Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Molecular Catalysis xxx (xxxx) xxxx

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

Waste cooking oil transformation into third generation green diesel


catalyzed by nickel – Alumina catalysts
Ioannis Nikolopoulosa, George Kogkosa, Eleana Kordoulia, Kyriakos Bourikasb,
Christos Kordulisa,b,c,*, Alexis Lycourghiotisa
a
Department of Chemistry, University of Patras, GR-26504, Patras, Greece
b
Hellenic Open University, Tsamadou 13-15, GR-26222, Patras, Greece
c
Foundation for Research and Technology, Institute of Chemical Engineering Science (FORTH/ICE-HT), Stadiou Str., Platani, P.O. Box 1414, GR-26500, Patras, Greece

A R T I C LE I N FO A B S T R A C T

Keywords: Three nickel-alumina catalysts were prepared with nickel content 60, 80 and 90 wt % (symbols: 60NiAl, 80NiAl,
Green diesel 90NiAl) following a co-precipitation methodology. The catalysts were characterized and evaluated in the se-
Renewable diesel lective deoxygenation of waste cooking oil for the production of third generation green diesel under solvent free
Waste cooking oils conditions.
Biofuels
Metallic Ni, NiO, and NiAl2O4 were detected in the 60NiAl catalyst. NiAl2O4 is hardly detected in the 80NiAl
Nickel catalysts
catalyst but not at all in the 90NiAl catalyst. The extent of reduction of NiO to metallic Ni upon activation is
considerably higher in the 80NiAl and 90NiAl samples with respect to the 60NiAl sample resulting in much
higher active site population. The increase in the nickel content causes a change in the porous structure and a
decrease in the specific surface area. The catalytic performance increases considerably from the 60NiAl catalyst
to the 80NiAl one due to the increase in the active site population. The higher catalytic performance of the
80NiAl catalyst with respect to that of the 90NiAl catalyst, exhibiting almost equal active site population, is due
to convenient porous structure of 80NiAl catalyst, which hinder the insertion of deleterious bulky compounds
present in the waste cooking oil, as well as to its strong acid sites.

1. Introduction waste cooking oils (WCO), for producing third generation green diesel is
extremely useful. The exploitation of WCO to produce fuels not only
The catalytic transformation of natural triglycerides into C15-C18 n- solves waste problem but also recycles energy [9]. Nowadays, WCO is
alkanes (green diesel) over nickel non-sulphided catalysts supported on readily available in large quantity and presents a cost effective resource
various carriers has gained much interest in the last decade [1–6]. The for biofuel production [10]. Nevertheless, studies dealing with the use
domain has been recently reviewed by our group [7]. This transfor- of WCO for the production of third generation green diesel are rela-
mation takes place by hydrotreatment in the temperature range tively few [7,10–18].
240–360 °C and hydrogen pressure 10–80 bar. The removal of oxygen In the last years we have contributed to this domain by developing a
without fragmentation of the side chains of triglycerides, selective co-precipitation methodology ensuring severe control of the precipita-
deoxygenation (SDO), may proceeds through decarboxylation (deCO2), tion parameters. Co-precipitation leads to Ni catalysts with high metal
decarbonylation-dehydration (deCO–deH2O) and/or dehydration surface area (desired for hydrogen splitting) and less amount of strong
(deH2O). Following a recently proposed terminology for biodiesel [8], acid sites (desired to avoid extended hydrocracking) [19]. Recently,
one can classify green diesel in four generations according to the raw Cao et al. [20] following a similar methodology have prepared high Ni-
triglyceride material used (first generation: edible vegetable oils, loading catalysts (50 mol %) with highly dispersed Ni nanoparticles,
second generation: non-edible vegetable oils, third generation: residual which showed excellent catalytic performance in benzonitrile hydro-
vegetable oils and animal fats, and fourth generation: algal oils). In the genation.
most of studies fresh edible vegetable oils are used as sources of natural Following co-precipitation we have also succeeded to prepare
triglycerides. However, the use of such oils compete the production of nickel-alumina [21] and nickel-zirconia [22] catalysts with high spe-
foods. Therefore, replacing fresh edible oils by residual materials, like cific surface area and small supported nickel nanocrystals even for


Corresponding author at: Department of Chemistry, University of Patras, GR-26504, Patras, Greece.
E-mail address: kordulis@upatras.gr (C. Kordulis).

https://doi.org/10.1016/j.mcat.2019.110697
Received 9 July 2019; Received in revised form 21 October 2019; Accepted 24 October 2019
2468-8231/ © 2019 Elsevier B.V. All rights reserved.

Please cite this article as: Ioannis Nikolopoulos, et al., Molecular Catalysis, https://doi.org/10.1016/j.mcat.2019.110697
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

catalysts with very high nickel content. Thus, catalysts with very high the catalyst surface. Then, a stream of NH3 was fed in the micro-reactor
active surface have been prepared. These catalysts exhibited high per- for 30 min at room temperature and then it was switched to He to re-
formance, even under solvent free conditions and very high vegetable move the physically adsorbed ammonia. The temperature was then
oil volume to catalyst mass ratio (100 mL/1 g). The activity was further increased linearly (10 °C/min) up to 500 °C. The amount of the des-
increased by applying the co-precipitation methodology for preparing orbed ammonia was determined by a thermal conductivity detector
molybdenum [18,23] and zinc [24] promoted nickel-alumina catalysts. (TCD).
In almost all cases sunflower oil (SO) has been used as source of natural CO-chemisorption was performed on the aforementioned labora-
triglycerides. tory-constructed equipment following pulse technique. 100 mg of the
The replacement of the SO by WCO is the main pursuit of the pre- pre-reduced catalyst were placed in the quartz micro-reactor and He
sent work. Previous catalytic tests had indicated that nickel catalysts was fed (40 mL min−1) until a constant base line to be obtained. The CO
with high nickel content and suitable pore size distribution are required uptake was measured at 30 °C by successive injection of 0.5 mL of CO
for obtaining high catalytic performance in the transformation of WCO samples (10% v/v in He) via calibrated loop of a six-port valve into He
into green diesel [22,23]. Therefore, the present study focuses on the carrier until saturated peak was detected by TCD in the reactor outlet.
preparation, characterization and evaluation of three nickel-alumina The catalyst surface was washed with He for half an hour and a second
catalysts with nickel loadings 60, 80 and 90 wt % symbolized by cycle of successive injections was performed. The chemically adsorbed
60NiAl, 80NiAl and 90NiAl, respectively. These catalysts were prepared CO was calculated by subtracting the CO adsorbed during the second
using the aforementioned co-precipitation procedure. Moreover, in cycle from that of the first one.
order to rationalize the catalytic behavior, important physiochemical Precursor samples (after Ar treatment at 400 °C and before reduc-
properties of WCO were investigated vis-à-vis to SO. tion) were characterized by H2-TPR experiments performed in the same
laboratory-constructed equipment. An amount of sample, 0.04 g, was
2. Experimental placed in the quartz reactor and the reducing gas mixture (H2/Ar: 5/95
v/v) was passed through it for 2 h with a flow rate of 40 mL min−1 at
The catalysts were prepared by co-precipitation at constant pH from room temperature. Then the temperature was increased to 1000 °C with
aqueous solutions containing Al(NO3)3*9H2O, (E. Merck) and Ni a constant rate of 10 °C min−1. Reduction leads to a decrease of the
(NO3)2*6H2O (Merck, kGaA). Ammonium hydroxide solution (NH4OH hydrogen concentration of the gas mixture, which was detected by the
30%, Carlo Erba Reagents) was used for regulating the pH. Proper TCD. The reducing gas mixture was dried in a cold trap (−95 °C) before
amounts of the above salts were dissolved in distilled water. The so- reaching the TCD.
lution was fed by a syringe pump to a 500 mL beaker containing 180 mL The catalysts were evaluated for transformation of WCO to green
of distilled water. The pH was initially adjusted to pH = 8 and main- diesel under solvent free conditions at 310 °C, 40 bar hydrogen pres-
tained constant during preparation using a pH-stat (Metrohom-645). A sure, WCO volume to catalyst mass ratio 100 mL/1 g and various re-
bare alumina sample was also prepared following the same procedure action times up to 9 h. A falling basket batch reactor (300 ml, Autoclave
without adding nickel nitrate. The solids obtained by filtration were Engineers), working in a semi-batch mode with hydrogen flow rate
dried at 110 °C overnight, heated under argon flow (30 mL/min) for 100 mL/min, has been used. The WCO used in this study, was obtained
40 min, required for the increase of temperature from room tempera- by a local company, “COLLECTOIL”. It was filtered three times using
ture to 400 °C, and then were reduced at 400 °C for 2.5 h under hy- cloth filters of different sizes and then it was centrifuged prior to its use.
drogen flow (30 mL/min). The activated catalysts were characterized The analysis of the liquid samples was performed by gas chromato-
using the following techniques. graphy and confirmed by gas chromatography-mass spectrometry.
Nitrogen adsorption-desorption isotherms were recorded in a More details concerning preparation, characterization and evaluation of
Micromeritics apparatus (Tristar 3000 porosimeter). Specific Surface the catalysts have been reported in references [8,10–13]. SO and WCO
Area (SSABET) calculation was based on BET equation using adsorption were characterized using TGA, FT-ATR, GC–MS whereas their im-
data recorded at relative N2 pressure in the range 0.03 < P/P0 < 0.5. portant physicochemical properties (acidity, viscosity, iodine number
Pore size distribution was determined using the BJH method and the N2 and water content) have been determined [25].
desorption curve.
X-ray powder diffraction (XRD) patterns were recorded in the range 3. Results and discussion
of 10° ≤ 2θ ≤ 80o by a Bucker D8 Advance diffractometer equipped
with nickel-filtered Cu Kα (1.5418 Å) radiation source working at 40 kV 3.1. Composition and physicochemical properties of WCO
and 20 mA. The step size and the time per step were respectively fixed
at 0.02o and 0.5 s. The mean crystallite size was estimated using It should be taken into account that the major portion of the WCO,
Scherrer’s equation. drawn from the Greek restaurants, is manly produced by cooking SO.
Scanning Electron Microscopy-Energy Dispersive Spectrometry The determination of characteristic parameters has shown that WCO
(SEM-EDS) was used for the determination of catalysts percentage exhibits higher static and dynamic viscosity (9.89 cSt, 0.088 P) than SO
composition in nickel. Scanning Electron Microscope (SEMJEOL (7.39 cSt, 0.066 P) as well as lower iodine number than SO (WCO: 67,
JSM6300) equipped with an Energy Dispersive Spectrometry accessory SO: 120–134). Both can be attributed to the acidic polymerization of a
has been used working with 20 kV accelerating voltage and 10 nA beam small portion of triglycerides upon cooking starting from the double
current. Microanalysis was performed on gold coated samples. The olefinic C]C bonds of their side chains. The increase in the viscosity
sample powders were mounted directly on the sample holder. upon cooking can be also partly attributed to the presence of potato
Transmittance Electron Microscopy (TEM) was used for recording starch or meat proteins in the WCO inserted in the SO upon cooking.
TEM images of the catalysts in a JEOL JEM-2100 system operated at The presence of polymeric compounds, starch and proteins in the WCO
200 kV (resolution: point 0.23 nm, lattice 0.14 nm) equipped with an is also investigated by comparing the TGA curves of SO and WCO il-
Erlangshen CCD Camera (Gatan, Model 782 ES500W). The specimens lustrated in Fig. 1. Inspection of this figure shows that the TGA curves
were prepared by dispersion in water and spread onto a carbon-coated are almost identical in the range 25–450 °C due to the triglyceride
copper grid (200 meshes). structure of both raw materials. This is also confirmed by the almost
NH3-TPD experiments were performed in laboratory-constructed identical FT-ATR spectra recorded for both of them (Fig. 2). However,
equipment. In a typical experiment, 100 mg of the pre-reduced catalyst returning to the TGA curves we are observing that the weight of SO
were placed in a quartz micro-reactor and helium was fed (flow rate takes a zero value at about 450 °C whereas a small amount of WCO
30 mL min−1) for 30 min in order to remove any adsorbed species from remains in the range 400–450 °C.

2
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

Fig. 1. TGA curves of SO (red curve) and WCO (black curve). (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article).

This indicates the presence of more difficultly sublimated com-


pounds, much probable polymers, starch and proteins. However, the
weight of WCO takes a zero value after heating the sample at 450 °C for
0.5 h. The determination of characteristic parameters has shown that
WCO exhibits higher acidity expressed in oleic acid (WCO: 0.83%, SO:
0.60%). This can be attributed to the partial dissociation of the natural
triglycerides to fatty acids upon cooking. In fact, small amounts of
stearic acid are identified in the GC chromatograms of the WCO (not
shown here) together with small amounts of mono and di-glycerides
produced by progressive dissociation of a small portion of triglycerides.
The water content of the two feedstocks was also determined and was
found to be different (WCO: 0.14 wt %, SO: 0.04 wt %). In view of the
above one can conclude that besides the differences in acidity and
water content, an important difference between SO and WCO is the
presence of bulky compounds in the latter.

3.2. Solid phases

The compositions of the catalysts determined by EDS are very close


to the nominal ones. The crystal phases of the catalysts studied were Fig. 3. XRD patterns of the activated (reduced) catalysts studied. The symbols
investigated by XRD and electron diffraction whereas the distribution of (#), (*) and (+) denote, respectively NiO, Ni and NiAl2O4.
the solid phases was investigated by TEM. The XRD patterns of the
catalysts studied are illustrated in Fig. 3.
Inspection of the XRD patterns shows the presence of metallic nickel

Fig. 2. FT – ATR spectra of SO (red curve) and


WCO (black curve). The assignments of dif-
ferent of stretching (s), bending (b) and
rocking (r) vibrations of the various groups of
triglycerides have been taken from the litera-
ture [26]. (For interpretation of the references
to colour in this figure legend, the reader is
referred to the web version of this article).

3
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

compatible to the mesoporous texture of the catalysts inferred by in-


specting Fig. 6.
The change of nickel content is accompanied by considerable
changes in the pore volume distributions. The 60NiAl catalyst exhibits a
double peak centered at 3 and 5 nm and a broad one in the range
12–100 nm. The catalyst 80NiAl exhibits its main peak at about 7 nm
and a smaller size peak in the range 18–100 nm. The no formation of
pores with 3 and 5 nm diameter in the 80NiAl catalyst is reflected in the
decrease of its SSABET whereas the smaller size of the peak in the range
above 10 nm is in accordance with its decreased SPV value (Table1).
Finally, the 90NiAl catalyst exhibits a broad pore volume distribution
with a main maximum in the range 10–20 nm which is extended to
higher than 100 nm justifying both the further decrease in SSABET and
the impressive increase in the value of MPD (Table 1).
It is well known that CO-chemisorption cannot be considered as an
absolute quantification measure of metallic sites. This is mainly caused
by the different ways CO-chemisorption occurs on the various sites. For
Ni catalysts, the CO-chemisorption is affected by the metal content, the
“Ni-support” interaction, the metal particles dispersion, and the pre-
sence of impurities. On the other hand, several possible modes of ad-
sorption such as linear, bridged and carbonyl type adsorption have been
observed [27] and metallic Ni centers can interact with two, even three
CO molecules [28]. However, the low temperature CO-chemisorption
Fig. 4. Electron diffraction pattern of the catalyst studied.
contacted in our case is expected to lead mainly in linear adsorption on
metallic Ni sites. Although, the quantification of chemisorbed CO with
[peaks at 2θ: 44.5, 51.8 and 76.3°, JCPDS 04-0850] and nickel oxide 1:1 stoichiometry per surface Ni atom is widely used in the recent lit-
[peaks at 2θ: 43.4, 37.3 and 62.9°, JCPDS 01–1239] on the catalysts erature for Ni surface area and dispersion determination [29], it is more
studied. Small amounts of nickel aluminate are detected in the sample realistic to consider the amount of chemisorbed CO (μmol/g catalyst) as
60NiAl [shoulders at 2θ: 45.0 and 65.5, JCPDS 01–1239]. Nickel alu- a reliable parameter for comparison [28,30]. In this sense, the CO
minate is very hardly detected, in the 80NiAl catalyst but not in the chemisorption results (Table 1) clearly show that the surface of metallic
90NiAl catalyst. The absence of peaks corresponding to alumina crys- nickel increases with the nickel content. The increase is abrupt when
tals strongly suggests that the alumina formed is amorphous. going from the 60NiAl catalyst to the 80NiAl one. In fact, whereas the
Comparison of the height of the peak at 2θ: 44.5° with that at 2θ: 43.4° weight ratio in total nickel is equal to 80/60 = 1.33, the corresponding
shows that the minimum ratio Ni/NiO is obtained in the 60NiAl sample ratio in the amount of the adsorbed CO is 209/15 = 13.93. This can be
and the maximum one in the sample 80NiAl. This indicates that in the explained considering that a great portion nickel in the 60NiAl catalyst
latter sample the highest extend of NiO reduction towards the cataly- is present as nickel oxide and nickel aluminate relatively not active in
tically active metallic nickel is achieved upon activation. The mean the CO adsorption. This is not the case for the 80NiAl catalyst where
crystal size of the metallic nickel calculated on the base of diffraction according to the XRD and electron diffraction results it is obtained the
peak at 2θ: 51.8° was found to be equal to 9.6, 11.2 and 10.9 nm for the highest extent of NiO reduction to metallic nickel and where nickel
60NiAl, 80NiAl and 90NiAl catalysts, respectively. aluminate is not present. Moreover, combining the XRD, electron dif-
The XRD results are confirmed by the electron diffraction results. fraction and CO adsorption results we incline to argue that nickel oxide
Thus, the electron diffraction pattern of the sample 60NiAl showed the and nickel aluminate relatively small nanocrystals presumably pre-
presence of Ni, NiO and NiAl2O4 nanocrystals whereas the electron dominate in regions of the 60NiAl catalyst pictured in the up- left image
diffraction patterns of the samples 80NiAl and 90NiAl showed the of Fig. 5. The increase in the nickel metallic surface follows the increase
presence of only Ni and NiO nanocrystals. A typical electron diffraction of the nickel content when going from the 80NiAl catalyst to the 90NiAl
pattern for the most active catalyst, 80NiAl, is illustrated in Fig. 4. one. In fact, the weight ratio in total nickel is equal to 90/80 = 1.12
The TEM images recorded for the catalysts studied depict the dis- and the corresponding ratio in the amount of the adsorbed CO is 227/
tribution of their solid phases at nanoscale. Typical examples are illu- 209 = 1.08. The slightly lower value in the latter ratio could be at-
strated in Fig. 5. The TEM images of the catalyst 60NiAl display two tributed to the smaller extend of NiO reduction in the 90NiAl catalyst
regions; one with relatively small nickel phase nanoparticles separated with respect to the 80NiAl catalyst deduced by XRD.
by areas comprised from very small alumina nanoparticles (Fig. 5a) and
another one with relatively big nickel phase nanoparticles separated by 3.4. Catalysts reducibility and acidity
mixed areas comprised from smaller and smaller nickel phase nano-
particles and very small alumina nanoparticles (Fig. 5b). The second The XRD, electron diffraction, TEM and CO adsorption findings are
picture predominates in the 80NiAl (Fig. 5c) and 90NiAl catalysts corroborated by the H2-TPR results illustrated in Fig. 7.
(Fig. 5d). The separating areas are wider in the 80NiAl catalyst. Before discussing the H2–TPR curves it is useful to mention that the
reduction time upon activation at 400 °C is much higher (2.5 h) than the
3.3. Textural properties and metallic nickel surface very small reduction time upon the transient H2-TPR procedure. This
allows us to assume that the hydrogen consumption up to roughly
Table 1 compiles important parameters of the catalysts studied 450 °C corresponds to the reduction of NiO nanocrystals to metallic
concerning their texture probed by N2 physisorption and metallic area nickel detected by XRD and electron diffraction in the reduced cata-
investigated by pulse CO adsorption experiments. lysts. The hydrogen consumption in the range 450–750 °C corresponds
Inspection of Table 1 shows that although the specific surface area to the reduction of NiO nanocrystals non-reducible upon activation and
decreases with nickel content, thanks to the preparation methodology thus they are detected by XRD and electron diffraction as nickel oxide.
applied it remains adequately high, even in the catalyst with nickel Finally, the hydrogen consumption in the range 750–800 °C corre-
content as high as 90 wt %. The relatively high SSABET values are sponds to the reduction of nickel aluminate [18,31–35]. Let’s

4
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

Fig. 5. Examples of TEM images recorded for the catalysts studied (a: 60NiAl, b: 60NiAl, c: 80NiAl, d: 90NiAl).

Table 1
Textural parameters of the catalysts studied and CO-chemisorption results.
Catalyst SSABET SPV* (cm3 g−1) MPD** (nm) COads*** (μmol/
(m2 g−1) gcat)

60NiAl 273 0.59 6.6 15


80NiAl 116 0.36 8.7 209
90NiAl 89 0.49 16.1 227

* Specific pore volume.


** Mean pore diameter.
*** CO-chemisorption.

Fig. 7. H2-TPR curves of the catalysts studied.

alumina [33,34] and resulting to better dispersed nickel nanocrystals is


relatively low compared to the other two samples. In contrast, the hy-
drogen consumption above 450 °C corresponding to NiO nanocrystals
non-reducible upon activation and nickel aluminate [35] is con-
siderably higher compared to those of the other two catalysts. The
above results are in full agreement to the XRD and electron diffraction
results and justify the relatively low metallic nickel surface inferred by
the CO adsorption. The increase in hydrogen consumption in the range
Fig. 6. Pore volume distribution of the catalysts studied. 350–450 °C with the nickel content is in line with the increase in the
amount of metallic nickel also inferred by the increase in the magnitude
concentrate our attention to the H2-TPR curve of the 60NiAl catalyst. of the XRD peak at 51.8° and the increase in the CO adsorption. The
We are observing a peak at about 350 °C corresponding to quite large decrease in hydrogen consumption in the range 500–750 °C with the
NiO nanocrystals weakly interacted with alumina [31,32] and thus nickel content reflects the decrease in the relative amount of NiO with
easily reducible to quite large nickel nanocrystals (low dispersion). respect to metallic nickel. Finally, nickel aluminate, reducible in the
Moreover, the hydrogen consumption in the range 350–450 °C corre- range 750–800 °C is indicated in the 60NiAl catalyst, hardly in the
sponding to smaller NiO nanocrystals moderately interacted with 80NiAl catalyst but not in the 90NiAl one, in agreement to the XRD

5
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

Fig. 9. Kinetic curves over the catalyst 80NiAl (310 °C, 40 bar hydrogen pres-
sure, reactant volume to catalyst mass ratio=100 mL/1 g).
Fig. 8. NH3-TPD curves of the catalysts studied and the bare alumina (0NiAl).
CO, propane and methane have been identified in the gas phase. The
results. above products are expected from the reaction network described just
The acidity of the catalysts was studied by NH3-TPD experiments. before. Typical kinetic curves for the catalyst 80NiAl are presented in
Fig. 8 illustrates the NH3-TPD curves. Fig. 9. It is seen that the conversion increases considerably even for
Fig. 8 illustrates the NH3-TPD curves of the catalysts studied and the reaction time of 1 h indicating that the first steps of the SDO network
bare alumina. All the samples exhibit high populations of weak acid (hydrogenation of olefinic bonds of the triglycerides, decomposition of
sites (desorption temperature < 300 °C), small populations of sites with the O‐C bonds in the glycerol side) are very rapid in agreement to the
moderate acidity (300 °C < desorption temperature < 450 °C) and few literature [7,17,18]. Moreover, the kinetic curves show maxima con-
strong acid sites, only in the sample 80NiAl (desorption tempera- cerning the acid and esters evolution indicating that these molecules
ture > 450 °C) [36]. We are observing that the bare alumina (0NiAl) are actually intermediate ones in agreement to the reaction network.
exhibits the highest population of weak acid sites. This population de- On the other hand, the evolution of the hydrocarbons, as final products,
creases with the nickel content following approximately the decrease of follows a rather linear kinetic curve. In almost all cases the amounts of
the alumina content and the parallel decrease in the specific surface n-alkanes with even number carbon atoms (principally n-C17) is much
area (Table 1). This is an indication that the weak acid sites are mainly higher than the amounts of the hydrocarbons with odd number of
situated on the alumina surface. The observation that the moderately carbon atoms, indicating that the process proceeds mainly through the
strong and very strong acid sites appear mainly on the surface of the Ni decarbonylation of the intermediate aldehydes which is the case of
catalysts indicates that these are situated on the supported nickel nickel metallic catalysts [7]. In conclusion, the replacement of the SO
phases. They could be related to the empty d orbitals of the metallic by WCO does not seem to affect generally the triglyceride SDO network.
nickel and the Ni2+ ions in the supported nickel oxide and nickel alu- The comparative evaluation of the catalysts studied is presented for
minate which can interact with the lone pair of electrons of ammonia the maximum reaction time (Fig. 10). As it is seen complete conversion
[37]. Taking into account the XRD results one can attribute the creation of WCO (100%) is achieved under reaction conditions used over all the
of moderately strong acid sites to all nickel phases while that of the very catalysts studied. However, the liquid product distribution obtained
few strong acid sites mainly to the metallic Ni. The latter phase was depends on the catalyst used.
found to be predominant in the 80NiAl catalyst. We shall return to this It is seen that the total amount of the hydrocarbons produced in the
point after the evaluation of the catalysts performance. diesel range, the most important evaluation parameter from the cata-
lytic point of view, follows a volcano-like trend maximized over the
80NiAl catalyst. The value obtained over this catalyst, equal to
3.5. Evaluation of the catalytic performance

Before dealing with the catalytic behavior of the catalysts studied it


is useful reminding the SDO of triglycerides network over nickel non-
sulphided catalysts [7]. It is proceeding through the hydrogenation of
olefinic bonds of the triglycerides followed by the decomposition of the
OCe bonds in the glycerol side and the formation of fatty acids and
propane. The fatty acids are reduced to aldehydes and alcohols. The
aldehydes are decarbonylated resulting to n‐C17 and n‐C15 alkanes and
CO. The alcohols are dehydrated and the produced olefins are hydro-
genated resulting to n‐C18 and n‐C16 alkanes. In parallel, the alcohols
react with the fatty acids producing long chain esters with both hy-
drocarbons parts corresponding to fatty acids of triglycerides. The es-
ters undergo SDO resulting to hydrocarbons. Thus fatty acids and esters
are intermediate products. The CO produced may react with hydrogen
producing methane.
The most important products identified in the liquid phase were
non-reacted triglycerides, long chain esters with both hydrocarbon Fig. 10. Evaluation parameters obtained for the SDO of WCO over the catalysts
parts corresponding to fatty acids of triglycerides, palmitic and stearic studied at reaction time equal to 9 h (310 °C, 40 bar hydrogen pressure, reactant
acid (mainly stearic acid) and normal alkanes, n-C15–n-C18, whereas volume to catalyst mass ratio =100 mL/1 g).

6
I. Nikolopoulos, et al. Molecular Catalysis xxx (xxxx) xxxx

36 ± 2% is quite high taking into account the very high ratio of WCO Declaration of Competing Interest
volume to catalyst mass adopted in the present work. Moreover, this is
considerably higher than that obtained over the 60NiAl catalyst The authors declare that they have no known competing financial
(13 ± 2%) and quite higher than that obtained over the 90NiAl cata- interests or personal relationships that could have appeared to influ-
lyst (29 ± 2%). The increase of the catalytic performance from the ence the work reported in this paper.
60NiAl catalyst to the 80NiAl can be mainly attributed to the increase
in the active site population inferred by the CO adsorption. It is well References
known that the reduction of intermediate free fatty acids to the corre-
sponding aldehydes is the rate determining step of the triglycerides SDO [1] I. Kubickova, D. Kubicka, Waste Biomass Valor. 1 (2010) 293–308.
network [7]. This step is catalyzed by active sites situated on the me- [2] S. Khan, A.N.K. Lup, K.M. Qureshi, F. Abnisa, W.M.A.W. Daud, M.F.A. Patah, J.
Anal. Appl. Pyrolysis 140 (2019) 1–24.
tallic nickel [7]. Thus, the highest green diesel yield achieved over the [3] J.K. Satyarthi, T. Chiranjeevi, D.T. Gokak, P.S. Viswanathan, Catal. Sci. Technol. 3
80NiAl catalyst is easily explained. In contrast, the change in the active (2013) 70–80.
site population cannot explain the decrease of the catalytic performance [4] J. Calero, D. Luna, E.D. Sancho, C. Luna, F.M. Bautista, A.A. Romero, A. Posadillo,
J. Berbel, C. Verdugo-Escamilla, Renew. Sust. Energ. Rev. 42 (2015) 1437–1452.
going from the 80NiAl catalyst to the 90NiAl one. Indeed, considering [5] E. Santillan-Jimenez, M. Crocker, J. Chem. Technol. Biotechnol. 87 (2012)
the chemisorbed CO (Table 1) as a relative measure of metallic nickel 1041–1050.
site population one can conclude that active site populations in these [6] M. Jin, M. Choi, Mol. Catal. 474 (2019) 110419, , https://doi.org/10.1016/j.mcat.
2019.110419.
two samples are almost equal. Therefore, two additional observations [7] C. Kordulis, K. Bourikas, M. Gousi, E. Kordouli, A. Lycourghiotis, Appl. Catal. B 181
should be taken into account in order to explain the achievement of the (2016) 156–196.
maximum catalytic performance over the 80NiAl catalyst. The first [8] D. Singh, D. Sharma, S.L. Soni, S. Sharma, D. Kumari, Fuel 253 (2019) 60–71.
[9] H. Wang, G. Li, K. Rogers, H. Lin, Y. Zheng, S. Ng, Mol. Catal. 443 (2017) 228–240.
concerns the presence of the very strong acid sites present exclusively in
[10] Z. Li, Z. Huang, S. Ding, F. Li, Z. Wang, H. Lin, C. Chen, Energy 157 (2018)
this catalyst (Fig. 8). These presumably involved in the decomposition 270–277.
of the bulky compounds being in WCO, mentioned in 3.1 subsection, [11] S. Bezergianni, S. Voutetakis, A. Kalogianni, Ind. Eng. Chem. Res. 48 (2009)
which would eventually adsorbed and thus deactivate active sites. The 8402–8406.
[12] S. Bezergianni, A. Dimitriadis, L.P. Chrysikou, Fuel 18 (2014) 300–307.
second observation concerns the catalysts texture. As already men- [13] H. Zhang, H. Lin, W. Wang, Y. Zheng, P. Hu, Appl. Catal. B 150−151 (2014)
tioned, pores of about 7 nm are the predominant ones present in the 238–248.
80NiAl catalyst. In contrast, pores with sizes extended from 15 up to [14] S. Liu, T. Simonetti, W. Zheng, B. Saha, ChemSusChem 11 (2018) 1446–1454.
[15] D. Singh, S.S. Sandhu, A.K. Sarma, Energ. Source Part A 40 (2018) 968–976.
higher than 100 nm are present in the 90NiAl catalyst (Fig. 6). One may [16] G. Abdulkareem-Alsultan, N. Asikin-Mijan, H.V. Lee, A.S. Albazzaz, Y.H. Taufiq-
assume that the WCO bulky compounds, probably acting as poisons, Yap, Energ. Convers. Manage. 151 (2017) 311–323.
cannot easily inserted in the small pores of the 80NiAl catalyst justi- [17] G. Xu, Y. Zhang, Y. Fu, Q. Guo, ACS Catal. 7 (2017) 1158–1169.
[18] E. Kordouli, L. Sygellou, C. Kordulis, K. Bourikas, A. Lycourghiotis, Appl. Catal. B
fying its very high performance in the SDO of WCO. 209 (2017) 12–22.
Closing this section we have to stress that Ni leaching determination [19] S. Janampelli, S. Darbha, Mol. Catal. 451 (2018) 125–134.
and catalysts reuse experiments are crucial for the application of such [20] Y. Cao, H. Zhang, J. Dong, Y. Ma, H. Sun, L. Niu, X. Lan, L. Cao, G. Bai, Mol. Catal
475 (2019) 110452, https://doi.org/10.1016/j.mcat.2019.110452.
catalysts in real SDO processes of WCO. SEM – EDS experiments per- [21] M. Gousi, C. Andriopoulou, K. Bourikas, S. Ladas, M. Sotiriou, C. Kordulis,
formed over fresh and used catalysts proved that the leaching of Ni A. Lycourghiotis, Appl. Catal. A Gen. 536 (2017) 45–56.
phase is not important. This is in agreement with our previous studies [22] G. Zafeiropoulos, N. Nikolopoulos, E. Kordouli, L. Sygellou, K. Bourikas,
C. Kordulis, A. Lycourghiotis, Catalysts 9 (2019) #210.
on relevant catalysts [18]. It seems that the co-precipitation metho-
[23] E. Kordouli, B. Pawelec, K. Bourikas, C. Kordulis, J.L.G. Fierro, A. Lycourghiotis,
dology used for catalysts preparation results to relatively strong metal- Appl. Catal. B 229 (2018) 139–154.
support interactions [38] preventing Ni leaching, ensuring also good [24] M. Gousi, E. Kordouli, K. Bourikas, E. Simianakis, S. Ladas, G.D. Panagiotou,
catalyst stability, as no appreciable aggregation is usually observed in C. Kordulis, A. Lycourghiotis, Catal. Today (2019), https://doi.org/10.1016/j.
cattod.2019.02.034.
the spent catalysts [20]. [25] I. Nikolopoulos, MSc Thesis, University of Patras, Chemistry Department, 2017, pp.
48–53.
4. Conclusions [26] E. Wembabazi, P.J. Mugisha, A. Ratibu, D. Wendiro, J. Kyambadde, P.C. Vuzi, J.
Spectroscopy 2015 (2015) 714396, , https://doi.org/10.1155/2015/714396.
[27] M.V. Alekseeva (Bykova), D.S. Otyuskaya, M.A. Rekhtina, O.A. Bulavchenko,
Nickel-alumina nanostructured catalysts with nickel content in the O. Stonkus, V.V. Kaichev, S.G. Zavarukhin, J.W. Thybaut, V. Alexiadis,
range 60–90 wt % (60NiAl, 80NiAl and 90NiAl) are proved to be active R.H. Venderbosch, V. Yakovlev, Appl. Catal. A Gen. 573 (2019) 1–12.
[28] A. Guerrero-Torres, C.P. Jiménez-Gómez, J.A. Cecilia, C. García-Sancho, F. Franco,
in the transformation of WCO into green diesel under very high WCO J.J. Quirante-Sánchez, P. Maireles-Torres, Top. Catal. 62 (2019) 535–550.
volume/catalyst mass ratio and solvent free conditions. The catalytic [29] A. Cross, A. Yeghishyan, K. Manukyan, J.T. Miller, D. Rensel, V. Danghyan,
performance follows a volcano-like trend maximized over the 80NiAl A.S. Mukasyan, E.E. Wolf, Appl. Catal. A Gen. 572 (2019) 124–133.
[30] M. Ameen, M.T. Azizan, A. Ramli, S. Yusup, M.S. Alnarabiji, Ultrason. Sonochem.
catalyst. The considerably high catalytic performance obtained over
51 (2019) 90–102.
this catalyst with respect to that of the 60NiAl one is mainly attributed [31] A. Tribalis, G.D. Panagiotou, K. Bourikas, L. Sygellou, S. Kennou, S. Ladas,
to its much higher metallic nickel surface compared to the latter cata- A. Lycourghiotis, C. Kordulis, Catalysts 6 (2016) #11.
[32] Q. Liu, J. Gao, F. Gu, X. Lu, Y. Liu, H. Li, Z. Zhong, B. Liu, G. Xu, F. Su, J. Catal. 326
lyst. On the other hand, the higher catalytic performance of the 80NiAl
(2015) 127–138.
catalyst with respect to the 90NiAl catalyst, exhibiting somewhat [33] E. Kraleva, M.-M. Pohl, A. Jurgensen, H. Ehrich, Appl. Catal. B 179 (2015) 509–520.
higher metallic nickel surface, is attributed not only to its convenient [34] S. Damyanova, B. Pawelec, K. Arishtirova, J.L.G. Fierro, Int. J. Hydrogen Energy 37
porous structure exhibiting pores with mean diameter about 7 nm, (2012) 15966–15975.
[35] M. Zaungouei, A.Z. Moghaddam, M. Arasteh, Chem. Eng. Res. Bull. 14 (2010)
which hinder the insertion of deleterious bulky compounds present in 97–102.
the WCO, but also to few but strong acid sites present exclusively in the [36] D. Zhang, W.-Q. Liu, Y.-A. Liu, U.J. Etim, X.-M. Liu, Z.-F. Yan, Chem. Eng. J. 330
80NiAl catalyst, which presumably are involved in the decomposition (2017) 706–717.
[37] S. Lycourghiotis, E. Kordouli, L. Sygellou, K. Bourikas, C. Kordulis, Appl. Catal. B
of these bulky compounds. 259 (2019) 118059, , https://doi.org/10.1016/j.apcatb.2019.118059.
The authors declare the following financial interests/personal re- [38] S. He, Z. An, M. Wei, D.G. Evans, X. Duan, Chem. Commun. (Camb.) 49 (2013)
lationships which may be considered as potential competing interests: 5912–5920.

You might also like