd/C. Reaction 0
conditions: coleic acid = 0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, 0 60 120 180 240 300 360
V_ carrier gas ¼ 25 ml=min. Reaction time (min)

Fig. 17. Deoxygenation of linoleic acid. Reaction conditions: clinoleic acid =


0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, V_ carrier gas ¼ 25 ml=min
= n-heptadecane and = other C17 products
ðAr þ H2 -mixtureÞ.
100

80
conversion after 6 h, 78% and 84%, respectively. Under-
standably, the conversion level was lower for linoleic acid,
Selectivity (%)

60 since an additional hydrogenation step, from diunsaturated


to monounsaturated, is included. Moreover, catalyst deac-
40 tivation by coke formation was more pronounced with lin-
oleic acid as the feed, according to TGA and physisorption
20
results (Fig. 4 and Table 3). The selectivity towards desired
products at 35% of converted total unsaturated feed was
0
0 4 8 12 16 20 enhanced in reactions conducted with oleic acid. Further-
Conversion of saturated methyl esters (%) more, the formation of higher products was substantially
Fig. 16. The formation of C17-products versus conversion of the higher for the two acids than for the methyl ester (Table 7).
saturated methyl-C18-ester, methyl stearate over 5% Pd/C. Reaction
conditions: coleic acid = 0.83 mol/l, mcatalyst = 1 g, T = 300 C, p = 15 bar, 3.2.5. Continuous deoxygenation of unsaturated feeds
V_ carrier gas ¼ 25 ml=min ðH2 Þ. The impact of using unsaturated feeds in an more indus-
trial like process was investigated by deoxygenating oleic
80% of saturated fatty acid ester was converted to n-hepta- acid in solvent free conditions over the same Pd/C catalyst
decane and other C17 products (1-, 3- and 8-heptadecene). in a continuous fixed bed system (residence time of 8 min).
As previously noticed in the semi-batch study the deoxy-
genation reaction proceeds via initial isomerisation and
3.2.3. Deoxygenation of diunsaturated fatty acid hydrogenation step following the subsequent deoxygen-
Deoxygenation of the diunsaturated acid, linoleic acid, ation of saturated acid. Simultaneously dehydrogenation
was performed under the same conditions under Ar + H2 of oleic acid is taking place. The deoxygenation is as well
atmosphere as in oleic acid and methyl oleate experiments carried out via direct formation of olefins from unsaturated
in order to obtain comparable results. The deoxygenation acid. The concentrations of saturated, mono- and diunsat-
of linoleic acid proceeded analogously to the previously urated acids versus time-on-stream are illustrated in
described feedstocks by hydrogenation to monounsatu- Fig. 18. The catalyst behaviour seems to be fairly stable,
rated and saturated acids and finally via deoxygenation hence no significant catalyst deactivation was observed
to hydrocarbons (Fig. 17). under these conditions.
The product distribution is similar to those obtained in
3.2.4. Comparison of different unsaturated feeds the batch operations. The intermediate, stearic acid, is the
Comparison of the deoxygenation of different feeds was most abundant product formed (selectivity <70%). The
performed at 300 C with Ar + H2-mixture (Table 7, runs selectivity towards dehydrogenation and higher boiling
IV, VII and VIII). The comparison illustrated that all feeds products is ca. 20%, whereas the formation of hydrocar-
were completely converted after 6 h. Nonetheless, when bons, mainly olefins and aromatics, is moderately low
considering the total unsaturated acids or esters converted, (Fig. 19).
the diunsaturated linoleic acid exhibited the lowest conver-
sion level (34%) after 6 h of reaction, while monounsatu- 3.2.5.1. Comparison of productivity in batch and continuous
rated oleic acid and methyl oleate displayed higher mode. A comparison of operation modes were performed
944 M. Snåre et al. / Fuel 87 (2008) 933–945

60 reaction became predominant. Additionally, isomerisation


monounsaturated acid
(double bond migration) of oleic acid occurred prior to
50 hydrogenation and deoxygenation. Analogous isomerisa-
Concentration (mol %)

tion, hydrogenation and deoxygenation trends were


40
observed in experiments conducted under H2 atmosphere
with other unsaturated feeds. Furthermore, the deoxygen-
30
diunsaturated acid
ation of unsaturated feeds was performed in continuous
20
mode under conditions of strong mass transfer influence,
confirming, however, catalyst stability and potential indus-
10 trial applicability.
saturated acid

0 Acknowledgements
0 60 120 180 240
Time-on-stream (min)
This work is part of the activities at the Åbo Akademi
Fig. 18. Concentration of acids as a function of time-on-stream in Process Chemistry Centre (ÅA-PCC) within the Finnish
continuous deoxygenation reaction over Pd/C. The reaction conditions:
Centre of Excellence Programme (2000–2011) appointed
T = 300 C, p = 1 bar (Ar), coleic acid = 3.15 mol/l (100%), mcatalyst = 0.2 g,
V_ liquid flow ¼ 0:05 ml=min. by the Academy of Finland. Financial support from
TEKES is gratefully acknowledged. The authors express
their gratitude to Markku Reunanen for his contribution
90
to GC–MS analysis and to Clifford Ekholm for his helpful
75 SEM characterization measurements.
Selectivity (%)

n-heptadecane
60
other C17-products
References
45 stearic acid
dehydrogenation and heavier products
30 [1] World statistics, 1998–2003, United Soybean Board.
[2] Fulton L, Howes T, Hardy J. Biofuels for transport: an international
15 perspective. Paris: International Energy Agency (IEA); 2004.
[3] Huber GW, Iborra S, Corma A. Synthesis of transportation fuels
0
0 60 120 180 240 from biomass: chemistry, catalysis and engineering. Chem Rev
Time-on-stream (min) 2006;106:4044–98.
[4] Ghadge SV, Raheman H. Biomass Bioenerg 2005;28:601–5.
Fig. 19. Selectivity of products as a function of time-on-stream in [5] Gübitz GM, Mittelbach M, Trabi M. Bioresource Technol
continuous deoxygenation of oleic acid over Pd/C. The reaction condi- 1999;67:73–82.
tions: T = 300 C, p = 1 bar (Ar), coleic acid = 3.15 mol/l (100%), mcatalyst = [6] Giannelos PN, Zannikos F, Stournas S, Lois E, Anastopoulos G. Ind
0.2 g, V_ liquid flow ¼ 0:05 ml=min. Crop Prod 2002;6:1–9.
[7] Karmee SK, Chadha A. Bioresource Technol 2005;96:1425–9.
at similar reaction conditions with the exception of total [8] Barnwal BK, Sharma MP. Renew Sust Energ Rev 2005;9:363–78.
[9] Graboski MS, McCormick RL. Prog Energ Combust 1998;24:
pressure. The pressure of the inert gas should not, however
125–64.
affect the catalytic performance significantly, hence the [10] Pryde EH. Northern Regional Research Center, US Department of
pressure effect is assumed to be negligible. The productivity Agriculture, Peoria, Illinois, The American Oil Chemists’ Society,
in semi-batch operations under solvent free conditions is Champaign, IL; 1979.
29.2 mmol/h gcat, whereas in continuous fixed bed opera- [11] Department of Chemical Engineering, Instituto Superior Técnico,
Lisbon, 1997 http://journeytoforever.org/biofuel_library/che-
tions, the productivity was 2.3 mmol/h gcat. The lower pro-
moils.html (04.05.2006).
ductivity of one order of magnitude detected in continuous [12] Snåre M, Kubickova I, Mäki-Arvela P, Eränen K, Murzin D Yu. Ind
mode could be attributed to mass transfer limitations in the Eng Chem Res 2006;45:5708–15.
tubular fixed bed reactor. [13] Kubičková I, Snåre M, Mäki-Arvela P, Eränen K, Murzin D Yu.
Catal Today 2005;106:197–200.
[14] Snåre M, Kubickova I, Mäki-Arvela P, Eränen K, Murzin DYu.
4. Conclusions Continuous deoxygenation of ethyl stearate – a model reaction for
production of diesel fuel hydrocarbons. Catal Org React 2006:415–25.
Deoxygenation of unsaturated renewables, such as oleic [15] Yu Murzin D, Kubickova I, Snåre M, Mäki-Arvela P. European
Patent Application, European patent 05075068.6, 2005.
acid, methyl oleate and linoleic acid, with high selectivity to [16] Indian Space Research Organisation, GB Pat. 1 524 781, 1978.
saturated diesel fuel range hydrocarbons, was successfully [17] Chow PW, US Pat. 5 233 109, 1993.
accomplished over Pd/C, proceeding in the studied param- [18] Twaiq FA, Zabidi NAM, Bhatia S. Ind Eng Chem Res
eter range via initial hydrogenation of double bonds and 1999;38:3230–7.
[19] Idem RO, Katikanieni SP, Bakhshi NN. Fuel Process Technol
subsequent deoxygenation of corresponding saturated
1997;51:101–25.
feeds. Parallel to hydrogenation formation of diunsatu- [20] Craig WK, Soveran DW. US Pat. 4 992 605, 1991.
rated acids occurred. Under a hydrogen rich atmosphere [21] Lima D, Soares VC, Ribeiro EB, Caravalho DA, Cardoso ECV,
the hydrogenation was enhanced and the deoxygenation Rassi FC, et al. J Anal Appl Pyrol 2004;71:987–96.
M. Snåre et al. / Fuel 87 (2008) 933–945 945

[22] Bertram SH. Chem Weekblad 1936:457–9. [31] Simonov PA, Romanenko AV, Prosvirin IP, Moroz EM, Boronin AI,
[23] Foglia TA, Barr PA. JAOCS 1976;53:737–41. Chuvilin AL, et al. Carbon 1997;35:73.
[24] Maier WF, Roth W, Thies I, Rague Schleyer Pv. Chem Ber [32] Gomez-Sainero LM, Seoane XL, Fierro JL, Arcoya A. J Catal
1982;115:808–12. 2002;209:279.
[25] Stern R, Hillion G, US Pat. 4 554 397, 1985. [33] Klingstedt F, Karhu H, Kalantar Neyestanaki A, Lindfors L-E,
[26] Halttunen ME, Niemelä M, Krause OI, Vaara T, Vuori AI. Appl Salmi T, Väyrynen J. J Catal 2002;206:248–62.
Catal A: Gen 2001;205:37. [34] Ramiréz E, Recasens F, Fernández M, Larroyoz M. AIChE
[27] Ramos A, Alves P, Aranda D, Schmal M. Appl Catal A: Gen 2004;50:7.
2004;277:71–81. [35] Bernas A, Laukkanen P, Kumar N, Mäki-Arvela P, Väyrynen J,
[28] Calvo L, Gilarranz MA, Casas JA, Mohedano AF, Rodriguez JJ. Ind Laine E, et al. J Catal 2002;210:354–66.
Eng Chem Res 2005;44:6661. [36] Smeds S, Salmi T, Murzin D Yu. Appl Catal A: Gen 1999;185:
[29] Auer E, Freund A, Pietsch J, Tacke T. Appl Catal A 131–6.
1998;173:259–71. [37] Harwood H. Chem Rev 1962;62:99–154.
[30] Tokarev A, Kustov L, Ivaska A, Murzin D Yu. Appl Catal A [38] Mäki-Arvela P, Kubičková I, Snåre M, Eränen K, Murzin DYu.
2006;309:52–61. Energy Fuels 2007;21:30–41.

Das könnte Ihnen auch gefallen