Hydrogenation of Olefins in Bitumen-Derived Naphtha over a

gases. Gases from the high-pressure separator and stabilizer column were analyzed by online gas chromatography. The reactor was packed with 10 mL of a...
1 downloads 9 Views 696KB Size
Subscriber access provided by Access provided by University of Liverpool Library

Catalysis and Kinetics

Hydrogenation of Olefins in Bitumen-Derived Naphtha over a Commercial Hydrotreating Catalyst Qin Xin, Anton Alvarez-Majmutov, Heather Dettman, and Jinwen Chen Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.8b00344 • Publication Date (Web): 10 Apr 2018 Downloaded from http://pubs.acs.org on April 10, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Hydrogenation of Olefins in Bitumen-Derived Naphtha over a Commercial Hydrotreating Catalyst Qin Xin, Anton Alvarez-Majmutov, Heather D. Dettman and Jinwen Chen* Natural Resources Canada, CanmetENERGY in Devon 1 Oil Patch Drive, Devon, Alberta, Canada T9G 1A8

Abstract Instability associated with the presence of olefins in bitumen that is thermally processed during partial upgrading is a major concern for pipeline transportation and downstream refining. A common strategy for stabilizing thermally processed oils is to selectively hydrogenate the olefinrich fractions, typically the naphtha fraction (IBP–204°C). In this paper, olefin hydrogenation was studied with hydrotreated bitumen-derived naphtha spiked with five model olefin compounds under mild hydrotreating conditions. The hydrogenation reactivities of the five model olefin/diolefin compounds are ranked in the order 1,3-hexadiene > allylbenzene > 1-heptene > 2methyl-2-pentene > 1-methyl-cyclopentene. The reactivity is largely determined by the position of the double bond, and to a lesser extent by the molecular structure of the olefin. The conjugated diolefin, 1,3-hexadiene, was the most reactive. The two terminal olefins, 1-heptene and allylbenzene, were observed to be more reactive than the two olefins with internal double bonds, 2-methyl-2-pentene and 1-methyl-cyclopentene. Results also show that temperature has a significant effect on olefin hydrogenation performance, with pressure and liquid hourly space velocity having relatively moderate effects. Meanwhile, flash calculations confirmed the presence of vapor-liquid equilibrium under the operation conditions used. When the reactor temperature is 150°C or less reactions take place mostly in the liquid phase, whereas when it is 200°C or higher the reactions take place in the vapor phase. A hydrogenation kinetics model is proposed that successfully describes the observed trends of olefin hydrogenation in the liquid phase. KEYWORDS: Hydrogenation; olefin; bitumen; naphtha; partial upgrading 1 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 34

1. Introduction Upgrading is a commercial processing pathway in Canadian oil sands development that converts bitumen extracted from oil sands into high-quality synthetic crude oil (SCO). SCO is either delivered to refineries for fuel production or used as diluent to transport bitumen by pipeline. However, bitumen upgrading is costly and requires a large energy input. In recent years, the concept of partial bitumen upgrading has emerged as a lower-cost alternative for processing bitumen while reducing environmental impacts and greenhouse gas (GHG) emissions. The goal of partial upgrading is to convert bitumen into a product that meets pipeline specifications (API gravity 19° and viscosity 350 cSt at 7°C), ideally without the need for dilution. Most partial upgrading technologies currently under development are based on some form of thermal cracking [1], which inevitably leads to the formation of olefins that can make the liquid product unstable and problematic. The presence of olefins in partially upgraded bitumen in quantities exceeding the permitted level for pipeline transportation (currently set at 1.0 wt%1-decene equivalent [2]) represents a technical concern due to potential fouling associated with olefins. Olefinic compounds tend to concentrate in the light fractions ( allylbenzene > 1-heptene > 2-methyl-2-pentene > 1-methylcyclopentene. An important observation is that the olefins with a terminal double bond (1,3hexadiene, allylbenzene, and 1-heptene) are more reactive that those having an internal double bond (2-methyl-2-pentene and 1-methyl-cyclopentene) over the tested range of operating conditions. 1,3-hexadiene is considered to fall in the terminal olefin category in this case because one of its two double bonds is on a terminal carbon. However, the mono-olefin product (3-hexene) from the first hydrogenation step of 1,3-hexadiene would be an internal olefin. Differences in reactivity within the terminal and internal olefin categories can be attributed to molecular structure. The terminal olefin, allylbenzene, is presumed to be more reactive than 1-heptene due to the preferential adsorption of the benzene ring group on the active sites of the catalyst particles. Comparing the two internal olefins, 1-methyl-cyclopentene is less reactive than 2-methyl-2pentene, most likely due to steric hindrance effects. Thus, olefin hydrogenation reactivity is largely determined by the position of the double bond, and to a lesser extent by the molecular structure of the olefins. These observations are consistent with other studies on hydrogenation of olefins [13, 15-16]. Chemical hydrogen consumption was estimated using a program developed in-house based on global mass balance for carbon and hydrogen in the gas and liquid phases. Hydrogen consumption is observed to increase within a relatively small range (4 and 27 NL/L) with respect to total olefin conversion, as presented in Figure 6. A more pronounced trend would have been observed using an actual cracked distillate fraction sustaining other reactions such as aromatics hydrogenation and hydrodesulfurization, which strongly contribute to hydrogen consumption. It is noted from Figure 6 that the calculated values of hydrogen consumption exhibit a certain degree of dispersion owing to measurement errors affecting carbon and hydrogen mass balance closure. The two points (25.5 and 26.3 NL/L) off the main trend are the most affected ones by such errors in the calculation method.

7 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 34

3.2. Vapor-Liquid Equilibrium Petroleum feedstocks have a tendency to vaporize when subjected to hydrotreating conditions. This is an important aspect to consider in analyzing hydrotreating reactions, as vaporization affects reaction rates by changing fluid flow rates and phase compositions in the reactor. Naphtha, the lightest petroleum distillate fraction, is particularly volatile at hydrotreating conditions. Hydrocarbons in this boiling range are completely vaporized at typical operating conditions for naphtha hydrotreating (250–360°C and 1–3 MPa). Therefore, naphtha hydrotreating reactions primarily occur in the vapor phase, as opposed to those of middle and heavy distillates fractions, which normally occur in the liquid phase, although liquid and vapor may co-exist. The range of temperatures (100–300°C) used in this study, however, is lower than that in conventional naphtha hydrotreating, resulting in incomplete vaporization of liquid and thus supporting the possible presence of vapor-liquid equilibrium (VLE) in the reactor. To verify this condition, flash calculations were performed using a validated in-house program developed specifically for hydrogen-petroleum systems [21-22]. The program uses the Peng-Robinson equation of state (EoS) to model the partition of hydrocarbon compounds between the vapor and liquid phases. Accordingly, the distribution of hydrocarbons and hydrogen at equilibrium is described by:  =

 (1)



where yi and xi are component mole fractions in the gas and liquid phases, respectively, and KVLEi is the VLE partition coefficient. The partition coefficients together with the phase compositions and properties are computed by the flash program with given reactor conditions (temperature, pressure, and H2/oil ratio) and feedstock composition. The distinctive characteristic of this flash model is that the binary interaction parameters between hydrogen and hydrocarbon components of the Peng-Robinson EoS were evaluated using extensive experimental VLE data sets for a variety of petroleum distillates over a wide range of conditions. Figures 7 and 8 show the predicted vaporization for the naphtha feed over the studied ranges of temperature and pressure, respectively. As seen in Figure 7, as expected, temperature has a major impact on feed vaporization. Between 100 and 150°C vaporization increases from 9 wt% to 26 wt%, while at 200°C vaporization is 63 wt%. Above 227°C the feed is completely in the vapor 8 ACS Paragon Plus Environment

Page 9 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

phase. Pressure on the other hand, has a moderate effect on vaporization, as evidenced by Figure 8. Vaporization is 20 wt% at the lowest and 37 wt% at the highest operating pressures at 150°C. The specific vaporization profiles of the model olefin compounds as function of temperature are presented in Figure 9. The least volatile compound is allylbenzene, followed by 1-heptene. 2methyl-2-pentene and 1-methyl-cyclopentene are the most volatile olefins, with both having a similar profile. It is noted that olefin vaporization does not surpass 50 wt% at 150°C. It can be concluded then that when reactor temperature is less than or equal to 150°C the olefin hydrogenation reactions take place mostly in the liquid phase, whereas at 200°C and above they occur in the vapor phase. Moreover, the present observations on the reactivities of terminal and internal olefins are in agreement with Lebedev’s rule [13]: olefin hydrogenation reactivity in the liquid phase decreases with the increase in the number of substituents on the double bond. However, this rule does not apply to gas-phase reactions as there is extensive double bond migration leading to complex olefin isomer distributions [13]. 3.3. Reaction Kinetics Hydrogenation kinetics in the liquid phase were analyzed for the four mono-olefins by using the LHSV, pressure, and temperature (≤150°C) data sets. The hydrotreating reactor is considered to behave as a pseudo-homogeneous plug-flow reactor. The catalyst particles are assumed to be fully wetted given that the reactor operates in up-flow mode [18]. Likewise, it is considered that only the reactants in the liquid phase have access to the catalyst particles [22]. With these assumptions, the continuity equation for model olefins becomes:  =   (2)  where CLi is the concentration of olefin i in the liquid phase, τ is space time, ri is the rate of hydrogenation of olefins, and ρb is the catalyst bulk density. Considering that feed vaporization reduces the liquid rate entering the reactor, space time is formulated in terms of the actual liquid rate corrected for vaporization: =

 (3)  (1 −  )

9 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 34

where Vb is the catalyst bed volume, FL is the total naphtha feed rate, and φV is the fraction of vaporized hydrocarbons at process conditions. Component concentrations in the liquid phase are given by:  = 

 (4) 

where ρL is the liquid phase density and ML is the liquid phase molecular weight. For each process condition, a flash calculation at the reactor inlet is executed to rectify the liquid rate and establish liquid phase composition and properties. The rate of olefin hydrogenation on the metal sites of the hydrotreating catalyst was expressed in the form of the following Langmuir-Hinshelwood (LH) rate equation [16, 23]:  = −

    !"#

$1 + ∑(  ' ' )

(5)

where ksi is the olefin hydrogenation rate constant, Kad is the component adsorption coefficient on metal sites, and !"# is hydrogen partial pressure. According to this equation, olefin hydrogenation reactions follow first-order kinetics and are irreversible under the studied conditions. The temperature dependence of the rate constants is represented by the Arrhenius equation. The denominator term describes competitive adsorption effects, with sub-index j referring to all the hydrocarbon components making up the feed, including olefins. Adsorption coefficients for olefins and hydrocarbon components are estimated from their molecular structure attributes using the following quantitative structure/reactivity correlation (QSRC) [23]:  = + +

,-./ 0,12 (6) 34

where a, b, c are model parameters, NAR is the number of aromatic rings, NSC is the number of saturated carbons in the molecule, R is the universal gas constant, and T is reaction temperature. Parameters a, b, c were taken from reference [23] with the assumption that they are valid for the hydrotreating catalyst used in this study. The competitive adsorption term of Equation 5 is evaluated using the calculated adsorption coefficients and the hydrocarbon component concentrations in the liquid phase, which are derived from the PIONA analysis of the base naphtha feed and the known concentrations of model olefins (Table 3). For simplicity, it is also

10 ACS Paragon Plus Environment

Page 11 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

assumed that the concentrations of naphtha components other than olefins remain constant throughout the reactor for each condition. Hydrogenation rate constants in Equation 5 were obtained by regression of the experimental data. In the model, olefin conversions at the reactor outlet are obtained considering the remaining olefins in the liquid phase together with the vaporized olefins that did not participate in the reactions. The pre-exponential factors (k0) and activation energies (Ea) for each olefin compound are listed in Table 4. The activation energies for the olefin hydrogenation reactions were found to fall between 18 and 28 kcal/mol. 1-methyl-cyclopentene had the highest activation energy (27.6 kcal/mol), meaning that it is the most sensitive to temperature. The parity plot presented in Figure 10 shows that the majority of model predictions are within the 10% error band, which confirms the adequacy of the fitted parameters. Simulated olefin conversions over the temperature range of 100–150°C are shown in Figure 11. As seen in Figure 11, the proposed kinetic model, combined with VLE calculations and QSRC for adsorption, provides a good description of the observed trends (Figure 3) of liquid-phase olefin hydrogenation. It should be noted that the range of applicability of the kinetic model is limited to liquid phase conditions and temperatures below 150°C. Additional experimentation will be needed to develop reaction kinetics in the gas phase. Also noted is that the developed model will require correction for hydrodynamic effects, such as catalyst wetting efficiency and/or liquid holdup, to simulate a commercial reactor operating in trickle-flow mode. The model can be extended to actual cracked distillate feedstocks by developing QSRC models to describe in detail the reactivity of olefin compound distributions and other hydrocarbon types.

11 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 34

Conclusions The results of this study demonstrate the feasibility of saturating model olefins under relatively mild hydrotreating conditions, and provide an improved understanding of olefin hydrogenation chemistry and reaction kinetics. The reactivities of five model olefin/diolefin compounds under such conditions are ranked in the order 1,3-hexadiene > allylbenzene > 1-heptene > 2-methyl-2pentene > 1-methyl-cyclopentene. Olefin hydrogenation reactivity is largely determined by the position of the double bond, and to a lesser extent by molecular structure. The conjugated diolefin 1,3-hexadiene is the most reactive compound, with the internal mono-olefins being the least reactive and the terminal mono-olefins being of intermediate reactivity. Of the process conditions employed in the experiments, operating temperature plays a greater role in olefin conversion than pressure or LHSV. Overall hydrogen consumption is between 4 and 27 NL/L. Unlike conventional gas-phase naphtha hydrotreating, olefin hydrogenation using a simulated cracked naphtha in the present work occurs in either liquid phase or gas phase, depending on the reaction temperature as determined by VLE flash calculations. The detailed hydrogenation kinetics of the four mono-olefins in the liquid phase were established. The kinetic model provides a good description of experimental olefin hydrogenation results over the temperature range of 100–150°C. The activation energies fall between 18 and 28 kcal/mol. The kinetic model can be extended to actual cracked distillate feedstocks through the application of QSRC models to capture in detail the reactivities of olefin compound distributions and other hydrocarbon types.

AUTHOR INFORMATION Corresponding Author Tel.: (+1) 780-987-8763; Fax: (+1) 780-987-5349; E-mail address: [email protected]

Notes The authors declare no competing financial interest.

12 ACS Paragon Plus Environment

Page 13 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

ACKNOWLEDGEMENTS Partial funding for this study was provided by Natural Resources Canada and government of Canada’s interdepartmental Program of Energy Research and Development (PERD). Comments and suggestions from Dr. Rafal Gieleciak on revising the manuscript are greatly appreciated. The authors are grateful to the pilot plant and analytical lab staff at CanmetENERGY Devon.

13 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 34

NOMENCLATURE a, b, c = Metal site adsorption parameters defined in Equation 6 CLi, CLj = Component concentrations in the liquid phase, mol/L Ea = Activation energy, kcal/mol FL = Naphtha feed rate, L/h k0 = Pre-exponential factor, mol/kgcat·h·MPa Kadi, Kadj = Component metal site adsorption coefficient, L/mol ksi = Surface rate coefficient of olefin hydrogenation reactions, mol/kgcat·h·MPa KVLEi = Component vapor-liquid equilibrium partition coefficient ML = Liquid phase molecular weight, kg/mol NAR = Number of aromatic rings NSC = Number of saturated carbons !"# = Hydrogen partial pressure, MPa

R = Universal gas constant, kcal/mol·K ri = Rate of olefin hydrogenation reactions, mol/kgcat·h T = Reactor temperature, K Vb = Catalyst bed volume, L xi = Component mole fraction in the liquid phase yi = Component mole fraction in the gas phase

Subscripts i = Olefin i j = Hydrocarbon component j L = Liquid phase

Greek symbols ρb = Catalyst bulk density, kg/L ρL = Liquid phase density, kg/L τ = Space time, h

φV = Fraction of vaporized hydrocarbons

14 ACS Paragon Plus Environment

Page 15 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

REFERENCES (1)

Colyar, J. Has the time for partial upgrading of heavy oil and bitumen arrived? Pet. Tech. Q. 2009, Q4, 43–56.

(2)

Maxxam Analytics Inc. & CAPP. Analytical method for the determination of olefin content of crude oils, condensates and diluents by 1H NMR. 2005.

(3)

Gray, M.R.; McCaffrey, W.C. Role of chain reactions and olefin formation in cracking, hydroconversion, and coking of petroleum and bitumen fractions. Energy & Fuels 2002, 16, 756–766.

(4)

Badawi, M.; Vivier, L.; Perot, G.; Duprez, D. Promoting effect of cobalt and nickel on the activity of hydrotreating catalysts in hydrogenation and isomerization of olefins. Journal of Molecular Catalysis A: Chemical 2008, 293, 53–58.

(5)

Jaffe, S.B. Kinetics of heat release in petroleum hydrogenation. Ind. Eng. Chem. Process Des. Develop. 1974, 13, 34–39.

(6)

Brunet, S.; Mey, D.; Perot, G.; Bouchy, C.; Diehl, F. On the hydrodesulfurization of FCC gasoline: a review. Applied Catalysis A: General 2005, 278, 143–172.

(7)

Lengyel, A.; Magyar, S.; Hancsok, J. Catalytic co-processing of delayed coker light naphtha with other refinery gasoline streams. Chemical Engineering 2009, 53/1, 3–7.

(8)

Lengyel, A.; Magyar, S.; Hancsok, J. Upgrading of delayed coker light naphtha in a crude oil refinery. Petroleum & Coal 2009, 51/2, 80–90.

(9)

Lengyel, A.; Magyar, S.; Kallo, D.; Hancsok, J. Catalytic coprocessing of delayed coker light naphtha with light straight-run naphtha/FCC gasoline. Petroleum Science and Technology 2010, 28, 946–954.

(10)

Meerbott, W.K.; Hinds, G.P. Reaction studies with mixtures of pure compounds. Ind. Eng. Chem. 1955, 47/4, 749–752.

(11)

Kraus, M. Organic substituent effects as probes for the mechanism of surface catalysis. Adv. Catal. 1980, 29, 151–196.

(12)

Marques Da Cruz, G.; Bugli, G.; Djega-Mariadassou, G. Cyclohexene hydrogenation on rhodium catalysts in the gas phase/Kinetics of the reaction and origin, mechanism and kinetics of the deactivation process. Appl. Catal. 1989, 46, 131–144.

15 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(13)

Page 16 of 34

Magyar, S.; Hancsok, J.; Kallo, D. Reactivity of several olefins in the HDS of full boiling range FCC gasoline over PtPd/USY. European Congress of Chemical Engineering, 2007, Sep, 16–20.

(14)

Toba, M.; Miki, Y.; Matsui, T.; Harada, M.; Yoshimura, Y. Reactivity of olefins in the hydrodesulfurization of FCC gasoline over CoMo sulfide catalyst. Appl. Catal. B: Envoronmental 2007, 70, 542–547.

(15)

Badawi, M.; Vivier, L.; Duprez, D. Kinetic study of olefin hydrogenation on hydrotreating catalysts. Journal of Molecular Catalysis A: Chemical 2010, 320, 34–39.

(16)

Ghosh, P.; Andrews, A. T.; Quann, R. J.; Halbert, T. R. Detailed kinetic model for the hydrodesulfurization of FCC naphtha. Energy Fuels 2009, 23, 5743–5759.

(17)

Yui, S.M. Removing diolefins from coker naphtha necessary before hydrotreating. Oil Gas J. 1999, 97, 64–68.

(18)

Mederos, F.S.; Ancheyta, J.; Chen, J. Review on criteria to ensure ideal behaviors in tricklebed reactors. Appl. Catal. A 2009, 355, 1–19.

(19)

Bremaud, M.; Vivier, L.; Perot, G.; Harle, V.; Bouchy, C. Hydrogenation of olefins over hydrotreating catalysts promotion effect on the activity and on the involvement of H2S in the reaction. Applied Catalysis A: General 2005, 289, 44–50.

(20)

Schug, K.A.; Santos, I.C. Recent advances and applications of gas chromatography vacuum ultraviolet spectroscopy. J. Sep. Sci. 2017, 40, 138–151.

(21)

Chen, J.; Wang, N.; Mederos, F.; Ancheyta, J. Vapor-liquid equilibrium study in trickle-bed reactors. Ind. Eng. Chem. Res. 2009, 48, 1096–1106.

(22)

Alvarez-Majmutov, A.; Chen, J. Modeling and simulation of a multibed hydrotreater with vapor-liquid equilibrium. Ind. Eng. Chem. Res. 2014, 53, 10566–10575.

(23)

Korre, S.C.; Klein, M.T.; Quann, R.J. Hydrocracking of polynuclear aromatic hydrocarbons. Development of rate laws through inhibition studies. Ind. Eng. Chem. Res. 1997, 36, 2041– 2050.

16 ACS Paragon Plus Environment

Page 17 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Table 1. Properties of the selected model olefin compounds.

Molecular

Boiling

Molecular

Amount

formula

point, °C

weight, g/mol

added, wt%

2-methyl-2-pentene

C6H12

67

84.16

0.43

1,3-hexadiene

C6H10

73.1

82.14

0.20

1-methylcyclopentene

C6H10

72

82.14

0.50

1-heptene

C7H14

94

98.19

0.50

C9H10

156

118.18

0.43

Compound

Structure

allylbenzene (3-phenyl-1-propene)

1 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 34

Table 2. Properties of the base naphtha feed.

Property Density, g/mL @15.6°C

0.7384

Carbon, wt%

84.14

Hydrogen, wt%

15.36

Nitrogen, wppm