You are on page 1of 8

Fuel 89 (2010) 25362543

Contents lists available at ScienceDirect

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

Effects of the operating variables on hydrotreating of heavy gas oil: Experimental, modeling, and kinetic studies
M. Mapiour a, V. Sundaramurthy a,1, A.K. Dalai a,*, J. Adjaye b
a b

Catalysis and Chemical Reaction Engineering Laboratories, Department of Chemical Engineering, University of Saskatchewan, Saskatoon, SK, Canada S7N 5A9 Syncrude Edmonton Research Centre, Edmonton, AB, Canada T6N 1H4

a r t i c l e

i n f o

a b s t r a c t
The effects of H2 purity, pressure, gas/oil ratio, temperature, and LHSV on hydrotreating activities were investigated in a micro-trickle bed reactor using a commercial NiMo/c-Al2O3 catalyst. Heavy gas oil (HGO) from Athabasca bitumen was used as feed. Due to their signicant effects on H2 partial pressure, H2 purity, pressure, and gas/oil ratio were chosen and used in a central composite design (CCD) method. Experimental conditions used were H2 purity, pressure, and gas/oil ratio were: 75100 vol.% (with the rest methane), 711 MPa, and 4001200 mL/mL, respectively. The effect of LHSV (0.652 h1) and temperature (360400 C) were studied in a separate set of experiments. Vapor/liquid equilibrium (VLE) calculations were performed to determine the inlet and outlet H2 partial pressures. It was observed that the enhancing effects of H2 purity on hydrodenitrogenation (HDN) and hydrodearomatization (HDA) activities were greater than that of gas/oil ratio; however, it was comparable to pressure. Hydrodesulphurization (HDS) activity was not considerably affected by H2 purity, pressure, or gas/oil ratio. Increasing LHSV led to a decrease in HDS, HDN, and HDA activities while increasing temperature resulted in an increase in HDS and HDN; HDA had maximum activity at about 385 C. Kinetic tting of the data to a pseudo-rstorder power law model suggested that conclusions on hydrotreating activities responses to a changing H2 pressure could be equally drawn from either inlet or outlet H2 partial pressure. However, from the catalyst deactivation standpoint, it is recommended that such conclusions are drawn from the outlet H2 partial pressure. Crown Copyright 2010 Published by Elsevier Ltd. All rights reserved.

Article history: Received 7 July 2009 Received in revised form 19 February 2010 Accepted 23 February 2010 Available online 5 March 2010 Keywords: H2 partial pressure Hydrotreatment H2 purity Heavy gas oil

1. Introduction Hydrotreating is a catalytic process in which oil feedstock is reacted with hydrogen to remove contaminants such as sulfur (S) and nitrogen (N) and saturate aromatics and olens [1]. The process of removal of S is known as hydrodesulfurization (HDS), the removal of N as hydrodenitrogenation (HDN), and the reduction of aromatics as hydrodearomatization (HDA). If these contaminants are not removed, they may have detrimental environmental and health effects. Commonly in the literature, hydrotreating is discussed in terms of the following operating variables: temperature, pressure, LHSV (Liquid hourly space velocity), and gas/oil [24]. There is rarely any mention of H2 purity even though it is known that H2 purity is an important variable [5]. H2 purity may have been ignored in literature as very pure H2 streams (>99.9%) were used in the research. However, in an industrial setting, H2 purity can be as low
* Corresponding author. Tel.: +1 306 966 4771; fax: +1 306 966 4777. E-mail addresses: mlm499@mail.usask.ca (M. Mapiour), sundaramurthy@src. sk.ca (V. Sundaramurthy), ajay.dalai@usask.ca (A.K. Dalai). 1 Present address: Saskatchewan Research Council, Saskatoon, SK, Canada S7N 2X8.

as 70 vol.% [5]. It is therefore necessary to conduct research at conditions similar to those found in practice. The efuent stream of a hydrotreater contains unreacted H2. For optimal economics this unreacted H2 must be recovered and recycled. The recycle H2 stream contains several impurities, mainly CH4 an H2S [6]. CH4 is produced by cracking and other reactions, and H2S is produced as a result of HDS. Methane is an important concern because it is difcult to remove and often builds up to high levels in the recycle stream [7]. In our previous work [8] it was found that methane was inert toward the commercial NiMo/cAl2O3 catalyst and its presence hindered hydrotreating activities by reducing the H2 partial pressure in the reactor. H2S is known to inhibit hydrotreating activities yet its presence is crucial to maintain the sulded active phase of the catalyst [9]. In our previous work [8], the effect of H2 purity on hydrotreating activities was studied. However, the interacting effects of H2 purity with other hydrotreating operating variables were not investigated. Moreover, no models were developed to correlate hydrotreating activities, namely: HDS, HDN, and HDA, with the operating conditions: namely H2 purity, pressure, and gas/oil ratio. The objective of this work therefore is to develop models for hydrotreating activities as functions of operating variables, specically

0016-2361/$ - see front matter Crown Copyright 2010 Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.fuel.2010.02.024

M. Mapiour et al. / Fuel 89 (2010) 25362543

2537

H2 purity, pressure, and gas/oil ratio, and to optimize these three variables for maximum hydrotreating activities. H2 purity, pressure, and gas/oil ratio were specically chosen because their effects on H2 partial pressure are greater than those of temperature and LHSV especially for heavier feedstock. The effects of temperature and LHSV on hydrotreating activities will also be presented in this paper. 2. Experimental Heavy gas oil (HGO) derived from oil sands bitumen supplied by Syncrude Canada Ltd. was used in this study as the feedstock. Some properties of this HGO are summarized in Table 1. The catalyst employed was a trilobe-shaped commercial NiMo/c-alumina (C424) from Criterion Catalysts with a diameter of 1.5 mm. The catalyst had a pore volume of 0.5 mL/g and a surface area of 160 m2/g. The experiments were carried out in a micro-trickle bed reactor using 5 mL of catalyst. A schematic of the experimental setup is given elsewhere [8]. The reactor had an internal diameter of 10 mm and a length of 240 mm. The catalyst, diluted with silicon carbide (SiC), was loaded into the reactor. This was to ensure that problems of hydrodynamics, poor catalyst wetting, wall effects, and back mixing were mitigated [3]. For loading the catalyst, the bottom end of the reactor was sealed with a Swagelok 60 micron stainless lter (Solon, OH, USA) and then packed from bottom to top in three parts. The bottom part was loaded with 22 mm of glass beads of size 3 mm diameter followed by 25 mm, 10 mm, and 10 mm of 16 mesh, 46 mesh, and 60 mesh SiC, respectively. In the middle part of the reactor, 5 mL of catalyst and 12 mL of 90 mesh SiC were loaded alternately; small quantity of each at a time, for a total number of 1012 layers. Finally, the top part was loaded with 8 mm of SiC of 60 mesh followed by 8 mm, 8 mm, and 20 mm of 46 mesh SiC, 16 mesh SiC, and 3 mm diameter glass beads, respectively. The top 20 mm of the reactor was kept empty. The sulding solution was made of 2.9 vol.% butanethiol, a commonly used sulding agent [10], in electrical insulating oil (VOLTESSO 35). The catalyst was introduced into the reactor in an oxide form, but is required to be activated by converting into a sulded form (hence the sulding agent); the electrical insulating oil acts as a carrier liquid. 100 mL of the sulding solution was pumped into the reactor at a high owrate (2.5 mL/min) to wet the catalyst. Subsequently, the owrate was adjusted to 5 mL/h and maintained. Gas/oil ratio was operated at 600 mL/mL. The catalyst bed was initially heated to 100 C and gradually increased from 100 to 193 C. The reactor temperature was then maintained for 24 h. Next, the temperature was further increased to 343 C in steps and the reactor was kept at this temperature for another 24 h. The catalyst was equilibrated (precoked) for seven days after catalyst suldation by owing HGO into the reactor at the rate of

5 mL/h. The temperature of the reactor was increased to 375 C. The purpose of catalyst precoking was to stabilize its activity to ensure uniform activity across the catalyst surface before the experiments were conducted [11]. Liquid products were collected every 24 h, stripped with nitrogen gas to remove dissolved NH3 and H2S, and analyzed for sulfur and total nitrogen contents. After the catalyst stabilization, the experiments were carried out as designed. Each experiment was run for three days, and product withdrawn every 24 h. A transient period of 24 h was allowed after a change in process conditions and samples taken within this period were discarded. Products collected in the second and third days were analyzed for sulfur, nitrogen, and aromatics conversions. To detect if there were any changes in the catalyst activity during a run, an experiment at the precoking condition (100% H2 purity, temperature of 375 C, pressure of 9 MPa, gas/oil of 600 mL/mL, and LHSV of 1 h1) was intermittently repeated before and after each experiment. It was assumed that the catalyst did not undergo deactivation if the hydrotreating conversions, at the abovementioned precoking condition, before and after an experiment were not signicantly different. Sulfur contents of the feed and the liquid products were determined using a combustion/uorescence technique according to ASTM 5463 procedure. Total nitrogen contents of the feed and the liquid products were measured using a combustion/chemiluminescence technique following ASTM D4629 procedure. Aromaticity, dened as the mole percent of carbon in a sample that is present as part of an aromatic ring structure, of the feed and the liquid products was determined by the carbon-13 nuclear magnetic resonance (13C NMR) spectroscopy. The spectra were obtained in the Fourier Transform mode operating at a frequency of 500 MHz. The instrumental conditions were as follows: a pulse delay of 2 seconds, a sweep width of 27.7 kHz and gated decoupling. Overall time for each sample was 56 minutes for 1056 scans. Deuterated chloroform, CDCl3, was used as a solvent. The HDS and HDN conversions were calculated as follows:

%conversion of speciesi

speciesi in feed speciesi in product speciesi in feed 100 1

Table 1 Properties of heavy gas oil feedstock. Physical properties Boiling range, C Density at 20 C, g/cm3 Sulfur content, ppm Nitrogen content, ppm Aromatics content, wt.% Simulated distillation Fraction Gasoline Kerosene Light gas oil Heavy gas oil Vacuum gas oil Boiling range (C) IBP to 205 205260 260315 315425 425600

where species(i) is sulfur, nitrogen, or aromatics. The concentrations of sulfur and nitrogen were measured by an S/N Analyzer; and the results directly substituted in Eq. (1). For HDA, the results were obtained from 13C NMR spectra. The spectrum is composed of two distinct zones separated by the solvent bar. Total saturated hydrocarbons were located between 0 and 50 ppm; whereas the total aromatics were observed between 120 and 150 ppm [12]. Eq. (2) was then used to determine the aromatics content (%) of each sample. The integrals of the saturated hydrocarbons zone, Isat, and the total aromatics zone, Iar, were determined using XWIN-NMR 3.5 software. The values from Eq. (2) were substituted in Eq. (1) to determine HDA conversions

C ar
258592 0.988 42,310 3156 31.4 Amount (wt.%) 0 1 4 40 55

Iar 100 Iar Isat

where Car, aromatics content; Iar, the integral of total aromatics; Isat, the integral of total saturates. 3. Results and discussion 3.1. Effects of pressure, H2 purity, and gas/oil ratio on hydrotreating conversions Often H2 purity is not included as an operating variable in hydrotreating studies even though, along with pressure and

2538

M. Mapiour et al. / Fuel 89 (2010) 25362543 Table 3 Results of test of the signicance of factors and interactions for HDS, HDN, and HDA models. Factor or interaction p-Value of factor or interaction HDS Purity Pressure Gas/oil (Purity)2 (Pressure)2 (Gas/oil)2 Purity x Pressure Purity x Gas/oil Pressure x Gas/oil Model 0.2916 0.6136 0.8961 0.6937 HDN <0.0001 <0.0001 0.0199 0.0030 <0.0001 HDA <0.0001 <0.0001 0.0005 <0.0001

gas/oil ratio, it has a signicant effect on H2 partial pressure. For this reason, H2 purity, pressure, and gas/oil ratio, were used in an experimental design using a central composite design method (available in Expert design 6.0.1) in an effort to study their effects on HDS, HDN, and HDA activities. H2 purity, pressure, and gas/oil ratio were varied within the range of 75100 vol.% (with the rest methane), 711 MPa, and 4001200 mL/mL, respectively. Temperature and LHSV were kept constant at 380 C and 1 h1, respectively. The experimental results obtained at conditions specied by the experimental design are summarized in Table 2. Analysis of the experimental results was carried out using DESIGN-EXPERT 6.0.1 to optimize the considered operating conditions with respect to HDS, HDN, and HDA conversions. Regression analysis of experimental data generated the following regression equations for HDS, HDN and HDA:

HDS 89:77468 0:046751 Purity 0:13798 Pressure 1:77717 104 gas=oil HDN 241:68506 5:87945 Purity 4:87321 Pressure 6:70444 103 gas=oil 0:037787 Purity
2

Table 4 R-Squared statistics for the develop models of HDS, HDN and HDA. Model HDS HDN HDA R2 0.0843 0.9262 0.9125 Adjusted R2 0.0875 0.9065 0.8961 Predicted R2 0.4553 0.8570 0.8496

HDA 2:07305 0:32246 Purity 1:95801 Pressure 4:98039 103 gas=oil 5


pendent of the values of the other two factors. R2 values of HDN and HDA developed models were 0.9262 and 0.9125, respectively. Eqs. (4) and (5) do not show straightway dependence of HDN and HDA conversions on the operating variables. Therefore, to clearly illustrate the dependence, surface response plots were developed, and are presented in Figs. 1 and 2 for HDN and HDA, respectively. These gures show that increasing pressure, H2 purity, and gas oil ratio led to increases in HDN and HDA conversions. As interpreting the 3-D surface response can be difcult, the perturbation plots of the effects of the variables on HDS (no surface response of HDS is shown), HDN and HDA are provided in Figs. 35. When interpreting a perturbation plot, one needs to be cautious since it looks only at one-dimensional paths through a multifactor surface. Therefore, it is recommended that perturbation plots are used in conjunction with the 3-D surface responses. Nonetheless, it is a powerful method of comparing the relative inuences of factors [14]. A perturbation plot shows the effect of each individual variable as the others are held constant [14]. Fig. 3 shows that effects of pressure, H2 purity, and gas/oil ratio on HDS are not considerably signicant. Nonetheless, Fig. 3 does show that the effects of pressure and H2 purity on HDS are slightly greater than that of the gas/oil ratio. Figs. 4 and 5 show that the effects of the pressure, H2 purity, and gas/oil ratio on HDN and HDA are noticeably significant, and that increasing these variables led to increases in HDN and HDA conversions. Moreover, these gures show that the effects of pressure and H2 purity on HDN and HDA are more signicant than that of the gas/oil ratio. In Figs. 4 and 5, it can also be observed that effects of the variables are greater on HDN than on HDA. This correlatively implies that the effect of H2 partial pressure is far greater on HDN than on HDA. However, by considering the mechanisms of HDN and HDA, one may expect the contrary. HDA reaction proceeds through hydrogenation, whereas HDN reaction proceeds through hydrogenation followed by hydrogenolysis, and hydrogenolysis is not affected by H2 partial pressure [1]. One explanation may be that at an operating temperature of 380 C HDA activity is close to optimum due to equilibrium thermodynamic limitation [1,15]. Thus, the effects of the other variables are not as signicant as they would have been at lower operating temperatures. A second explanation may be the fact that the overall rate of HDN is frequently

where HDS, HDN, HDA are in percentage conversions (%); purity, pressure, gas/oil are in vol.%, MPa, and mL/mL, respectively. The equations are valid within the operating conditions studied. Tables 3 and 4 contain the test of signicance (f-test) and R2 test results, respectively, of HDS, HDN and HDA models. Table 3 shows that the operating variables have no signicant effects on HDS, i.e. p-value > 0.05 [13]. Moreover, Table 4 shows that R2 value for HDS model is very small, 0.0843, which means that HDS developed model poorly represents the experimental data. The f-test of HDN and HDA data shows that pressure, H2 purity, and gas/oil ratio have signicant effects on HDN and HDA activities. Moreover, it shows that these three factors do not interact as they affect HDN and HDA activities, meaning that the effect of each factor is inde-

Table 2 Experimental results at the conditions specied by the CCD experimental design. Run Experimental conditions # Pressure Gas/oil Purity (MPa) (mL/ (%) mL) Response HDN (%) HDS (%) HDA (%) Inlet H2 partial pressure (MPa) 6.7 6.2 8.1 6.2 8.1 7.8 7.8 6.1 7.8 7.8 9.6 7.8 7.8 7.8 7.8 7.3 7.3 9.5 9.6 8.9 Outlet H2 partial pressure (MPa) 5.3 5.4 5.8 4.6 7 5.2 6.9 5.6 6.8 6.8 8.3 6.8 6.8 6.8 6.8 6.8 6.2 8.1 8.9 8.2

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

9 7.8 10.2 7.8 10.2 9 9 7 9 9 11 9 9 9 9 7.8 7.8 10.2 10.2 9

800 1038 562 562 1038 400 1200 800 800 800 800 800 800 800 800 1038 562 562 1038 800

75 80 80 80 80 88 88 88 88 88 88 88 88 88 88 95 95 95 95 100

63.2 59.1 67.4 55.1 67.6 59.9 65 54.4 66.7 65.7 72.8 68.2 66.6 69.6 66.3 69.3 65.7 78.5 83.9 78.8

95.8 95.4 93.6 94.9 93.2 96 95.3 94.5 95.5 94 95 95.6 92.9 95.8 95 93.8 94 96 96.7 96.6

48.9 48.1 51.2 46 52.6 47.9 53 46.1 51.8 52.4 54.9 52.4 53.1 52.6 51.8 53.9 52.3 55.5 58 55.4

M. Mapiour et al. / Fuel 89 (2010) 25362543

2539

79.09 73.47 67.85 62.22 56.60

56.62 54.25 51.89 49.53 47.16

HDN, %

HDA, %

10.19 9.59 9.00 Pressure , MPa 8.41 7.81 80.07 87.50 83.78 91.22

94.93

10.19 9.59 9.00 Pressure , MPa 8.41 7.81 80.07 87.50 83.78 91.22

94.93

H2 purity, vol.%

H2 purity, vol. %

74.89 71.37 67.85 64.32 60.80

55.47 53.68 51.89 50.10 48.31

HDN, %

94.93 91.22 87.50 H2 purity, vol. % 83.78 80.07

1037.84 918.92 800.00 681.08 Gas/oil, mL/mL 562.16

94.93 91.22 87.50 H2 purity, vol.% 83.78 80.07

HDA, %

1037.84 918.92 800.00 681.08 Gas/oil, mL/mL 562.16

Fig. 1. Surface response of the effects of pressure, H2 purity, and gas/oil ratio on HDN activity.

Fig. 2. Surface response of the effects of pressure, H2 purity, and gas/oil ratio on HDA activity.

determined by the hydrogenation rate rather than by hydrogenolysis [1], and hydrogenolysis is not affected by H2 partial pressure. Thus, pressure, H2 purity, and gas/oil ratio can only affect hydrogenation processes in HDN and HDA reactions. Since, hydrogenation of aromatic rings with heteroatoms is easier than of those which lack a heteroatom, HDN is more affected than HDA. A possible and plausible third explanation is that the initial concentration of aromatics is about 100 folds that of the nitrogen concentration in the feed, indicating that no fair comparison can be made by looking at HDN and HDA conversions. Difference among the HDS, HDN, and HDA mechanism may offer an explanation as to why there are dissimilarities in the effects of the variables on HDN and HDA versus those on HDS. As previously mentioned in this section, HDN reaction takes place via hydrogenation followed by hydrogenolysis, and HDA reaction occurs via hydrogenation. Increasing H2 purity, pressure, and gas/ oil ratio result in increases in H2 partial pressure. This increase in H2 partial pressure directly affects the hydrogenation process. As a result, changes in HDA and HDN conversions were observed as H2 purity, pressure, and gas/oil ratio were varied. On the other hand, HDS reaction can proceed via two pathways: (1) hydrogenation followed by hydrogenolysis or (2) direct hydrogenolysis [4,16]. Consequently, HDS conversions were only very slightly affected by H2 partial pressure since HDS reaction has the option of taking place directly via hydrogenolysis. Consequently, no signicant effects of pressure, H2 purity, and gas/oil ratio on HDS conversions were observed. The rate of the HDS two pathways depends on a number of factors: catalysts properties, feeds type, and reaction conditions. It was suggested that the hydrogenation

96.7

95.675 HDS conversion A 94.65 C B A B C

93.625

92.6 -1.000 -0.500 0.000 0.500 1.000 Deviation from Reference Point

Fig. 3. HDS perturbation plot: (A) is H2 purity, (B) is pressure, and (C) is gas/oil ratio.

step of the HDS took place at the S-edge while the SC scission step occurred at both S-edge and Mo-edge; where S-edge and Mo-edge are active sites. At lower H2 partial pressure, S-edge is inactive and the catalytic reactions therefore occur at Mo-edge; hence, SC is favored [17]. The optimal operating conditions were calculated based on constraints in which HDS, HDN, and HDA conversions were to be

2540

M. Mapiour et al. / Fuel 89 (2010) 25362543

3.2. Effects of temperature and LHSV on hydrotreating conversions

83.9
Effects of temperature and LHSV on the hydrotreating conversions were also studied. Temperature and LHSV ranges were 360400 C and 0.652 h1, respectively. H2 purity, pressure and gas/oil ratio were kept constant at 100%, 9 MPa, and 800 mL/mL, respectively. LHSV was kept constant at 1 h1 when the effect of temperature was studied. Temperature was kept constant at 380 C when the effect of LHSV was studied. LHSV, which is the inverse of residence time, is an indication of the time spent in the reactor by the reactants [4]. It was observed that decreasing LHSV led to increases in HDS, HDN, and HDA conversions (see Fig. 6). However, one needs to bear in mind that for HDA this observation is only true for the conditions employed in this work, especially temperature. For example, Mann et al. [18] found that HDA is independent of LHSV (between 0.5 and 4 h1) at the temperature of 450 C and pressure of 6.99 MPa. The reason is that HDA maximum conversion is achieved between 370 and 400 C (usually 375385 C) due to the interrelation between thermodynamic equilibrium and reaction rates [15]. The authors observed similar results for HDS and HDN to those found in this work. Increasing the temperature generally leads to increases in hydrotreating conversions. Nevertheless, excessive temperature may impose thermodynamic equilibrium limitations leading to decreases in hydrotreating conversions. In the cases of HDS and HDN, this hindering effect of temperature is observed at temperatures higher than those used in practice (i.e. >425 C) [1]. In the case of HDA, the hindering effect of temperature is observed at lower temperatures than that of HDS and HDN. As mentioned earlier in this section, the maximum HDA conversion usually occurs at temperature range of 370385 C [15]; this range could be little higher if the H2 partial pressure is substantially increased. In this work it was found that both HDS and HDN conversions increase with increasing temperature, however, HDN shows superior increases than that of HDS (see Fig. 7). This superior effect of temperature on HDN could not be because HDN has higher reaction rate than HDS. The bond energy of C@N {147 kcal/mol} is higher than that of C@S {114128 kcal/mol} (i.e. CN scission in more difcult than that of CS). Moreover, N (0.75 ) has smaller atomic radius than S (1.09 ), therefore is more difcult to remove N than S [19]. The radius effect may be explained as follow: the surface reaction steps can be simplied as follow: (i) adsorption onto the active site, (ii) reaction, and (iii) desorption. The adsorption is via lone pairs or election clouds. For the nitrogen atom

HDN conversion

76.525 A B 69.15 C A B C

61.775

54.4 -1.000 -0.500 0.000 0.500 Deviation from Reference Point 1.000

Fig. 4. HDN perturbation plot: (A) is H2 purity, (B) is pressure, and (C) is gas/oil ratio.

58

55 HDA conversion

BA C

52
C A B

49

46 -1.000 -0.500 0.000 0.500 Deviation from Reference Point 1.000

Fig. 5. HDA perturbation plot: (A) is H2 purity, (B) is pressure, and (C) is gas/oil ratio.

Conversions, %

70 60 50 40 30 20 0 0.5 1 1.5
-1

56.0

52.0

Table 5 Comparison between the predicted and observed values of hydrotreating at optimal conditions: pressure of 10.1 MPa, H2 purity of 95 vol.%, and gas/oil ratio of 1037 mL/ mL. Temperature and LHSV were 380 C and 1 h1, respectively. Reactions HDS HDN HDA Predicted by models (%) 95.4 80.1 57.8 Observed experimentally (%) 96.1 82.0 57.9 Percentage differences (%) 0.7 2.3 0.2

48.0 2 2.5 LHSV, h

Fig. 6. Effect of LHSV on hydrotreating conversions. Pressure, temperature, H2 purity, and gas/oil ratio were 9 MPa, 380 C, 100%, and 800 mL/mL, respectively.

Conversions, %

maximal within the ranges of the operating variables studied. The collective optimum operating conditions for HDS, HDN, and HDA were determined to be: pressure of 10.2 MPa, H2 purity of 95 vol.%, and gas/oil ratio of 1037 mL/mL. Experiments were conducted under these conditions, and the experimental data was compared to those predicted (see Table 5). As shown in Table 5, the percentage differences of HDS, HDN, and HDA for the experimental results versus the predicted results were 0.7%, 2.3%, and 0.2%, respectively. It may be noted that the results for HDS may not be reliable since the R2 of the developed model was only 0.0843.

HDS 100 90 80

HDN

HDA 60.0

M. Mapiour et al. / Fuel 89 (2010) 25362543

2541

HDS

HDN

HDA

100 90 80 70 60 50 40 30 20 350

360

370 380 390 Temperature ,C

400

57 56 55 54 53 52 51 50 49 48 410

Fig. 7. Effect of temperature on hydrotreating conversions. Pressure, LHSV, H2 purity, and gas/oil ratio were 9 MPa, 1 h1, 100%, and 800 mL/mL, respectively.

the radius is smaller thus valence electrons have higher interaction with the nucleus; consequently, nitrogen has lesser adsorption strength in comparison to sulfur. Thus in theory, HDS should be more signicantly affected by temperature than HDN; however, this is not the case. An explanation may be that, at the temperature range under study, the effect of temperature on HDS starts to subside, while the effect of temperature on HDN starts to become more pronounced. In a study by Mann et al. [18] using NiMo/c-Al2O3 as a catalyst and HGO as a feed, it was found that in the temperature range of 300350 C, HDS and HDN percentage conversions per C (degree Celsius) were 0.25% and 0.10%/C, respectively. However, in the temperature range of 350400 C HDN has a higher percentage conversion per C (0.36%/C) than HDS (0.22%/C). Fig. 7 also shows that HDA conversion passes through a maximum with respect to the temperature, which is in agreement with the literature. By taking the rst derivative the empirical equation (see Eq. (6)), maximum HDA conversion was determined to have occurred at 385 C. From the foregoing discussion it seems that temperature is most critical of all of the variables.

the catalysts deactivation standpoint, however, the authors did not support this suggestion with experimental evidences. A possible way to validate this suggestion is to rst determine the parameters in a kinetic expression, containing H2 partial pressure as a variable, using both inlet and outlet H2 partial pressure. Next, the two resulting kinetic expressions, one with inlet H2 partial pressure and another with outlet H2 partial pressure, can then be compared through statistical analysis to determine which has a better prediction ability with higher R2 and adjusted R2. Finally, predicted data from these two expressions can be compared against experimental data generated at conditions different than those used in the models development. In an attempt to evaluate the above suggestion, experiments were conducted under the following conditions: temperature range of 360400 C, pressure range of 711, LHSV range of 0.652 h1, and gas/oil range of 4001200 mL/mL. Inlet and outlet H2 partial pressures were determined using HYSYS and their values in Table 2. The kinetic parameters in Eq. (7) were then determined using the inlet and outlet H2 partial pressure data along with the rest of the variables. Eq. (7) is the power law model [16], and Eqs. (7.b) and (7.c) are the solutions for Eq. (7) depending on the value of n (reaction order)

HDS and HDN Conversions, wt.%

HDA Conversion, wt.%

dC n ki P m H C dt

7 7:a

where ki T ko eE=RT
For Eq. (1), the solution is:

ln

Cf ko eE=RT Pm H ; Cp LHSVC

n1

7:b

" # 1 1 1 ko eE=RT Pm H n1 ; n 1 n 1 Cp Cf LHSVC

n>1

7:c

HDA 0:0089 Temperature2 6:8447 Temperature 1258 6

3.3. Kinetics of HDS, HDN, and HDA The effects of the operating variables on hydrotreating performance can be predicted by a suitable kinetic expression or model [16,20]. H2 partial pressure often appears as a variable in many hydrotreating kinetic expressions but one must decide whether to use inlet or outlet H2 partial pressure in the calculations. McCulloch and Roeder [21] suggested that more meaningful results are attained when outlet H2 partial pressure is used especially from

where Cf and Cp, initial and nal concentrations of S, N, or aromatics, respectively, wt.%. LHSV, liquid hourly space velocity, h1; R, gas constant, 8.314; T, temperature, K; ki, apparent rate constant; ko, pre-exponential factor; PH, inlet or outlet H2 partial pressure; n, reaction order; E, activation energy; m and c, empirical regression factors. The experimental data was analyzed using non-linear regression model in Polymath software [20]. The parameters for HDS, HDN and HDA are presented in Table 6. HDS, HDN, and HDA were assumed to be pseudo-rst order, i.e. n = 1 [1], and consequently Eq. (7.b) was used in the determination of the kinetic parameters. The R2 and adjusted R2, presented in Table 6, suggest that for HDS and HDN kinetic interpretation of the data using either inlet or outlet H2 partial pressure will generate models with the same predictive ability. However for HDA, kinetic interpretation of data using outlet H2 partial pressure generates models with better predictive ability than using inlet H2 partial pressure. For example, for the

Table 6 Power law model kinetic parameters for HDS, HDN, and HDA (temperature range is: 360400 C). Parameter Using inlet H2pp HDS ko E m c R2 Adjusted R2 171 24.3 0.21 0.58 0.74 0.71 HDN 2564 56.8 1.33 1.16 0.90 0.88 HDAa 37.2 27.7 0.57 0.21 0.82 0.79 HDAb 0.0126 56.5 0.58 0.22 0.81 0.74 Using outlet H2pp HDS 164 23.3 0.16 0.59 0.74 0.71 HDN 11012 60.1 0.99 1.21 0.90 0.89 HDAa 150 33.6 0.46 0.24 0.90 0.89 HDAb 0.0075 20.1 0.46 0.24 0.90 0.88

E = kJ/mol. a Temperature 6 380 C. b Temperature P 380 C.

2542

M. Mapiour et al. / Fuel 89 (2010) 25362543 Table 8 Results of the catalyst deactivation testing (temperature, gas/oil ratio, LHSV were 380 C, 800 mL/mL, and 1 h1, respectively). Activity For H2 partial pressure of 8.1 MPa Before HDN (%) HDA (%) 75 53 After 75 54 For H2 partial pressure of 4.5 MPa Before 74 53 After 68 50

temperature range of 360380 C, the R2 and adjusted R2 for the model generated using inlet H2 partial pressure are 0.82 and 0.79, respectively, whereas for that generated using outlet H2 partial pressure they were 0.90 and 0.89, respectively. After models were developed new experimental data were used to validate the models. Three experiments were conducted in which pressure, temperature, LHSV, gas/oil ratio were kept constant at 9 MPa, 380 C, 1 h1, and 800 mL/mL, respectively, and H2 purity was varied as follows: 50, 80, and 90 vol.% (with the rest methane). None of these chosen conditions were used in the development of the kinetic models. Also, note that H2 purity of 50 vol.% is an extrapolated condition, i.e. falls outside the range of the conditions originally used to develop the models. The experimental results from these experiments were then compared against those predicted by the models and the results are presented in Table 7. Table 7 shows that the predictions from the developed kinetic models for HDS and HDN were in close agreement with the experimental values regardless of whether inlet or outlet H2 partial pressure was used. For HDA, the kinetic model developed using outlet H2 partial pressure produced ever-so-slightly better predictions than those produced using inlet H2 partial pressure. Thus, it seems that it does not matter whether hydrotreating conclusions were drawn from inlet or outlet H2 partial pressure. However, these conclusions do not address catalyst deactivation. Catalyst deactivation is the most important concern in any catalytic process. In hydrotreating it is well known that increasing H2 partial pressure results in a decrease in deactivation rate. Due to H2 consumption, the reactors outlet H2 partial pressure can be considerably lower than its inlet H2 partial pressure, especially for heavy feedstock (see Table 2). Therefore the portion of the catalyst bed at and near the reactor outlet may to experience higher deactivation rate as a results of lower H2 partial pressure environment. Botchwey et al. [22] and Alvarez and Ancheyta [23] have shown experimentally that the largest portion of hydrotreating conversions take place in the rst 30% of the catalyst beds length. Consequently this is also where most of the hydrogen consumption takes place, leaving a large portion of the catalsyts bed at a H2 partial pressure level signicantly lower than that at the inlet. Another noteworthy point that can be deduced from Botchwey et al. [22] and Alvarez and Ancheyta [23] ndings is that for about 70% of the catalyst beds length the H2 partial pressure level is closer in value to the outlet H2 partial pressure than it is to the inlet H2 partial pressure. In other words, outlet H2 partial pressure level resembles that experienced by most parts of the catalysts bed. Therefore, it is very critical that the outlet H2 partial pressure is determined so more complete and meaningful conclusions are drawn. To show the relationship between the catalyst deactivation and H2 partial pressure levels, a commercial NiMo/c-Al2O3 catalyst bed was subjected to two inlet H2 partial pressures of 4.5 MPa and 8.1 MPa for a period of three days while hydrotreating experiments were being carried out. The temperature and LHSV were 380 C and 1 h1, respectively for all experiments. Before the catalyst bed was subjected to either of the two H2 partial pressure levels
Table 7 Comparison of the prediction results using inlet versus outlet H2 partial pressure (temperature, pressure, gas/oil ratio, LHSV were 380 C, 9 MPa, 800 mL/mL, and 1 h1, respectively). Purity (vol.%) Determined experimentally (%) Determined from the models (%) Using inlet H2 partial pressure HDS 93 95 95 HDN 42 64 69 HDA 41 50 53 Using outlet H2 partial pressure HDS 93 95 95 HDN 38 63 70 HDA 39 50 53

a designated experiment with inlet H2 partial pressure of 9 MPa, named control, was conducted. The same control experiment was then repeated after the catalyst bed had been subjected to inlet H2 partial pressure of 4.5 MPa or 8.1 MPa. The hypothesis was that if the before and the after hydrotreating conversions of the control experiment were different it can be concluded that the catalyst underwent some deactivation. The results are presented in Table 8. Table 8 shows that for the experiement at 4.5 MPa inlet H2 partial pressure, the after HDN and HDA conversions of the control are lower than those of the before conversions, meaning that the catalyst suffered deactivation due to low H2 partial pressure levels. No signicant differences were observed between the after and before HDN and HDA conversions in the experiment conducted at H2 partial pressure of 8.1 MPa. Thus, it can be seen that lower H2 partial pressure may cause severe catalyst deactivation in a very short time. It is therefore important that the entire catalyst bed is maintained at a high enough H2 partial pressure to avoid deactivation. One way of ensuring that the entire catalyst bed is at sufcient H2 partial pressure level is to maintain high outlet H2 partial pressure, the reactor point with the lowest H2 partial pressure, to avoid untimely catalyst deactivation. 4. Conclusion Within the range of the experimental conditions considered in this study the following conclusions were made: Increasing pressure, H2 purity, and gas/oil ratio led to increases in HDN and HDA activities. Effects of these variables on HDS activity are not considerably signicant. The positive effects of H2 purity on HDN and HDA activities were greater than those of gas/oil ratio and comparable to those of pressure. The optimal conditions for HDS, HDN, and HDA are: pressure of 10.1 MPa, H2 purity of 95 vol.%, and gas/oil ratio of 1037 mL/mL. This is achieved at temperature and LHSV of 380 C and 1 h1, respectively. Decreasing LHSV led to increases in HDS, HDN, and HDA activities, while increasing temperature led to increases in HDS and HDN. HDA passed through a maximum of 380 C as the temperature was varied. A low H2 partial pressure environment can accelerate the catalysts deactivation Information on H2 partial pressure effects on hydrotreating activities can be equally satisfactorily obtained using either inlet or outlet H2 partial pressure. However, from the catalyst deactivation standpoint it is vital to use outlet H2 partial pressure, since it is the reactor point with the lowest H2 partial pressure.

HDS 50 80 90 92 95 96

HDN 38 64 71

HDA 38 48 49

References
[1] Girgis Michael J, Gates Bruce C. Reactivities, reaction networks, and kinetics in high-pressure catalytic hydroprocessing. Ind Eng Chem Res 1991;30:20212.

M. Mapiour et al. / Fuel 89 (2010) 25362543 [2] Speight JG. The desulfurization of heavy oils and residua. New York: Marcel Dekker Inc.; 1981. [3] Bej K, Ajay Dalai, John Adjaye. Comparison of hydrodenitrogenation of basic nonbasic nitrogen compounds present in oil sands derived heavy gas oil. Energ Fuel 2001;15:377. [4] Christian Botchwey, Ajay Dalai, John Adjaye. Product selectivity during hydrotreating and mild hydrocracking of bitumen-derived gas oil. Energ Fuel 2003;17:1372. [5] Adrian Gruia. In: David SJ Jones, Peter R Pujad, editors, Handbook of petroleum processing. The Netherlands: Springer; 2006 [Chapter 8]. [6] Peramanu S, Cox BG, Pruden BB. Economics of hydrogen recovery processes for the purication of hydroprocessor purge and off gases. Int J Hydrogen Energ 1999;24:405. [7] Turner J, Reisdorf M. Consider revamping hydrotreaters to handle higher H2 partial pressures. Hydrocarbon Process 2004;March:6170. [8] Mapiour M, Sundaramurthy V, Dalai AK, Adjaye J. Effect of hydrogen purity on hydroprocessing of heavy gas oil derived from oil-sands bitumen. Energ Fuel 2009;23(4):212935. [9] Christian Botchwey, Ajay Dalai, John Adjaye. Two-stage hydrotreating of Athabasca heavy gas oil with interstage hydrogen sulde removal: effect of process conditions and kinetic analyses. Ind Eng Chem Res 2004;43(18): 5854. [10] Andari Mounif K, Abu-Seedo Fatima, Stanislaus Anthony, Qabazard Hassan M. Kinetics of individual sulfur compounds in deep hydrodesulfurization of Kuwait diesel oil. Fuel 1996;75(14):166470. [11] Speight JG. The desulfurization of heavy oils and residua. New York: Marcel Dekker Inc.; 2000. [12] Owusu-Boakye, Abena. Two-stage aromatics hydrogenation of bitumenderived light gas oil. Master Thesis, University of Saskatchewan; 2005.

2543

[13] Montgomery DC. Design and analysis of experiments, 4th ed. USA: John Wiley & Sons; 1997. [14] Mark J Anderson, Patrick J Whitcomb. RSM simplied: optimizing processes using response surface methods for design of experiments. New York, NY: Productivity Press; 2005. [15] James H Gary, Glenn E Handwerk, Mark J Kaiser. Petroleum rening: technology and economics, 5th ed. New York, NY: CRC Press; 2007. [16] Knudsen KG, Cooper BH, Topsoe H. Catalyst and process technologies for ultra low sulfur diesel. Appl Catal 1999;189:205. [17] Poul Georg Moses, Berit Hinnemann, Henrik Topse, Jens K Nrskov, The hydrogenation and direct desulfurization reaction pathway in thiophene hydrodesulfurization over MoS2 catalysts at realistic conditions: a density functional study. J Catal 2007;248:188203. [18] Mann RS, Sambi IS, Khulbe K. Catalytic hydroning of heavy gas oil. Ind Eng Chem Res 1987;26:410. [19] Landau MV. Deep desulfurization of diesel fuels: kinetic modeling of model compounds in trickle-bed. Catal Today 1997;36:393. [20] Deena Ferdous, Ajay Dalai, John Adjaye. Hydrodenitrogenation and hydrodesulfurization of heavy gas oil using NiMo/Al2O3 catalyst containing boron: experimental and kinetic studies. Ind Eng Chem Res 2006;45:544552. [21] Donald C McCulloch, Roeder RA. Find hydrogen partial pressure. Hydrocarbon Process, 1976;February:81. [22] Christian Botchwey, Ajay Dalai, John Adjaye. Simulation of a two-stage micro trickle-bed hydrotreating reactor using Athabasca bitumen-derived heavy gas catalyst: effect of H2S on oil over commercial NiMo/Al2O3 hydrodesulfurization and hydrodenitrogenation. Int J Chem React Eng 2006;4:A20. [23] Anton Alvarez, Jorge Ancheyta. Modeling residue hydroprocessing in a multixed-bed reactor system. Appl Catal A: General 2008;351(2):148.

You might also like