Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

Catalysis Communications 12 (2011) 559–562

Contents lists available at ScienceDirect

Catalysis Communications
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / c a t c o m

Hydrotreating and hydrocracking catalysts for processing of waste soya-oil and


refinery-oil mixtures
Rashmi Tiwari, Bharat S. Rana, Rohit Kumar, Deepak Verma, Rakesh Kumar, Rakesh K. Joshi,
Madhukar O. Garg, Anil K. Sinha ⁎
Indian Institute of Petroleum, Dehradun-248005, India

a r t i c l e i n f o a b s t r a c t

Article history: Mesoporous SiO2–Al2O3 and Al2O3 were used as supports to prepare hydrocracking (sulfided Ni–W/SiO2–
Received 30 October 2010 Al2O3) and hydrotreating (sulfided Ni–Mo/Al2O3) catalysts. These hydroprocessing catalysts were used under
Received in revised form 1 December 2010 typical hydroprocessing conditions to convert waste soya-oil mixtures with refinery-oil into saturated
Accepted 6 December 2010
hydrocarbons. The hydrocracking catalyst was more selective for the kerosene range (140–250 °C)
Available online 13 December 2010
hydrocarbons while the less acidic hydrotreating catalyst was more selective for the diesel range (250–
380 °C) hydrocarbons. The hydrodeoxygenation pathway for oxygen removal from triglycerides seems to be
Keywords:
Hydrotreating
favored over the hydrotreating catalyst, while decarboxylation + decarbonylation pathway is favored over the
Hydrocracking hydrocracking catalyst.
Biofuel © 2010 Elsevier B.V. All rights reserved.
Soya-oil
Hydroprocessing catalysts

1. Introduction pore size catalysts to overcome diffusion limitations. Hydroprocessing


has been reported [3–18] to produce straight chain alkanes from fatty
Hydroprocessing is used in the petroleum refinery to crack larger acid triglycerides of animal fat, tall oil, and other vegetable oils. But, not
molecules and/or to remove S, N and metals from petroleum derived much focus has been given to developing larger pore size catalysts for
feed stocks such as, gas-oil and heavy oil [1]. Sulfided hydrocracking this process.
(such as Ni–W/SiO2–Al2O3) and hydrotreating (such as Ni–Mo/Al2O3, Here we report on the catalysts based on moderately acidic
Co–Mo/Al2O3,) catalysts are typically used. Similar catalysts could also mesoporous silica–alumina and weakly acidic mesoporous alumina
be used for deoxygenation of glyceride and fatty acid molecules in the supports to hydroprocess mixtures of gas-oil and waste soya-oil to
oils from vegetables and animals such as animal fat, tall oil, waste obtain pure hydrocarbon mixtures in diesel and kerosene range.
restaurant oil, jatropha oil, algae oil etc. into pure hydrocarbons. Such a
process would result in fuels with desired viscosity, low oxygen content 2. Experimental
and improved atomization and lubricity [2]. It can also contribute to
meeting increased demand for clean fuel. Such a process is more suitable Mesoporous silica–alumina support was obtained by the reported
than Fatty Acid Methyl Esters (FAME) produced by transesterification of sol–gel process [19]. Mesoporous γ-alumina was obtained from Sasol.
triglycerides. Because neat use of FAME requires some modification in The porous properties of catalysts were examined by N2 adsorption–
engine, and additionally, it gives poor performance in cold weather and desorption isotherms at 77 K and the related data (surface area, SBET; pore
poor emission. Transesterification route also produces large quantities volume, VP; pore diameter, Dp; BEL, Japan, Belsorp-max) were calculated.
of by-product glycerol. Hydroprocessing of renewable oils is compatible Mesopore size distribution was obtained by Barret–Joyner–Halenda
with a petroleum refinery. But there are two issues with using typical (BJH) method from the adsorption branch of the isotherms. All samples
hydroprocessing catalysts for deoxygenation reactions of triglycerides: were degassed before actual measurements. Scanning electron micro-
(1) unlike hydrodesulfurization reaction, high oxygen content of the scope (SEM) images were obtained on FEI, Quanta 200F, scanner,
glyceride feedstock would result in continuous S leaching from the Netherlands. Elemental composition was determined by ICP-AES
catalyst surface and hence its gradual deactivation; (2) unlike petroleum measurement. The amount and strength of the acid sites were measured
derived feedstocks, bulky triglyceride molecules would require larger by ammonia adsorption–desorption technique using a chemical adsorp-
tion instrument, Micrometrics 2900 with a thermal conductivity detector.
About 0.20 g sample was saturated with NH3 at 120 °C, then flushed with
⁎ Corresponding author. Tel.: + 91 1352525842. helium to remove the physically adsorbed NH3, finally the desorption of
E-mail address: asinha@iip.res.in (A.K. Sinha). NH3 was carried out at a heating rate of 10 °C/min in helium flow.

1566-7367/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.catcom.2010.12.008
560 R. Tiwari et al. / Catalysis Communications 12 (2011) 559–562

Waste soya-oil and gas-oil mixtures were processed in a fixed bed glycerol part, which is around 0.6 nm), the pore size of the mesoporous
reactor with sulfided Ni–W/SiO2–Al2O3 and Ni–Mo/Al2O3 catalysts support is 3–5 times larger than the largest triglyceride dimension (along
listed in Table 1. The catalysts diluted with SiC were loaded into a the fatty acid chain) and no restrictions to their access inside the pores are
stainless steel tubular reactor (1.3-cm I.D and 30 cm in length). The expected. The micropore volume is very low for both the supports
catalysts were presulfided using a mixture of dimethyl disulfide and (0.011 ml/g from the mesoporous silica–alumina and 0.004 ml/g for the
gas-oil at atmospheric pressure and 350 °C for 9 h. The reaction mesoporous alumina). So the concentration of micropores (with pore size
condition for catalytic hydrotreating experiment were as follows : b2 nm) which are inaccessible to bulky substrates is very low and it can be
temperature 340–380 °C, pressure 50 bar , LHSV 2, 4 h− 1, and H2 to feed concluded that to a greater extent, surface as well as pores are accessible
ratio of 1500 ml H2 gas / ml liquid feed. Gas-oil was supplied by Mathura to the reactant triglycerides. Ni–W/SiO2–Al2O3 catalyst has a typical
Refinery, India. The carbon yield reported in this paper is based on the composition of a hydrocracking catalyst [1] and has considerable acidity.
simulated distillation results. Waste restaurant soya-oil was used for The relative ratio of weak, moderate and strong acidities (both Lewis as
this study after filtration to remove solid residues and was characterized well as Brönsted) is roughly 6.4:5.4:1. 0.77 ml/g of ammonia desorbed up
by different techniques. to 100 °C is attributed to weak acid sites (mostly Lewis), 0.0.65 ml/g of
The liquid products were withdrawn after stabilization of reaction ammonia desorbed from 100 to 300 °C is attributed to moderate acid sites
conditions (6 h) in two-hour intervals (at each temperature) and (mostly Brönsted) and 0.12 ml/g of ammonia desorbed in the range 300–
analyzed by off-line gas chromatography after separation of the water 400 °C is attributed to strong acid sites (Lewis as well as Brönsted) [20].
phase. We analyzed the liquid products thrice during stabilization Second catalyst Ni–Mo/Al2O3 has a typical composition of hydrotreating
period by GC to monitor constant activity. The reaction gases were catalyst with mesoporous alumina as the support and weak (Lewis)
analyzed using a Varian 3800-GC equipped with a flame ionization acidity. These mesoporous supports would be advantageous for the
detector (FID) and a thermal conductivity detector (TCD). Liquid transformations of bulky triglyceride feedstock with better diffusion of
samples were analyzed with a Varian 3800-GC for alkanes. Simulated reactants and products. Deactivation by pore blockage by coking and by
distillation of the hydrotreated products were carried out using a waxy product intermediates formed during prolonged use of the catalysts
Varian 3800-GC according to the ASTM-2887-D86 procedure. In would also be lower.
analyzing the stimulate distillation results it was assumed that the Fig. 1 shows the nitrogen sorption isotherms and BJH pore size
area of the each distillation fraction were proportional to the amount distribution obtained from adsorption branch of the isotherms.
of carbon in that fraction. This assumption is valid for experiments in a
this paper which was at complete triglyceride conversion. 200
0.7
The concentration of sulfur in the feed and liquid products was 180 0.6
determined by XRF analysis. The hydrodesulphurization conversion (HDS
dVp/dlog(rp)

0.5
conversion) was calculated by subtraction the sulfur in the feed minus 160
0.4
Volume adsorbed (cm3/g

the sulfur in the product and divided by the sulfur in the feed. Product 140 0.3
density (@15 °C) was determined according to ASTM D4052 method. 0.2
Cetane number was determined according to ASTM D613 method. 120
0.1

100 0.0
0 20 40 60 80 100
rp/nm
3. Results and discussion 80

60
The physicochemical properties of catalysts and mesoporous supports
are presented in Table 1. Mesoporous supports had considerably high 40
surface area and large mesopores, 8.6 nm for SiO2–Al2O3 and 12.9 nm for 20
Al2O3, which would readily allow the diffusion of bulky triglyceride
molecules. Comparison of the mean pore diameter of mesoporous 0
supports (8–13 nm) with the dimensions of the triglycerides could give 0.0 0.2 0.4 0.6 0.8 1.0
an idea regarding diffusion related issues and the access of the voluminous p/p0
triglyceride molecules to the active sites within the pores of the support.
The triglyceride dimension can be estimated from the dimension of for
instance methyl oleate (around 2.5 nm length) and of glycerol (around
b 500
1.2
0.6 nm). Even by assuming that the relevant dimension of triglyceride
1.0
molecule is the dimension along the fatty acid chain (though triglyceride
dVp/dlog(rp)

can also go through the pores along the smallest dimension, that of the 400 0.8
Volume adsorbed (cm3/g)

0.6

Table 1 0.4
300
Physicochemical properties of mesoporous supports and catalysts.
0.2

SiO2–Al2O3 Al2O3 Ni–W/SiO2–Al2O3 Ni–Mo/Al2O3 0.0


0 20 40 60 80 100
Surface Area, m2/g 236 243 157.8 209.3 200
rp/nm
Total pore volume 0.39 0.9 0.26 0.61
ml/g
Micropore volume – – 0.011 0.004 100
Mean pore 8.6 12.9 7.8 12.2
diameter, nm
Surface acidity, – – 0.77 (100 °C) 0.18
mmol/g NH3 0.44(200 °C) (100–400 °C) 0
0.21(300 °C) 0.0 0.2 0.4 0.6 0.8 1.0
0.12(400 °C)
p/p0
Chemical 49.38(SiO2), 96.7 26.7(WO3), 3.3 (NiO),
composition 47.09(Al2O3) (Al2O3) 9.8(NiO) 14.7(MoO3)
Fig. 1. Nitrogen sorption isotherm and pore size distribution of mesoporous (a) Ni–Mo/
wt.%
Al2O3 and (b) Ni–W/SiO2–Al2O3.
R. Tiwari et al. / Catalysis Communications 12 (2011) 559–562 561

Fig. 2. Scanning Electron Micrographs (SEM) of mesoporous silica–alumina.

Nitrogen sorption studies showed isotherms typical for mesoporous Using sulfided Ni–W catalyst on acidic silica–alumina support the
materials with a predominantly type IV isotherm and a clear step in diesel range (250–380 °C cut) product yield is lower than that over
the adsorption curve at partial pressure p/p0 of 0.0.6–0.9 due to well- sulfide Ni–Mo/Al2O3 catalyst, about 45% at 2 h− 1 and about 80% at 4 h− 1.
defined mesoporosity. There is no significantly N2 uptake in the region The lower diesel yield for the Ni–W catalyst than for the Ni–Mo catalysts
p/p0 b0.1 indicating absence of microporosity in all the samples. BJH is due to more cracking of hydrocarbons over the former due to its higher
analysis of adsorption isotherm clearly shows that all these mesopor- acidity. The yield of the 140–250 °C distillation fraction (kerosene range)
ous materials have a very narrow pore size distribution with mean is nearly 50% at 2 h− 1 and about 15% at 4 h− 1 for the hydrocracking of
pore diameter of 8.6 nm for SiO2–Al2O3 and 12.9 for Al2O3. The mean soya-oil—gas-oil mixtures over Ni–W/SiO2–Al2O3 catalyst.
pore diameter of samples after metal impregnation was slightly Thus, for the Ni–W catalyst the 250–380 °C fraction is much lower
reduced, 7.8 nm for Ni–W/SiO2–Al2O3 and 12.2 for Ni–Mo/Al2O3. than that for the Ni–Mo catalyst which is attributed to more cracking
Scanning electron micrograph of mesoporous silica–alumina showed of this fraction into lower components due to higher acidity of the
elongated particles made up of fibrous aggregate (Fig. 2). former catalyst. Very high yield (85–95%) for 250–380 °C fraction is
Mixtures of soya-oil and gas-oil were hydroprocessed at reaction obtained using Ni–Mo/Al2O3 catalysts (Table 2) due to very little
temperatures ranging from 340 °C to 380 °C at two different space cracking ability of this catalyst.
velocities over the sulfided hydrotreating (Ni–Mo/Al2O3) and hydro- It is necessary to analyze the effect of catalyst composition, on the
cracking (Ni–W/SiO2–Al2O3) catalysts and the results are shown in relative rate of different oxygen removal pathways. Deoxygenation
Table 2. All the results were compared at near complete conversion of glycerides and FFA can follow different reaction pathways—
(N99%) of soya-oil in the mixture as observed from GC analysis. The hydrodeoxygenation, decarbonylation and decarboxylation. The
majority of the products produced from vegetable oil hydrotreatment oxygen elimination pathways influence the distribution of hydro-
are CO, CO2, methane, H2O, propane and the straight chain alkanes carbons in the products [5–17]. Hydrocarbons with one carbon atom
ranging from C8 to C18. The amount of CO formed during co-processing less than the original fatty acid chain i.e. heptadecane and
was nearly 5 to 8 times more than CO2. To recover the purge gas in pentadecane are the product of oxygen removal from triglyceride
industrial process it would be necessary to treat the CO in a methanator by decarboxylation + decarbonylation (with CO and CO2 as the side-
where it will react with hydrogen to form methane using a suitable products). Hydrocarbons with same carbon number as the original
methanation catalyst. Alternatively, it could be removed along with fatty acid i.e. octadecane and hexadecane, are the product of
other impurity components by pressure swing adsorption. hydrodeoxygenation reaction (with water as the side-product). The
During hydrotreating, the yield of product in the 250–380 °C extent of both pathways can be consequently elucidated from the
distillation range (which is the fraction corresponding to the diesel liquid hydrocarbon distributions. We have taken the C17/C18 ratio of
range—C15 to C18 hydrocarbons,) based on simulated distillation major product i.e. n-heptadecane (C17) and n-octadecane (C18)
results vary between 85 and 95% at different reaction temperature (they make up more than 89% of total n-alkanes from triglycerides),
and different space velocities for the Ni–Mo based hydrotreating as a measure of relative ratio of decarboxylation + decarbonylation
catalysts as shown in Table 2. The yield of alkanes in the 140–250 °C versus hydrodeoxygenation and the results for the two catalysts are
cut is less than 10% over the Ni–Mo catalysts. shown in Fig. 3. Different extents of two different pathways have

Table 2
Hydroprocessing of waste soya-oil and gas-oil mixtures on mesoporous Ni–Mo/Al2O3 (Ni–Mo) and Ni–W/SiO2–Al2O3 (Ni–W) catalysts.

Catalyst Feed Reaction conditions 140–250 Kerosene 250–380 Diesel 380–FBP bC15 C15–C18 N C18

Ni–Mo 10% Soya + 90% G.O. 380 °C 4 h− 1 5 95.0 – 6.66 92.42 0.9
Ni–Mo 25% Soya + 75% G.O. 360 °C 2 h− 1 3.3 86.1 10.6 7.37 82.96 8.82
Ni–Mo 25% Soya + 75% G.O. 380 °C, 2 h− 1 4.5 92.5 3.0 5.02 92.6 2.38
Ni–Mo 100% Soya-oil 380 °C, 2 h− 1 3.0 94.4 2.6 2.3 96.4 1.3
Ni–W 40% Soya + 60% G.O. 360 °C, 4 h− 1 15.2 79.5 5.3 8.8 80.8 10.4
Ni–W 25% Soya + 75% G.O. 360 °C, 2 h− 1 50.3 44.5 5.2 32.5 58.9 3.5
Ni–W 25% Soya + 75% G.O. 370 °C, 2 h− 1 51.4 43.2 5.4 39.2 55.7 4.9
Ni–W 100% Soya-oil 380 °C, 2 h− 1 21.8 74.5 4.7 19.6 76.8 3.6
562 R. Tiwari et al. / Catalysis Communications 12 (2011) 559–562
Decarbon.+Decarbox.

soya-oil did not decrease the HDS activity for gas-oil at complete
triglyceride conversion indicating that deoxygenation of soya-oil and
1.6 hydrodesulphurization are not competing reactions. The active sites for
both HDS and HDO can be expected to be similar but the number of
active sites would be sufficient enough for both the reactions to occur
1.4 simultaneously without inhibiting each other. The sulfur removal is 86–
NiMo, 4h-1
NiMo, 2h-1 93% for Ni–Mo catalyst but comparatively lower (60–85%) for Ni–W
NiW, 2h-1 catalyst. In comparison, complete deoxygenation of triglycerides was
1.2 observed indicating that HDO is a more favourable reaction than HDS.
Hydrodeoxygen. C17/18

Density of vegetable oil (0.92 g/cc) is much higher than that of the
gas-oil (0.8583 g/cc). After hydrotreating, densities of the products
1.0
decreased due to hydrodesulfurization, hydrodeoxygenation and
partial hydrogenation reactions and were in the range specified for
diesel fuel i.e. 0.82–0.86 g/cc (at 15 °C) . The product produced by
0.8
hydrotreating of soya-oil and gas-oil mixtures has much better cetane
number (54–67) than that for hydrotreated gas-oil (43).
340 345 350 355 360 365 370 375 380 385
4. Conclusions
Temperature (OC)
Mixtures of non-edible waste soya oil and refinery gas-oil can be
Fig. 3. Decarboxylation + decarbonylation versus hydrodeoxygenation measured as
C17/C18 ratio for 25% soya-oil in gas-oil over sulfide Ni–Mo/Al2O3 and Ni–W/SiO2– easily deoxygenated into pure hydrocarbons under hydroprocessing
Al2O3 catalysts. conditions using typical hydroprocessing catalysts supported on
mesoporous silica and silica–alumina. Plant-oils can be mixed with
petroleum fractions and treated in a conventional hydrotreater with no
been observed and reported as a result of temperature variation, adverse impact on hydrodesulfurization activity. Ni–W catalyst sup-
feed-ratio variation and due to different catalysts used (NiMo, Ni) ported on mesoporous silica–alumina could be used for selective
[16]. In terms of hydrogen consumption, hydrodeoxygenation is cracking along with deoxygenation to produce more kerosene. Ni–Mo
more demanding (4–12 times more) [21] than decarboxylation + catalyst supported on mesoporous alumina has very low cracking ability
decarbonylation and is therefore less preferred at higher hydrogen and has high selectivity for diesel range product due to deoxygenation of
pressures. Also, more gaseous products are formed during decar- soya-oil into C15–C18 hydrocarbons. The more hydrogen consuming
boxylation + decarbonylation. The hydrogen consumption is an hydrodeoxygenation pathway for oxygen removal from triglyceride is
important issue from the industrial point of view; a process with favored over the Ni–Mo/Al2O3 catalyst, while decarboxylation +
lower hydrogen demands would be more economical. The selectivity decarbonylation pathway is favored over the Ni–W/SiO2–Al2O3 catalyst.
to decarbonylation + decarboxylation and hydrodeoxygenation pro- This study has demonstrated that hydrotreating and hydrocracking of
ducts varies with type of catalyst used. The C17/C18 ratio for Ni–W vegetables oils with petroleum fractions is a viable process for
catalyst is more than 1 (1.5–1.6 at 350–370 °C but 1.2 at 380 °C). The production of liquid fuels with high cetane value and acceptable density
C17/C18 ratio is lower than 1 (0.5–0.8) at lower reaction tempera- using mesoporous catalyst supports.
tures of 350–370 °C but increases to ~ 1.0 at 380 °C. These differences
in the C17/C18 ratio for the two types of catalysts indicate that Acknowledgements
hydrodeoxygenation is more favored over the Ni–Mo catalyst than
that over Ni–W catalyst. Hence Ni–W seems to be a better catalyst for The authors thank the Analytical Sciences Division, IIP for
reducing the hydrogen consumption. analytical services. DST, India is acknowledged for the artial research
It is necessary to check if deoxygenation inhibits desulfurization funding. DV thanks UGC, India for research fellowship.
when both S- and O- containing compounds are present. Fig. 4 shows
the HDS activity of feed with 25% soya-oil mixture in gas-oil. Addition of
References
[1] J.G. Speight, Catal. Today 98 (2004) 55.
95 [2] B. Freedman, M.O. Bagby, J. Am. Oil Chem. Soc. 66 (1989) 1601.
[3] S. Bezergianni, A. Kalogianni, Biores. Technol. 100 (2009) 3927.
[4] Y. Liu, R. Sotelo-Boyás, K. Murata, T. Minowa, K. Sakanishil, Chem. Lett. 38 (2009) 552.
90 [5] L. Rodolfi, G.C. Zittelli, N. Bassi, G. Padovani, N. Biondi, G. Bonini, M.R. Tredici,
Biotechnol. Bioeng. 102 (2009) 100.
85 [6] D. Kubicka, P. Simacek, N. Zilkova, Top. Catal. 52 (2009) 161.
[7] S. Bezergianni, A. Kalogianni, I.A. Vasalos, Biores. Technol. 100 (2009) 3036.
[8] A.A. Lappas, S. Bezergianni, I.A. Vasalos, Catal. Today 145 (2009) 55.
HDS%

80 NiMo, 4h-1 [9] B. Donnis, R.S. Gottschalck Egeberg, P. Blom, K.G. Knudsen, Top. Catal. 52 (2009) 229.
[10] I. Kubicková, M. Snåre, K. Eränen, P. Mäki-Arvela, D.Yu. Murzin, Catal. Today 106 (2005)
NiMo, 2h-1
197.
75 NiW, 2h-1 [11] M. Stumborg, A. Wong, E. Hogan, Biores. Technol. 56 (1996) 13.
NiW, 4h-1 [12] S. Bezergianni, S. Voutetakis, A. Kalogianni, Ind. Eng. Chem. Res. 48 (2009) 8402.
[13] S. Bezergianni, A. Kalogianni, Biores. Technol. 100 (2009) 3927.
70
[14] I. Sebos, A. Matsoukas, V. Apostolopoulos, N. Papayannakos, Fuel 88 (2009) 145.
[15] T. Kalnes, T. Marker, D.R. Shonnard, Int. J. Chem. React. Eng. 5 (2007) A48.
65 [16] G.W. Huber, P. O'Connor, A. Corma, Appl. Catal. A Gen. 329 (2007) 120.
[17] P. Simacek, D. Kubicka, G. Sebor, M. Pospisil, Fuel 88 (2009) 456.
[18] S. Melis, S. Mayo, B. Leliveld, Biofuels Technol. 1 (2009) 43.
350 360 370 380 [19] P.R. Aravinda, P. Mukundana, P. Krishna Pillaia, K.G.K. Warrier, Micro. Meso.
O Mater. 96 (2006) 14–20.
Temperature ( C) [20] N.-Y. Topsoe, K. Pedersen, E.G. Derouane, J. Catal. 70 (1981) 41.
[21] Jeremy G. Immer, H. Henry, Lamb Energy Fuels 24 (2010) 5291–5299.
Fig. 4. Hydrodesulurization (HDS%) over sulfided Ni–Mo/Al2O3 and Ni–W/SiO2–Al2O3
catalysts with 25% soya-oil in gas-oil.

You might also like