Vous êtes sur la page 1sur 9

Applied Catalysis A: General 441442 (2012) 99107

Contents lists available at SciVerse ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Preparation and evaluation of hydrotreating catalysts based on activated carbon derived from oil sand petroleum coke
Yu Shi a, , Jinwen Chen a , Jian Chen b , Robb A. Macleod b , Marek Malac b
a b

Natural Resources Canada, CanmetENERGY-Devon, One Oil Patch Drive, Devon, AB, T9G 1A8, Canada National Institute for Nanotechnology, National Research Council Canada, 11421 Saskatchewan Drive, Edmonton, AB, T6G 2M9, Canada

a r t i c l e

i n f o

a b s t r a c t
Novel NiMo/activated carbon (AC) hydrotreating catalysts were prepared and evaluated for upgrading heavy vacuum gas oil (HVGO). The AC supports were derived from Alberta oil sand petroleum coke, i.e. uid coke and/or delayed coke, hereafter referred to as OSP coke, through a chemical process. The BET surface area was as high as 2194 m2 /g for the uid coke derived AC and 2357 m2 /g for the delayed coke derived AC. Both ACs contained a large number of micropores with pore volume as high as 1.2 cm3 /g. Ni and Mo based active component precursors could be easily loaded on the activated carbon supports by chemical impregnation of nickel nitrate and ammonium molybdate followed by calcination in nitrogen at 773 K without further modication or oxidation treatment to the activated carbons. Scanning electron microscopy (SEM) observation showed highly porous surface structure of the bare activated carbon supports and well dispersed metal (oxide) precursor nanoparticles of 3050 nm loaded on the AC supports. For comparison, two reference catalysts were also prepared by the same procedure but using commercial activated carbon and porous alumina as supports. After catalyst activation by sulding, the hydrotreating performance of the prepared catalysts was evaluated in a magnetically stirred autoclave with a HVGO feedstock to examine their hydrodesulfurization (HDS) and hydrodenitrogenation (HDN) activities. Two commercial hydrotreating catalysts were also tested and compared under similar conditions with the same feed. The results showed that the catalysts based on the activated carbon supports prepared from OSP coke had better hydrotreating performance than the other catalysts. Scanning transmission electron microscopy (STEM) characterization of the catalysts after activation showed that small particles of nanostructure (25 nm in size) were evenly embedded in the carbon matrix except for some bigger particles that were located on the catalyst surface. Energy dispersive X-ray (EDX) spectroscopy revealed that these particles were composed of Ni, Mo and S elements. The dispersed nanoparticles formed the active sites and were responsible for the observed high HDS and HDN activity. Elemental analysis and surface characterization of the spent catalysts showed that the formation of coke precursors was favored on the alumina supported catalyst, which resulted in catalyst deactivation. Crown Copyright 2012 Published by Elsevier B.V. All rights reserved.

Article history: Received 19 May 2012 Received in revised form 13 July 2012 Accepted 14 July 2012 Available online 20 July 2012 Keywords: Oil sand petroleum (OSP) coke Activated carbon (AC) Hydrotreating Heavy vacuum gas oil (HVGO)

1. Introduction As the world remaining accessible crude oil become heavier and more sour, there is an urgent need for improved technologies to process such crudes to produce clean transportation fuels [1]. Catalytic hydroprocessing technology is well established in conventional reneries worldwide. Due to the high contents of sulfur, nitrogen, asphaltenes and heavy metals (nickel and vanadium) in heavy feedstocks [2], existing catalysts and technologies have to be modied or new ones have to be developed to keep pace with more stringent environmental regulations, including

Corresponding author. Tel.: +1 780 987 8703; fax: +1 780 987 5349. E-mail address: YuShi@dal.ca (Y. Shi).

emissions from hydrocarbon fuel use, such as SOx , NOx , and CO2 [3,4]. Hydrotreating catalysts are used in reneries to catalytically remove S, N and metals, and to saturate aromatic compounds. Currently, alumina-supported hydrotreating catalysts are commonly used because of the good mechanical and textural specics of alumina [5,6]. However, suldation of alumina supported metal oxides is always incomplete due to the strong metalsupport interactions (SMSI) present in the catalyst suldation step, which is a significant drawback of alumina [79]. In addition, alumina supported catalysts suffer from deactivation caused by coking and nitrogen compounds, and heavy metal deposition when heavy oil is treated [2,1014]. In the past decades much effort has been paid either to modify existing catalysts by developing new synthesis methods through addition of new promoting species, or to develop new supports based catalysts to improve hydrotreating performance for heavy feedstocks [1520].

0926-860X/$ see front matter Crown Copyright 2012 Published by Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.apcata.2012.07.014

100

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

Activated carbon (AC) possesses high microporosity and surface area. It has attracted considerable attention recently as a possible candidate for replacing conventional alumina support for upgrading heavy gas oil [2,2123]. Activated carbon supported hydrotreating catalysts also have less sensitivity to nitrogen compounds, good resistance to coke deposition and easy recovery of metal components from spent catalysts, giving it advantages over -Al2 O3 for processing heavy petroleum feedstocks [13,2325]. Previous studies have demonstrated good hydrotreatment activities toward sulfur removal with AC supported Mo, NiMo or CoMo hydrotreating catalysts [2426]. However they have not been widely applied in commercial operations. Other promising hydrodesulfurization (HDS) results were also reported either by applying activated carbon (or mesoporous carbon) through hybridization with alumina support or by using activated carbon solely as catalyst support [19,2628]. Since activated carbon contains much less polar oxygen containing species than alumina, during suldation, the much weaker interaction between the support surface and metals favors the formation of the Type II high active sites (NiMoS II) [2931]. Furthermore, since cost is an important factor in catalyst selection for reneries, the relatively low production cost of activated carbon as hydrotreating catalyst support and easy recovery of metals by burning off the spent catalyst provide additional advantages. Recently, although activation of petroleum coke has been reported by several research groups with different activation processes, no efcient method has yet been demonstrated for converting Alberta OSP coke (both uid coke and delayed coke) into porous carbon materials with relatively high surface area and porosity [3236]. Since OSP coke contains more impurities, such as S (>6 wt%) and heavy metals, than conventional coke, the activation and application of it are more challenging [10,37]. Sulfur content in the AC can potentially limit its application as a hydrogenation catalyst support for supporting active metal components, especially for precious metals. On the contrary, transition metal sulde Ni(Co)S or Mo(W)S hydroprocessing catalysts usually require pre-suldation with sulfur containing compounds (such as H2 S) to activate the catalyst precursor. Therefore, sulfur is helpful, rather than harmful, in the activation process of this type of hydrotreating catalysts. Although activated carbon has a number of properties which are favorable for hydrotreating heavy petroleum feedstocks, its low mechanical strength and weak surface functionality might be the potential hurdles for industrial/commercial application. In the mean time alumina, due to its high mechanical strength and surface functionality, and mature commercial production technology, will continue to be the dominating commercial hydrotreating catalyst support in the foreseeable future. This paper explores the feasibility of developing low cost hydrotreating catalysts based on activated carbon derived from Alberta OSP coke and examines their hydroprocessing performance using heavy vacuum gas oil (HVGO) as the testing feed. The catalytic activity of the prepared catalysts and the effects of catalyst support have been evaluated by conducting comparative performance tests with a commercial AC supported catalyst, an alumina supported catalyst, and two commercial hydrotreating catalysts supported on -Al2 O3 .

mesoporous commercial -Al2 O3 (PURALOX HP 14/150) powder was provided by Sasol Germany GmbH and used as another reference catalyst support without any treatment. Since HVGO was used as feed in this study to evaluate catalyst activity, the mesoporous CAC and -Al2 O3 were expected to provide good diffusivity for large hydrocarbon molecules in the feed. In addition, two reference commercial NiMo/-Al2 O3 catalysts were obtained from two different catalyst vendors. Potassium hydroxide (ACS grade) and hydrochloric acid (ACS grade) used for coke activation and the product washing were purchased from Fisher Scientic Canada. Other chemicals used for catalyst preparation and activation, Ni(NO3 )2 6H2 O, (NH4 )6 Mo7 O24 4H2 O and dimethyl disulde (DMDS), were from SigmaAldrich with >99.9% purity (99% for DMDS). The HVGO feed was provided by a North American petroleum renery. 2.2. Activation of OSP coke Both coke samples were activated through a chemical activation process with KOH as activation agent. The raw delayed coke sample was ground and screened to a particle size range of 60200 mesh to match that of the uid coke as it was received without grinding. The coke sample was mixed with KOH powder with specied KOH/coke mass ratios of 2/1, 3/1 and 4/1. The mixture was transferred into a stainless steel combustion boat, which was then put into a quartz tube furnace with nitrogen gas ow at a rate of 50 ml/min. Thereafter, the mixture was heated to 673 K at a ramping rate of 10 K/min. After pre-activation for 1 h, the sample was heated again to a desired temperature of 9231123 K at a ramping rate of 5 K/min, and maintained at this temperature for 15 h. After activation and cooling down to room temperature, the product was washed with 1 M diluted hydrochloric acid and deionized (DI) water through ltration until the pH value of the ltrate was constant at 7. The obtained activated carbon was dried in an oven at 393 K overnight. In this paper, the activated carbon prepared from delayed coke is denoted as DAC and the one prepared from uid coke is denoted as FAC. 2.3. Catalyst preparation NiMo/AC supported catalysts were synthesized by incipient wetness metal impregnation of the prepared ACs. It is noted that both FAC and DAC were directly used for loading metals without any pretreatment for functionalization of the carbon surface. Ammonium hepta-molybdate and nickel nitrate aqueous solutions with specied metal concentrations were used to sequentially impregnate the AC supports. Mo was rst impregnated at a temperature of 353 K. After overnight drying and subsequent calcination at 723 K for 4 h with N2 purging, Ni was impregnated followed by drying and calcination under the same conditions. The metal concentrations of the solutions were so determined that the oxide metal concentrations in the calcined catalysts reached 15 wt% for MoO3 and 5 wt% for NiO. Similarly, 15 wt% of MoO3 and 5 wt% of NiO over the commercial AC and Al2 O3 supports were also loaded for comparison purpose. 2.4. Catalyst characterization

2. Experimental 2.1. Materials Dry uid coke (F-coke) and delayed coke (D-coke) were used directly without further pretreatment. A reference mesoporous commercial AC (CAC) was purchased from Aldrich Chemical Company, Inc. and ground into similar particle size range (60200 mesh) as that of AC samples converted from the OSP coke. A The C/H/S/N/O elemental composition and metal contents of the raw coke samples, the prepared AC samples and the spent catalysts were analyzed according to ASTM standard methods (ASTM D5291 and ASTM D5708). Specic surface area and porosity analyses of the samples were performed with a Micromeritic ASAP 2020 instrument. The BET surface area was obtained from N2 adsorption isotherm at 77 K. The pore size was analyzed from the desorption isotherm following the BJH method. The total pore volume was

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107 Table 1 Main properties of the HVGO feed. Density (g/cm3 , ASTM D4052) Carbon (wt%, ASTM 5291M) Hydrogen (wt%, ASTM 5291M) Sulfur (wt%, ASTM 4294) Nitrogen (wppm, ASTM 4629) Conradson carbon residue (CCR, wt%, ASTM D4530) Hydrocarbon type composition (ASTM D2786/D3239) Saturate (wt%) Aromatic (wt%) Polars (wt%) Simulated distillation ( C, ASTM 7169) IBP 10 30 50 70 90 FBP 0.9336 86.26 11.60 2.06 1359 0.16 44.4 49.8 5.78 182.6 295.0 372.8 415.0 455.0 509.6 590.0 Table 2 Element and ash analyses of the coke and AC samples. Element Carbon (wt%, ASTM 5373M) Hydrogen (wt%, ASTM 5373M) Oxygen (wt%) Nitrogen (wt%, ASTM 5373M) Sulfur (wt%, ASTM 4239M) Nickel (wppm, ICP-MS) Molybdenum (wppm, ICP-MS) Ash (wt%, ASTM D5142) F-coke 76.73 1.82 1.56 1.98 6.90 601 99.1 9.52 D-coke 83.98 3.90 0.88 1.67 6.33 398 89.5 3.49 FAC 87.22 0.51 2.76 0.10 <0.30 320 27.4 5.22 DAC

101

93.61 0.20 1.95 0.10 <0.30 172 30.7 0.23

chromatograph (GC) equipped with a thermal conductivity detector. The liquid product was collected and analyzed after separating the spent catalyst from the hydrogenated oil. 2.5.3. Treatment and characterization of spent catalysts Since the prepared catalysts were used for hydroprocessing of HVGO, coking on catalyst surface and pore structure changes might have occurred, which could lead to catalyst deactivation. To investigate the changes of the catalyst samples before and after the hydroprocessing tests, the spent catalysts were treated with reuxing toluene to extract any residue oil and other contaminants deposited on catalyst surface. The extraction process (soxhlet extraction) was performed over a 5-h time period. The treated catalysts were dried at 423 K in N2 for 5 h. Elemental analysis and surface characterization were then performed with the catalyst samples. 3. Results and discussion 3.1. Activation of OSP coke The yield of AC based on raw coke was calculated as the percentage of total mass of AC product over the total mass of the raw coke. The element and ash analyses of the raw cokes and the ACs (activated at typical conditions and used as supports for preparation of the NiMo/AC catalysts) were listed in Table 2. As Ni, Mo and S are desired constituents for hydrotreating catalyst [39], particular attention was paid to determine their contents in cokes and the prepared ACs. Both of the uid and delayed cokes contain as high as 6.3 wt% of sulfur, but relatively low amounts of Ni (<0.01 wt%) and Mo (<0.001 wt%). After activation and the subsequent washing, less than half of the original metals were removed, whereas more than 95% of the S was removed. Considering that the NiMo hydrotreating catalysts to be prepared is aimed at containing 15 wt% of MoO3 and 5 wt% of NiO, the metal contents left in the activated carbon after the activation of coke are considerably low and their contribution to the catalytic performance is negligible. The concentrations of the other elements and ashes were also reduced after activation except for those of oxygen and carbon. The oxygen content was increased by over 1 wt% in the activated carbons compared to the raw cokes, the reason for which will be discussed in the next section. The carbon content was increased by 10 wt% after the activation, which is expected since much of the heteroatoms and inorganic substances have been removed. It should be mentioned that the yields of the activated carbon obtained at all of the tested activation conditions were in the range of 6770 wt%. 3.2. Selection of activation conditions The catalytic activity of hydrotreating catalysts for heavy gas oil conversion depends on the surface texture and porosity of the catalyst support. Therefore, various activation conditions were investigated to achieve better AC textural properties. The BET surface characteristics of the ACs obtained at different conditions are

derived from the amount of N2 adsorbed at a relative pressure of P/P0 = 0.9677 (desorption branch). The morphology and structure characterization of the supports and NiMo oxide loaded catalyst precursors were carried out by scanning electron microscopy (SEM), transmission electron microscopy (TEM), and X-ray diffraction. SEM images of the samples were recorded on Hitachi S-4800 high resolution eld emission instrument. A thin layer of the powder sample was uniformly distributed onto a piece of adhesive conductive carbon tape that was subsequently mounted on alumina stubs for observation under SEM. Scanning transmission electron microscopy (STEM) and high resolution micrographs were acquired using JOEL 2200S eld emission TEM with an accelerating voltage of 200 kV. Energy dispersive X-ray (EDX) elemental analysis was also recorded from the TEM as follows: a small amount of powder sample, ground in an agate mortar, was dispersed into ethanol by ultrasonication to get a mixture suspension, which was placed over a lacy carbon coated copper grid for TEM imaging. XRD analysis was conducted on Rigaku Ultima IV diffractometer with Co K radiation ( = 0.1789 nm) and graphite monochromator. The XRD patterns of the samples were scanned in the 2 range 2090 with a scanning step of 0.05 . 2.5. Catalyst activity evaluation 2.5.1. Pre-suldation All the catalyst samples, including the commercial ones, were pre-activated by sulding with DMDS. Typically, 5 g of catalyst and 2 ml of DMDS were charged into a 300 ml magnetically stirred autoclave (Autoclave Engineers) at ambient temperature. The reactor was purged with 0.5 MPa H2 to remove air and then pressurized with 3.45 MPa H2 at ambient temperature prior to heating up for pre-suldation. The pre-suldation of the catalysts was conducted in two steps at 503 K and 613 K respectively, with each step lasting for 1 h. 2.5.2. Catalytic activity test A HVGO feed was used to evaluate the hydrotreating activity of the catalysts using the same magnetically stirred autoclave right after completion of pre-suldation. The main characteristic properties of the HVGO are given in Table 1. 100 g of HVGO feed was charged under nitrogen gas protection at ambient temperature into the autoclave containing the sulded catalyst. All the evaluation tests were carried out at 643 K under a hydrogen pressure of 3.455.50 MPa, with magnetic stirring speed at 400 rmp. After 2 h of reaction period at 643 K, the autoclave reactor was cooled down naturally. Before releasing the autoclave pressure a gas sample was taken and analyzed with Agilent 3000 micro gas

102

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

Table 3 Surface characterization of the coke and AC samples obtained at different activation conditions. Sample F-coke FAC FAC FAC FAC FAC FAC FAC FAC FAC FAC FAC D-coke DAC DAC DAC CAC -Al2 O3 NiMo/FAC NiMo/DAC NiMo/CAC NiMo/Al2 O3 Temperature (K) 923 923 973 1023 1023 1023 1023 1073 1123 1023 1023 973 1023 1073 Time (h) 1.5 5 1.5 1 1.5 2 4 1.5 1.5 1.5 1.5 1.5 1.5 1.5 Mass ratio of KOH/coke 3:1 3:1 3:1 3:1 3:1 3:1 3:1 3:1 3:1 2:1 4:1 3:1 3:1 3:1 BET surface area (m2 /g) 8.8 1686 1566 1926 1987 2194 2133 2030 1827 1689 1174 1878 0.17 1716 2357 2257 559 143 1681 1830 323 108 Pore volume (cm3 /g) 0 0.95 0.73 1.11 1.15 1.21 1.20 1.13 0.99 0.93 0.65 1.03 0 0.91 1.24 1.19 0.76 1.04 0.88 0.92 0.48 0.76 Average pore diameter (nm) 2.1 1.9 2.1 2.1 2.1 2.1 2.1 2.2 2.2 2.2 2.2 2.1 2.2 2.1 5.3 27.0 2.1 2.0 6.1 28.6

shown in Table 3. For comparison, the BET surface area and porosity characteristics of the raw cokes, CAC, and -Al2 O3 were also measured and are listed in Table 3. At most of the tested conditions, both F-coke and D-coke were successfully activated with high surface area (>1600 m2 /g) and large pore volume (>0.93 cm3 /g), except for KOH/coke mass ratio of 2. Compared with CAC and Al2 O3 , although the average pore diameters of the prepared ACs are around 2 nm, their extremely high surface area and pore volume would still make them good catalyst support to achieve high catalyst activity. Fig. 1 shows the surface characteristics of the FACs obtained at different activation temperatures and activation times. The surface area and pore volume of the FACs increased from 1686 m2 /g and 0.95 cm3 /g to 2194 m2 /g and 1.21 cm3 /g, respectively, when the activation temperature was increased from 900 to 1023 K (Fig. 1A) with activation time xed at 1.5 h. When the temperature was increased further from 1023 K, the surface area and pore volume decreased sharply since extremely high temperature might have caused texture sintering. Fig. 1B shows the BET surface areas and pore volumes at different activation times at xed activation temperature of 1023 K. The highest surface area and pore volume were achieved with activation time of 1.5 h. Similar observations were made for DACs (not shown here). DAC with the highest surface area of 2357 m2 /g and pore volume of 1.24 cm3 /g was obtained at 1023 K from D-coke. In addition to

activation temperature and time, mass ratio of KOH/coke also affected the textural properties of the prepared AC. Under the same activation conditions, KOH/F-coke mass ratio of 3/1 resulted in higher surface area and pore volume than 2/1 and 4/1. In the mean time, another batch of F-coke was chemically processed at 923 K for 5 h to cover a wider range of operation conditions. The surface measurement shows that neither BET surface area nor pore volume is higher than the one treated at the same temperature for a short activation time (1.5 h). Therefore, KOH/coke mass ratio of 3/1, activation temperature of 1023 K, and activation time of 1.5 h were identied as the best activation conditions for converting the OSP coke into activated carbon. 3.3. Characterization of OSP coke and activated carbon Fig. 2 shows the representative high resolution SEM imaging characterization of F-coke and D-coke (Fig. 2A and B, respectively) and their corresponding activated carbon after activation (Fig. 2C and D, respectively). The insets are the corresponding low magnication SEM images. The high magnication images show at or rough surface of the raw cokes (Fig. 2A and B), while the corresponding activated carbons show highly porous structure (Fig. 2C and D). In addition to micro-sized pores, meso-pores and macro-pores are also seen in both ACs, indicating good structural characteristics as catalyst supports for hydrotreating heavy gas oil.

1.22 2200

BET surface area (m /g)

BET surface area (m /g)

2100 2000 1900 1800 1700 1600 900 950 1000 1050

1.2

2200

1.20 1.18 1.16 1.14 1.12

3 Pore volume (cm /g)

3 Pore volume (cm /g)

2150 2100 2050 2000 1950

1.1

1.0

1100

0.9 1150

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Temperature (K)

Time (h)

Fig. 1. BET surface and pore volume of FAC at different activation temperatures (A) and activation times (B).

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

103

Fig. 2. SEM images of F-coke (A), D-coke (B), FAC (C) and DAC (D) (insets are the corresponding low magnication SEM images).

The structure and crystallinity of the obtained ACs were also analyzed with XRD. The typical XRD patterns presented in Fig. 3 (curves a and b) reveals that the prepared ACs are almost amorphous. Only two weak peaks are observed and no obvious impurity can be detected. The two diffraction peaks corresponding to (0 0 2) and (1 0 1) are those of graphitic carbon, indicating the disordered AC structure. Since activated carbon is normally composed of randomly stacked graphite planes, such a disordered structure leads to a high porosity of the carbon materials, which is also conrmed by SEM characterization. 3.4. Characterization of NiMo/AC catalyst precursors The XRD patterns of the calcined NiMo/AC precursors are also shown in Fig. 3 (curves c and d). In comparison with curves a and b for the AC supports, other than the two weak peaks originated from
800

600

a
C (002) C (101)

400

b c
200

d
0 20 40 60 80

2 Theta (deg.)
Fig. 3. XRD patterns of the AC supports (a, DAC; b, FAC) and their corresponding calcined NiMo/AC catalyst precursors (c, NiMo/DAC; d, NiMo/FAC).

carbon, no additional diffraction peaks are observed from the XRD patterns. Although the metal loading contents are 15 wt% for MoO3 and 5 wt% for NiO for the calcined samples, neither MoO3 and NiO, nor NiMoOx diffraction peaks were recorded. Therefore, the XRD patterns suggest that very good dispersions of both MoO3 and NiO on the AC supports are achieved. Fig. 4A and B shows the SEM characterization of the calcined catalyst precursors NiMo/FAC and NiMo/DAC, respectively. Highly dispersed nanoparticles (white dots) of 3050 nm in size can be clearly seen on both of the porous AC supports. Compared with the bare AC supports (Fig. 2C and D), these nanoparticles should be the loaded metal oxides. Such a very good distribution of metal oxides accounts for the related XRD patterns with an indication of good catalytic activity for hydrotreatment. As mentioned earlier that the oxygen content was increased by more than 1 wt% after activation. The addition of oxygen to the activated carbon during chemical treatment is attributed to the use of KOH as active agent for converting the coke. Normally, the presence of various oxygenated groups of either acidic or basic nature on the carbon surface is of great importance for metal adsorption and catalyst dispersion [4042], since the surfaces of carbon supports are nearly neutral and heteroatoms can modify the surface properties of carbon [43,44]. The oxygen containing functional groups generated in situ on carbon surface during the activation process lead to an effective and uniform dispersion/distribution of active metals during impregnation [38]. To examine the differences in surface properties of the AC supports and the prepared catalyst precursors, BET surface area and pore volume were also measured with the latter. The results are shown in Table 3 as well. It was observed that both the surface area and pore volume of the catalyst precursors were about 25% lower than those of their corresponding AC and alumina supports. The drop with the CAC catalyst precursor was even higher (40%). This indicates that no matter which catalyst support is used, considerable decease in surface area is expected after metal loading and calcination. However, little differences were observed with the pore size distribution as seen in Table 3.

Intensity (a.u.)

104

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

0.94

Density (mg/l)

0.93

0.92

0.91

0.90
3 2 1 C C C ed Fe M/FA M/DA M/CA /Al2O Com Com N N N NM
Fig. 5. Density (ASTM D4052) of the feed and hydrogenated liquid products obtained with different catalysts (temperature: 643 K; pressure: 3.45 MPa; reaction time: 2 h).

80

Sulfur conversion (%)

60

40

Fig. 4. SEM images of the calcined NiMo/AC catalyst precursors (A: NiMo/DAC; B: NiMo/FAC).

20

3.5. Hydrotreating activities of the NiMo/AC catalysts

0
All the hydrotreating evaluation experiments were carried out in a 300 ml autoclave reactor with a magnetic stirrer under the similar operating conditions. The catalysts, including the prepared NiMo/FAC, NiMo/DAC, NiMo/CAC, NiMo/-Al2 O3 , and the two commercial reference hydrotreating catalysts were activated using DMDS with the same pre-sulding procedure under the similar conditions before activity tests. Since the two commercial NiMo hydrotreating catalysts were in extruded form, for comparison they were ground into the same size range as the prepared AC catalysts. Table 4 lists the main properties of the hydrotreated liquid products obtained with different catalysts. It is observed that all the catalysts show some extent of hydrotreating activity indicated by lower oil density, and lower contents of S, N and Conradson carbon residue (CCR) than those of the HVGO feed. The specic hydrotreating performance, in terms of density change, HDS, HDN, and CCR conversion are demonstrated in Figs. 58, respectively. Fig. 5 presents the densities of the hydrogenated HVGO with the 6 different catalysts. The hydrogenated products obtained with NiMo/FAC and NiMo/DAC have lower density than those obtained with other catalysts. Fig. 6 shows the HDS conversions obtained with different catalysts. Over 70 wt% of the sulfur in the HVGO feed was removed after hydrotreating with either FAC or DAC supported NiMo catalyst, suggesting a higher HDS activity of the developed AC catalysts. It is worth noting that in agreement with S content of the hydrotreated liquid products, gas analysis showed that the AC-supported catalysts gave both higher hydrogen consumption and hydrogen sulde production than other catalysts, which are also indications of HDS

3 2 1 C C C /FA M/DA M/CA /Al2O Com Com M N N N NM


Fig. 6. HDS activity obtained with different catalysts (temperature: 643 K; pressure: 3.45 MPa; reaction time: 2 h).

Nitrogen conversion (%)

30 25 20 15 10 5 0
2 3 1 C C C /FA M/DA M/CA /Al2O Com Com M N N N NM

Fig. 7. HDN activity obtained with different catalysts (temperature: 643 K; pressure: 3.45 MPa; reaction time: 2 h).

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107 Table 4 Main properties of the HVGO feed and liquid products with different catalysts. Density (g/ml, ASTM D4052) Feed NiMo/FAC NiMo/DAC NiMo/CAC NiMo/Al2 O3 COM1 COM2 0.9336 0.918 0.9159 0.9226 0.9228 0.9232 0.9205 S (wt%, ASTM 4294) 2.06 0.56 0.45 0.76 0.71 0.74 1.16 N (wppm, ASTM 4629) 1359 1036 954.3 1324 1255.5 1257 1140

105

CCR (wt%, ASTM D4530) 0.16 0.01 <0.01 0.08 0.05 0.11 0.06

activity. Although the HDN conversions shown in Fig. 7 for all of the catalysts are low (<30 wt%), NiMo/FAC and NiMo/DAC performed better than the other four reference catalysts. Fig. 8 exhibits the CCR conversions obtained with different catalysts. Again, the NiMo catalysts supported on activated carbon derived from OSP coke (NiMo/FAC and NiMo/DAC) yielded higher CCR conversion (>92 wt%). It is also noted that between the two catalysts based on OSP coke derived activated carbon, NiMo/DAC showed higher hydrotreating activity than NiMo/FAC in terms of density reduction, HDS, HDN and CCR conversion, indicating that the delayed coke used in this study is more suitable for making activated carbon catalyst support. The higher hydrotreating activity of NiMo/DAC is attributed to the higher BET surface area and pore volume of the activated carbon support (Table 3). Note that the CAC and the -Al2 O3 used for catalyst preparation in this study are mesoporous supports, which favour the diffusion of large hydrocarbon molecules, such as those in HVGO. Therefore the corresponding catalysts should have demonstrated higher hydrotreating activity than the FAC and DAC based microporous catalysts. However, experimental observations as discussed above show the opposite trend. Therefore the surface properties of FAC and DAC, which allow uniform distribution of well-dispersed small metal oxide particles might have had signicant contribution to the observed high hydrotreating activity. It should be pointed out that the activity tests conducted in this study with all the catalyst samples lasted only a few hours and the hydrotreating activity data represent their initial values. Due to the limitation of the autoclave batch reactor, catalyst stability and deactivation over a long time period could not be evaluated. The large number of micropores in the DAC and FAC supports might suffer from the deposition of coke and contaminants over time, resulting in catalyst deactivation. The testing of the prepared catalysts in a continuous ow reactor system is undergoing to evaluate the

catalyst stability and deactivation. The results will be published in the future. 3.6. Aromatics conversion In addition to HDS and HDN activities, the conversion of aromatic molecules is another important indication of catalyst activity for hydrotreating heavy feeds. Fig. 9 presents the contents of different types of hydrocarbons in the feed and in the hydrotreated liquid products. All the liquid products show at least 10 wt% higher content of saturates and 10 wt% lower content of aromatics than the feed, indicating their hydrogenation activity for large aromatic molecules. As seen in Fig. 9, the prepared NiMo/FAC and NiMo/DAC do not show any advantage in aromatics hydrogenation over other catalysts. Instead, the catalysts prepared with mesoporous CAC and alumina show slightly higher aromatics hydrogenation activity than others. The lowest contents of polars in the liquid products obtained with NiMo/FAC and NiMo/DAC catalysts conrm the superior HDS and HDN activities. 3.7. Structure characterization of NiMo/DAC In order to understand the observed higher hydrotreating activity of the two catalysts prepared with activated carbon supports derived from OSP coke, further study on metal distribution and microstructure of the NiMo/DAC catalyst after suldation (with DMDS in H2 atmosphere) was conducted with STEM. Fig. 10 shows high-resolution STEM images of the sulded NiMo/DAC catalyst. At low magnication (Fig. 10A), particles of 3050 nm in size can be clearly seen dispersed on the AC support from the typical bright eld STEM image, which is similar to the SEM observation. High angle annular dark eld (HAADF) STEM imaging technique was

100

70 60

CCR conversion (wt%)

Saturates Aromatics Polars

80

Content (wt%)

50 40 30 20 10 0
Fe ed FAC O 3 om 1 om 2 AC AC / /D M/C /Al2 C C NM NM M N N

60 40 20 0

1 2 3 C C C /FA M/DA M/CA /Al2O Com Com M N N N NM

Fig. 8. CCR conversion obtained with different catalysts (temperature: 643 K; pressure: 3.45 MPa; reaction time: 2 h).

Fig. 9. Hydrocarbon type composition (ASTM D2786/D3239) of the feed and hydrotreated liquid products obtained with different catalysts (temperature: 643 K; pressure: 3.45 MPa; reaction time: 2 h).

106

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

Fig. 10. STEM images (A and B) and EDX elemental analysis (C) of the MoNi/AC catalyst after suldation. A: low-magnication bright eld view; B: high-magnication HAADF view of the marked edge area in image A; C: EDX elemental analysis of a selected nanoparticle imaged in B.

applied to explore the nanostructure feature of the ne oxide particles on the catalyst support. Since the contrast in HAADF-STEM depends on the Z-number of the constituent atoms with brighter contrast for heavier constituents, it is a powerful tool for characterization of supported catalyst. Fig. 10B shows the high resolution HAADF-STEM image taken from the circled area in Fig. 10A. The magnied image acquired from the selected thin edge of the specimen exhibits large numbers of small particles embedded in the support matrix with size ranging from 2 to 5 nm. EDX element analysis performed with the nanoparticles, as shown in Fig. 10C, conrms that the composition of the nanostructures is Ni, Mo and S. The carbon peak comes from the background of the AC support. With respect to the hydrotreating activity of the prepared FAC and DAC catalysts, the high surface area and porous structure definitely have a positive effect on HDS and HDN. Further study on activation of OSP coke is underway to enlarge the micropores in the activated carbon, thereby further enhancing its supported catalyst activity for conversion of large hydrocarbon molecules in heavy feeds. Meanwhile, because of the absence of strong metalsupport interactions in activated carbon supported NiMo catalysts, metalsuldes particles tend to be dispersed more uniformly in a smaller size on the porous surface, which is another important factor for their observed higher hydrotreating activity.
Table 5 Elemental analysis and surface characterization of the spent catalysts. Sample Spent NiMo/FAC Spent NiMo/DAC Spent NiMo/Al2 O3 Carbon content (wt%) 72.2 75.6 8.1

3.8. Catalyst deactivation As mentioned earlier, to better understand the deactivation mechanism and its effect on catalyst life, three spent catalysts, i.e. NiMo/FAC, NiMo/DAC and NiMo/Al2 O3 were analyzed after soxhlet extraction. Table 5 shows the carbon contents and surface properties of the spent catalyst samples. After hydroprocessing, 8.1 wt% of carbon was found on the alumina catalyst while carbon content increase was much lower for the two AC-based catalysts, 3 wt% for NiMo/FAC and <1 wt% for NiMo/DAC, in comparison with the calculated values for the fresh ones. Evidently, coke precursor was formed during the hydroprocessing tests and the alumina support was more favorable for coke deposition than the two AC supports. Surface characterization shows that the AC supported catalysts lost more than half of the surface area and pore volume after the hydroprocessing tests, while the alumina supported one retained almost all the surface area and 90% of the pore volume, compared with their counterparts (fresh catalysts before suldation). However, after the hydroprocessing tests, the two AC supported catalysts still had similar average pore size as the fresh catalysts while average pore size of the alumina supported catalyst dropped from 28.6 nm to 22.2 nm. Indicatively, although the mesopores in alumina were not blocked as much as the micropores in

BET surface area (m2 /g) 781 866 115

Pore volume (cm3 /g) 0.40 0.42 0.62

Average pore diameter (nm) 2.0 1.9 22.2

Y. Shi et al. / Applied Catalysis A: General 441442 (2012) 99107

107

the AC supports after the hydroprocessing tests, they shrank considerably due to coke deposition. In combination with the carbon elemental analysis data, it is suggested that coke precursor particles were formed on the walls/surface of the mesopores in the alumina support during the hydroprocessing tests, which prevented the active sites on the catalyst surface from contacting hydrocarbon molecules. This might partially account for the observed lower hydrotreating activity of the alumina supported catalyst compared with the AC supported catalysts. 4. Conclusions Two types of OSP coke were successfully activated through a chemical process with an average yield of 68.5 wt%. The obtained activated carbons have high BET surface areas, over 2194 m2 /g for the uid coke and over 2357 m2 /g for the delayed coke. The activated carbons contain a large number of micropores with pore volumes as high as 1.2 cm3 /g. High resolution SEM characterization was performed with the AC samples to conrm their structure and morphological features. Two NiMo hydrotreating catalysts were prepared by using the AC as supports. Good dispersion of the metal oxides was observed on the catalyst surface. Both XRD and SEM analyses consistently conrmed the uniform distribution of the metal particles. The prepared catalysts were evaluated in a batch autoclave for hydrotreating heavy vacuum gas oil and were compared with catalysts prepared by using a commercial AC and an alumina as supports, and with two reference commercial hydrotreating catalysts. Among all the catalysts evaluated, the two prepared with ACs derived from OSP coke showed signicantly improved hydrotreating performance in terms of HDS, HDN and CCR conversion for HVGO hydrotreating. High resolution HAADF-STEM imaging indicated that 25 nm ne nanostructures are embedded in the AC matrix with a composition of NiMoS after suldation of the catalysts, which is the essential constituent of active site for hydrotreating. Therefore, the high surface area and porous structure, the ability for resistance to coke deposition of the AC supports, and the highly dispersed small metalsulde particles account for the observed hydrotreating activity of the prepared catalysts. Acknowledgements Partial funding for this study was provided by the Canadian Interdepartmental Program of Energy Research and Development (PERD 1.1.3). The authors are grateful to the CanmetENERGY-Devon pilot plant staff for conducting the experiments and to the analytical lab staff for performing the analyses. Analytical support from the National Institute for Nanotechnology (National Research Council, Canada) is greatly appreciated. Dr. Edward Littles valuable inputs and suggestions on revising this paper are greatly appreciated.

References
[1] J.G. Speight, The Chemistry and Technology of Petroleum, fourth ed., Taylor and Francis Group, Florida, 2007. [2] H. Fukuyama, S. Terai, M. Uchida, J.L. Cano, J. Ancheyta, Catal. Today 98 (2004) 207215. [3] J.H. Gary, G.E. Handwerk, Petroleum Rening Technology and Economic, fourth ed., Marcel Dekker, New York, 2001. [4] C. Song, Catal. Today 86 (2003) 211263. [5] H. Topse, B.S. Clausen, Appl. Catal. 25 (12) (1986) 273293. [6] H. Shimada, T. Sato, Y. Yoshimura, J. Hiraishi, A. Nishijima, J. Catal. 110 (1988) 275284. [7] J.C. Duchet, E.M. van Oers, V.H.J. de Beer, R. Prins, J. Catal. 80 (1983) 386402. [8] J.P.R. Vissers, B. Scheffer, V.H.J. de Beer, J.A. Moulijn, R. Prins, J. Catal. 105 (1987) 277284. [9] B. Scheffer, P. Arnoldy, J.A. Moulijn, J. Catal. 112 (1988) 516527. [10] M.R. Gray, Upgrading Petroleum Residues and Heavy Oils, Marcel Dekker, New York, 1994. [11] F. Diez, B.C. Gates, J.T. Miller, D.J. Sajkowski, S.G. Kukes, Ind. Eng. Chem. Res. 29 (1990) 19992004. [12] S.C. Korre, M.T. Klein, R.J. Quann, Ind. Eng. Chem. Res. 34 (1995) 101117. [13] H. Fukuyama, S. Terai, Petrol. Sci. Technol. 25 (2007) 231240. [14] M. Absi-Halabi, A. Stanislaus, D.L. Trimm, Appl. Catal. 72 (1991) 193215. [15] M.S. Rana, V. Smano, J. Ancheyta, J.A.I. Diaz, Fuel 86 (2007) 12161231. [16] Y.D. Liu, L. Gao, L.Y. Wen, B.N. Zong, Recent Pat. Chem. Eng. 2 (2009) 2236. [17] M. Breysse, C. Geantet, P. Afanasiev, J. Blanchard, M. Vrinat, Catal. Today 130 (2008) 313. [18] E. Puello-Polo, J.L. Brito, J. Mol. Catal. A 281 (2008) 8592. [19] F. Liu, S.P. Xu, Y.W. Chi, D.F. Xue, Catal. Commun. 12 (2011) 521524. [20] H.L. Yin, T.N. Zhou, Y.Q. Liu, Y.M. Chai, C.G. Liu, Reac. Kinet. Mech. Cat. 102 (2011) 405416. [21] C.B. Xu, S. Hamilton, A. Malik, M. Ghosh, Energy Fuels 21 (2007) 34903498. [22] P. Serp, J.L. Figueiredo, Carbon Materials for Catalysis, John Wiley & Sons, New Jersey, 2009. [23] L.A. Rankel, Energy Fuel 7 (1993) 937942. [24] Y. Sugimoto, 16th Saudi ArabiaJapan Joint Symposium, Dhahran, Saudi Arabia, November 56, 2006, 2006. [25] C. Glasson, C. Geantet, M. Lacroix, F. Labruyere, P. Dufresne, J. Catal. 212 (2002) 7685. [26] P. Vazquez, L. Pizzio, M. Blanco, C. Caceres, H. Thomas, R. Arriagada, S. Bendezu, R. Cid, R. Garcia, Appl. Catal. A 184 (1999) 303313. [27] I. Eswaramoorthi, V. Sundaramurthy, Nikhil Das, A.K. Dalai, J. Adjaye, Appl. Catal. A 339 (2008) 187195. [28] E.J.M. Hensen, H.J.A. Brans, G.M.H.J. Lardinois, V.H.J. de Beer, J.A.R. van Veen, R.A. van Santen, J. Catal. 192 (2000) 98107. [29] H. Topse, B.S. Clausen, F.E. Massoth, Hydrotreating Catalysis, Springer, Berlin, 1996. [30] J.A.R. van Veen, H.A. Colijn, P.A.J.M. Hendriks, A.J. van Welsenes, Fuel Process. Technol. 35 (1993) 137157. [31] S.M.A.M. Bouwens, F.B.M. Vanzon, M.P. Vandijk, A.M. Vanderkraan, V.H.J. Debeer, J.A.R. Vanveen, D.C. Koningsberger, J. Catal. 146 (1994) 375393. [32] R. DiPanlo, N.O. Egiebor, Fuel Process.Technol. 46 (1996) 157169. [33] S. Mitani, S.I. Lee, S.H. Yoon, Y. Korai, I. Mochida, J. Power Sources 133 (2004) 298301. [34] S. Koutcheiko, T. McCracken, J. Kung, L. Kotlyar, Petrol. Sci. Technol. 25 (2007) 12151224. [35] C.L. Lu, S.P. Xu, Y.X. Gan, S.Q. Liu, C.H. Liu, Carbon 43 (2005) 22952301. [36] M. Kubota, A. Hata, H. Matsuda, Carbon 47 (2009) 28052811. [37] L.F. Jia, E.J. Anthony, I. Lau, J.S. Wang, Fuel 85 (2006) 635642. [38] C. Moreno-Castilla, Carbon 42 (2004) 8394. [39] M. Brorson, A. Carlsson, H. Topse, Catal. Today 123 (2007) 3136. [40] G. de la Puente, J.A. Menendez, Solid State Ionics 112 (1998) 103111. [41] B.R. Puri, Chemistry and Physics of Carbon, Marcel Dekker, New York, 1970. [42] J.P. Donnet, Carbon 6 (1968) 161176. [43] F. Rodriguez-Reinoso, C. Salinas-Martinez de Lecea, A. Sepulveda-Escribano, J.D. Lopez-Gonzalez, Catal. Today 7 (1990) 287298. [44] H.P. Bohem, Carbon 32 (1994) 759769.

Vous aimerez peut-être aussi