Uploaded by chud.44

2

advertisement
See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/314716075
The Increase in the Yield of Light Fractions During the Catalytic Cracking of C13C40 Hydrocarbons
Article in Current Organic Synthesis · March 2017
DOI: 10.2174/1570179413666161031121659
CITATIONS
READS
6
516
5 authors, including:
Nataliya Belinskaya
Elena Nikolaevna Ivashkina
Tomsk Polytechnic University
Tomsk Polytechnic University
41 PUBLICATIONS 135 CITATIONS
99 PUBLICATIONS 445 CITATIONS
SEE PROFILE
All content following this page was uploaded by Elena Nikolaevna Ivashkina on 29 April 2019.
The user has requested enhancement of the downloaded file.
SEE PROFILE
Send Orders for Reprints to reprints@benthamscience.ae
Current Organic Synthesis, 2017, 14, 1-12
1
RESEARCH ARTICLE
The Increase in the Yield of Light Fractions During the Catalytic Cracking of C13-C40
Hydrocarbons
Elena Ivashkinaa, Galina Nazarovaa*, Emiliya Ivanchina a, Nataliya Belinskaya3a and Stanislav Ivanovb
a
Institute of Natural Resources, Department of Fuel Engineering and Chemical Cybernetics, National Research Tomsk Polytechnic
University, Tomsk, Russia; bChemical and Biochemical Engineering Department, Faculty of Engineering, Western University,
London ON, Canada
ARTICLEHISTORY
Received: November 27, 2015
Revised: December 03, 2015
Accepted: January 12, 2016
DOI:
Abstract: The work performed constituted thermodynamic kinetic analysis of the chemical transformations of
C13–C40 hydrocarbons during the catalytic cracking process. Laboratory research to determine the structural
grouped and individual composition of feedstock and products of catalytic cracking allowed a list to be made of
catalytic cracking process reactions. Using Density Functional Theory, the thermodynamic parameters of the
process reactions were determined and a formalized scheme of hydrocarbons transformations compiled based
upon which the kinetic model of the process was documented and implemented programmatically. The determination of kinetic parameters of the reactions was carried out by solving the inverse kinetic problem using experimental data from an industrial plant and laboratory studies. The use of a kinetic model of catalytic cracking allowed changes in the concentration of the reactants to be calculated as well as the yield and composition of the
catalytic cracking products and ensured the selection of optimum conditions for increasing the yield of gasoline
and light gas oil fractions based on group composition of raw material for catalytic cracking.
Keywords: Advanced petroleum refining, catalytic cracking, group composition of raw materials, kinetic model, light fraction yield, resource
efficiency.
1. INTRODUCTION
Currently, increasing the depth of oil refining and developing
highly efficient catalyst compositions are the main targets for oil
refining globally [1-4], since the proportion of hard-to-extract bituminous oil as well as the need for motor fuels (gasoline and diesel)
is increasing every year [5, 6].
At the same time, plants have to adjust to significant changes in
the properties of raw materials, which is a kind of challenge arising
in the design and construction of the new technologies of deep oil
processing and in the optimization of processing units in operation.
Catalytic cracking is the complex physical and chemical process of converting high molecular weight hydrocarbons into valuable
products such as gasoline fraction, light gas oil, and gas through
contact with powdered zeolite catalyst in the technological conditions of the process [7].
Catalytic cracking being used in refineries around the world, as
well as other processes aimed at the deepening of oil refining, results in a high degree of conversion of raw petroleum products –
85–95% [8-13].
It is important to note that the production of commercial gasoline in large Russian and foreign plants involves 30–40% of catalytic cracking gasoline [14]. At the same time, the content of unsaturated hydrocarbons, benzene, and sulfur in the resulting product
is one of the defining criteria in blending gasoline of different
grades, the amount of which in commercial gasolines largely depends on the quality of catalytic cracking gasolines.
The purpose of this research is to increase the yield of light
fractions during the catalytic cracking of C13–C40 hydrocarbons.
*Address correspondence to this author at the Institute of Natural Resources, National
Research Tomsk Polytechnic University, P.O. Box: 634050, Tomsk, Russia; Tel/Fax:
+7-952-181-8788; E-mail: silko@tpu.ru
1570-1794/17 $58.00+.00
The thermodynamic and kinetic laws of the catalytic cracking
process can be reflected in a formalized schema of hydrocarbon
transformations and serve as a basis for a method that increases the
conversion rate and depth of oil refining. Thus, the need to satisfy
the needs of the industrial sector in the solution of complex, interrelated, and multi-variable problems in predicting the structure and
properties of the products of catalytic cracking, depending on the
hydrocarbon composition of the feedstock and technological mode
catalytic reactor. A number of mathematical models based on a
formalized mechanism for catalytic cracking reactions are commonly used for this purpose.
There are different approaches to the formalization of the
chemical reactions in the catalytic cracking process with varying
levels of detail that determine the adequacy of the calculations performed using the model and depend directly on the stages of the
technological process development and the evolution of the catalyst
compositions used in catalytic cracking (fixed bed reactor using an
amorphous catalyst vs. riser type reactor using a microspherical
zeolite catalyst).
In the development of a formalized scheme for hydrocarbon
conversion in the deep processing of crude oil, it is important to
take into account the many factors that equally affect the composition, quantity, and quality of the target products (technological
mode of process, feedstock being multi-component, the type of
catalyst, the pairing of devices comprising the technological
scheme, etc.) [15-17]. In describing the hydrocarbon transformation
scheme, some sort of technological aggregation on certain parameters is used.
Weekman and Nace [18] on the basis of work of Wei and Prater
[19] being among the first to establish a formalized hydrocarbon
transformations scheme, identified three groups of substances:
process feedstock, gas oil, and residues (dry gas and coke). The
conversion of the cracked feedstock has been studied under isothermal conditions in fixed bed, moving bed, and fluidized bed
© 2017 Bentham Science Publishers
2 Current Organic Synthesis, 2017, Vol. 14, No. 0
reactors. This approach allowed the yield of gas oil in catalytic
cracking to be estimated, and the kinetic parameters were determined using experimental data.
The most popular approach to formalizing schemes of chemical
transformations in these processes is technological aggregation into
pseudo-components, based mainly on fractional composition [2024], but the hydrocarbon composition of the fractions in each group
is diverse, and this approach does not take into account the varying
reactivity of the hydrocarbons within the selected groups. For example, Corella and Francés in [25] designed a five-lump component scheme in which aggregation was performed, based on boiling
points accounting for hydrodynamic effects and catalyst deactivation. Lee et al. in [26] developed a four-lump hydrocarbon transformation scheme by flows in a catalytic cracking unit with a separate catalyst regeneration step.
In the group approach, components of the reaction system are
formalized in accordance with the molecular structure of the hydrocarbons, and the least probable reactions are excluded from the
scheme in order to simplify it. This approach is characterized by
information about the reaction mechanism with the averaged reactive ability of hydrocarbon groups. The complexity of the group
approach to the processes of deep refining of petroleum feedstock
lies in determining the group composition of heavy fractions of the
feed stream (fractional composition of 350–570°C). Therefore,
preference is given to hybrid models that take into account not only
the pseudo-components interaction but also the chemical conversion of main hydrocarbon groups in the feed stream of the catalytic
cracking process unit [27-30].
The first efforts in this direction were made by Jacob et al. [31].
The hydrocarbons conversion scheme contains paraffins, naphthenes, aromatic rings, substituted aromatic groups for light and heavy
petroleum fractions, and coke+gas and gasoline groups. The model
accounts for the dependence of the catalyst activity on time, adsorption of aromatic compounds, and nitrogen poisoning of the catalyst.
Subsequently, the hybrid and flow schemes of hydrocarbon
conversion in the catalytic cracking process were supplemented and
enhanced by dividing the components into different reaction
groups, introducing coke deposits on the catalyst into the scheme
and hydrodynamic calculations for the flow in the reaction system
[32-39]. For example, Oliveira, Biscaia and Gao et al [40, 41] introduced into the scheme a separation of the aromatic crude oil
component into resins, asphaltenes, saturated hydrocarbons, and
aromatics as well as a separation of the coke formation stage into
the two stages of gas formation and a description of the turbulent
flow of gas-solid particles and reactions in the reactor.
Ivashkina et al.
calculation of the thermodynamic parameters of the reactions, it is
important to take into account the technological operating conditions of the industrial plant and the effects of the interaction of hydrocarbon molecules with the molecules of the reaction mixture.
For that reason, quantum chemical calculations are widely used
in studies of oil refining and petrochemical processes [51]—for
example, to calculate the electronic structure of molecules, transition states, entropy and enthalpy of substances, heat of adsorption,
their connection with the activation energies, etc. [52, 53]. The
application of quantum chemistry methods helps to identify mechanisms for the substances' formation on the surface of the catalyst
and the stages of transitional states formation [54, 55].
In this paper, thermodynamic and kinetic analysis of the transformations of C13–C40 hydrocarbons in the catalytic cracking process has been performed. The thermodynamic parameters of the
catalytic cracking reactions are determined using Density Functional Theory, and the kinetic model is recorded in accordance with
a formalized hydrocarbon conversion scheme, developed on the
basis of thermodynamic analysis and laboratory research.
2. ABOUT CATALYTIC CRACKING
2.1. Catalytic Cracking Technology
The catalytic cracking process is used for the production of
gasoline and diesel fractions via catalytic cracking of vacuum distillate hydrocarbons or a mixture of residues of secondary processes
(350–570°C fractions).
The technological process is implemented in a riser reactor. The
main parameters that characterize the reactor operation are the
consumption of feedstock by the plant (160–365 m3/h), the pressure
in the reactor sludge zone (0.8–1.5) kgf/cm2), the temperature of
raw feed (240–350°C), and the product temperature at the exit of
the ballistic separator (495–535°C).
The catalytic cracking plant products are rich gas and gasoline
fraction, light gas oil of catalytic cracking origin (195–310°C fraction), and heavy gas oil (310–420°C fraction) of catalytic cracking
processes.
Note that the feedstock is characterized by a high concentration
of paraffins and naphthenes and promotes high yields of high octane gasoline and liquefied gases due to an intense hydrogen redistribution reaction, in which naphthenic and olefinic hydrocarbons
interact to form iso-paraffins and aromatic hydrocarbons (Scheme
1). At the same time, it is important to optimize the speed of hydrogen transport, as it contributes to the formation of aromatic hydrocarbons and, accordingly, coke.
Hydrocarbon conversion scheme [31] has also been supplemented by Barbosa et al. in [42] in 3D CFD modeling and studying
the effect of gas particles in a turbulent mixture stream on the heat
effect of the reaction and the quality and yield of gasoline.
Most of today’s models take into account the catalyst deactivation feature [43-47]. Such models allow estimation of the conversion, yield of catalytic cracking products, and amount of coke on
the catalyst. A detailed description of the chemical transformations
during catalytic cracking, including the catalyst composition and
prediction of coke amount on the catalyst surface, is found in works
of Froment and Dewachtere et al. [48, 49]. Bidabehere and Sedran
[50] developed a catalytic cracking model, taking into account the
theory of instantaneous adsorption of diffusion effects, the chemical
reaction, and the theory of deactivation.
It should be noted that the formalized scheme of hydrocarbon
conversion should take into account the thermodynamic parameters
of the catalytic cracking reactions, the values of which are practically impossible to find in literature in the case of reactions involving high molecular weight hydrocarbons. At the same time, in the
H2
C
C
H2
H2
C
C
H2
H2
CH3
H
+H C C C C CH
3
3
H
CH3
H2
C
C
H2
CH3
H2
C
C
H2
H
CH3
CH3 + 2H
C
+ HC
2
C
3
H2
CH3
Scheme 1.
2.2. Catalytic Cracking Reactions
Chemical transformations of the raw feed being cracked occur
according to a carbonium ion mechanism through chemisorption of
the raw product hydrocarbon molecules to the surface of the microspherical acid catalyst [56-61]. Carbonium ions are formed by heterolytic C-H bond breaking with the tertiary carbonium ion, which
is highly reactive being the most stable (Scheme 2).
The Increase in the Yield of Light Fractions During the Catalytic Cracking
H+
C
CH3
+ HX
H2C
olefin
carbonium
ion
CH3
carbonium ion
R2
H+
C
RH +
C
H2
suturated
hydrocarbon
C
H2
R1
suturated
hydrocarbon
carbonium
ion
- Cracking reactions of carbonium ion by -bond in relation to
the positively charged carbon atom (Scheme 3);
R
H2
C
R1
C
H2
R
CH2
C
H
+ +H2C
R1
Scheme 3.
- Reaction of ion isomerization by shifting the hydrogen atom
and skeletal isomerization caused by a shift of the methyl group
with the primary carbocation isomerized faster than the decomposition by -bond occurs (Scheme 4);
H2
C
CH+
H3C
C
H2
+H+
H
C
CH+
H3C
H
C
H3C
CH3
H2
C
H-shift
H2
C
C
H2
H-shift
C
H3C
CH3
H2
C
CH+
H3C
C
H
CH3
C
H
CH3
CH+
H
CH3
H2
C
H3C
C+
To determine the group and the structural-group composition of
raw materials and light and heavy gas oil of catalytic cracking origin, a liquid-adsorption chromatographic separation on ASK silica
gel with a grain size of 0.2–0.3 mm according to VNII NP (AllRussia Research Institute for Oil Processing) methodology was
conducted. Hydrocarbon fractions were sequentially extracted by
desorption using solvents with different polarity (hexane, hexane
and toluene mixture, and ethyl alcohol and toluene mixture). Hydrocarbons were separated by refractive index using an IRF-22
refractometer into a paraffin-naphthenic fraction and an aromatic
hydrocarbons fraction. The structural-group composition of paraffinic-naphthenic fractions was determined using the n-d-m method
and, for the aromatic hydrocarbons fraction, using the Hazelwood
method.
The molecular weight of the feedstock and products of catalytic
cracking was determined using CRYETTE WR laboratory equipment, which registers the freezing point of the samples.
To determine the individual composition of the gasoline fraction, laboratory studies were conducted via gas-liquid chromatography using a CHROMATEC-CRYSTAL 5000 mod. 2 gas chromatographer with a flame ionization detector running Chromatec Analyst software for the management, collection, and processing of
chromatographic data, capillary column DB-1, 100·0.25·0.5.
H-shift
H2
C
H-shift
CH3
CH+
R2
As analytical equipment to determine the qualitative composition of raw materials used in the catalytic cracking, a Hewlett Packard 6890 Gas Chromatograph System gas chromatography-mass
spectrometer with a 5973 Mass Selective Detector running GC
ChemStation software was used.
CH3
C+
R1
The total sulfur content in the feedstock was determined using
the energy dispersive X-ray fluorescence sulfur analyzer SPECTRO-SCAN SL.
CH3
-H+
H2
C
+H+
-H+
H2
C
CH3
H3C
H2
C
CH3
C
H2
H2C
CH3
The development of a formalized scheme for hydrocarbon conversion during the catalytic cracking process requires information
about the group, quantitative and qualitative composition of the
plant feed stream, individual composition of the gasoline fraction,
and the structural-group and group composition of light and heavy
gas oil.
R2
The main reactions of catalytic cracking with carbonium ions
H2
C
CH3
3. MATERIALS AND METHOD
are:
CH+
R2
+
Scheme 5.
Scheme 2.
H2
C
R2 + H+
+
R+ + LH- + H2
carbonium ion
RH + L
paraffin Lewis
acid
R+ + R1
CH
R+ + X- + H2
paraffin Bronsted
acid
H2
C
CH3
+ X-
carbonium ion
RH + HX
H R1
CH
R1
H3C
acid catalyst
Current Organic Synthesis, 2017, Vol. 14, No. 0 3
H2
C
CH3
CH3
Scheme 4.
- Reactions of aromatic hydrocarbons dealkylation to form
aromatic hydrocarbon and carbonium ion (Scheme 5);
- Hydrogen redistribution reaction, in which one hydrocarbon is
saturated and an increase in the degree of unsaturation of the second
hydrocarbons occurs (Scheme 5).
The thermodynamic parameters of the catalytic cracking reactions were calculated using Gaussian software, which implements
quantum chemistry methods for calculating the electronic structure
of molecules. To calculate the thermodynamic parameters of the
catalytic cracking reactions, the DFT (Density Functional Theory)
method, based on the methods of quantum mechanics, and the
semi-empirical PM3 method, which uses parameters obtained from
experimental data to simplify calculations, was utilized.
The DFT method was chosen as the main method of calculation. The B3LYP (Becke’s density functional theory (B3) model
using Lee-Young-Parr (LYP) electron correlation) with basis 3-21G
were used as theoretical approximation. Models of the substances
involved in the reactions have been built using GaussView software.
4 Current Organic Synthesis, 2017, Vol. 14, No. 0
Ivashkina et al.
Table 1. Results of determining molecular weight and the group composition of catalytic cracking feedstock and products.
Hydrocarbon groups, % wt.
Unit flow
Molecular weight, g/mol
Paraffins
Resins/
Aromatics
Naphtenes
asphaltenes
Raw materials
382.43
65.4
31.10
3.5/0.00
Light gasoil
164.71
21.55
76.14
1.68/0.00
Heavy gasoil
234.15
18.9
703.09
6.00/2.00
Table 2. The results of structural-group composition of catalytic cracking raw materials and products, %.
Raw Materials
Light gasoil
Heavy gasoil
Parameter
P+N
A
b
P+N
a
A
b
P+N
a
Ab
Ca
2.59
27.52
17.10
56.81
1.42
50.79
Cn
31.27
55.73
53.84
43.19
31.56
49.21
Cp
66.14
16.76
29.05
0.00
67.01
0.00
Ka
0.14
1.40
0.44
1.20
0.062
1.90
Kn
2.05
3.90
1.55
2.60
1.53
4.90
2.59
5.3
1.96
3.8
1.59
6.8
Ko
a
a
b.
Paraffin and naphtenic fractions; Aromatic fractions
Fig. (1). Examples of naphthenic hydrocarbon structure. (methylalkylcyclohexanes, dimethylalkylcyclohexanes bicycloalkanes, et al.).
4. EXPERIMENTAL PART
4.1. Laboratory Investigation of Catalytic Cracking Raw
Materials and Products
The results of determining the group and structural-group composition of feedstock and products of catalytic cracking using gasadsorption chromatography on silica gel are shown in Tables 1 and
2.
The feedstock of the catalytic cracking process has a high content of saturated hydrocarbons, about 61.2% by weight. According
to the data on the structural-group composition of the paraffinicnaphthenic fraction of the feedstock obtained by the n-d-m method,
the fraction contains some aromatic structures, whose presence in
the saturated part is explained by the fact that their sorbability by
silica gel is almost equal to that of the naphthenes, and therefore
they are washed away with solvent during adsorptive separation.
The carbon content in the aromatic structures of the paraffinicnaphthenic fraction is 2.59%. The content of carbon in the
naphthenic rings is 31.27%, and the average number of naphthene
rings is 2.05. The sulfur content in the feed is low, 0.0461% by
weight.
According to the results of chromato-mass spectrometry, the
paraffinic hydrocarbons of the feedstock are presented by hydro-
carbons with the number of carbon atoms of 13–40. Naphthenic
hydrocarbons are mono- and poly-cycloalkanes with long alkyl
substituents of normal and iso-structure with the number of carbon
atoms in the alkyl substituents C1–C25 (Fig. 1).
The content of aromatic hydrocarbons in the raw materials for
catalytic cracking is 35.57% by weight, and alcohol-benzene resin –
3.23% by weight. Analysis of aromatic concentrates by the Hazelwood method showed that aromatic concentrates are represented
predominantly by hybrid structures. Aromatic fractions extracted
from the raw material are characterized by a number of naphthene
rings (3.9) higher than the average number of aromatic rings (1.4),
with the total number of rings in raw materials being 5.3. The carbon content of the paraffin fragments of aromatic concentrates
(16.76%) indicates the presence of alkyl-substituted aromatic hydrocarbons and naphthenes with carbon content of aromatic and
naphthenic fragments of 27.52% and 55.73%. The simplest representatives of aromatic hydrocarbons in catalytic cracking raw materials are monoalkylbenzenes, methylalkylbenzenes, methylnaphthalenes and substituted naphthalenes, etc.
Light gas oil of catalytic cracking origin has a high aromatic
content (76.14 wt.%), with a tar content of 1.68 wt.%. According to
the structural-group composition of paraffin-naphthenic fractions of
light gas oil from catalytic cracking, the carbon content in
The Increase in the Yield of Light Fractions During the Catalytic Cracking
Current Organic Synthesis, 2017, Vol. 14, No. 0 5
Table 3. The results of group composition of catalytic cracking gasoline fraction, %.
Hydrocarbon Group
Concentration, % wt.
Paraffins
4.42
Isoparaffins
31.16
Olefins
19.13
Naphthenes
10.56
Monoromatics
34.70
Table 4. The compression results of calculation values of Gibbs energy with table values.
G, kJ/mol
Reaction
Table
DTF,
PM3
C8H18 C4H 10+ C4H8
– 55.12
– 60.79
– 66.62
n-716 2- methylhexane
– 2.4
–2.35
10.40
i-C4H10 CH4 + C3H6
– 44.83
– 39.37
–36.93
510 36 + 2 4
–8.56
–13.56
5.82
918 6 12 + 36
–17.97
–28.65
– 45.32
612 66 + 3 2
–87.81
–93.27
–311.9
naphthenic fragments of light gas oil is 53.1%. Naphthene hydrocarbons of light gas oil are mainly represented by monocyclic hydrocarbons since the average number of naphthene rings is 1.55.
According to the Hazelwood method, carbon content in aromatic structures of light gas oil is 56.81%. The carbon content of
naphthenic fragments of light gas oil is 43.19%. The average number of naphthene rings exceeds the average number of aromatic
rings, which indicates that the aromatic concentrates are represented
by hybrid structures.
Heavy gas oil from catalytic cracking has a high aromatic content of 70.24 wt.%, with the content of alcohol-benzene resins in
heavy gas oil being 8.05 wt.%, and reported presence of asphaltenes
- 1.76 wt.%) According to the structural-group composition, the
content of naphthenic rings exceeds the number of aromatic rings.
The average number of rings is 6.8.
The group composition of gasoline fractions, as determined by
gas chromatography, is presented in Table 3.
Gasoline fraction of catalytic cracking origin has a high content
of isoparaffins (31.16 wt. %), aromatic hydrocarbons (34.70 wt. %),
and olefins (19.13 wt.%).
3.2. Thermodynamic Analysis of Catalytic Cracking Reactions
Calculations to determine the thermodynamic parameters of individual hydrocarbon formation involved in the reactions of crude
oil catalytic cracking were performed, using quantum chemistry
methods.
The magnitude of change in the Gibbs energy (G) of the reaction characterizes the probability of their occurrence and is important to justify the level of formalization of the mechanism of the
chemical hydrocarbon transformation scheme.
The list of forecast reactions in the catalytic cracking process is
made based on the analysis of experimental data obtained from an
industrial plant, the group composition of raw materials and the
products of catalytic cracking process defined by chromatographic
separation on silica gel, structural-group composition using n-d-m
and Hazelwood methods, and the results of chromato-mass spectrometry and chromatography.
After comparing the results of calculation using non-empirical
DFT and semi-empirical PM3 methods with table values [62, 63]
(Table 4), it was determined that the Gibbs energy values of the
catalytic cracking reaction and enthalpy values calculated by DFT
present a more adequate result.
The list of reactions and the resulting thermodynamic parameters of reactions from the list are shown in Table 5.
According to the reported results of the cracking reaction of
paraffin wax (Gavg = –74.86 kJ/mol), redistribution of hydrogen
(Gavg = –111.76 kJ/mol), dehydrogenation of naphthenes (Gavg =
–124.63 kJ/mol), aromatic dealkylation hydrocarbons (Gavg = –
89.04 kJ/mol), and naphthenes (Gavg = –120.4 kJ/mol), and the
coke formation reactions (Gavg = –597.2 kJ/mol) are characterized
by the highest thermodynamic probability. When creating a formalized scheme of hydrocarbon transformations during the catalytic
cracking process, it is important to take into account the contribution of these reactions to the formation of the final products.
4. RESULTS AND DISCUSSIONS
4.1. Formalized Hydrocarbon Conversion Scheme
According to the results of laboratory research that aimed to determine the group, structural-group, and qualitative composition of
raw materials, group and structural-group composition of light and
heavy gas oil, the individual composition of the gasoline fraction,
and the results of the thermodynamic analysis, a summary table of
the hydrocarbon groups of formalized conversion scheme participating in the catalytic cracking reactions has been compiled (Table
6).
Figure 2 shows a detailed diagram of the hydrocarbon conversion with feedstock components and light and heavy gas oil. The
area indicated by the dotted line includes gasoline fraction hydrocarbons.
6 Current Organic Synthesis, 2017, Vol. 14, No. 0
Ivashkina et al.
Table 5. The calculation results of thermodynamic parameters of reactions (temperature 504 C, pressure 1.08 MPa)
Reaction
H, kJ/mol
G, kJ/mol
Cracking of high molecular weight n-paraffins C13–C40 (C16H 34 C8H18+C8H16)
69.4
–74.8
Cracking of high molecular weight i-paraffins C13–C40 (C16H 34 i-8 18 +8 16)
70.5
–70.8
Cracking of medium molecular weight n-paraffins (C7H16 C4H8 +C3H 6)
69.8
–62.3
Isomerization of medium weight paraffins (n-C7H16 i-C7H16)
–1.9
–2.3
Cracking of medium molecular weight i-paraffins (3-(3)-(H2)3-3) i-4 10+3 6)
62.1
–63.2
Cracking of olefins (714 5 10 +24)
94.2
–28.3
Hydrogen transfer ((3)3-69+ 5 10 (3)3-63 + i-512+2 H2)
99.3
–111.7
Dealkylation of naphthenes ((1021) 2-610 (10 21)2-64 + 32)
156.0
–120.4
Dealkylation of monoaromatic hydrocarbons ((1021)2-6 4 6 6+2·10 20)
157.8
–89.0
Dehydrogenation of naphthenes ((1021)2-610 (1021)2-6 4 + 32)
221.9
–124.6
Coke formation (polycondensation) (12 C10H8) 5 C24H12+18H2)
104.9
–597.2
Cyclization of olefins to naphthenes (714 714)
–53.8
–7.5
Table 6. Summary table of formalized hydrocarbon conversion scheme groups involved in the FCC reactions.
Hydrocarbon Groups
Qualitative Characteristics
Source
13–40
Feedstock, light and heavy gas oil
5–11+
Gasoline fraction
Isoparaffinic hydrocarbons
5–11+
Gasoline fraction
Olefin hydrocarbons
5–9+
Gasoline fraction
5–10+
Gasoline fraction
Cycloparaffinic hydrocarbons
Mono-and polynaphthenes with long substituents 1–25 (average number of naphthene
Paraffinic hydrocarbons
rings – 1.53–2.05)
6–11+
Aromatic hydrocarbons
Gasoline fraction
Mono and poly structures with long substituents (average number of aromatic rings – 1.20–
1.90).
Resins
Feedstock, light and heavy gas oil
Naphthene-aromatic structures (total number of rings – 3.8–6.8).
Feedstock, light and heavy gas oil
Feedstock, light and heavy gas oil
Fig. (2). Scheme of hydrocarbon conversion in the catalytic cracking process: kj – the rate constant of direct chemical reaction; k-j – the rate constant of the
reverse chemical reaction.
4.2. Kinetic Model Development and Determination of Kinetic
Parameters
The kinetic model is built on the basis of a formalized scheme
of hydrocarbon conversion in the catalytic cracking process and is
represented by a system of differential equations (1). The expres-
sions for the reaction rate constants are written according to the law
of mass action. The kinetic model of the catalytic cracking process
describes the change in concentration of the reactants depending on
contact time with the initial conditions: = 0, Ci = Ci0, where i is the
corresponding hydrocarbon. The differential equations system is
solved by the 4th order Runge-Kutta method.
The Increase in the Yield of Light Fractions During the Catalytic Cracking
dC paraffinsHMW
= k1C paraffinsHMW k2C paraffinsHMW ;
d
dC paraffinsMMW
= k1C paraffinsHMW k3C paraffinsMMW k4C paraffinsMMW +
d
+ k4Cisoparaffins ;
dÑisoparaffins = k C
2 paraffinsHMW + k 4 C paraffinsMMW k 4 Cisoparaffins d
k C
+ k7 Colefins Cnaphthenes ;
5 isoparaffins
dÑolefins
= k1C paraffinsHMW + k2C paraffinsHMW k6Colefins + k6C 2 gas d
k7 Cnaphthenes Ñî lefins + k8CnaphthenesHMW + k9Caromatics k13Cî lefins k9Cmonoaromatics Ñolefins + k13Cnaphthenes + k14Cmonoaromatics k14Cî lefins Ñmonoaromatics ;
dÑ
gas = 2k3C paraffinsMMW + 2k5Cisoparaffins + 2k6Cî lifins 2k6C 2 gas ;
d
dÑ
naphthenes = k C
8 naphthenesHMW k7 Cnaphthenes Ñî lefins + k13 C olefins d
k13 C naphthenes ;
dÑ
monoaromatics = k7 Cnaphthenes Ñî lefins + k9Caromatics k14 C monoaromatics d
k9Cmonoaromatics Ñî lefins + k 14 Cî lefins Ñmonoaromatics ;
dÑnaphthenesHMW
= k8CnaphthenesHMW k10CnaphthenesHMW ;
d
dÑaromatics = k C
9 aromatics + k 9 Cmonoaromatics Ñî lefins + k10 CnaphthenesHMW d
2
2k11C aromatics ;
dÑ
re sin s = k11C 2 aromatics k12Cre sin s ;
d
dÑcoke
= k12Cre sin s
d
dÑÍ 2
= 3k10CnaphthenesHMW + k11C 2 aromatics + 2k7 Colefins Ñnaphthenes .
d
Current Organic Synthesis, 2017, Vol. 14, No. 0 7
carbon group; – the duration of contact, sec; Wj – the rate of direct
chemical reaction; W-j – the rate of reverse chemical reaction, paraffins HMW and paraffins MMW – paraffins of high (C13–C40) and
medium (C5–C11+) molecular weight, and naphthenes HMW – high
molecular weight naphthenes.
The differential equations system (1):
The determination of kinetic parameters of reactions involves
finding the reaction rate constants. The determination of kinetic
parameters is determined by solving the inverse kinetic problem of
the forward chaining method. The activation energy values for the
above reactions are selected on the basis of the literature [64-69].
The input data for the inverse kinetic problem are presented in
Tables 1, 3, 7, and 8, and include a material balance of the process,
the technological conditions, and the composition of the raw materials and the product of the catalytic cracking unit.
The rate constants of chemical reactions that are included in the
kinetic model are a combination of the constants of all intermediate
stages (Table 9, where), and the resulting kinetic model is formalized and quasi-homogeneous.
According to Table 9, the reactions of hydrogen redistribution
(kdr= 25.48 l·s-1·mol-1), cracking of paraffin (kdr = 4.19·10-1 s-1) and
olefin hydrocarbons (kdr =1.51·10-1 s-1), dealkylation of naphthenic
(kdr =3.71·10-1 s-1) and aromatic hydrocarbons (kdr =1.82·10-1 s-1) as
well as the condensation of aromatic hydrocarbons (kdr = 1.42 l·s1
·mol-1) and coking (kdr =3.92·10-1 s-1) occur at the highest rate.
However, less intense reactions are isomerization of paraffinic hydrocarbons (kdr =2.90·10-4 s-1) and cyclization of olefins (kdr
=2.90·10-4 s-1).
(1)
The system of differential equations (1) uses the following notations: dCi – change in concentration of the respective i-th hydro-
4.3. Evaluate the Adequacy of Kinetic Models
Check of the model for adequacy was performed by determining the margin between the calculated and experimental data on
product yields and concentration of and hydrocarbon groups that
Table 7. Material balance of a catalytic cracking unit.
Flows
kg/h
%
Input: Vacuum distillate
272624.00
100.00
160456.32
58.86
Consumption of light gas oil
30851.00
11.32
Consumption of heavy oil
25137.00
9.22
Consumption of wet gas
44089.76
16.17
Coke
11668.30
4.28
Losses
421.60
0.15
Total
272624.00
100.00
Yield:
Unstable gasoline consumption
Table 8. The parameters of the technological process in the catalytic cracking reactor.
Process Parameters
3
Consumption of feedstock by the plant, m /h
Density g/cm
3
378.22
0.904
Total steam consumption in the reaction zone of the riser reactor, kg/h
Temperature at the reactor inlet, °C
Pressure, kgf/cm
Value
2
Ratio of catalyst to feedstock
7898.04
303.7
1.44
5.56
8 Current Organic Synthesis, 2017, Vol. 14, No. 0
Ivashkina et al.
Table 9. The kinetic parameters of catalytic cracking reactions (temperature 504 C, pressure 1.08 MPa).
Rate Constant, s-1
Ea,
Reaction
kJ/mol
Cracking of HMW n-paraffins
Cracking of HMW paraffins to form iso-paraffins
krr
kdr
-1
125.4
1.09·10
125.4
4.19·10 –
-1
–
Cracking of MMW n-paraffins
126.0
9.17·10 –
Isomerization of MMW n-paraffins
110.0
2.90·10-4
2.05·10-4
Cracking of MMW i-paraffins
126.0
9.33·10-3
–
Cracking of olefins
160.2
1.51·10 2.08·10-3
Redistribution of hydrogen
144.0
25.48*
–
119.0
3.71·10 –
Dealkylation of aromatics
152.0
1.82·10 1.50·10-5
Dehydrogenation of HMW naphthenes
144.0
6.28·10-2
–
Dealkylation of naphthenes
Condensation reactions of aromatics
-2
-1
-1
-1
*
127.0
1.42
127.0
3.92·10 –
Cyclization of olefins
127.0
2.90·10 2.50·10-3
Dealkylation of monoaromatics
152.0
2.03·10-2
2.00·10-6
Formation of coke (polycondensation)
–
-1*
-4
* - l·s-1 ·mol-1
Table 10. Comparison of the calculated and experimental group composition of gasoline for values 1.
Hydrocarbon Groups
Exp.value
Calculated value
Discrepancy (abs.)
Paraffins
4.42
4.12
0.30
Isoparaffins
31.16
31.9
0.73
Olefins
19.12
19.81
0.68
Naphthenes
10.56
11.78
1.21
Monoaromatics hydrocarbons
34.70
33.02
1.68
Table 11. Process conditions of catalytic cracking reactor.
Process Conditions
3
Feedstock consumption, m /h
Density of feedstock g/sm
3
Values 1
Values 2
Values 3
378.22
369.99
371.52
0.904
0.9009
0.9002
Total hydrogen consumption in the reaction zone of riser reactor, kg / h
7898.04
7799.9
7803.7
Reactor inlet temperature, °
303.7
303.2
313.3
Product temperature at the outlet from the ballistic separator, °
521.4
522.9
521
Pressure, kgs/sm2
1.44
1.33
1.47
The ratio of catalyst: raw materials
5.56
5.50
5.58
Paraffins and naphthenes
65.4
58.6
61.2
Aromatics
31.1
38.6
35.58
Resins/asphaltenes
3.5
2.8
3.23
are part of gasoline fraction (Table 10, 12). Process conditions for
calculation are presented in table 11.
3.4. Investigation of Technological Mode Influents on the Light
Fraction Yields
According to results presented in Tables 10, 11 an maximum
absolute calculation error on the kinetic model between calculation
and experimental of group composition is 1.68 % wt. and product
yield is not more than 0.62 % wt., that is comparable with the error
of experimental determination method of these parameters.
In order to increase the yield of light fractions from catalytic
cracking, calculations were performed using a software-based kinetic model. For example, the duration of contact between the raw
material and the catalyst has a significant effect on the products’
yield.
The Increase in the Yield of Light Fractions During the Catalytic Cracking
Current Organic Synthesis, 2017, Vol. 14, No. 0 9
Table 12. Calculation error between calculation and experimental of product yields.
Values 1
Exp.
Calculated
value
value
Unstable gasoline
58.86
58.24
Unit flow
Values 2
Exp.
Calculated
value
value
0.62
55.99
56.39
Error
Values 3
Exp.
Calculated
value
value
0.4
57.15
57.45
0.30
Error
Error
Light gasoil
11.32
11.41
0.09
12.72
13.13
0.31
11.38
11.56
0.18
Heavy gasoil
9.22
8.98
0.24
10.2
9.5
0.6
9.14
9.1
0.04
Rich gas
16.17
16.55
0.38
16.03
15.95
0.08
16.82
16.97
0.15
Coke
4.28
4.36
0.08
4.55
4.47
0.08
4.98
4.45
0.53
Loses
0.15
0.46
0.31
0.52
0.56
0.04
0.53
0.47
0.06
By increasing the contact duration, the depth of feedstock conversion, gasoline yield, and octane number pass through a maximum (Figs. 3-5), which corresponds to the theoretical concept of
the process described in literature [70, 71].
The depth of conversion in this case means the total yield
of gas, gasoline, and coke. Octane number of unstable gasoline was calculated on the kinetic model taking into account
non-additive depending on the hydrocarbon composition of
gasoline [14].
Fig. (5). Dependence of gasoline octane number on contact time.
cracking, and the gasoline yield initially increases (to a temperature of 535°C), and then gasoline yield decreases, “overcracking”
occurs, and the output of rich gas increases (Fig. 8). The maximum
theoretical yield of gasoline is 59.41%.
Fig. (3). Dependence of the depth of feedstock conversion on contact time.
Fig. (6). Dependence of unstable gasoline yield on reactor temperature.
Fig. (4). Dependence of unstable gasoline yield on contact time.
Studies on the effect of temperature on the performance of the
catalytic cracking process were carried out. The composition of
feedstock and products of catalytic cracking are shown in Tables 1
and 3. In the calculations, the other process parameters were not
changed and are consistent with the data presented in Table 8. The
calculation results are presented in Figures 6-11.
According to the calculations, increasing the process temperature increases the depth of conversion of crude oil from catalytic
Fig. (7). Dependence of octane number of unstable gasoline on reactor
temperature.
10 Current Organic Synthesis, 2017, Vol. 14, No. 0
Ivashkina et al.
the coke, light and heavy gas oil yield are presented in Figures 911.
To obtain a high yield of unstable gasoline, the process conditions should be organized so as to maintain the temperature at the
exit of the ballistic separator at 530°C; this allows for obtaining a
high yield of high-octane (94.83) gasoline fraction (59.3%), with a
coke yield of 4.51%.
Fig. (8). Dependence of rich gas yield on reactor temperature.
The yield of light gas oil decreases with increasing temperature
in the reactor due to the cracking of gas oil and the formation of
gasoline fraction hydrocarbons and gaseous products. When the
reactor temperature is maintained at 510°C, there is a high yield of
light gas oil (16.53%) and a low coke yield (3.82%). Since the coke
yield is an important characteristic that determines the process temperature, yield of the light fractions and the temperature mode of
the reactor - regenerator system, the resulting value output of coke
will not provide the desired temperature, so this mode is not desirable.
A suitable temperature for the technological process aimed at a
high yield of light gas (12.09%) will be 520°C, which maintains the
yield of gasoline and coke at 4.22% and 57.84%.
CONCLUSION
Fig. (9). Dependence of light gas oil yield on reactor temperature.
This paper describes the development of a formalized scheme
for hydrocarbon conversion in the course of vacuum distillate catalytic cracking, based on the results of laboratory research, to determine the quality and quantity of raw materials and products of the
catalytic cracking process as well as the thermodynamic parameters
of the reactions and the reactivity of the hydrocarbons determined,
using quantum chemistry calculation methods.
Based on a formalized hydrocarbon conversion scheme, the kinetic model of the process was recorded and implemented
programmatically. Kinetic regularities of hydrocarbon conversion,
found by solving the inverse kinetic problem, adequately describe
the existing laws of the catalytic cracking process.
Using the kinetic model of the process, a study was conducted
to determine the temperature in the catalytic cracking reactor aimed
at achieving a process that yields gasoline fraction and light gas oil
fraction.
To achieve an optimum yield of high-octane (94.8) gasoline
from catalytic cracking (59.30%), it is necessary to maintain the
temperature at the exit of the ballistic separator at 530°C. To
achieve an optimum yield of light gas oil from catalytic cracking
(12.09%), it is necessary to maintain the
Fig. (10). Dependence of heavy gas oil yield on reactor temperature.
CONFLICT OF INTEREST
The authors confirm that this article content has no conflict of
interest.
ACKNOWLEDGEMENTS
The work was supported by a the Government Contract "Science".
REFERENCES
[1]
[2]
Fig. (11). Dependence of coke yield on reactor temperature.
The yield of light and heavy gas oil decreases with temperature
increase since the light gas oil hydrocarbons are further cracked to
form light products thus increasing the reaction rate of coke formation and, consequently, the yield of coke. The calculation results on
[3]
[4]
Ramakanta Sahu; Byung Jin Song; Ji Sun Im; Young-Pyo Jeon; Chul Wee
Lee. A review of recent advances in catalytic hydrocracking of heavy
residues. Journal of Industrial and Engineering Chemistry, 2015, 27, 12–24.
Lappas, A.A.; Iatridis, D.K.; Papapetrou, M.C.; Kopalidou, E.P.; Vasalos
I.A. Feedstock and catalyst effects in fluid catalytic cracking – Comparative
yields in bench scale and pilot plant reactors. Chemical Engineering Journal,
2015, 278,140–149.
Corma, A.; Serra, J.M. Heterogeneous combinatorial catalysis applied to oil
refining, petrochemistry and fine chemistry. Catalysis Today, 2005, 107-108,
3-11.
Abbasov, V.; Mammadova, T.; Andrushenko, N.; Hasankhanova, N.;, Y.;
Abdullayev, E. Halloysite–Y-zeolite blends as novel mesoporous catalysts
The Increase in the Yield of Light Fractions During the Catalytic Cracking
[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]
for the cracking of waste vegetable oils with vacuum gasoil. Fuel, 2014, 117
Part A, 552–555.
Jian Chaia; Zhong-Yu Zhang; Shou-Yang Wang; Kin Keung Lai; John Liu
Aviation fuel demand development in China. Energy Economics, 2014, 46,
224–235.
Chèze B.; Gastineau P.; Chevallierd J. Forecasting world and regional
aviation jet fuel demands to the mid-term (2025). Energy Policy, 39(9),
5147–5158.
Doronin, V.P.; Lipin, P.V.; Sorokina, T.P. Effect of Process Conditions on
the Composition of Products in the Conventional and Deep Catalytic Cracking of Oil Fractions. Catalysis in Industry, 2012, 4 (2), 100–104.
Guo-Qiang Chen; Zheng-HongLuo. New insights into intra particle transfer,particle kinetics, and gas–solid two-phase flow in polydisperse fluid catalytic cracking riser reactors under reaction conditions using multi-scale modeling. Chemical Engineering Science, 2014, 109, 38-52.
Gladys Jiménez-García; Ricardo Aguilar-López; Rafael Maya-Yescas. The
fluidized-bed catalytic cracking unit building its future environment. Fuel,
2011, 90, 3531-3541.
Sildir, H.; Arkun, Y.; Cakal, B.; Gokce, D.; Kuzu, E. A dynamic nonisothermal model for a hydrocracking reactor: Model development by the
method of continuous lumping and application to an industrial unit. Journal
of Process Control, 2012, 22, 1956-1965.
Martínez, J.; Ancheyta, J.; Kinetic model for hydrocracking of heavy oil in a
CSTR involving short term catalyst deactivation. Fuel, 2012, 100, 193-199.
Mehlberg, R.L.; Upson, L.L.; Glavin, J.P. Fluid catalytic cracking system
and process. US 2015/0065774 A1, March 5, 2015.
Palmas, P. Process and apparatus for fluid catalyticcracking. US
2014/0364300 A1, December, 11, 2014.
Ivanchina, E.D.; Kirgina, M.V.; Chekantsev, N.V.; Sakhnevich, B.V.;
Sviridova, E.V.; Romanovskiy R.V. In: Chemical Engineering Journal, 2015,
282, pp. 194-205. In: Proceeding of the International Conference on Chemical Reactors “CHEMREACTOR-21”, Netherlands, Delft, September 22-25,
2014.
Frantsina, E.V.; Ivanchina, E.D.; Ivashkina, E.N.; Romanovskii, R.V. Developing of the mathematical model for controlling the operation of alkane dehydrogenation catalyst of linear alkyl benzene. Chemical Engineering Journal, 2014, 238 (1), 129-139
Zagoruiko, A.N.; Alexander S. Belyi, A.S.; Smolikov, M.D.; Noskov, A.S.
Unsteady-state kinetic simulation of naphtha reforming and coke combustion
processes in the fixed and moving catalyst beds. 2014. Catalysis Today, 220222, 168-177.
Belinskaya, N.; Ivanchina, E.; Ivashkina, E.; Frantsina, E.; Silko, G. Mathematical model of straight run diesel catalytic hydroisomerization. IOP Conference Series: Earth and Environmental Science, 2014, 21(1), 1-7.
Weekman Jr, V.W.; Nace, D.M. A Model of Catalytic Cracking Conversion
in Fixed, Moving, and Fluid -Bed Reactors, Industrial and Engineering
Chemistry Process Design and Development, 1968, 7(1), 90–95.
Wei, J.; Prater, .D. The structure and analysis of complex reaction systems.
Advances in Catalysis, 1962, 13, 203-215.
Zhang, J.; Wang, Z.; Jiang, H.; Chu, J.; Zhou, J.; Shao, S. Modeling fluid
catalytic cracking risers with special pseudo-components. Chemical Engineering Science, 2013, 102, 87-98.
Baudrez, E.; Heynderickx, G.J.; Marin, G.B. Steady-state simulation of Fluid
Catalytic Cracking riser reactors using a decoupled solution method with
feedback of the cracking reactions on the flow. Chemical Engineering Research and Design, 2010, 88, 290-303.
Dagde, K.K.; Puyate, Y.T. Six-lump kinetic modelling of adiabatic plug-flow
riser reactor in an industrial FCC unit. IJERA, 2012, 2 (6), 557-568.
Kai Xiong; Chunxi Lu; Zhifeng Wang; Xionghou Gao. Kinetic study of
catalytic cracking of heavy oil over an in-situ crystallized FCC catalyst, Fuel,
2015, 142, 65–72.
Dupain, X.; Gamas, E.D.; Madon, R.; Kelkar, C.P.; Makkee, M.; Moulijn,
J.A. Aromatic gas oil cracking under realistic FCC conditions in a microriser
reactor, Fuel, 2003, 82, 1559-1569.
Corella, J.; Francés, E. in: Fluid Catalytic Cracking II; Mario L. Occelli, Ed.;
American Chemical Society, ACS Symposium Series: Washington, DC,
1991, Vol. 452, 165-182.
Lee, L.; Chen, Y.; Huang, T.; Pan, W. Four-lump kinetic model for fluid
catalytic cracking process. The Canadian Journal of Chemical Engineering,
1989, 67, 615-619.
Das, A.K.; Baudrez, E.; Marin, G.B.; Heynderickx, G.J. Three-Dimensional
simulation of fluid catalytic cracking riser reactor. Ind. Eng. Chem.
Res., 2003, 42 (12), 2602–2617.
Kang, X.; Guo, X.; You, H. An Introduction to the lump kinetics model and
reaction mechanism of FCC gasoline / // Energy Sources. – Part A, 35:1921–
1928, 2013. – 343-358.
Radu,S.; Ciuparu. D. Modelling and Simulation of an Industrial Fluid Catalytic Cracking Unit. Revista de Chimie, 2014, 1, 113-119.
Lan, X.; Xu, C.; Wang, G.; Wu, L.; Gao J. CFD modeling of gas–solid flow
and cracking reaction in two-stage riser FCC reactors. Chemical Engineering
Science, 2009, 64, 3847 –3858.
Jacob, S.M.; Gross, B.; Voltz, S.E.; Weekman, V.W jr. A lumping and
reaction scheme for catalytic cracking. AIChE J., 1976, 22, 701-713.
Current Organic Synthesis, 2017, Vol. 14, No. 0 11
[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]
Gross, B.; Jacob, S.M.; Nace, D.M. Voltz, S.E. Simulation of catalytic cracking process. US 3960707, June 1, 1976.
Coxon, P.G.; Bischoff, K.B. Lumping strategy, Introduction to techniques
and application of cluster analysis, Ind. Eng. Chem. Res. 1987, 26, 12391248.
Theologos K.N.; Markatos N.C. Advanced modeling of fluid catalytic cracking riser-type reactors, AIChE Journal, 1993, 39 (6), 1007-1017.
Pitault, I.; Nevicato, D.; Forissier, M.; Bernard. J. Kinetic model based on a
molecular description for catalytic cracking of vacuum gas oil. Chem. Eng.
Sci, 1994, 49(24A), 4249-4262.
Derouin, C.; Nevicato, D.; Forissier, M.; Wild, G.; Bernard, J.R. Hydrodynamics of riser units and the impact on FCC operation. Industrial and Engineering Chemistry Research, 1997, 36, 45044515.
Sa, Y.; Liang, X.; Chen, X.; Liu J. Study of 13-lump kinetic model for residual catalytic cracking. Petrochem. Eng. Cor. 1995, 145-152.
Berry, T. A.; McKeen, T. R.; Pugsley T. S.; Dalai, A. K. Two dimensional
reaction engineering model of the riser section of a fluid catalytic cracking
unit. Industrial & Engineering Chemistry Research, 2004, 43(18), 55715581.
Carella, J. On the modelling of the kinetics of the selective deactivation of
catalysts. Application tj the fluid catalytic cracking process. Industrial & Engineering Chemistry Research, 2004, 43, 4080-4086.
Oliveira, L.; Biscaia. E.C. Catalytic cracking kinetic models parameter
estimation and model evaluation. Industrial & Engineering Chemistry Research. 1989. 28, 264-271.
Gao, J.; Xu, C.; Lin, S.; Yang, G.; Guo, Y. Advanced model for turbulent
gas-solid flow and reaction in FCC riser reactor. AIChE Journal, 1999,
45, 1095-1113.
Barbosa, A.C.; Lopes, G.C.; Rosa, L.M.; Mori, M.; Martignoni, W.P. Three
dimensional simulation of catalytic cracking reactions in an industrial scale
riser using a 11-lump kinetic. Chemical engineering transactions, 2013, 32,
637-642.
Sedighi, M.; Keyvanloo, K.; Towfighi, J. Kinetic study of steam catalytic
cracking of naphtha on a Fe/ZSM-5 catalyst. Fuel. 2013, 109, 432–438.
Bartholomew, C.H., Fuentes, G.A. Catalyst deactivation /Elsevier, 1997, p.
695.
Froment G. On fundamental kinetic equations for chemical reactions and
processes. Current Opinion in Chemical Engineering, 2014, 5, 1–6
Lopez-Isunza. F. Deactivation of fluid catalytic cracking catalysts: modelling
approach. Studies in Surface Science and Catalysts. 2001, 134,187–199.
Varshney, P.; Kunzru, D.; Gupta, S.K. Modelling of the riser reactor in a
resid fluidised-bed catalytic cracking unit using a multigrain model for an active matrix-zeolite catalyst. Indian Chemical Engineer. 2014, 21.
Froment, G.F.; Bartholomew C.H.; Fuentes G. (eds.) Catalyst Deactivation
1997, Elsevier, 53.
Dewachtere, N.V.; Froment, G.F.; Vasalos,I.; Markatos, N.; Skandalis, N.
Applied thermal engineering, 1997, 17(8-10), 837-844.
Bidabehere, C.M.; Sedran, U. Simultaneous diffusion, adsorption, and reaction in fluid catalytic cracking catalysts. Industrial & Engineering Chemistry
Research, 2001, 40, 530-535.
Deglmann, P.V.; Schafer, A.; Lennartz, C. Application of quantum Calculations in the Chemical Industry – an Overview. International Journal of
Quantum Chemistry, 2015, 115, 107-136.
Deglmann, P.; Schafer, A.; Lennartz, . Application of quantum
Calculations in the Chemical Industry – an Overview. International Journal
of Quantum Chemistry, 2015, 115, 107-136.
Roque-Malherbe, R.; Diaz-Castro. F.; Calculationof the energy adsorption of
n-paraffins in nanoporous crystalline and ordered acid catalysts, and its
relationships with the activation energy of the monomolecular catalytic
cracking reaction. Journal of Molecular Catalysis A: Chemical, 2008, 280,
194-202.
Tranca, D.C.; Hansen, N.; Swisher, J.A.; Smit, B.; Keil, F.J. Combined
Density Functional Theory and Monte Carlo Analysis of Monomolecular
Cracking of Light Alkanes Over H-ZSM-5. The Journal of Physical
Chemistry C, 2012, 116, 23408–23417.
Swisher, J.A.; Hansen, N.; Maesen, T.; Keil, F.J.; Smit, B.;Bell, A.T.
Theoretical Simulation of n-Alkane Cracking on Zeolites. The Journal of
Physical Chemistry C, 2010, 114, 10229-10239.
Doronin, V.P.; Lipin, P.V.; Potapenko, O.V.;
Sorokina, T.P.;
Korotkova, N.V.; Gordenko, V.I. Advanced developments: Cracking
catalysts and additives to them. Catalysis in Industry, 2014, 6(4), 307-311.
Awayssa, O.; Al-Yassir, N.; Aitani, A.; Al-Khattaf. S. Modified HZSM-5 as
FCC additive for enhancing light olefins yield from catalytic cracking of
VGO. Applied Catalysis A: General, 2014, 477, 172-183.
Siddiqui, M.A.B.; Aitani, A.M.; Saeed, M.R.; Al-Yassir, N.; Al-Khattaf, S.
Enhancing propylene production from catalytic cracking of Arabian Light
VGO over novel zeolites as FCC catalyst additives. Fuel, 2011, 90, 459–466.
Han, D.; Sun, N.; Liu, J.; Li, C.; Shan, H.; Yang. C. Synergistic effect of W
and P on ZSM-5 and its catalytic performance in the cracking of heavy oil.
Journal of Energy Chemistry, 2014, 23(4), 519–526.
Liu, X.; Yang, T.; Bai, P.; Han. L.Y/MCM-41 composites assembled from
nanocrystals. Microporous and Mesoporous Materials, 2013, 181, 116–122.
12 Current Organic Synthesis, 2017, Vol. 14, No. 0
[61]
[62]
[63]
[64]
[65]
[66]
Rahimi, N.; Karimzadeh, R.; Catalytic cracking of hydrocarbons over
modified ZSM-5 zeolites to produce light olefins: A review. Applied
Catalysis A: General, 2011, 398, 1–17.
Stull, D; Westrum, E; Sinke, G. The Chemical Thermodynamics of Organic
Compounds: New York; 1969.
Tatevskii, V. Physico-chemical properties of the individual hydrocarbons:
State scientific and technical PH; 1960.
Xian; X.; Liu, G.; Zhang, X.; Wang, L.; Mi. Z. Catalytic cracking of ndodecane overHZSM-5zeolite under supercritical conditions: Experiments
and kinetics. Chemical Engineering Science, 2010, 65(20), 5588-5604.
Nakasaka, Y.; Okamura, T.; Konno, H.; Tago, T.; Masuda, T. Crystal size of
MFI-type zeolites for catalytic cracking of n-hexane under reaction-control
conditions. Microporous and Mesoporous Materials, 2013, 182, 244–249.
Hong, Y.; Gruver, V.; Fripiat, J.J. Pentane conversion on dealuminated H-Y
and HZSM-5. Journal of Catalysis, 1996, 161(2), 766–775.
Ivashkina et al.
[67]
[68]
[69]
[70]
[71]
Li, Y.; Zhu, J.; Liu, H.; He, P.;, Wang, P.; Tian, H. Theoretical study of 1hexene activation over La/ZSM-5 zeolite. Journal of Beijing University of
Chemical Technology (Natural Science Edition), 2011, 38(3), 6-11.
Fierroa, V.;Duplan, J.L.; Verstraete, J.; Schuurman, Y.; Mirodatos, C. A nonstationary kinetics approach for the determination of the kinetic parameters
of the protolytic cracking of methylcyclohexane. Studies in Surface Science
and Catalysis, 2001,133, 341-348.
Konno, H.; Ohnaka,R.; Nishimura, J.; Tago, T.;Nakasaka, Y.; Masuda, T.
Kinetics of the catalytic cracking of naphtha over ZSM-5 zeolite: effect of
reduced crystal size on the reaction of naphthenes. Catal. Sci. Technol.,
2014, 4, 4265-4273.
Sadeghbeigi, R. Fluid Catalytic Cracking Handbook, 2rd ed.; Gulf Professional Publishing: United States of America, 2000.
The new handbook for Chemist and technologist. Raw materials and products of industrial organic and inorganic substances. Part 1, Publishing: St. Petersburg «Professional», 2002.
DISCLAIMER: The above article has been published in Epub (ahead of print) on the basis of the materials provided by the author. The Editorial Department reserves the right to make minor modifications for further improvement of the manuscript.
View publication stats
Download