Light Cycle Oil Upgrading to Benzene, Toluene, and Xylenes by

Sep 7, 2017 - To facilitate the study of the effect of the experimental conditions, a review regarding the hydrogenation-hydrocracking (HYD-HCK) proce...
2 downloads 13 Views 1MB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Review

Light Cycle Oil Upgrading to Benzene, Toluene and Xylenes by Hydrocracking: Studies Using Model Mixtures Georgna C Cecilia Laredo, Patricia Pérez-Romo, Jose Escobar, JOSE LUIS GARCIA-GUTIERREZ, and Pedro M. Vega-Merino Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.7b02827 • Publication Date (Web): 07 Sep 2017 Downloaded from http://pubs.acs.org on September 8, 2017

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 free 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 accessible to all readers and 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.

Industrial & Engineering Chemistry Research 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 35

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

Industrial & Engineering Chemistry Research

Light Cycle Oil Upgrading to Benzene, Toluene and Xylenes by Hydrocracking: Studies Using Model Mixtures

Georgina C. Laredo*, Patricia Pérez-Romo, José Escobar, José Luis Garcia-Gutierrez, and Pedro M. Vega-Merino

a

Instituto Mexicano del Petróleo, Lázaro Cárdenas 152, México 07730 D.F., México.

*Corresponding author. Tel.: +52 55 9175 6615 E-mail addresses: [email protected] (G.C. Laredo); [email protected] (P. Pérez-Romo); [email protected] (J. Escobar); [email protected] (J.L. Garcia-Gutierrez); [email protected] (P.M. Vega-Merino).

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

Abstract The regular use of light cycle oil (LCO) for diesel fuel production by hydrotreatment (HDT) procedures has been facing difficulties for complying with the currently stricter environmental regulations due to the low quality of this middle distillate with high sulfur, nitrogen and aromatic contents. An interesting alternative is to obtain valuable petrochemicals from this feedstock. LCO presents a high percentage of di-aromatic hydrocarbons (naphthalene derivatives). In order to facilitate the study of the effect of the experimental conditions, a review regarding the hydrogenation-hydrocracking (HYD-HCK) procedures for obtaining enriched streams containing benzene, toluene and xylene (BTX) from model mixtures (naphthalene, methylnaphthalenes and tetralin) is presented. In addition to the economic advantages of using a middle distillate of reduced marketability, this research work opens the door to the development of technologies for obtaining valuable chemicals like BTX from oil sources. The focusing on model mixture studies facilitated the understanding of the involved kinetics and mechanisms and the effect of the experimental conditions on the chemical composition of the products as well.

Keywords: Light cycle oil, hydrogenation, hydrocracking, benzene, toluene and xylene.

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

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

Industrial & Engineering Chemistry Research

INTRODUCTION Light cycle oil (LCO) is a middle distillate from the FCC (fluid catalytic cracking) refinery unit. LCO is a very low quality feedstock for diesel fuel production due to its high sulfur (up to 4 wt.%), nitrogen (up to 400 mg/kg), and aromatic contents (up to 88 wt.%)

1,2

,

which make this middle distillate difficult to process in order to comply with worldwide stringent environmental regulations (Table 1)3-6. According to the US EPA’s (Environmental Protection Agency) diesel program standards, it is required that6 : • After 2010, all highway diesel fuel supplied to the market must be ultra low sulfur diesel (ULSD) and all highway diesel vehicles must use ULSD. • After 2014, all non-road, locomotive, and marine (NRLM) diesel fuel must be ULSD, and all NRLM engines and equipment use this fuel (with some exceptions for older locomotive and marine engines). These standards could reduce harmful emissions from highway vehicles and non-road engines and equipment by more than 90%. The chemical nature of LCO is highly aromatic. LCO from Mexican refineries contains up to 75 wt.% of mono-, di- and tri-aromatic compounds (Table 2) 7. One of the alternatives for LCO upgrading is to obtain a benzene, toluene, and xylene (BTX) fraction after hydrotreating (HDT) and hydrocracking (HCK). Reaction pathways for HDT and HCK of naphthalene, which well-represents diaromatics present in LCO, are shown in Figure 18. A commercial technology by UOP named LCO-X is already available9. To obtain a fraction rich in BTX, the process configuration requires the combination of HDT and HCK. An additional step, named aromatic maximization, where may be BTX range naphtha species can be converted into other aromatic compounds through dehydrogenation (DEHYD).

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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 4 of 35

According to Choi et al.8 the production of BTX from LCO firstly requires an HDT procedure to yield 1,2,3,4-tetrahydronaphthalene (tetralin) derivatives and also to decrease the amount of sulfur and nitrogen contaminants, followed by HCK. The HCK units allow the upgrading of heavy feedstocks by breaking large molecules into smaller ones by means of hydrogen at high temperatures and pressures. In order to make this process economically feasible, a deep understanding of the involved reactions is a must to overcome the difficulties associated to the very demanding experimental conditions or with expensive catalysts. Besides, dealing with real feedstocks makes the chemical characterization of the products nearly impossible. Therefore, very often, model compounds have been used for studying in detail the kinetics of LCO transformation into a valuable BTX fraction. However, when a catalyst or a process is almost ready for scaling up or commercial application, experiments with real feedstocks are mandatory, for process design and optimization studies10. It has been suggested8,11 that catalysts play a key role in the hydrogenation (partial saturation) of naphthalene followed by the selective cracking of naphthenic structures producing one-ring-aromatic hydrocarbons with alkyl chains. Bifunctional catalysts with acid (support) and metal (hydrogenation-dehydrogenation) functions are frequently used. However, an excess of hydrogenation and strong acidity cracks molecules excessively into light

gaseous

hydrocarbons

and

also

accelerates

coke

deposition.

The

hydro-dehydrogenation function of hydrocracking catalysts is generally created by sulfides of group VI and VIII metals; for the hydrocracking of heavy oil feeds, metallic-sulfidesupported catalysts are considered more effective than metals12.

ACS Paragon Plus Environment

Page 5 of 35

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

Industrial & Engineering Chemistry Research

1. Selective hydrogenation of naphthalene to tetralin The main findings regarding naphthalene HYD as a model reaction is shown in Table S1 (supporting information). In order to find a technology for producing tetralin, in 1969, Krichko et al.13 claimed to hydrogenate naphthalene to tetralin to an extent of 75-80% using a CoMo/Al2O3 catalyst, at pressure of 2 MPa, 350-370 °C and feed rate of 0.25-1.0 kg/l h-1. Additionally, Krichko et al.13 reported that the HYD conversion of 2-methylnaphthalene was 1.5 times less efficient than that of naphthalene. An improved naphthalene HYD conversion of 90-95% can be achieved at 4 MPa using a nickel tungsten sulfide catalyst. Naphthalene and pyrene were used as polyaromatic models for fuel liquefaction by Dutta and Shobert14. In a micro autoclave reactor, 3 g of naphthalene were mixed with 5 mg of ammonium tetrathiomolybdate (ATTM) and heated to 350, 400 and 450 °C at 6.3-7.1 MPa. The authors concluded that high temperatures were desirable for the first 40 min of reaction, because the reaction rate increased with the reaction temperature, but after that the thermodynamic equilibrium limits the conversion. Therefore, it is advisable to drop the reaction temperature to 350 °C after that time, in order to obtain higher amounts of tetralin. The kinetics of liquid phase HYD of naphthalene and tetralin in decane on a Ni/Al2O3 catalyst was studied by Rautanen et al.15 in order to comply with future environmental restrictions regarding aromatic contents in fuels. Product distributions of naphthalene, tetralin, decalin (decahydronaphthalene) and coke were obtained by changing the experimental conditions. The best experimental conditions were found when they used an 8% mol naphthalene feed, hydrogenated at 130 °C with a H2 pressure of 2 MPa.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

For the improvement of petroleum fractions such as middle distillates and in order to satisfy increasingly stringent regulations, Jacquin et al.16 studied the HYD of naphthalene using rhodium, platinum, iridium and ruthenium (Rh, Pt, Ir and Ru, respectively) supported on mesoporous aluminosilicates as novel catalysts at atmospheric pressure operating at 200 and 300 °C, and at 6 MPa, at 220 and 340 °C. At atmospheric pressure and 200 °C, all four catalysts showed more than 95-100% of conversion. Under these conditions, only the 2PtSiAl20 catalyst produced more than 97% of decalin. The rest of the catalysts produced mostly tetralin (83-98%). At 300 °C and atmospheric pressure, the naphthalene conversion decreased severely for all the catalysts, even for the 2PtSiAl20 catalyst, from the 95-100% obtained at 200 °C, to 9-14%. Additionally, although the naphthalene conversion decreased, the tetralin selectivity remained fairly good 75-94%. In this case, the high hydrogenolysis activity of the 2PtSiAl20 catalyst provided a tetralin selectivity of 94 %. At that temperature (300 °C) other compounds like cyclohexane, toluene, xylene, alkylbenzene, decadiene and C1-C4 hydrocarbons were produced in small amounts. Except for the platinum catalyst, toluene and xylene were produced in a 2.1-5.6%. It was demonstrated that of the four catalysts, Rh, Ru and Ir showed higher selectivity towards hydrogenolysis and/or ring-opening products than the Pt-containing catalyst in the hydrogenation of naphthalene at 300 °C and 0.1 MPa. Cheng et al.17 studied the selectivity improvement of the hydrogenation of naphthalene to tetralin through a greener method with a cheap Fe-Mo-based catalyst (which was activated with S powder in order to maintain it in the sulfide form) and formic acid as H2 and water source, using a batch reactor from 274 to 333 °C. The study was focused on the impact of water on the hydrogenation of naphthalene to produce tetralin. Experiments demonstrated

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

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

Industrial & Engineering Chemistry Research

that the addition of water could increase the yield of tetralin and suppress the formation of coke effectively; however, decalins were also produced. The influence of water density, temperature, reaction time and catalyst loading was evaluated. It was found that the conversion of naphthalene increased with the temperature, water density, reaction time and catalyst loading at a maximum of 333 °C and 0. 082 g mL-1, 5 h and 6.6 mg of catalyst and then decreased. The maximum reached tetralin yield was 85% in the presence of water while in the absence of water it was only of about 60%. A MoP catalyst supported on HZSM-5, HBeta and HY catalysts was tested in the conversion of naphthalene into tetralin by Usman et al.18. The hydrogenation of naphthalene was carried out using a fixed bed reactor at 300 °C, LHSV of 3 h-1 and hydrogen pressure of 4.0 MPa. The results showed that the naphthalene conversion and selectivity depended on the acidity, pore size and interaction of the metals with the zeolite support. MoP/HBeta exhibited the highest naphthalene conversion but low selectivity to tetralin while MoP/HZSM5 showed low activity with low selectivity, and MoP/HY presented the highest selectivity (99%) with the highest naphthalene conversion (85%). MoP/HY was tested at different temperature and the highest conversion was found at 360 °C with low selectivity (65%), however, the selectivity was 100% from 240 to 300 °C. The low surface area, high acidity of HZSM5 and high metal-support interaction caused the low performance of MoP/HZSM5 while the lowest pore volume, high dispersion, small metal particles and weak acid sites on MoP/HY allowed optimal conversion with the highest selectivity. Choi et al.8 presented their results on the hydroconversion of naphthalene derivatives into light aromatics using a down-flow fixed bed reactor at 3 MPa. They prepared eight HYD

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

catalysts, being the best, regarding the naphthalene conversion (90-92%) and tetralin selectivity (100%), the Mo2C(20)/γ-Al2O3 catalyst at 350 °C and LHSV of 0.7 h-1. The distinctive reactivity found in the first-ring hydrogenation of naphthalene to tetralin as to the subsequent total saturation to decalin (the former generally being one order of magnitude higher) suggests different reaction mechanisms for di- and monoaromatic conversion. According to the previously proposed aromatic hydrogenation mechanism15,19, naphthalene and tetralin could be adsorbed on metallic sites in three different configurations. Then, they could be adsorbed through a π/σ transition state in equilibrium with π- and σ- bonded ones. However, not all the three adsorbed species could react with dissociatively adsorbed H2 to produce the corresponding saturated compounds. The π-bonded species could be lying flat adsorbed through the aromatic ring requiring an array of sites whereas the σ-bonded ones could be vertically adsorbed on a single site in a one-point adsorption mode. The π/σ transition complex could behave in an intermediate manner between those already mentioned. It has been proposed15 that naphthalene is transformed into tetralin mainly through π/σ- adsorption, producing dihydronaphthalene as an intermediary requiring a single site. Tetralin, on the other hand, could be converted through the π- adsorbed species demanding an array of metallic sites. Consequently, naphthalene and tetralin saturations could be then classified as structure insensitive and structure sensitive reactions, respectively. Also, differences in the reaction mechanisms which occurred during naphthalene and tetralin saturation could be explained by their distinctive aromaticity15. The π-electron density of the aromatic ring of the latter is higher than that of the former. Thus, increased resonance energy of the aromatic ring in tetralin as to that in naphthalene contributes to explain the lowest reactivity of the former.

ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35

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

Industrial & Engineering Chemistry Research

Following the aforementioned way of reasoning, enhanced metallic dispersion could exert the opposite effect on the mentioned reactions as it could be beneficial to naphthalene hydrogenation as single adsorption sites are required (whose number directly augments by diminishing the metallic particle size) whereas tetralin transformation requires flat basal planes whose proportion decreases in very small supported crystals20. Then, the second ring saturation could require a larger metallic atom array to be carried out21. Another point to be considered is the possible inhibiting effect of strong naphthalene adsorption on tetralin hydrogenation. In this regard, the increased transformation of the latter could be obtained after significantly high conversion of the diaromatic compounds22,23. Clearly, limited tetralin saturation required for its transformation into valuable BTX fraction components could be then dictated by both thermodynamics and kinetics. Performances of several catalysts tested in naphthalene hydrogenation are shown in Figure 2. It can be seen that the best catalyst for tetralin formation either considering yield or selectivity is the Mo2C(20)/Al2O3 catalyst.

2. Hydrogenation of naphthalene derivatives The main findings regarding the HYD of methylnaphthalene derivatives as model compounds is shown in Table S2 (supporting information). The hydrocracking mechanism of heavy oils was studied by Landau et al.24 using 1-methylnaphthalene as a model compound. 1-methylnaphthalene hydrocracking resulted in isomerization to 2-methylnaphthalene as well as HYD to 2-methyltetralin. The HYD of

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

methyltetralins to methyldecalins was an order of magnitude slower than the HYD of methylnaphthalenes and comparable to the naphthene ring-opening reactions at 350 °C and 19.3 (6.9 H2) MPa. The observed behavior may be due to the higher steric hindrance of the methyl moiety that the intermediate may present during the HYD process, and to the difficulty for the resulting cis-decalin to perform the thermodynamic ring shift. Transdecalin derivatives are not able to perform that shift25. 1-methylnaphthalene in tetradecane, pentadecane and hexadecane mixtures was used as a model reaction for regeneration of the hydrogen donor solvent for liquefaction and hydrotreating of aromatic sources of gas oil by Ishigaki and Goto26 who used a CoMo/Al2O3 catalyst at 5-8 (4.1-7.7 H2) MPa and 310-350 °C. The products were mixtures of methyltetralin isomers with a small amount of unconverted 1-methylnaphthalene. Higher hydrogen partial pressures increased the HYD rate; however, the maximum amount of methyltetralin isomers was independent of the hydrogen pressure when working at 330 °C. The yield of methyltetralin isomers decreased as the temperature increased. Finally, Choi et al.8, hydrogenated 2-methylnaphthalene using the same catalyst (Mo2C(20)/γ-Al2O3) developed for the hydrogenation of naphthalene, using the same down-flow fixed bed reactor (3 MPa, LHSV of 0.7 h-1) at 250 and 300 °C, obtaining conversions of 97 and 82%, respectively. The product was a mixture of 6-methytetralin (6MT, 72.4 and 63.6% at 250 and 300 °C, respectively) and 2-methyltetralins (2MT, 24.8 and 25.8% at 250 °C and 300 °C, respectively). It is interesting to observe that the ring without an alkyl group was selectively more hydrogenated than the branched one. Therefore, the 6MT/2MT ratio was almost 3.

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

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

Industrial & Engineering Chemistry Research

Nevertheless, none of these authors tried to use a chemical mixture similar to that found in a LCO real feedstock excepting Bouchy et al.27 who in order to study the upgrading of LCO feedstocks to diesel and naphtha prepared a synthetic LCO model containing 40% of 1-methylnaphthalene, 2% of dimethyldisulfide and 0.25% of n-butyl amine in n-heptane and treated the synthetic mixture using a NiMo/Al2O3 catalyst at 2.8 MPa, 400 °C and LHSV of 1 m3 h-1 m-3 cat. Under these conditions, the thermodynamic equilibrium for the HYD of the first ring was reached before any hydrocracking occurred. The hydrocracking yield was 12%. The main products were alkyl indans produced by the isomerization of the methyltetralins, species from the opening of the saturated ring to give alkyl benzenes, and those from demethylation and methylation reactions on the naphthalene ring to give naphthalene and dimethylnaphthalenes as well. The chemical distribution of the products can be seen in Figure 3. Almost half of the 1-methylnaphthalene remained as naphthalene or methylnaphthalene isomers. The formation of tetralin derivatives only reached 39 %. Again, the amount of produced 5-methyltetralin was twice as much as that of 1-methyltetralin. These results indicate a predisposition for HYD of the aromatic ring that does not have an alkyl group. There were some decalin isomers formed from HYD of tetralins, and indans from isomerization. A little bit of alkyl benzenes and cyclohexanes were also present. Therefore, it seems that careful studies on the effect of experimental conditions on the HYD of the naphthalene isomers in real feedstocks are required. A comparison among various catalysts performances in the naphthalene-derivatives hydrogenation at several operating conditions is presented in Figure 4. It can be seen in this figure, that the selection of both catalyst and experimental conditions for the HYD of 1-

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

and 2-methylnaphthalene in order to get almost pure methyltetralin derivatives is not that difficult8,26. However, according to Bouchy et al.27, when other chemical components like organic sulfur and nitrogen compounds are added, the reaction does not proceed that smoothly. Firstly, it does not reach completeness and secondly, gives way to the formation of byproducts.

3. Hydrocracking of naphthalene and methylnaphthalene derivatives The main findings regarding the HCK of naphthalene and naphthalene derivatives is shown in detail in Table S3 (supporting information).

The hydrocracking of aromatic hydrocarbons was studied by Chareonpanich et al.28 using USY zeolite without metal loading for the naphthalene, tetralin and 2-methylnaphthalene hydrocracking. As the temperature increased from 400 to 600 °C at 5 MPa, the percentage of conversion of the three hydrocarbons studied increased and reached more than 100% due to the light gas formation. The naphthalene conversion increased from 50 to 105%. Notwithstanding, the BTX fraction was always less than half of the formed product (35% at 600 °C). The rest of the products were C1 to C7 gases (70%). The same could be said for the HCK of tetralin (38% of BTX and 67% of C1 to C7 gases) and 1-methylnaphthalene (45% of BTX and 60% of C1 to C7 gases). When the temperature was maintained at 600 °C, 1-methylnaphthalene required only 2 MPa for reaching almost the same conversion (100%) although a small decrement in the BTX formation was observed, from 45 to 40%.

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35

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

Industrial & Engineering Chemistry Research

With the objective of converting large polyaromatic compounds into BTX, Kim et al.29 studied the HCK of naphthalene to BTX using a novel Ni2P/zeolite (ZSM-5, Beta and USY) catalyst. The best catalysts were found to be active only at 3 MPa and 400 °C. The catalytic activity went from the highest to the lowest: Ni2P/Beta>Ni2P/USY>Ni2P/ZSM-5. They reported 94.4% BTX selectivity using a Ni2P/Beta catalyst (yield from naphthalene of 99%). Nevertheless, it is likely that Kim et al.29 did not consider the gas fraction from the cracking of naphthalene and the solvent and just calculated the yield by using the liquid fraction, because reported 94.4% conversion seemed unrealistic. A gas fraction is always formed during HCK and it is required to be taken into account when properly calculating the corresponding yield to various products. HCK reactions of naphthalene were investigated in a slurry-type reactor with different catalyst compositions consisting of iron-based compounds, metal oxides, and elementary sulfur in order to evaluate the most efficient catalyst composition by Akmaz and Caglayan30 at 400-450 °C and 5, 10 and 12 MPa. The most effective catalyst composition for the HCK reaction of naphthalene was found to be a mixture of FeSO4-H2O (20 wt.%) Fe2O3 (20 wt.%), Al2O3 (20 wt.%) and sulfur (40 wt.%) at 450 °C. First of all, tetralin formation occurred by HYD of naphthalene while 1-methylindane was formed by tetralin isomerization. The amounts of tetralin and 1-methylindane decreased and those of ethylbenzene, toluene, and benzene increased with ring-opening and dealkylation reactions. By increasing the temperature and reaction time, the amount of remaining naphthalene decreased while that of light products such as toluene and benzene grew after the HCK reaction. The highest conversion of naphthalene was achieved (89%) at 450 °C at 90 min reaction time. By raising the pressure of hydrogen, the amount of remaining naphthalene decreased while that of the light products (toluene and benzene) increased.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

Very recently, it has been reported by Kim et al.31 that textural properties and particle morphology could also play a decisive role in both the activity and selectivity of catalysts aimed at producing valuable BTX fractions from naphthalene derivatives. In this line, studies were carried out31 on Ni2P catalysts supported on two types of beta-zeolites having different crystal sizes namely based on nano and micron scales (β-N or β-M, respectively) applied in the hydrocracking of 1-methylnaphthalene (1-MN) at 6.0 MPa and 380 °C in a three-phase fixed bed reactor. The former material showed improved activity and stability in the hydrocracking of 1-MN exhibiting promoting BTX yields, (42.3% versus 30.5% for Ni2P/β-M). Additionally, after 160 h time on stream this material produced 8.8% of coke against 13.6% of the micrometer sized one. A comparison between the product selectivity using both catalysts is shown in Figure 5. The hydrocracking of 1-MN over a Ni2P/β catalyst was found to take place through the partial hydrogenation of an aromatic ring followed by cracking of the saturated ring to produce alkyl-substituted benzenes. The intercrystalline mesoporosity of β-N was reflected in enhanced Ni2P dispersion and accessibility to acidic sites. These features resulted in augmented activity in polyaromatic HCK (15 wt.% phenanthrene in 1-MN). Even more, that material maintained a stable structure under the reaction conditions. On the other hand, Ni2P/β-M suffered from increasing coke deposition with time on stream. A comparison of performance of various catalysts performances in the hydrocracking of naphthalene and 1-methylnaphthalene (1MN) considering only the composition of the liquid phase is shown in Figure 6, although gas formation, which in some cases depends on the catalyst and experimental conditions were not taken into account. In this case, the catalysts prepared by Kim et al.31 presented the best behavior in terms of selectivity and catalyst life.

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

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

Industrial & Engineering Chemistry Research

4. Hydrocracking of tetralin The effect of various reaction conditions on the HCK of tetralin as a model molecule is shown in detail in Table S4 (supporting information). Townsend and Abbot studied the catalytic cracking of tetralin on HY32 and HZSM-533 at 400 °C. In the first case, they found that the initial reactions of tetralin included isomerization to methylindans and a set of complex bimolecular processes involving both cracking and hydrogen transfer. Products derived from naphthalene, benzene and methylindans were the major species resulting from these bimolecular processes. However, by using this zeolite, the selectivity towards BTX was affected due to the presence of more than 20 products that were also formed. Under the same experimental conditions, but using HZSM-5, naphthalene, methylindans and benzene were the major products. These results may be related to the both acid strength related with the aluminum content and the structural properties that can affect the hydrocarbons diffusion through the zeolites framework34. HZSM-5 is a zeolite with low acidity and small pore size while HY presented higher acidity with major pore size. It can be suggested that the size of the pores may have an influence in the product distribution. A larger cage allows the reaction to proceed to a wider array of products. The HCK of aromatic hydrocarbons and tetralin was studied using a USY zeolite without metal loading by Chareonpanich et al.28 . As the temperature increased from 400 to 600 °C at 5 MPa, the tetralin conversion increased. Nonetheless, benzene, toluene and xylene (BTX) formation always corresponded to less than half of the products (38 %). The rest of them were C1 to C7 gases (68 %).

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

By studying the HCK of diphenylmethane and tetralin using a bifunctional NiW sulfide catalyst over three types of zeolites (USY, HY and mordenite), Sato et al.35 found that the conversion of tetralin required relatively strong acid sites. Ultra stable Y (USY) zeolite with strong acidity exhibited the best performance, while mordenite showed the worst. A 45% conversion was found using USY at 600 °C with 18% BTX fraction selectivity. The loaded NiW promoted enhanced catalytic activity by supplying active hydrogen to saturate the dehydrogenated products. By using NiW/USY, the corresponding conversion was of about 45% with 24% BTX selectivity. Continuing with their studies on the upgrading of heavy oils, Sato et al.36 presented the effect of USY and NiW/USY on the HCK of tetralin. They concluded that the presence of NiW sulfide contributes to hydrogenate the aromatic compounds so that the produced cycloparaffins are easily cracked over the acid sites. However, NiW sulfide cannot supply hydrogen to the acid sites fast enough to stabilize the intermediate cations. Therefore, the formation of polycyclic aromatic compounds, which are coke precursors, was unavoidable. Choi et al.8 tested the capabilities of a mono-functional H-Beta and bifunctional Ni/H-Beta catalysts in the HCK of tetralin to BTX aromatics. For the high per pass yield of BTX in the hydrocracking of tetralin in which chemical equilibrium limits both conversion and selectivity, the bifunctional Ni/H-Beta catalyst was promising as compared to the monofunctional H-Beta catalyst. The bifunctional H-Beta catalyst presented a yield and selectivity to BTX in the liquid product of 69.5 and 40.7%, respectively, at 99.5% tetralin conversion at 450 °C and 4 MPa. The Ni/H-Beta behavior suggested that the BTX yield may be improved by properly controlling the HYD power of the metallic sites, the acidity of the H-Beta and the balance between both capacities.

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

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

Industrial & Engineering Chemistry Research

Following the last idea, Lee et al.37 studied the HCK of tetralin for BTX conversion by supporting various phases (Ni, Ni-Sn, CoMo-S) over H-Beta in order to control the HYD activity. From these catalysts, Ni-Sn/H-Beta showed remarkable HYD-HCK capability for BTX formation (48%) at tetralin conversion over 99%. Figure 7 shows the outstanding selectivity of this catalyst. BTX and other alkyl aromatic compounds were formed over 81%. The CoMo-S/H-Beta catalyst also showed good selectivity to BTX (47.4%). These yields represent approximately 75% of the theoretical BTX formation from tetralin. Other findings were that the H2/tetralin molar ratio should be greater than 4 and H-Beta should have moderate acidity. Recently, Upare et al.38 studied the selective hydrocracking of tetralin on Mo/ß and compared its activity with a promoted CoMo/ß catalyst. Reactions were carried out in a high pressure continuous flow fixed bed reactor at 370 °C, 80 bar, LHSV of 1.6 h-1 and hydrogen flow rate of 100 cm3/min. The results showed that the incorporation of Co as a promoter improved the catalytic activity in hydrocracking reactions by reducing coke deposition and hence improving the catalyst stability. The effect of different SiO2/Al2O3 ratios in ß zeolite was also investigated. It is important to note that the catalysts were activated with dimethyl disulfide to improve their stability and prevent sulfur from poisoning during hydrocracking reactions. CoMo/ß with SiO2/Al2O3 of 25 showed 99.5% conversion of tetralin with selectivity of 62.5% to monocyclic aromatic hydrocarbons such as benzene, toluene and xylene for more than 140 h of time on stream. Additionally, the authors found that tetralin conversion improved by increasing support acidity. The improvement of the desired monoaromatic compounds was achieved by the synergy between Co and Mo.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

A comparison of performance of several catalysts used for the hydrocracking of tetralin at various operating conditions considering only the liquid phase products is shown in Figure 8. It can be seen that the materials designed by Choi et al.8 and Lee et al.37, based on HBeta zeolite, were the best to improve the BTX selectivity and yield when tetralin was used as model compound.

5. Yields and economic advantage Considering the chemical characterization of the LCO and just taking into account the average of mono and diaromatics (70.6%) and since average yield to tetralins was 77.4 and 66.2% into BTX, the final yield results in approximately 36 % of product under the best experimental conditions. If a LCO price of 36 USD/ton39 is considered meanwhile the average one of BTX is 750 USD/ton40-42, a 36 % conversion yields a value of 271 USD for the 360 kg of BTX fraction obtained from a ton of LCO. Therefore, a payback of 235 USD for each ton of processed LCO may be expected. Of course, this is a very simplistic calculation procedure involving just the commercial prices of both raw material and aimed products. In more realistic estimations, the costs of both required HDT and HCK processes must be taken into account.

6. Conclusion Refiners will have to consider investments in technology to upgrade distressed refinery streams with limited disposition options (e.g. LCO), to higher valuable chemical products (e.g. BTX), as they prepare for increased market demand in the future. In this review, the

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35

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

Industrial & Engineering Chemistry Research

most recent advances in the field of hydrogenation and hydrocracking of model mixtures containing naphthalene, metylnaphthalenes and tetralins were compiled because better understanding of reaction mechanisms, optimum operating conditions and catalyst performance will be required to the development of novel technologies. The process economics may show that the investment in an LCO hydrocracking unit to produce BTX yields higher net present values (NPV) and provides good payout compared to a conventional hydrotreating solution.

Author contributions The article was written through contributions from all authors. All authors have given approval to the final version of the manuscript.

Notes The authors declare no competing financial interest.

Acknowledgements

The authors are grateful for the financial support provided by the Instituto Mexicano del Petróleo (IMP) via the research project D.61065.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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

Supporting Information Hydrogenation of naphthalene under various reaction conditions and over different catalysts (Table S1), hydrogenation of naphthalene derivatives under various reaction conditions and over different catalysts. (Table S2), hydrocracking of naphthalene and naphthalene derivatives under various reaction conditions and over different catalysts. (Table S3), hydrocracking of tetralin under various reaction conditions and over different catalysts. (Table S4).

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35

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

Industrial & Engineering Chemistry Research

Table 1. Diesel fuel typical specifications.

Properties

ASTM

Low Sulfur

Ultra Low Sulfur

Specification

Diesel (LSD)

Diesel (ULSD)

D4052

36

36

Flash Point (°C)

D93

52

52

Kin. Viscosity at 40 °C (cSt)

D445

2.5

3.1

Corrosion test 3 h at 50 °C

D130

1A

1A

Sulfur (mg/kg)

D5453

500

15

Pour point (°C)

D97

-32

-32

Cloud point (°C)

D2500

-10

-10

Ash (wt.%)

D482

NIL

NIL

Distillation at 760 mm Hg (°C)

D86

IBP

185

186

50%

260

255

90%

313

321

FBP

343

339

D2709

NIL

0.05

D613/D6890

40

45

API Gravity

Water and sediments (vol.%) Cetane Number Ramsbottom Carbon residue

D524

0.35

(wt.%) Lubricity (µm)

D6079

520

IBP: Initial boiling point FBP: Final boiling point

ACS Paragon Plus Environment

430

Industrial & Engineering Chemistry Research

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 22 of 35

22 Table 2. Properties of LCO from typical FCC units in Mexico. Property Sp. Gravity (20/4 °C) Sulfur, %p Total nitrogen, ppm Basic nitrogen, ppm Cetane index Flash point, ºC Pour point, ºC Aniline point, ºC Refraction index Color Total aromatics, wt.% HPLC chemical characterization, wt.% Paraffins Monoaromatics Diaromatics Poliaromatics Polars Total Distillation at 760 mm Hg, °C IBP 10% 30% 50% 70% 90% FBP

Method ASTM D-70 ASTM D-4294 ASTM D-4629 UOP-313 ASTM D-976 ASTM D-92 ASTM D-97 ASTM D-611 ASTM D-1218 ASTM D-1500 ASTM D-2459

Sample 1 0.9373 2.14 566 230 30 93 3 34.2 0 L 2.0 68.8

Sample 2 0.9573 3.52 972 70 25 127 0 24 0 L 3.0 74.7

Sample 3 0.9442 2.93 445 38 25 112 -21 14.8 1.5841 L 2.0 75

27.44 17.33 33.77 12.19 9.27 100

20.23 15.21 38.91 15.94 9.71 100

20.1 17.03 53.01 6.52 3.34 100

0 179 255 285 314 341 368 402

0 216 259 284 302 327 355 380

0 226 264 284 297 308 333 362

IP-391

ASTM D-86

ACS Paragon Plus Environment

Page 23 of 35

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

Industrial & Engineering Chemistry Research

23 Figure captions Figure 1. Reaction scheme for obtaining BTX, LPG and naphtha from naphthalene as model LCO compound8. Figure 2. Yield to different products over several catalysts tested in naphthalene hydrogenation at various operating conditions. Figure 3. Chemical components in products of the hydrogenation of 1-methylnaphthalene in presence of sulfur and nitrogen compounds27. Figure 4. Comparison of the catalysts employed

for the hydrogenation of

methylnaphthalene derivatives. Figure 5. Comparison of the product selectivity of 1-methylnaphthalene hydrocracking at 380 °C, 6 MPa, and LHSV of 0.5 h-1 using Ni2P/Beta nano and Ni2P/ Beta micro sized supports. Figure 6. Comparison of various catalysts performances in the hydrocracking of naphthalene and 1-methylnaphthalene (1MN). Figure 7. Chemical composition of products from hydrocracking of tetralin in presence of a Ni(5)-Sn(5)/H-Beta catalyst37. Figure 8. Comparison of performances of several catalysts employed for the hydrocracking of tetralin at various operating conditions.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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 24 of 35

24 Figure 1

CH3 CH3

Hydrocracking

Hydrogenation

Benzene

Toluene

H3C

CH3

Tetralin

Ethylbenzene

CH3

CH3

Naphthalene

H2C

CH3

Hydrogenation

H

m-Xylene

H

Hydrocracking

H

cis-Decalin

p-Xylene

H

trans-Decalin

ACS Paragon Plus Environment

LPG, Naphtha

CH3

o-Xylene

Page 25 of 35

25 Figure 2 100 90

Chemical Composition, %

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

Industrial & Engineering Chemistry Research

80 70 60 50 40 30 20 10 0

Naphthalene

Tetralin

*Ammonium tetrathiomolybdate

ACS Paragon Plus Environment

Decalin

Industrial & Engineering Chemistry Research

26 Figure 3 40

Naphthalenes 46.5 %

35

Chemical composition, %

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 26 of 35

30 25

Tetralins 39.0 %

20 15 10

Alkylbenzenes and Alkylcyclohexanes 4.0 % Indans 2.0 %

Decalins 5.2 %

5 0

ACS Paragon Plus Environment

Page 27 of 35

27 Figure 4 From 1-Methylnaphthalene

From 2-Methylnaphthalene

100

Chemical Composition, %

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

Industrial & Engineering Chemistry Research

90 80 70 60 50 40 30 20 10 0

1-Methylnaphthalene Methyltetralins Alkylbenzenes

ACS Paragon Plus Environment

2-Methylnaphthalene Methyldecalins Others

Industrial & Engineering Chemistry Research

28 Figure 5 25 Alkanes

Ni2P/Beta Ni nano 2P/Betanano Ni2P/Beta micro Ni2P/Beta micro

BTX and C10Alkylaromatics

20

Chemical composition, %

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 28 of 35

Phenanthrenes and Unidentified HC

15

10

Decalins, Tetralins and C1-C4 Naphthalenes Indans and C10Alkylaromatics

5

0

ACS Paragon Plus Environment

Page 29 of 35

29 Figure 6

100 90

Chemical Composition, %

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

Industrial & Engineering Chemistry Research

80 70 60 50 40 30 20 10 0 [29] FeSO4 H2O, Ni2P/Beta USY USY from USYfrom from USY from 1MN Ni2P/Beta [29] FeSO44 H2O, FeSO4H2O, H2O, FeSO Ni4P/Beta nano nano Ni2P/Beta [28] [28] [31] Naphthalene [28] Fe2O3, Al2O3, Fe2O3, Al2O3, from 1MN Fe O , Al O , Naphthalene 1MN Fe2O3, Al2O3, 2 3 2 3 from 1MN[31] [28] and S (450 °C, and S (450 °C, and S (450 °C, and S (450 °C, 10 MPa) [30] MPa) [30] [30] [30] 12 12MPa)

10MPa)

Naphthalene Xylenes

Tetralin Ethylbenzene

Benzene Methylindans

ACS Paragon Plus Environment

Toluene Other

Industrial & Engineering Chemistry Research

30

Figure 7

35 BTX and C9Alkylaromatics 81.4% 30

Chemical composition, %

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 30 of 35

25 20 15 Tetralins 10 Naphtha 7.8 % 5

Indans and C10Alkylaromatics 2.9 %

0

ACS Paragon Plus Environment

Naphthalenes and C11Alkylaromatics 7.9 %

Page 31 of 35

31

Figure 8

Chemical Composition, %

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

Industrial & Engineering Chemistry Research

100 90 80 70 60 50 40 30 20 10 0

Tetralin and Methyltetralins Toluene Alkylbenzenes Indan and Methylindans Others

Benzene Xylenes Naphthalene and Methylnaphthalenes Decalin and Methyldecalins

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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 32 of 35

32 References

(1) Stanislaus, A.; Marafi A.; Rana M.S. Recent advances in the science and technology of ultra low sulfur diesel (ULSD) production. Catal. Today 2010, 153, 1-68. (2) https://www.acea.be/uploads/publications/Worldwide_Fuel_Charter_5ed_2013.pdf. (3) NORMA Oficial Mexicana de Emergencia NOM-EM-005-CRE-2015, Especificaciones de calidad de los petrolíferos. Diario Oficial de la Federación. 30/10/2015. (4) http://www.cenex.com/~/media/cenex/files/fuels/data%20sheets%20and%20brochures/ 2%20ultra%20low%20sulfur%20diesel%20fuel.ashx (5) http://www.cenex.com/~/media/cenex/files/fuels/data%20sheets%20and%20brochures/ ruby%20se_roadmaster%20se.ashx (6) https://www.epa.gov/diesel-fuel-standards/diesel-fuel-standards-rulemakings. (7) Laredo G.C.; Figueroa Y.; Cano J.L.; Mares M.T.; Castillo J. Estudio de la composición del aceite cíclico ligero provenientes de crudos mexicanos. J. Mex. Chem. Soc. 2002, 46, 115-119. (8) Choi Y.; Lee J.; Shin J.; Lee S.; Kim D.; Lee J.K. Selective hydroconversion of naphthalenes into light aromatics. Appl Catal. A. 2015, 492, 140-150. (9) Bisht D.; Petri J. Considerations for upgrading light cycle oil with hydroprocessing technologies. Indian Chem, Eng. 2014, 56, 321-335. (10) Ancheyta J. Chemical Reaction Kinetics: Concepts, Methods and Case Studies. Wiley 2017. (11) Park J-I.; Ali S. A.; Alhooshani K.; Azizi N.; Miyawaki J.; Kim T.; Lee Y.; Kim H-S.; Yoon S-H.; Mochida I. Mild hydrocracking of 1-methyl naphthalene (1-MN) over alumina modified zeolite. J. Ind. Eng. Chem. 2013, 19, 627-632. (12) Baudon A.; Lemberton J.L.; Guisnet M.; Marchal N.; Mignard S. Hydrocracking of nheptane on a sulfided NiMo-Y zeolite catalyst: effect of the sulfidation method. Catal. Letter 1996, 36, 245-247. (13) Krichko A.A.; Skvortsov S.K.; Titova T.A.; Filippov V.S.; Dogadkina N.F. Production of tetraline by the hydrogenation of naphthalene-containing fractions. Petro. Gas Proc. Chem. Tech. Fuels Oils 1969, 5, 18-22. (14) Dutta R. P.; Shobert H.H. Hydrogenation/Dehydrogenation of multicyclic compounds using ATTM as catalyst precursor. ACS 40-4 Chicago 1995, 950-955. https://web.anl.gov/PCS/acsfuel/preprint%20archive/Files/40_4_CHICAGO_0895_0950.pdf (15) Rautanen P.A.; Lylykangas M.K.; Aittamaa J.R.; Krause A.O.I. Liquid phase hydrogenation of naphthalene on Ni/Al2O3: Kinetic model. Ind. Eng, Chem. Res. 2002, 41, 5966-5973. (16) Jacquin M.; Jones D.J.; Roziére J.; Albertazzi S.; Vaccari A.; Lenarda M.; Storato L.; Canzeria R. Novel supported Rh, Pt, Ir and Ru mesoporous aluminosilicates as catalysts for the hydrogenation of naphthalene. Appl. Catal. A. 2003, 251, 131-141.

ACS Paragon Plus Environment

Page 33 of 35

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

Industrial & Engineering Chemistry Research

33

(17) Cheng Y.; Fan H., Wu S.; Wang Q.; Guo J.; Gao L.; Zong B.; Han B. Enhancing the selectivity of the hydrogenation of naphthalene to tetralin by high temperature water. Green Chem., 2009, 11, 1061-1065. (18) Usman M.; Li D.; Pazzq R.; Yaseen M. Novel MoP/HY catalyst for the selective conversion of naphthalene to tetralin. J. Ind. Eng. Chem. 2015, 23, 21-26. (19) Tjandra S.; Zaera F. A Surface Science Study of the Hydrogenation and Dehydrogenation Steps in the Interconversion of C6 Cyclic Hydrocarbons on Ni(100). J. Catal. 1996 164, 82-93. (20) Desai P.H.; Richardson, J.T. Crystallite size effects in nickel catalysts: Cyclohexane dehydrogenation and hydrogenolysis. J. Catal. 1986, 98, 392-400. (21) Vít Z.; Gulková D.; Kaluža L.; Kupčík J. Pd–Pt catalysts on mesoporous SiO2–Al2O3 with superior activity for HDS of 4,6-dimethyldibenzothiophene: Effect of metal loading and support composition. Appl. Catal. B 2015, 179, 44-53. (22) Escobar J.; Barrera M.C.; Santes V.; Terrazas J.E. Naphthalene hydrogenation over Mg-doped Pt/Al2O3. Catal. Today 2017; http://dx.doi.org/10.1016/j.cattod.2017.04.064 (23) Jongpatiwut S.; Li Z.; Resasco D.E.; Alvarez W.E.; Sughrue E.L.; Dodwell G.W. Competitive hydrogenation of poly-aromatic hydrocarbons on sulfur-resistant bimetallic PtPd catalysts. Appl. Catal. A 2004, 262, 241-253. (24) Landau R.N.; Korré S.C.; Neurock M.; Klein M.T.; Quann R.J. Hydrocracking of heavy oils: development of structure/reactivity correlations for kinetics. ACS 37-4 Washington, DC 1992, 1871-1877. https://web.anl.gov/PCS/acsfuel/preprint%20archive/Files/37_4_WASHINGTON%20DC_ 08-92_1871.pdf (25) http://www.masterorganicchemistry.com/2014/09/02/bredts-rule-and-summary-ofcycloalkanes/ (26) Ishigaki S.; Goto S. Kinetics studies in liquid phase hydrogenation of 1methylnaphtalene. J. Chem, Eng. Jpn. 1994, 27, 309-313. (27) Bouchy M.; Dufresne P.; Kasztelan S. Hydrogenation and hydrocracking of a model light cycle oil feed. 1. Properties of a NiMo hydrocracking catalyst. Ind. Eng. Chem. Res. 1992, 31, 2661-2669. (28) Chareonpanich M.; Zhang Z.; Tomita A. Hydrocracking of aromatic hydrocarbons over USY-zeolite. Energy Fuels 1996, 10, 927-931. (29) Kim Y.S.; Yun G.N.; Lee Y.K. Novel NiP/zeolite catalyst for hydrocracking of naphthalene to BTX. Chem. Comm. 2014, 45, 133-138. (30) Akmaz S.; Caglayan P.A. Effect of catalyst, temperature and hydrogen pressure on slurry hydrocracking reactions of naphthalene. Chem. Eng. Techn. 2015, 38, 917-930. (31) Kim Y.-S.; Choi K.-S.; Lee Y.-K. Morphology effect of β-zeolite supports for Ni2P catalysts on the hydrocracking of polycyclic aromatic hydrocarbons to benzene, toluene, and xylene. J. Catal. 2017, 351, 67-78.

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

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 34 of 35

34

(32) Townsend A.T.; Abbot J. Catalytic reactions of tetralin on HY zeolite. Appl. Catal. A 1992, 90, 97-115. (33) Townsend A.T.; Abbot J. Catalytic reactions of tetralin on HZSM-5 zeolite. Appl. Catal. A 1993, 95, 221-236. (34) Gonçalves de Almeida J.L.; Dufaux M.; BenTaarit; Naccach Cl. Effect of pore size and aluminium content on the production of linear alkylbenzenes over HY,H-ZSM-5 and H-ZSM-12 zeolites : Alkylation of benzene with 1-dodecene. Appl. Catal. A 1994, 114, 141-159. (35) Sato K.; Iwata Y.; Yoneda T.; Nishijima A.; Miki Y.; Shimada H. Hydrocracking of diphenylmethane and tetralin over bifunctional NiW sulfide catalyst supported in three types of zeolites. Catal. Today 1998, 45, 367-374. (36) Sato K.; Iwata Y.; Miki Y.; Shimada H. Hydrocracking of tetralin over NiW/USY zeolite catalyst for the improvement of heavy oils upgrading catalyst. J. Catal. 1999, 186, 45-56. (37) Lee J.; Choi Y.; Shin J.; Lee J.K. Selective hydrocracking of tetralin for light aromatic hydrocarbons. Catal. Today 2016, 265, 144-153. (38) Upare D.P.; Park S.; Kim M.S.; Kim J.; Lee D.; Lee J.; Chang H.; Choi W.; Choi S.; Jeon Y.-P.; Park Y.-K.; Lee C.W. Cobalt promoted Mo/beta zeolite for selective hydrocracking of tetraline and pyrolysis fuel oil into monocyclic aromatic hydrocarbons. J. Ind, Eng. Chem., 2016, 35, 99-107. (39) http://www.opisnet.com/Images/ProductSamples/IFI-sample.pdf (40) http://www.platts.com.es/news-feature/2014/petrochemicals/pgpi/benzene (41) http://www.platts.com.es/news-feature/2014/petrochemicals/pgpi/toluene (42) http://www.platts.com.es/news-feature/2016/petrochemicals/global-aromatics-pricinganalysis/aromatics-xylenes-prices

ACS Paragon Plus Environment

Page 35 of 35

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

Industrial & Engineering Chemistry Research

338x190mm (96 x 96 DPI)

ACS Paragon Plus Environment