You are on page 1of 48

Subscriber access provided by Binghamton University | Libraries

Article
Stochastic Modeling and Simulation Approach
for Industrial Fixed-Bed Hydrocrackers
Anton Alvarez-Majmutov, and Jinwen Chen
Ind. Eng. Chem. Res., Just Accepted Manuscript • Publication Date (Web): 31 May 2017
Downloaded from http://pubs.acs.org on June 5, 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 47 Industrial & Engineering Chemistry Research

1
2
3
4
5
Stochastic Modeling and Simulation
6
7
8
Approach for Industrial Fixed-Bed
9
10
11 Hydrocrackers
12
13
14
15 Anton Alvarez-Majmutov* and Jinwen Chen
16
17 Natural Resources Canada, CanmetENERGY-Devon, One Oil Patch Drive, Devon, AB,
18
19
T9G 1A8, Canada
20 © Her Majesty the Queen in Right of Canada, as represented by the Minister of Natural Resources,
21 2017.
22
23
24
25 Keywords
26
27 Stochastic simulation; vacuum gas oil; hydrocracking; adiabatic reactor
28
29
30
31 Abstract
32
33 In this study, we developed a comprehensive model to simulate vacuum gas oil
34
hydrocracking. The model incorporates knowledge of feedstock composition and reaction
35
36 chemistry at the molecular level and is tuned with pilot plant data. The molecular makeup
37
38 of the feedstock is derived from its analytical characterization via statistical modeling.
39
40 Hydrocracking reactivity is described based on the molecular structure of the reactants.
41
42
Progression of chemical reactions is simulated using a kinetic Monte Carlo algorithm. A
43 stochastic heat balance equation is integrated into the model to represent adiabatic
44
45 operation. Vapor-liquid equilibrium calculations are executed in parallel with an in-house
46
47 flash program. Hydrocracker simulations were conducted covering the typical range of
48
49 conversion levels observed in commercial operation. The model generates detailed
50
information on product distribution as well as product quality. In addition, the model is
51
52 able to track the evolution of hydrogen consumption, hydrocarbon vaporization, and
53
54 temperature along the reactor.
55
56
57
58
59
60 1
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 47

1
2
3
1. Introduction
4
5
6
7 The configuration of petroleum refineries nowadays is largely driven by market
8
9 requirements. Hydrocracking has become a major conversion technology for the
10
11 production of high-quality fuels and lubricants. Commercial hydrocracking units are able
12
13 to take feedstocks comprised of a variety of difficult streams, including heavy vacuum
14 gas oil (VGO), light cycle oil, coker gas oil, and deasphalted oil. Depending on the
15
16 refining objectives and product specifications, hydrocracker designs can have one or two
17
18 reaction stages, with once-through or recycle flow modes.1
19
20 The exothermic nature of hydrocracking reactions makes heat management an
21
important aspect of the process. Large amounts of heat are produced through the
22
23 chemical addition of hydrogen during the saturation of aromatic hydrocarbons and
24
25 stabilization of cracking products.2 This heat accumulates in the reacting stream, resulting
26
27 in sharp evolution of temperature along the catalyst bed. Maintaining the temperature rise
28
29
within specific boundaries is crucial for optimal operation, as too-high temperatures can
30 lead to loss of product selectivity, higher hydrogen consumption, and coke build-up on
31
32 the catalyst surface.3 In commercial practice, hydrocracking reactors are designed with
33
34 multiple catalyst beds for the purpose of injecting quench gas in between, thereby
35
36 reducing temperature rise.4 Hydrocracking reactor design is an elaborate undertaking
37 requiring reliable models with the capability to predict not only product composition and
38
39 properties, but also the behavior of operational parameters.
40
41 Because of the complexity of hydrocracking chemistry and feedstock
42
43 composition, process models often rely on simplified reaction schemes formulated in
44
45
terms of broad chemical lumps defined by bulk property measurements.5 To simulate
46 adiabatic reactor operation, it is also common to use a lumped heat of reaction in the heat
47
48 balance equation.6 This term is essentially a model parameter obtained by regression of
49
50 plant data7 and is considered an approximate average of the net heat release of all
51
52 reactions. A few examples of adiabatic hydrocracker simulations using this approach are
53 available in the literature.8–11 The global heat of reaction reported in these studies is about
54
55 10 kcal per mole of H2 consumed, which corresponds to the range of normal alkane
56
57 hydrocracking.2 Since the total heat evolved is determined by the extent of hydrogenation
58
59
60 2
ACS Paragon Plus Environment
Page 3 of 47 Industrial & Engineering Chemistry Research

1
2
3
of the various hydrocarbon classes in the reacting mixture, models without detailed
4
5 process chemistry have limited capacity to respond to variations in feedstock reactivity
6
7 and operating conditions.12 Furthermore, other factors that affect hydrocracker
8
9 performance but are usually ignored, for example vapor-liquid equilibrium (VLE), are
10
11
needed to expand model predictive capabilities.13
12 Molecular-level modeling of refinery processes enhances simulation capability
13
14 through more fundamental descriptions of the process. To overcome the difficulties of
15
16 petroleum compositional analysis, composition modeling techniques have been
17
18 developed and applied to derive consistent representations of refinery feedstocks based
19 on thousands of molecules. These are subsequently used to model chemical reactions.14
20
21 The resulting reaction network in most cases is of considerable size, making its
22
23 implementation through deterministic rate equations quite challenging. As an alternative,
24
25 stochastic simulation algorithms have been proved to be capable of facilitating the task of
26
27
dealing with complex chemical processes.15 Detailed composition-reaction models for
28 various refinery hydroprocesses have recently been developed following this approach.
29
16–18
30 While the methodology has demonstrated sufficiently robust capabilities, the focus
31
32 so far has been limited to bench-scale isothermal reactors.
33
34 In previous work18 we developed a molecular-level process model for vacuum gas
35 oil hydrocracking using experimental data from a bench-scale reactor. The model
36
37 accounts for acid- and metal-catalyzed reactions, chemical equilibrium, adsorption on
38
39 catalytic sites, and VLE, but is restricted to isothermal operation. In the present study, we
40
41 extend this model by introducing heat release kinetics to simulate adiabatic
42
43
hydrocracking reactor operation with inter-bed gas quenching.
44
45
46 2. Modeling Approach
47
48
49
50 The developed modeling approach comprises two main components: feed
51
52 composition modeling and process modeling. The first component consists of deriving a
53
54 consistent molecular representation that mimics the chemical composition of the
55 hydrocracker feedstock. The second component is for simulating chemical reactions at
56
57
58
59
60 3
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 47

1
2
3
the molecular level, phase equilibrium, and thermal behavior of the hydrocracking
4
5 reactor. The various elements of the hydrocracker model are described below.
6
7
8
9 2.1. Feed composition modeling
10
11 The purpose of the composition model is to simulate the molecular composition
12
13 of complex oil fractions by systematically generating virtual mixtures of representative
14
15 molecules. The basic concept behind this approach is that the massive number of
16 hydrocarbon molecules making up petroleum can be broken down into a limited number
17
18 of structural fragments such as alkyl chains, benzene rings, cyclohexane rings, and
19
20 heterocyclic rings, and that the frequency of occurrence of each of these fragments can be
21
22 modeled using continuous statistical distributions.19 Thus, a wide spectrum of
23
24
hydrocarbon molecules can be assembled from a limited number of fragments by
25 computational sampling of their statistical distributions. The use of continuous
26
27 distribution functions to represent oil composition has its origin in Boduszynski’s model,
28 20,21
29 in which petroleum composition is a continuum in molecular weight and structure.
30
31 In the present work, compositional simulation of VGO feedstock is performed
32 using a methodology described in detail in our previous publications.22,23 Briefly, the
33
34 VGO is represented by paraffins, cycloparaffins, aromatics, and their heterocyclic
35
36 homologues (e.g. organosulfur and organonitrogen compounds). Specific building
37
38 pathways and chemical rules are implemented to determine the molecular structure of
39
each hydrocarbon class. For example, aromatics are constructed by first setting the
40
41 number of benzene and cyclohexane rings and then establishing the number and length of
42
43 the alkyl branches that are attached to the aromatic core structure. This process is
44
45 executed through Monte Carlo sampling of a set of probability functions describing the
46
47
distribution of each structural fragment or feature. The simulated feedstock is built on a
48 large set (104) of computational molecules and optimized by applying simulated
49
50 annealing and entropy maximization methods to match the available analytical data.
51
52 The methodology was developed and validated by using middle distillate (204–
53
54 343°C) and vacuum distillate (343–550°C) samples having diverse geographic origins
55 and processing histories. Input for the simulations was obtained from routine and
56
57 advanced characterization techniques: liquid density, elemental analysis, simulated
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 47 Industrial & Engineering Chemistry Research

1
2
3 13
distillation, C NMR spectroscopy, SARA fractionation, and hydrocarbon type and
4
5 sulfur compound speciation by two-dimensional gas chromatography. Table 1 provides
6
7 an example of the application of the composition model to one VGO feedstock. It is
8
9 observed that the simulated VGO compares well with the real sample in terms of physical
10
11
and chemical properties. The molecular representation contains sufficient information on
12 the composition and chemical structure of the feed to enable detailed process modeling.
13
14
15
16 2.2. Reaction modeling
17
18 The set of molecules describing the VGO feedstock forms the basis to simulate
19
20 the hydrocracking process. Instead of using deterministic equations to simulate the
21
22 progression of chemical reactions, a variable time step kinetic Monte Carlo algorithm is
23
24
employed.16,24 By this method, reactions of the discrete population of VGO molecules are
25 simulated one at a time, meaning that through each reaction step one reactant molecule is
26
27 replaced by new product molecules. For example, the change of unimolecular reaction
28
29 system “a → b” from the current reaction step s to the next one (s+1) is executed as
30
31 follows:
32  =  − 1 (1)
33
34   =   + 1 (2)
35
36 where ma and mb are the number of reactant “a” molecules and product “b” molecules,
37
38 respectively. The reaction event to be executed is selected by random sampling of a
39
40 cumulative probability distribution, which is formed by accumulating the rates of all
41 potential reactions. To this end, it is necessary to define the possible reactions (e.g.,
42
43 hydrogenation, cracking, and isomerization) for each molecule present in the population.
44
45 After executing the selected reaction, a time increment is set to advance the reaction time.
46
47 The time increment (∆ts) between the current reaction step s and the next one is defined
48
as:
49
ln()
∆ = −
50 (3)
51 ∑ 
52
53 where rn is a random number uniformly distributed between 0 and 1 and ∑  is the
54
55 summation of all the potential reaction probabilities, or in this case reaction rates ri. Since
56
57 the molecular population evolves with each reaction step, a new distribution of potential
58
59
60 5
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 47

1
2
3
reaction probabilities is generated each time to define the corresponding reaction event
4
5 and time increment.
6
7 To simulate hydrocracking reactions in the axial reactor coordinate, an
8
9 appropriate form of equation 3 is formulated. Assuming that the hydrocracking reactor
10
11
behaves as a pseudo-homogeneous plug-flow reactor, the mass continuity equation for
12 hydrocarbon component i is:

13

=   
14 (4)
15 
16 
17 where CLi is the component concentration in the liquid phase, τ is space time, rik is the
18
19 component rate to form hydrocracking product k, and ρb is the catalyst bulk density.
20
21 Space time can be defined in terms of liquid phase velocity (uL) and axial reactor distance
22
23 (z) as follows:

=
24
(5)

25
26
27 Considering that the time coordinate used in equation 3 is in essence equivalent to space
28
29 time in a plug-flow reactor, the stochastic axial position increment (∆zs) between the
30
31 current reaction step s and the next one becomes:
32
ln()
∆ = −
(6)
1
33
∑ ∑
    
34
35
36 The stochastic approach is well-suited for handling large-scale systems where
37
38 numerous molecular species participate in many highly coupled chemical reactions. A
39
40 particular advantage over deterministic modeling is that the hydrocracking reaction
41
42 network is generated while reactions are moving forward.16 However, we must note that
43
44
since the model is derived from a pseudo-homogeneous plug-flow reactor equation, the
45 effects of external mass transfer limitations are excluded by default. This means that this
46
47 aspect of the process is implicit in the model parameters. On the other hand, given that
48
49 the pilot plant unit used in this study operates in up-flow mode (section 3), external mass
50
51 transfer is expected to improve, particularly at the liquid-solid interphase.25 Another
52 related assumption in the model is full catalyst wetting. This is justified by the fact that
53
54 up-flow operation provides complete wetting of the catalyst bed. Therefore, it can be
55
56 considered that correction by wetting efficiency to represent commercial-scale reactor
57
58 operation under full catalyst wetting conditions is not necessary. Likewise, it is implicitly
59
60 6
ACS Paragon Plus Environment
Page 7 of 47 Industrial & Engineering Chemistry Research

1
2
3
assumed that diffusional limitations inside the catalyst particle are minimal owing to the
4
5 wide pore structure of the zeolite catalyst matrix.26,27 It should be noted that pilot plant
6
7 experiments were conducted using commercial size catalyst particles, meaning that the
8
9 measured kinetics reflect catalyst performance in its working state including any possible
10
11
diffusional effects.
12
13
14 2.2.1. Reaction kinetics
15
16 Hydrocracking reactions are promoted in the presence of dual-functional
17
18 catalysts. The metal phase catalyzes hydrogenation and dehydrogenation reactions, while
19
20 the acid function of the catalyst governs cracking and isomerization. Figure 1 illustrates
21
22 the possible reaction pathways for aromatics hydrocracking, with naphthalene as the
23
24
representative compound.28–31 Naphthalene is first saturated on a terminal ring via
25 naphthalenic-type hydrogenation to form tetralin. The saturated ring of the tetralin
26
27 molecule is subsequently isomerized into a five-membered ring with a methyl group
28
29 substituent, followed by ring opening into butylbenzene. The final step is the removal of
30
31 the butyl branch attached to the benzene ring. The tetralin molecule from the initial step
32 can also be saturated into decalin through a benzenic-type hydrogenation, which then
33
34 follows a similar sequence of isomerization and cracking steps. In an analogous way,
35
36 paraffins undergo various skeletal isomerizations to form mono- and multibranched
37
38 structures before being cracked into smaller fragments via β-scissions (Types A, B1, and
39
B2), as shown in Figure 2.32,33 Given the recursive pattern of the hydrocracking network,
40
41 reactions can be efficiently organized into families according to the types of metal- and
42
43 acid-catalyzed transformations34: aromatic ring hydrogenations (benzenic, naphthalenic,
44
45 and phenanthrenic)28, paraffin isomerization, paraffin cracking, ring isomerization, ring
46
47
opening, and ring dealkylation.
48 The rate of hydrocracking reactions on metal and acid sites is formulated in terms
49
50 of the Langmuir-Hinshelwood-Hougen-Watson (LHHW) formalism as follows:35

51
! "# $%" & '() *# − $ # -
52
+,"#
(7)
 =
53 Metal sites:
1 + ∑ $%" 
54
55
56
57
58
59
60 7
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 47

1
2
3 
4 ! ./ $%. & − $ / -
+,./
(8)
, =
5 Acid sites:
6 1 + ∑ $%. 
7
8 where ksMij and ksAiq are the rate constants for metal and acid site reactions, respectively;
9
10 CLi, CLj, CLq are component concentrations in the liquid phase; '() is hydrogen partial
11
12 pressure and nij is the hydrogen stoichiometry; KeqMij and KeqAiq are chemical equilibrium
13
14 constants for metal and acid site reactions, respectively; KadMi and KadAi are the metal and
15
16 acid site adsorption equilibrium constants, respectively.
17 Considering the size of the hydrocracking reaction network, applying Equations 7
18
19 and 8 involves an overwhelming number of model parameters (rate constants, adsorption
20
21 and chemical equilibrium parameters). Strategic simplifications are needed to reduce the
22
23 number of model parameters to a manageable level. A consistent way of doing this is to
24
use quantitative structure/reactivity correlations (QSRC) whereby reactivity parameters
25
26 are correlated with molecular descriptors pertaining to each reaction family.36,37 Korre et
27
28 al.29,38 developed robust QSRCs for hydrocracking of aromatic hydrocarbons in a
29
30 comprehensive study using model compounds. These correlations enable systematic
31
32
descriptions of component reaction rate, adsorption, and equilibrium parameters for metal
33 and acid sites. Table 2 lists the QSRCs for the hydrocracking reaction families. As
34
35 observed, reactivity parameters are described by using molecular structure descriptors
36
37 and thermodynamic quantities that can be estimated through semi-empirical quantum
38
39 chemistry calculations (e.g. AMPAC software).
40 The expressions used for paraffin reactions are given in Table 3. Paraffin
41
42 reactivity is assumed to increase as a function of paraffin chain length.39,40 Structural
43
44 proportionality factors (NPCi - aiq) indicating the number of reactive paraffinic carbons are
45
46 introduced for approximating isomerization and cracking rate parameters.41 The rates of
47
48
B1- and B2-type cracking are related to that of type A using factors derived from
49 experimental data.42 Isomerization and cracking reactions are assumed to have equal
50
51 activation energies.33,43 Adsorption on acid sites is considered to be invariant with respect
52
53 to paraffin chain length, as suggested by experimental evidence44,45, and therefore is
54
55 represented by a single fitting coefficient. Paraffin isomerization equilibrium coefficients
56 are estimated using Benson’s group contribution method. Reference rate and adsorption
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 47 Industrial & Engineering Chemistry Research

1
2
3
parameters are obtained from model compound studies43 of representative long-chain
4
5 paraffins.
6
7
8
9 2.2.2. Vapor-liquid equilibrium
10
11 Hydrocarbon vaporization is a key factor influencing conversion of reactants,
12
13 fluid dynamics, and the heat balance in a hydroprocessing reactor. Flash calculations are
14
15 therefore necessary to quantify the extent of feed vaporization in the reactor as well as the
16 changes in phase compositions and properties.
17
18 Assuming that the vapor and liquid phases achieve equilibrium at any position
19
20 (i.e. reaction step) inside the hydrocracking reactor, the distribution of hydrocarbon
21
22 components and hydrogen between the vapor and liquid phases is given by:
1
= $34
23 (9)
2
24
25
26
27 where yi and xi are component mole fractions in the gas and liquid phases, respectively,
28 and $34 is the component VLE partition coefficient. Mole fractions in each phase are
29
30 defined in terms of the discrete population of reacting molecules as follows:
5,
31
1 =
32 (10)
33 ∑ 5,
34
,
2 =
35
(11)
∑ ,
36
37
38
39 where mGi,s and mLi,s are the number of component i molecules in the gas and liquid
40
41 phases, respectively, at current reaction step s. Component liquid phase concentrations
42
defined in rate equations 7 and 8 are evaluated using the following expression:

43
 = 1
44 (12)
45 7
46
47 where ρL is the liquid phase density and ML is the liquid phase molecular weight.
48
49 Calculation of VLE partition coefficients for hydrocarbons and hydrogen and phase
50
51 compositions at reactor conditions is done by solving the Peng-Robinson equation of
52
53 state (EoS). Computations are executed using an in-house flash calculation program
54
55
designed specifically for hydrogen-petroleum systems. The critical properties of
56 hydrocarbons and other related parameters used in the EoS are estimated from their
57
58
59
60 9
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 47

1
2
3
boiling point, density, and molecular weight. For hydrogen, such information is available
4
5 the literature. Liquid and gas phase thermophysical properties, such as enthalpies and
6
7 heat capacities, are estimated by means of various correlations. The key feature of the
8
9 flash program is that the binary interaction coefficients between hydrogen and
10
11
hydrocarbons of the EoS were determined from experimental VLE data sets for a variety
12 of middle and heavy distillates over wide ranges of temperatures and pressures. A
13
14 number of previous studies documenting the development of this program and its
15
16 application to hydroprocessing reactor modeling are available in the literature.13,46–48
17
18
19
20 2.2.3. Heat balance
21
22 Hydrocracking reactions are exothermic overall since the heat generated by the
23
24
hydrogenation of aromatic rings and unsaturated cracking products exceeds the heat
25 consumed by C-C bond scission during cracking reactions.2 The consequent temperature
26
27 rise along the hydrocracking catalyst bed can be modeled with the following pseudo-
28
29 homogeneous energy balance equation for fixed-bed adiabatic reactors:
30 < @3 <
895 : 5 + 9 :  ; + =3 >? −  8−ΔBC*D ;   = 0
(13)
 < 
31
 
32
33
34 where FL and FG are the liquid- and gas-phase superficial mass velocities, respectively,
:  and : 5 are the liquid- and gas-phase heat capacities, respectively, T is reactor
35
36
37
38 temperature, λV is the heat of vaporization, G0 is the superficial mass velocity of the
hydrocarbon feed, φV is the fraction of vaporized hydrocarbons, and ΔBC*D is the heat
39
40
41
42
release of hydrocracking reaction ik, which comprises any possible metal and acid center
43 reaction.
44
45 The first term in equation 13 represents convective transport, the second term is
46
47 the heat consumed by vaporization, and the third term is net heat generation. Heat
48
49 capacities of each phase, heats of vaporization, and the vaporized fraction at process
50 conditions are calculated by the VLE program described in section 2.2.2. Heats of
51
52 reaction conventionally are estimated by thermodynamic calculations using the molecular
53
54 structure of reactants and products. However, when taking into account the number of
55
56 reactions involved, it is preferable to reduce the problem in a systematic manner. Jaffe2
57
identified from various sets of thermochemical data of model hydrocracking reactions49
58
59
60 10
ACS Paragon Plus Environment
Page 11 of 47 Industrial & Engineering Chemistry Research

1
2
3
that the heat generated per mole of hydrogen consumed is practically constant throughout
4
5 distinct reaction families, namely paraffin cracking, ring opening, aromatics
6
7 hydrogenation, and olefin saturation. Also, isomerization reactions were determined to
8
9 have a minor contribution, as there is no hydrogen addition involved. Such a pattern
10
11
indicates that the heat released in the hydrocracking process can be efficiently described
12 using representative values for each reaction family that consumes hydrogen. The heat
13
14 release data for each reaction family considered in the model is listed in Table 4 (from
15
16 Stull et al.49). Reported values are relative to the moles of hydrogen consumed in each
17
18 reaction family, hence the hydrogen stoichiometric coefficients (nik) in the third term of
19 equation 13. Also noted is that the heat release of paraffin cracking, ring opening, and
20
21 dealkylation is an overall pathway value considering cracking and product saturation.
22
23 Deterministic equation 13 is reformulated into the corresponding stochastic
24
25 expression. The following approximation is proposed to calculate the temperature step
26
27
increase (∆Ts) between reaction steps:
28 8−ΔBC*D ;
∆< = F∆     G
(14)
@3
29
H95 : 5 + 9 :  + =3 >? I
 
<
30
31
32 Here, the stochastic axial position increment ∆zs substitutes the differential dz. Since
33
34 reactions are executed on a one-by-one basis, the heat release of the current reaction is
35
36 used instead of the net heat release. Thermophysical properties in the denominator
37
38 including the derivative term are computed by the flash program.
39
40
An additional energy balance equation is required for describing inter-bed gas
41 quenching. The following equation is introduced to calculate the required quench gas rate
42
43 (fQ) to cool the effluent from the outlet bed temperature (Tout) of the upper bed to a
44
45 specified inlet temperature (T0) of the lower bed:
M M
J5 LM N : 5 < + J LM N :  <
46
(15)
JK = − OPQ OPQ
47
M
LM N : K <
48
49 R
50
51
52
53 2.2.4. Stochastic reactor simulation
54
55 The simulation is set up by specifying model inputs: feedstock properties, reactor
56
57 geometry, and operating conditions. The feed composition model generates a set of
58
59
60 11
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 47

1
2
3
10,000 representative VGO molecules along with their concentrations and physical and
4
5 chemical properties. Reactor simulations start with a flash calculation to establish the
6
7 distribution of VGO molecules in the liquid and vapor phases at the reactor inlet, along
8
9 with phase compositions and properties. Adsorption, equilibrium, and rate parameters at
10
11
process conditions are estimated using the equations presented in Tables 2 and 3.
12 Component reaction rates are evaluated at process conditions using the liquid phase
13
14 concentrations and reactivity parameters. The rates of all possible hydrocracking
15
16 reactions are accumulated to generate a distribution function of reaction probabilities.
17
18 The reaction event to be carried out is then randomly selected from this distribution
19 function.
20
21 The molecular population evolves through each reaction step in a discrete manner,
22
23 as exemplified by equations 1 and 2. Accordingly, one hydrocarbon reactant molecule in
24
25 the liquid phase is consumed in any reaction. Hydrogenation, isomerization, and ring
26
27
opening reactions generate one product molecule, whereas ring dealkylation and paraffin
28 cracking reactions give two product molecules. Hydrogen consumption is determined by
29
30 the hydrogen stoichiometry of each reaction.
31
32 Reactions of the molecular population are executed one-by-one until reaching the
33
34 reactor outlet. Reactor axial distance and temperature are moved forward by estimating
35 the corresponding increments at each reaction step using equations 6 and 14. Product
36
37 composition and properties, VLE, and reactivity parameters are updated based on each
38
39 new molecular population and process conditions. Quench zone calculations are
40
41 performed between catalyst beds to establish the new gas rate and process temperature. A
42
43
flash calculation is also conducted to update phase compositions and properties at the
44 entrance of each catalyst bed.
45
46
47
48 3. Summary of Experiments
49
50
51
52 In our previous study18, hydrocracking experiments covering a wide range of
53
54 conversions (10–70 wt%) were conducted in a bench-scale unit to obtain data for model
55 parameter estimation. The unit comprises an up-flow isothermal fixed-bed reactor (length
56
57 100 cm, internal diameter 9 mm) with a four-zone heating furnace and multiple
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 47 Industrial & Engineering Chemistry Research

1
2
3
thermocouples for temperature control, a high-pressure gas-liquid separator, an
4
5 atmospheric stabilizer column to remove dissolved gases in liquid products, and an online
6
7 gas chromatograph (HP5880 GC) for refinery gas analysis. The reactor was packed with
8
9 10 cm3 of commercial zeolite hydrocracking catalyst in the form of extrudates. The
10
11
catalyst bed was diluted with an equal volume of 0.2-mm glass beads to reduce axial
12 dispersion and provide good catalyst wetting. The reactor packing scheme also included
13
14 layers of glass beads with different sizes on top and below the catalyst bed to ensure
15
16 proper distribution of reactants. These measures are essential for accurate scale-up of
17
18 experimental results. The catalyst was activated in situ with hydrogen sulfide (H2S) and
19 stabilized by running the feed for 200 h.
20
21 The feedstock was a vacuum gas oil that had previously been hydrotreated to meet
22
23 the zeolite catalyst specification for sulfur and nitrogen content. The physical and
24
25 chemical properties of the feed are presented in Table 1. Experiments were conducted at
26
27
liquid hourly space velocities (LHSVs) of 0.75–3.0 h-1 and temperatures of 380–400°C,
28 while maintaining pressure and H2/oil ratio constant at 11 MPa and 4500 scf/bbl,
29
30 respectively. Liquid products were analyzed for liquid density, elemental composition,
31
32 boiling point distribution, and hydrocarbon type using standard characterization methods.
33
34 Chemical hydrogen consumption was estimated using an in-house procedure based on a
35 global hydrogen balance in the gas and liquid products.
36
37 The data collected from the bench-scale tests were used to fit the rate parameters
38
39 of the hydrocracking reaction families. A comparison of the model predictions versus the
40
41 experimental data set is given in Figure 3. Overall, the model is able to adequately
42
43
capture observed trends for commercial product distribution, hydrocarbon type, API
44 gravity, and hydrogen consumption.
45
46
47
48 4. Results and Discussion
49
50
51
52 The developed model was applied to simulate a commercial-scale multi-bed VGO
53
54 hydrocracker with inter-bed gas quenching. The feed was assumed to be the VGO
55 described in the Experimental section (Table 1). The VGO feed is rich in cycloparaffins
56
57 (61.3 wt%) and aromatics (35.4 wt%), and low in paraffinic content (3.2 wt%). The
58
59
60 13
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 14 of 47

1
2
3
aromatic fraction consists primarily of mono- and di-aromatics (18.6 and 9.1 wt%,
4
5 respectively) plus a moderate amount of polyaromatics (7.7 wt%) comprising structures
6
7 of 3–5 aromatic rings.
8
9 Reactor geometry was taken from Pacheco and Dassori9: total catalyst bed volume
10
11
124.7 m3 and catalyst bed diameter 3.7 m. Catalyst bed distribution was adjusted during
12 the simulations to achieve equal temperature rise values per catalyst bed. The following
13
14 conditions were used for the simulations: LHSV 1.5 h-1, feed temperature 380–395ºC,
15
16 quench gas temperature 70ºC, pressure 11 MPa, and H2/oil ratio 5000–5300 scf/bbl. The
17
18 specified H2/oil ratio represents the total amount of hydrogen fed to the reactor, including
19 quench gas. Since the product gas coming out of the hydrocracking reactor is normally
20
21 purified by amine treatment and mixed with high-purity make-up hydrogen, it can be
22
23 reasonably assumed that the recycle/quench gas composition is 100% hydrogen.
24
25
26
27 4.1. Verification of the stochastic heat balance equation
28
29 Before implementing the developed model, we first verified the mathematical
30
31 consistency of the proposed heat balance equation (equation 14) by comparing the
32 stochastic simulation results with that from deterministic simulation based on ordinary
33
34 differential equations. To this end, we developed a simplified case study considering
35
36 selected hydrocracking reactions shown in Figure 1: (1) hydrogenation of naphthalene
37
38 into tetralin, (2) hydrogenation of tetralin into decalin, (3) isomerization of tetralin into
39
methylindane, and (4) ring opening of methylindane into n-butylbenzene. The reaction set
40
41 was chosen to illustrate a system comprising reactions in series and parallel under non-
42
43 isothermal conditions. Since the starting reactants, naphthalene and tetralin, are present in
44
45 low concentrations in the VGO feedstock, their initial concentrations were adjusted to
46
47
obtain a visible reaction performance. The reactor was specified to have one catalyst bed
48 and the thermophysical properties of the gas and liquid phases were kept constant along
49
50 the reactor for simplicity. Deterministic mass and heat balance equations were developed
51
52 for the case study and the solution was compared to the respective stochastic simulation.
53
54 Concentration profiles obtained with the two approaches are presented in Figure
55 4. The relative concentration values on the y-axis represent the component concentrations
56
57 rescaled by the total initial concentration of the starting reactants. Temperature profiles
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 47 Industrial & Engineering Chemistry Research

1
2
3
generated at two inlet temperatures are shown in Figure 5. It is observed that the
4
5 stochastic simulation, for the most part, converges to the deterministic solution,
6
7 confirming the validity of the developed approach. The slight deviations from the
8
9 deterministic solution in various sections along the reactor are caused by the inherent
10
11
random noise of stochastic simulation methods. Such divergences can be corrected by
12 averaging repeated stochastic simulations, as pointed out by de Oliveira et al.50 Therefore,
13
14 multiple repetitions of the hydrocracker simulations were required in order to eliminate
15
16 any defects associated with the simulation method.
17
18
19
20 4.2. Hydrocracker simulations
21
22 Reactor simulations were conducted at different 343°C+ conversion levels with
23
24
the purpose of describing the typical range of commercial hydrocracking operations,
25 ranging from mild to high severity (∼20–80 wt%)1. Three scenarios were developed: low
26
27 conversion (35 wt%), intermediate conversion (50 wt%), and high conversion (65 wt%).
28
29 The total catalyst volume was split into four catalyst beds with three inter-bed quench
30
31 points. Catalyst bed lengths were adjusted via trial-and-error simulations to ensure that
32
the four catalyst beds had the same temperature rise.
33
34 Table 5 presents the simulation results for the three scenarios. The specified
35
36 conversions of 343°C+ material are achieved by controlling the inlet reactor temperature
37
38 between 381–393°C. Conversion is observed to change significantly in a narrow range of
39
40 weight average bed temperature (WABT): 35 wt% at 386.0°C, 50 wt% at 393.1°C, and
41 65 wt% at 398.3°C. The temperature rise (delta-T) provoked by the exothermicity of
42
43 hydrocracking reactions is an increasing function of conversion. Delta-Ts are within the
44
45 range of 30–37°C, which is in agreement with available commercial operation data
46
47 presented in the literature (36°C from Bhutani et al.10 and 30°C from Zhou et al.11). It
48
should be pointed out that the delta-Ts presented in Table 5 are model predictions based
49
50 on actual thermochemical data, as opposed to previous studies8–11 where temperature
51
52 profiles are matched to industrial data by tuning the global heat of reaction.
53
54 Another important feature of the developed model is its ability to track hydrogen
55
56
consumption from the hydrogen stoichiometry of hydrocracking reactions. From the set
57
58
59
60 15
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 16 of 47

1
2
3
of simulations shown in Table 5 it is seen that the model predicts the expected trend for
4
5 hydrogen consumption with respect to conversion.
6
7 Also given in Table 5 are the commercial product distributions and properties of
8
9 the total liquid product (TLP). It is noted that naphtha selectivity is very sensitive to
10
11
reactor temperature. Increments of ~9–10 wt% in naphtha yield are observed for the
12 relatively small temperature steps evaluated in the simulations. Diesel yield, on the other
13
14 hand, is virtually fixed (36.7–39.0 wt%) in the range of conditions studied, indicating that
15
16 it is an intermediate product. LPG (C3 + C4) is found to vary in a fairly small range (1.4–
17
18 2.5 wt%). Inspection of hydrocarbon type distributions shows a significant increase in
19 paraffinic content from 3.2 wt% in the feed (Table 1) to 17.6–29.5 wt% in the TLPs. The
20
21 correspondence between the trends for paraffins and naphtha indicates that paraffins
22
23 preferentially concentrate in the boiling range of naphtha. Aromatics content is brought
24
25 down via hydrogenation from 35.4 wt% to 19.8–23.2 wt%. Naphthenic content is reduced
26
27
as well, but to a much lesser extent. The model also estimates product density from the
28 individual densities of the molecules in the liquid product. Results show that liquid
29
30 density decreases as more heavy material is converted into light distillates with higher
31
32 saturated hydrocarbon contents, consistent with our bench-scale experimental data.
33
34 Axial reactor profiles of temperature, product yield, H2/oil ratio, and oil
35 vaporization for the 65 wt% conversion scenario are presented in Figures 6 through 12.
36
37 Figure 6 illustrates temperature profile along the reactor. For the proposed catalyst bed
38
39 layout, the temperature rise in each catalyst bed is 9.2ºC. Since the heat release is faster
40
41 in the front segment of the reactor, shorter catalyst beds are required in this segment to
42
43
equalize temperature distribution. The three temperature drops correspond to the quench
44 gas injection points located at relative reactor lengths of 0.09, 0.29, and 0.59. Quench gas
45
46 rates are adjusted to cool the products to the inlet temperature (T0 = 393ºC). It should be
47
48 pointed out that the simulated temperature profile needs to be verified using actual plant
49
50 data.
51 Figure 7 shows the evolution of product yields of different fractions along the
52
53 reactor. The yields are obtained by grouping the molecules in the reacting stream by
54
55 boiling range. The VGO fraction is progressively consumed to generate diesel, naphtha,
56
57 and LPG. Naphtha and LPG yields increase steadily down through the catalyst beds,
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 47 Industrial & Engineering Chemistry Research

1
2
3
while diesel production reaches a plateau in the last catalyst bed (relative reactor length
4
5 of 0.59–1.0).
6
7 The evolution of aromatic classes along the reactor is presented in Figure 8. The
8
9 shift from one aromatic class to another one proceeds via sequential hydrogenation.
10
11
Saturation of tetraaromatics is rapid and achieves equilibrium quickly, while
12 hydrogenation of tri- and di-aromatics is more gradual. Monoaromatics reach a maximum
13
14 in the second catalyst bed (relative reactor length 0.09–0.29) followed by a continuous
15
16 decline in the last two catalyst beds. Figure 9 illustrates the axial reactor profiles for
17
18 cycloparaffinic classes. Cracking of cycloparaffinic ring cores occurs in a ring-by-ring
19 manner. It is seen that the process is selective towards mono- and di-cycloparaffin
20
21 formation. These products are also stripped from their alkyl branch fragments as reactions
22
23 move forward, becoming light enough to contribute to naphtha and diesel. Figure 10
24
25 shows the evolution of paraffin reactions. n-Paraffins are generated through ring
26
27
dealkylations and are subsequently isomerized to mono- and multibranched paraffins.
28 The latter are cracked into smaller n-paraffins and monobranched isoparaffins, as
29
30 schematized in Figure 2. From these profiles it is noted that the paraffinic fraction of the
31
32 cracked product is largely composed of linear and monobranched paraffins plus moderate
33
34 quantities of multibranched paraffins.
35 H2/oil ratio is a process variable with a characteristic profile, as shown in Figure
36
37 11. The constant decline in H2/oil ratio along each catalyst bed is a result of chemical
38
39 hydrogen consumption. Quench gas addition, consisting primarily of hydrogen, is
40
41 reflected in the H2/oil ratio profile as a pronounced jump in each quench point. Besides
42
43
its cooling effect, gas quenching also contributes to maintaining an adequate hydrogen
44 concentration/partial pressure along the reactor.
45
46 The hydrocarbon vaporization profile of the hydrocracker is displayed in Figure
47
48 12. It is observed that the incoming feed is partially vaporized (18 wt%) at the reactor
49
50 inlet. The temperature rise in each catalyst bed together with the generation of more
51 volatile hydrocarbons causes vaporization to increase continuously, reaching 62.5 wt% at
52
53 the reactor outlet. The three small jumps in vaporization corresponding to the quench
54
55 points (relative reactor lengths of 0.09, 0.29, and 0.59) are prompted by the large volumes
56
57 of quench gas being injected, as discussed in Figure 11. Hydrocarbon condensation
58
59
60 17
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 18 of 47

1
2
3
would normally be expected after cooling with quench gas, but in this case the effect of
4
5 H2/oil ratio on vaporization is stronger than that of temperature, resulting in additional
6
7 vaporization.
8
9
10
11 4.3. Effect of latent heat of vaporization
12
13 To evaluate the effect of latent heat of vaporization (LHV), simulations were
14
15 conducted without considering this term in the energy balance equation. Simulated
16 temperature profiles with and without LHV for the 65 wt% conversion scenario (see
17
18 Table 5) are compared in Figure 13. It is noted that the model predicts a much greater
19
20 temperature rise (45.3ºC vs 36.8ºC) when vaporization is excluded from the energy
21
22 balance. This indicates that vaporization removes a meaningful portion of the heat
23
24
released by hydrocracking reactions. Moreover, the amount of heat absorbed is not
25 constant since vaporization has a strong dependence on reactor temperature, pressure, and
26
27 H2/oil ratio. From this observation it can be deduced that the net heat release in a
28
29 hydrocracking reactor is determined by a complex interaction between chemical reactions
30
31 and phase equilibrium. Therefore, reactor models accounting for liquid vaporization are
32 well suited for responding to variations in process conditions.
33
34
35
36 Conclusions
37
38 The simulation approach for commercial vacuum gas oil hydrocrackers developed
39
40 in this study is based on a molecular-level representation of feedstock composition and
41
42 the reaction network while accounting for vapor-liquid equilibrium and heat release
43
44
kinetics. A distinctive feature of the model is the implementation of a stochastic heat
45 balance equation for simulating the thermal behavior of the hydrocracking reactor. The
46
47 model was applied to analysis of hydrocracker performance over a wide range of
48
49 conversion levels (35, 50, and 65 wt%). Simulations showed that the model produces the
50
51 expected trends for product distribution and properties as well as for process parameters
52 such as hydrogen consumption and temperature rise along the catalyst beds. Reactor
53
54 temperature was observed to have a major impact on process performance. It was also
55
56 determined that hydrocarbon vaporization has a marked effect on the heat balance of the
57
58 hydrocracking reactor, confirming its importance in the modeling of hydroprocessing
59
60 18
ACS Paragon Plus Environment
Page 19 of 47 Industrial & Engineering Chemistry Research

1
2
3
reactors. The model has potential to be used in the design and optimization of commercial
4
5 hydrocrackers. Nevertheless, the ability of the model to make accurate predictions needs
6
7 to be validated using actual plant data.
8
9
10
11 Corresponding author
12
13 *Tel.: (+1) 780-987-8348; fax: (+1) 780-987-5349; email: anton.alvarez-
14
15 majmutov@canada.ca
16
17
18
19 Notes
20
21 The authors declare no competing financial interest.
22
23
24
25
Acknowledgments
26
27
Partial funding for this study was provided by Natural Resources Canada and the
28 government of Canada’s interdepartmental Program of Energy Research and
29
30 Development (PERD). Comments and suggestions from Dr. Mohamed Ali on revising
31
32 the manuscript are greatly appreciated. The authors are grateful to the pilot plant and
33
34 analytical lab staff at CanmetENERGY in Devon.
35
36
37
Nomenclature
38
39
aadA, badA, cadA = Acid center adsorption parameters
40
41 aadM, badM, cadM = Metal center adsorption parameters
42
43 aadP = Paraffins adsorption parameter
44
45 aeqA, beqA = Acid center equilibrium parameters
46
47 aeqM, beqM = Metal center equilibrium parameters
48 aiq = Paraffins structural rate factor
49
50 asM, bsM, csM = Hydrogenation surface rate parameters
51
52 asRD, bsRD = Dealkylation surface rate parameters
53
54 asRI, bsRI, csRI = Ring isomerization surface rate parameters
55
asRO, bsRO, csRO = Ring opening surface rate parameters
56
57 CG = Total concentration of the gas phase, kmol/m3
58
59
60 19
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 20 of 47

1
2
3
CGi = Component concentration in the gas phase, kmol/m3
4
5 CL = Total concentration of the liquid phase, kmol/m3
6
7 CLi, CLj, CLq = Component concentrations in the liquid phase, kmol/m3
: 5 , :  , : K = Heat capacities, kcal/kg·K
8
9
10
11 FG, FL = Superficial mass velocities, kg/m2·s
12
13 fG, fL, fQ = Mass flow rates, kg/s
14 G0 = Hydrocarbon feed superficial mass velocity, kg/m2·s
15
16 KadAi = Component acid center adsorption coefficient, m3/kmol
17
18 KadMi = Component metal center adsorption coefficient, m3/kmol
19
20 kAiq = Combined rate coefficient of paraffins, m3/kgcat·s
21
22
KeqAiq = Equilibrium constant of acid center reactions, kmolj/kmoli
23 KeqMij = Equilibrium constant of metal center reactions, kmolj/kmoli·atmni
24
25 kRefA = Type A cracking reference rate coefficient
26
27 kRefMB = Mono-branched paraffins isomerization reference rate coefficient
28
29 kRefDB = Di-branched paraffins isomerization reference rate coefficient
30 kRefTB = Tri-branched paraffins isomerization reference rate coefficient
31
32 ksAiq = Surface rate coefficient of acid center reactions, kmol/kgcat·s
33
34 ksMij = Surface rate coefficient of metal center reactions, kmol/kgcat·s·atmni
35
36 KVLEi = Component vapor-liquid equilibrium partition coefficient
37
ma = Number of reactant “a” molecules
38
39 mb = Number of product “b” molecules
40
41 mGi = Number of component i molecules in the gas phase
42
43 mLi = Number of component i molecules in the liquid phase
44
45 ML = Liquid phase molecular weight, kg/kmol
46 ni = Hydrogen stoichiometry
47
48 NARi = Number of aromatic rings
49
50 NPCi = Number of paraffinic carbon atoms
51
52 NSCi = Number of saturated carbons
53
P = Reactor pressure, MPa
54
55 '() = Hydrogen partial pressure, atm
56
57 R = Universal gas constant, kcal/mol·K
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 47 Industrial & Engineering Chemistry Research

1
2
3
rij = Rate of metal center reactions, kmol/kgcat·s
4
5 riq = Rate of acid center reactions, kmol/kgcat·s
6
7 T = Reactor temperature, °C or K
8
9 uL = Liquid phase superficial velocity, m/s
10
11
xi = Component mole fraction in the liquid phase
12 yi = Component mole fraction in the gas phase
13
14 z = Axial reactor coordinate, m
15
16
17
18 Subscripts
19 G = Gas phase
20
21 i = Component i
22
23 ij = Metal center reaction ij
24
25 iq = Acid center reaction iq
26
27
ik = Hydrocracking reaction ik
28 j = Metal center reaction product j
29
30 k = Hydrocracking reaction product k
31
32 L = Liquid phase
33
34 out = Outlet
35 Q = Quench gas
36
37 q = Acid center reaction product q
38
39 s = Reaction step s
40
41 0 = Inlet
42
43
44 Greek Symbols
45
46 ∆HC+iq = Heat of formation of carbenium ions of acid center reactions, kcal/mol
47
48 ∆Hrij = Heat of reaction of metal center reactions, kcal/mol
49
50 ∆Hriq = Heat of reaction of acid center reactions, kcal/mol
51 ΔBC*D = Heat release of hydrocracking reactions, kcal/kmol of H2
52
53 ∆ts = Time step, s
54
55 ∆Ts = Temperature step, K
56
57 ∆zs = Axial position step, m
58
59
60 21
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 22 of 47

1
2
3
4
λV = Heat of vaporization, kcal/kg
5 ρb = Catalyst bulk density, kg/m3
6
7 ρL = Liquid phase density, kg/m3
8
9 τ = Space time
10
11 φV = Fraction of vaporized hydrocarbons, kg/kg
12
13
14 Abbreviations
15
16 EoS = Equation of state
17
18 IBP = Initial boiling point
19
20 FBP = Final boiling point
21
LHHW = Langmuir-Hinshelwood-Hougen-Watson
22
23 LHSV = Liquid hourly space velocity
24
25 LHV = Latent heat of vaporization
26
27 LPG = Liquefied petroleum gas
28
29
QSRC = Quantitative structure/reactivity correlation
30 TLP = Total liquid product
31
32 VGO = Vacuum gas oil
33
34 VLE = Vapor-liquid equilibrium
35
36 WABT = Weight average bed temperature
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 22
ACS Paragon Plus Environment
Page 23 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 References
5
6 (1) Robinson, P.R.; Dolbear, G.E. Hydrotreating and hydrocracking: fundamentals. In
7
8
Practical advances in petroleum processing Vol. 1; Hsu, C.S., Robinson, P.R., Eds.;
9 Springer: New York, 2006; Chapter 7.
10
11 (2) Jaffe, S.B. Kinetics of heat release in petroleum hydrogenation. Ind. Eng. Chem.,
12
13 Process Des. Develop. 1974, 13, 34–39.
14
15 (3) Yan, T.Y. Dynamics of a trickle-bed hydrocracker with a quenching system. Can. J.
16 Chem. Eng. 1980, 58, 259–266.
17
18 (4) Gruia, A. Recent advances in hydrocracking. In Practical advances in petroleum
19
20 processing Vol. 1; Hsu, C.S., Robinson, P.R., Eds.; Springer: New York, 2006;
21
22 Chapter 8.
23
24
(5) Ho, T.C. Kinetic modeling of large scale systems. Catal. Rev. Sci. Eng. 2008, 50,
25 287–378.
26
27 (6) Ancheyta, J.; Alvarez-Majmutov, A.; Leyva, C. Hydrotreating of oil fractions. In
28
29 Multiphase catalytic reactors: Theory, design, manufacturing, and applications;
30
31 Onsan, Z.I.; Avci, A.K., Eds.; John Wiley & Sons: Hoboken, N.J., 2016; Chapter
32 13.
33
34 (7) Döhler, W.; Rupp, M. Comparison of performance of an industrial VGO-treater
35
36 with reactor model predictions. Chem. Eng. Technol. 1987, 10, 340–352.
37
38 (8) Mohanty, S.; Saraf, D.N.; Kunzru, D. Modeling of a hydrocracking reactor. Fuel
39
Process. Technol. 1991, 29, 1–17.
40
41 (9) Pacheco, M.A.; Dassori, C.G. Hydrocracking: An improved kinetic model and
42
43 reactor modeling. Chem. Eng. Comm. 2002, 189, 1684–1704.
44
45 (10) Bhutani, N.; Ray, A.K.; Rangaiah, G.P. Modeling, simulation, and multi-objective
46
47
optimization of an industrial hydrocracking unit. Ind. Eng. Chem. Res. 2006, 45,
48 1354–1372.
49
50 (11) Zhou, H.; Lu, J.; Cao, Z.; Shi, J.; Pan, M.; Li, W.; Jiang, Q. Modeling and
51
52 optimization of an industrial hydrocracking unit to improve the yield of diesel or
53
54 kerosene. Fuel 2011, 90, 3521–3530.
55 (12) Quann, R.J. Modeling the chemistry of complex petroleum mixtures. Environ.
56
57 Health Persp. 1998, 106, 1441–1448.
58
59
60 23
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 24 of 47

1
2
3
(13) Chen, J.; Mulgundmath, V.; Wang, N. Accounting for vapor-liquid equilibrium in
4
5 the modeling and simulation of a commercial hydrotreating reactor. Ind. Eng.
6
7 Chem. Res. 2011, 50, 1571–1579.
8
9 (14) Froment, G.F. On fundamental kinetic equations for chemical reactions and
10
11
processes. Curr. Opin. Chem. Eng. 2014, 5, 1–6.
12 (15) de Oliveira, L.P.; Hudebine, D.; Guillaume, D.; Verstraete, J.J. A review of kinetic
13
14 modeling methodologies for complex processes. Oil Gas Sci. Technol. 2016, 71, 45.
15
16 (16) de Oliveira, L.P.; Verstraete, J.J.; Kolb, M. A Monte Carlo modeling methodology
17
18 for the simulation of hydrotreating processes. Chem. Eng. J. 2012, 207–208, 94–
19 102.
20
21 (17) de Oliveira, L.P.; Verstraete, J.J.; Kolb, M. Simulating vacuum residue
22
23 hydroconversion by means of Monte-Carlo techniques. Catal. Today 2014, 220–
24
25 222, 208–220.
26
27
(18) Alvarez-Majmutov, A.; Chen, J.; Gieleciak, R. Molecular-level modeling and
28 simulation of vacuum gas oil hydrocracking. Energy Fuels 2016, 30, 138–148.
29
30 (19) Neurock, M.; Libanati, C.; Nigam, A.; Klein, M.T. Monte Carlo simulation of
31
32 complex reaction systems: Molecular structure and reactivity in modelling heavy
33
34 oils. Chem. Eng. Sci. 1990, 45, 2083-2088.
35 (20) Boduszynski, M.M. Composition of heavy petroleums. 1. Molecular weight,
36
37 hydrogen deficiency, and heteroatom concentration as a function of atmospheric
38
39 equivalent boiling point up to 1400°F (760°C). Energy Fuels 1987, 1, 2–11.
40
41 (21) Boduszynski, M.M. Composition of heavy petroleums. 2. Molecular
42
43
characterization. Energy Fuels 1988, 2, 597–613.
44 (22) Alvarez-Majmutov, A.; Chen, J.; Gieleciak, R.; Hager, R.; Heshka, N.; Salmon, S.
45
46 Deriving the molecular composition of middle distillates by integrating statistical
47
48 modeling with advanced hydrocarbon characterization. Energy Fuels 2014, 28,
49
50 7385–7393.
51 (23) Alvarez-Majmutov, A.; Gieleciak, R.; Chen, J. Deriving the molecular composition
52
53 of vacuum distillates by integrating statistical modeling with detailed hydrocarbon
54
55 characterization. Energy Fuels 2015, 29, 7931–7940.
56
57
58
59
60 24
ACS Paragon Plus Environment
Page 25 of 47 Industrial & Engineering Chemistry Research

1
2
3
(24) Gillespie, D.T. A general method for numerically simulating the stochastic time
4
5 evolution of coupled chemical reactions. J. Comput. Phys. 1976, 22, 403–434.
6
7 (25) Mederos, F.S.; Ancheyta, J.; Chen, J. Review on criteria to ensure ideal behaviors in
8
9 trickle-bed reactors. Appl. Catal. A 2009, 355, 1–19.
10
11
(26) Landau, R.N.; Korre, S.C.; Neurock, N.; Klein, M.T.; Quann, R.J. Hydrocracking of
12 heavy oils : Development of structure/reactivity correlations for kinetics. Am.
13
14 Chem. Soc. Div. Fuel Chem. Prepr. 1992, 37, 1871–1877.
15
16 (27) Kumar, H.; Froment, G.F. A generalized mechanistic kinetic model for the
17
18 hydroisomerization and hydrocracking of long-chain paraffins. Ind. Eng. Chem.
19 Res. 2007, 46, 4075–4090.
20
21 (28) Korre, S.C.; Klein, M.T.; Quann, R.J. Polynuclear aromatic hydrocarbons
22
23 hydrogenation. 1. Experimental reaction pathways and kinetics. Ind. Eng. Chem.
24
25 Res. 1995, 34, 101–117.
26
27
(29) Korre, S.C.; Klein, M.T.; Quann, R.J. Hydrocracking of polynuclear aromatic
28 hydrocarbons. Development of rate laws through inhibition studies. Ind. Eng.
29
30 Chem. Res. 1997, 36, 2041–2050.
31
32 (30) Du, H.; Fairbridge, C.; Yang, H.; Ring, Z. The chemistry of selective ring-opening
33
34 catalysts. Appl. Catal. A 2005, 294, 1–21.
35 (31) McVicker, G.B.; Daage, M.; Touvelle, M.S.; Hudson, C.W.; Klein, D.P.; Baird,
36
37 W.C.; Cook, B.R.; Chen, J.G.; Hantzer, S.; Vaughan, D.E.W.; Ellis, E.S.; Feeley,
38
39 O.C. Selective ring opening of naphthenic molecules. J. Catal. 2002, 210, 137–148.
40
41 (32) Martens, J.A.; Jacobs, P.A.; Weitkamp, J. Attempts to rationalize the distribution of
42
43
hydrocracked products. I Qualitative description of the primary hydrocracking
44 models of long chain paraffins in open zeolites. Appl. Catal. 1986, 20, 239–281.
45
46 (33) Froment, G.F. Kinetics of the hydroisomerization and hydrocracking of paraffins on
47
48 a platinum containing bifunctional Y-zeolite. Catal. Today 1987, 1, 455–473.
49
50 (34) Quann, R.J.; Jaffe, S.B. Structure–oriented lumping: Describing the chemistry of
51 complex hydrocarbon mixtures. Ind. Eng. Chem. Res. 1992, 31, 2483–2497.
52
53 (35) Korre, S.C.; Klein, M.T. Development of temperature-independent quantitative
54
55 structure/reactivity relationships for metal- and acid-catalyzed reactions. Catal.
56
57 Today 1996, 31, 79–91.
58
59
60 25
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 26 of 47

1
2
3
(36) Mochida, I.; Yoneda, Y. Linear free energy relationships in heterogeneous catalysis.
4
5 I. Dealkylation of alkylbenzenes on cracking catalysts. J. Catal. 1967, 7, 385–392.
6
7 (37) Neurock, M.; Klein, M.T. Linear free energy relationships in kinetic analyses:
8
9 Applications of quantum chemistry. Polyc. Arom. Compd. 1993, 3, 231–246.
10
11
(38) Korre, S.C.; Neurock, M.; Klein, M.T.; Quann, R.J. Hydrogenation of polynuclear
12 aromatic hydrocarbons. 2. Quantitative structure/reactivity correlations. Chem. Eng.
13
14 Sci. 1994, 49, 4191–4210.
15
16 (39) Sie, S.T.; Senden, M.M.G.; van Wechem, H.M.H. Conversion of natural gas to
17
18 transportation fuels via the Shell Middle Distillate Synthesis Processes (SMDS).
19 Catal. Today 1991, 8, 371–394.
20
21 (40) Denayer, J.F.; Baron, G.V.; Souverijns, W.; Martens, J.A.; Jacobs, P.A.
22
23 Hydrocracking of n-alkane mixtures on Pt/H-Y zeolite: Chain length dependence of
24
25 the adsorption and the kinetic constants. Ind. Eng. Chem. Res. 1997, 36, 3242–
26
27
3247.
28 (41) Sie, S.T. Acid-catalyzed cracking of paraffinic hydrocarbons. 3. Evidence for the
29
30 protonated cyclopropane mechanism from hydrocracking/hydroisomerization
31
32 experiments. Ind. Eng. Chem. Res. 1993, 32, 403–408.
33
34 (42) Möller, K.; le Grange, P.; Acolla, C. A two-phase reactor model for the
35 hydrocracking of Fischer-Tropsch-derived wax. Ind. Eng. Chem. Res. 2009, 48,
36
37 3791–3801.
38
39 (43) Rosetti, I.; Gambaro, C.; Calemma, V. Hydrocracking of long chain linear paraffins.
40
41 Chem. Eng. J. 2009, 154, 295–301.
42
43
(44) Laxmi Narasimhan, C.S.; Thybaut, J.W.; Martens, J.A.; Jacobs, P.A.; Denayer, J.F.;
44 Marin, G.B. A unified single-event microkinetic model for alkane hydroconversion
45
46 in different aggregation states on Pt/H-USY-Zeolites. J. Phys. Chem. B 2006, 110,
47
48 6750–6758.
49
50 (45) Li, B.; Calemma, V.; Gambaro, C.; Baron, G.V.; Denayer, J.F. Competitive
51 adsorption of C20–C36 linear paraffins on the amorphous microporous silica–
52
53 alumina ERS-8 in vapor phase and liquid phase. Ind. Eng. Chem. Res. 2010, 49,
54
55 7541–7549.
56
57
58
59
60 26
ACS Paragon Plus Environment
Page 27 of 47 Industrial & Engineering Chemistry Research

1
2
3
(46) Chen, J.; Wang, N.; Mederos, F.; Ancheyta, J. Vapor-liquid equilibrium study in
4
5 trickle-bed reactors. Ind. Eng. Chem. Res. 2009, 48, 1096–1106.
6
7 (47) Munteanu, M. C.; Chen, J. Vapor-liquid equilibrium (VLE)-based modeling for the
8
9 prediction of operating regimes in a heavy gas oil hydrotreater. Energy Fuels 2012,
10
11
26, 1230–1236.
12 (48) Alvarez-Majmutov, A.; Chen, J. Modeling and simulation of a multibed
13
14 hydrotreater with vapor-liquid equilibrium. Ind. Eng. Chem. Res. 2014, 53, 10566–
15
16 10575.
17
18 (49) Stull, D.R.; Westrum, E.F.; Sinke, G.C. The chemical thermodynamics of organic
19 compounds; Wiley: New York, 1969.
20
21 (50) de Oliveira, L.P.; Verstraete, J.J.; Kolb, M. Development of a general modelling
22
23 methodology for vacuum residue hydroconversion. Oil Gas Sci. Technol. 2013, 68,
24
25 1027–1038.
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 27
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 28 of 47

1
2
3
Table 1. Experimental and simulated properties of the vacuum gas oil feedstock
4
5 Property Experimental Simulated
6 3
7 Density at 15.6ºC, g/cm 0.9115 0.9136
8 C, wt% 87.48 87.29
9
10 H, wt% 12.60 12.60
11 S, wt% 0.09 0.10
12
N, wppm 129 98
13
14 Hydrocarbon classes, wt%
15
16 Paraffins 3.2 3.9
17 Cycloparaffins 61.3 62.7
18
19 Monoaromatics 18.6 17.1
20 Diaromatics 9.1 8.3
21
22 Polyaromatics 7.7 8.0
23 Simulated distillation, °C
24
25 IBP 168 177
26 10 wt% 292 286
27
28 30 wt% 352 353
29 50 wt% 396 397
30
70 wt% 440 439
31
32 90 wt% 494 495
33
FBP 567 577
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 28
ACS Paragon Plus Environment
Page 29 of 47 Industrial & Engineering Chemistry Research

1
2
3
Table 2. QSRCs for aromatics hydrocracking reaction families
4
5 Reactivity parameter Correlation Descriptor
6
7 Adsorption
8 8T%" U.V + W%" UXY ;
ln $%" = S%" +
Z<
9 Metal sites
NARi - number of aromatic rings
10
8T%. U.V + W%. UXY ;
ln $%. = S%. +
11 NSCi - number of saturated carbons
Z<
Acid sites
12
13
Chemical equilibrium
14
1
ln $+,"# = S+,"  + H − T+," I ∆BC#
15 ni - hydrogen stoichiometry
Z<
16 Hydrogenation
∆Hrij - heat of reaction
17
1
ln $+,./ = S+,. + HT+,. − I ∆BC/
18 Ring isomerization
Z<
19 ∆Hriq - heat of reaction
Ring opening
20
21 Surface rate
22 ∆BC#
ln ! "# = S " + T "  + W " & -
ni - hydrogen stoichiometry
Z<
23 Hydrogenation
24 ∆Hrij - heat of reaction
25 T V[ W V[
ln ! ./ = S V[ + + \1 − ] ∆BY^/
Z< Z<
26 Ring isomerization
∆HC+iq - heat of formation of
T V_ W V_
27
ln ! ./ = S V_ + + \1 − ] ∆BY^/
28 carbenium ion intermediate
Z< Z<
Ring opening
29
T V`
30
ln ! ./ = S V` + ∆BY^/
Ring dealkylation ∆HC+iq - heat of formation of
Z<
31
32 leaving alkyl carbenium ion
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 29
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 30 of 47

1
2
3
Table 3. Reactivity parameters for paraffin hydrocracking reactions
4
5 Reactivity parameter Correlation Descriptor
6
7 Adsorption
8 Acid sites ln $%. = S%.:
9
10 Rate
11 Isomerization
!./ = 8U:Y − S, ;!V+abc , S, = 3
12
13 Monobranched paraffins

!./ = 8U:Y − S, ;!V+aec , S, = 4


14
15 Dibranched paraffins

!./ = 8U:Y − S, ;!V+agc , S, = 5


16
17 Tribranched paraffins
18 Cracking NPCi - paraffin chain length
19
20 Type A !./ = 8U:Y − S, ;!V+ai , S, = 7 aiq - non-reacting carbons
!V+ai
21
!./ = 8U:Y − S, ; ,S = 6
214 ,
22 Type B1
23
!V+ai
!./ = 8U:Y − S, ; ,S = 6
24
600 ,
25 Type B2
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 30
ACS Paragon Plus Environment
Page 31 of 47 Industrial & Engineering Chemistry Research

1
2
3
Table 4. Heat release of cracking and hydrogenation reactions
4
5 Reaction family ∆Hrn, kcal/kmol of H2 H2 stoichiometry
6 3
7 Paraffin cracking −10.4×10 1
8 Ring opening −7.4×103 1
9 3
10 Ring dealkylation −7.0×10 1
3
11 Benzenic-type hydrogenation −15.4×10 3
12
Naphthalenic-type hydrogenation −16.7×103 2
13
3
14 Phenanthrenic-type hydrogenation −14.9×10 1
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 31
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 32 of 47

1
2
3
Table 5. Hydrocracker simulation results
4
5 Scenario Low conversion Intermediate conversion High conversion
6
7 35 wt% 50 wt% 65 wt%
8 Operational parameters
9
10 Inlet temperature, ºC 381.5 388.5 393.0
11 WABT, °C 386.0 393.1 398.3
12
Total delta-T, °C 30.4 34.2 36.8
13
14 H2 consumption, scf/bbl 447 584 709
15
Product distribution
16
17 LPG, wt% 1.4 1.9 2.5
18
Naphtha (IBP–204°C), wt% 14.4 24.5 33.1
19
20 Diesel (204–343°C), wt% 36.7 37.7 39.0
21 Unconverted oil (343°C+), wt% 48.3 37.0 26.6
22
23 TLP properties
24 Liquid density, g/cm3 0.8658 0.8474 0.8314
25
26 Paraffins, wt% 17.6 24.6 29.5
27 Naphthenes, wt% 59.2 54.2 50.7
28
29 Aromatics, wt% 23.2 21.2 19.8
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 32
ACS Paragon Plus Environment
Page 33 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 Figure captions
5
6 Figure 1. Key reactions for aromatics hydrocracking.
7
8
Figure 2. Paraffin hydrocracking reactions.
9 Figure 3. Parity plots comparing experimental data and model predictions. (a) Product
10
11 distribution by hydrocarbon type; (b) product distribution by boiling point, API gravity,
12
13 and hydrogen consumption.
14
15 Figure 4. Concentration profiles at T0 = 380ºC, LHSV = 1.5 h-1, P = 11 MPa, and H2/oil =
16 5300 scf/bbl. (Lines) deterministic simulation; (circles) stochastic simulation.
17
18 Figure 5. Reactor temperature rise profile at LHSV = 1.5 h-1, P = 11 MPa, and H2/oil =
19
20 5300 scf/bbl. (Lines) deterministic simulation; (circles) stochastic simulation.
21
22 Figure 6. Reactor temperature profile at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa, and
23
24
H2/oil = 5300 scf/bbl.
25 Figure 7. Evolution of commercial products at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa,
26
27 and H2/oil = 5300 scf/bbl.
28
29 Figure 8. Evolution of aromatic classes at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa, and
30
31 H2/oil = 5300 scf/bbl.
32 Figure 9. Evolution of cycloparaffinic classes at T0 = 393ºC, LHSV = 1.5 h-1, P = 11
33
34 MPa, and H2/oil = 5,300 scf/bbl.
35
36 Figure 10. Evolution of paraffinic classes at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa,
37
38 and H2/oil = 5,300 scf/bbl.
39
Figure 11. H2/oil profile at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa, and H2/oil = 5,300
40
41 scf/bbl.
42
43 Figure 12. Oil vaporization profile at T0 = 393ºC, LHSV = 1.5 h-1, P = 11 MPa, and
44
45 H2/oil = 5,300 scf/bbl.
46
47
Figure 13. Reactor temperature profiles with and without LHV at T0 = 393ºC, LHSV =
48 1.5 h-1, P = 11 MPa, and H2/oil = 5,300 scf/bbl.
49
50
51
52
53
54
55
56
57
58
59
60 33
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 34 of 47

1
2
3
Naphthalenic-type
4 hydrogenation Ring isomerization Ring opening Dealkylation
5
6 + 2H2 + H2 + H2
7 +
8
9
Benzenic-type + 3H2 + 3H2 + 3H2
10
hydrogenation
11
12
13 + H2 + H2
+
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 Figure 1
58
59
60 34
ACS Paragon Plus Environment
Page 35 of 47 Industrial & Engineering Chemistry Research

1
2
3
Isomerization Isomerization Isomerization
4
5
6
7
8 Isomerization + H2 + H2
9
10
11 + +
12
13 B1-type cracking A-type cracking
+ H2
14
15
16
+
17
18 B2-type cracking
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 Figure 2
57
58
59
60 35
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 36 of 47

1
2
3
4 70
5 (a)
6 60
7
8 50
9
Model, wt%

10
11 40
12
13 30
14 n-Paraffins
Isoparaffins
15 20 Naphthenes
16 Monoaromatics
17 Diaromatics
18 10
Triaromatics
19
20 0
21 0 10 20 30 40 50 60 70
22
23 Experimental, wt%
24
25
26 Experimental, scf/bbl
27
28 0 200 400 600 800 1000 1200 1400
29 70 1400
30 (b)
31
60 1200
32
33
Model, wt% or °API

34 50 1000 Model, scf/bbl


35
36 40 800
37
38
30 600
39 LPG
40 Naphtha
41 20 Diesel 400
42 VGO
43 10 API Gravity 200
44 H2 consumption
45
46 0 0
47 0 10 20 30 40 50 60 70
48 Experimental, wt% or °API
49
50
51
52
53
54
55
56 Figure 3
57
58
59
60 36
ACS Paragon Plus Environment
Page 37 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 0.8
5 Naphthalene
6 0.7 Tetralin
Methylindane
7
Relative concentration

0.6 Decalin
8 n-Butylbenzene
9
10 0.5
11
12 0.4
13
14 0.3
15
16 0.2
17
18 0.1
19
20 0.0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 4
58
59
60 37
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 38 of 47

1
2
3
4 +11
5 +10 T0 = 380°C
6 T0 = 390°C
7 +9
Temperature rise, °C

8 +8
9
10 +7
11 +6
12
13 +5
14 +4
15
16 +3
17 +2
18
+1
19
20 T0
+0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 5
58
59
60 38
ACS Paragon Plus Environment
Page 39 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 410
5
6
7 405
8
Temperature, °C

9
10 400
11
12
13 395
14
15
16
17 390
18
19
20 385
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 6
58
59
60 39
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 40 of 47

1
2
3
4 100
5 90 LPG
6 Naphtha
7 80 Diesel
8 VGO
70
9
Yield, wt%

10 60
11
12 50
13
40
14
15 30
16
17 20
18 10
19
20 0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 7
58
59
60 40
ACS Paragon Plus Environment
Page 41 of 47 Industrial & Engineering Chemistry Research

1
2
3
4
30
5 Monoaromatics
6 Diaromatics
25
7 Triaromatics
8 Tetraaromatics
9 20
Yield, wt%

10
11
12 15
13
14
10
15
16
17 5
18
19
20 0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 8
58
59
60 41
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 42 of 47

1
2
3
4
30
5 Monocycloparaffins
6 Dicycloparaffins
25 Tricycloparaffins
7
8 Tetracycloparaffins
9 20
Yield, wt%

10
11
12 15
13
14
10
15
16
17 5
18
19
20 0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 9
58
59
60 42
ACS Paragon Plus Environment
Page 43 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 16
5 n-Paraffins
6 14 Monobranched
7 Dibranched
8 12 Tribranched
9
10
Yield, wt%

10
11
12 8
13
14 6
15
16 4
17
18 2
19
20 0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 10
58
59
60 43
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 44 of 47

1
2
3
4 5500
5
6
7 5000
H2/oil ratio, scf/bbl

8
9
10 4500
11
12
13
4000
14
15
16
17 3500
18
19
20 3000
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 11
58
59
60 44
ACS Paragon Plus Environment
Page 45 of 47 Industrial & Engineering Chemistry Research

1
2
3
4 70
5
6 60
7
Oil vaporization, wt%

8
50
9
10
11 40
12
13 30
14
15
20
16
17
18 10
19
20 0
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 12
58
59
60 45
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 46 of 47

1
2
3
410
4
5 without LHV
6 with LHV
7 405
8
Temperature, °C

9
10 400
11
12
13 395
14
15
16
17 390
18
19
20 385
21 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
22
23 Relative reactor length
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 Figure 13
58
59
60 46
ACS Paragon Plus Environment
Page 47 of 47 Industrial & Engineering Chemistry Research

1
2
3
For Table Contents Only
4
5
6
7
8
9
10
11
12 Oil + Hydrogen
13 1.0
14
15 0.8
Concentration
16
17 0.6
18
19 0.4
20
21 0.2
22
23
0.0
24
25 Products 0.0 0.2 0.4 0.6 0.8 1.0
26 Relative reactor length
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 47
ACS Paragon Plus Environment

You might also like