You are on page 1of 50

energies

Review
A Review of Modelling of the FCC Unit–Part I: The Riser
Thabang W. Selalame, Raj Patel, Iqbal M. Mujtaba and Yakubu M. John *

Department of Chemical Engineering, Faculty of Engineering and Informatics, University of Bradford,


Bradford BD7 1DP, UK; t.selalame@bradford.ac.uk (T.W.S.); R.Patel@bradford.ac.uk (R.P.);
I.M.Mujtaba@bradford.ac.uk (I.M.M.)
* Correspondence: Y.John@bradford.ac.uk

Abstract: Heavy petroleum industries, including the fluid catalytic cracking (FCC) unit, are useful
for producing fuels but they are among some of the biggest contributors to global greenhouse gas
(GHG) emissions. The recent global push for mitigation efforts against climate change has resulted in
increased legislation that affects the operations and future of these industries. In terms of the FCC
unit, on the riser side, more legislation is pushing towards them switching from petroleum-driven
energy sources to more renewable sources such as solar and wind, which threatens the profitability of
the unit. On the regenerator side, there is more legislation aimed at reducing emissions of GHGs from
such units. As a result, it is more important than ever to develop models that are accurate and reliable,
that will help optimise the unit for maximisation of profits under new regulations and changing
trends, and that predict emissions of various GHGs to keep up with new reporting guidelines. This
article, split over two parts, reviews traditional modelling methodologies used in modelling and
simulation of the FCC unit. In Part I, hydrodynamics and kinetics of the riser are discussed in terms of
experimental data and modelling approaches. A brief review of the FCC feed is undertaken in terms
of characterisations and cracking reaction chemistry, and how these factors have affected modelling
approaches. A brief overview of how vaporisation and catalyst deactivation are addressed in the
FCC modelling literature is also undertaken. Modelling of constitutive parts that are important to the
 FCC riser unit such as gas-solid cyclones, disengaging and stripping vessels, is also considered. This
 review then identifies areas where current models for the riser can be improved for the future. In Part
Citation: Selalame, T.W.; Patel, R.; II, a similar review is presented for the FCC regenerator system.
Mujtaba, I.M.; John, Y.M. A Review of
Modelling of the FCC Unit–Part I: Keywords: fluid catalytic cracking; riser; complex mixtures; hydrodynamics; modelling
The Riser. Energies 2022, 15, 308.
https://doi.org/10.3390/en15010308

Academic Editor: Dmitri A. Bulushev


1. Introduction
Received: 16 October 2021
Accepted: 28 December 2021
The aim of refinery operations is the transformation of the complex crude oil mixture
Published: 3 January 2022
into useful products. The crude oil is distilled into products such as liquefied petroleum gas
(LPG), gasoline, kerosene, jet fuel, diesel and some fuel oils. The bottoms of the distillation
Publisher’s Note: MDPI stays neutral
products such as gas oil are usually heavy or high molecular weight products which are of
with regard to jurisdictional claims in
low value. These bottoms are then sent to the fluid catalytic cracking (FCC) unit where they
published maps and institutional affil-
are catalytically cracked into lighter and more valuable products. Several configurations
iations.
(such as the Exxon Models, Universal Oil Products (UOP) staked units and Kellogg’s Heavy
Oil Cracker (HOC) unit) of the FCC unit have been developed and used commercially
over the more than six decades that the unit has existed, but they are generally similar
Copyright: © 2022 by the authors.
in that they are essentially a two-reactor system [1] consisting of the riser-reactor and
Licensee MDPI, Basel, Switzerland. the regenerator. Typical operating conditions of the riser are shown in Table 1. In the
This article is an open access article riser-reactor, hot porous catalyst particles from the regenerator are brought into contact
distributed under the terms and with the gas oil feed and the lifting steam. Upon contact, the gas oil is vaporised and
conditions of the Creative Commons cracked into lower molecular weight compounds [2]. Additionally, coke is also produced
Attribution (CC BY) license (https:// as a by-product of the cracking reactions, which lowers the activity of the catalyst as it
creativecommons.org/licenses/by/ blocks access of the vaporised gas oil to the active site on the catalyst. The deactivated
4.0/). catalyst is sent to the regenerator where coke on the catalyst is burned off. The regenerator

Energies 2022, 15, 308. https://doi.org/10.3390/en15010308 https://www.mdpi.com/journal/energies


Energies 2022, 15, 308 2 of 50

then has two main functions: to burn off the coke on the catalyst to regain the catalyst’s
activity, and to provide the heat needed to drive the endothermic cracking reactions in the
riser-reactor via the exothermic combustion reactions [2].
The ability to model the FCC unit has been the subject of a lot of research during the
lifetime of the unit. The prediction of the behaviour of any chemical reactors requires that
information on reaction stoichiometry, thermodynamics, heat and mass transfer, reaction ki-
netics and lastly flow or mixing patterns of the material inside the reactor be available [3,4].
Furthermore, the FCC unit is notoriously difficult to model due to the size of the process,
complicated hydrodynamics, and complex reaction kinetics, especially in the riser [5,6].
Both the riser and the regenerator are fluidised bed reactors; Table 1 shows the normal
operating conditions of the FCC riser. The regenerator operates in the bubbling/turbulent
regime of fluidisation while the riser is a circulating fluidised bed vessel in fast fluidisa-
tion [7]. Flow in these regimes for Geldart group A particles has been investigated through
experimental studies published in the literature [8–12]. Consequently, these studies inform
the assumptions made during modelling of the unit [7]. Typically, two types of model of
FCC units are seen in the literature [13]: (i) one or two dimensional (1D/2D) models from
traditional reaction engineering approaches; and (ii) models employing computational
fluid dynamics (CFD). In both cases, modelling involves development of differential equa-
tions for mass, momentum and thermal balances. In the former the flow field is usually
averaged over the cross-sectional area or the volume to create a 1/2D system; however,
the latter are much more involved and more computationally expensive describing flow at
the micro- scale in all three dimensions (3D) compared to the former which are interested
in the macro scale. For this reason, CFD techniques usually require specialist software
such as Ansys Fluent. Nevertheless, both modelling approaches are being used in the
present climate. Riser models published since the 1960s have varied in complexity and
with varying objectives. Early papers on FCC riser modelling were mostly concerned with
lumped reaction scheme kinetics [14–18]. In the 1990s, attention turned to the modelling
of the FCC unit to investigate the bifurcation or multiplicity of the steady state [19,20]
during steady state simulations. This is also the period when mechanistic kinetic models
of FCC riser reactions were developed using understanding of the reaction chemistry of
the complex riser mixture [21–23]. Post-2000 saw an explosion in the adoption of CFD to
model and simulate the FCC riser and its constitutive parts in order to study flow patterns
and yields in reactive CFD simulations. Recently, there have also been papers published on
FCC riser optimisation to meet changing global demand trends [24–26], such as increased
demand for propylene [27], and modelling of the riser using unconventional feeds such
as bio-oils [28–30]. Evidently, the literature on the modelling and simulation of the riser is
now vast and varied, and this is the subject of the present work.
There currently exist only a few reviews on the modelling of FCC units present in the
literature. Grace [31] produced a review of the critical areas in fluid bed reactor modelling.
They summarised different hydrodynamic models of reactor flow, and gas mixing in the
dense phase of bubbling fluidised beds, an area they identified as needing further work
since due to the lack of agreement about how it should be treated. They also reviewed the
state of modelling of interphase gas mass transfer.
Energies 2022, 15, 308 3 of 50
Energies 2022, 15, 308 3 of 52

Table 1. Dimensions and operating conditions of fluid catalytic cracking (FCC) unit, [1,5,7,32,33].
Table 1. Dimensions and operating conditions of fluid catalytic cracking (FCC) unit, [1,5,7,32,33].
Riser
Riser
Dimensions
Dimensions
HeightHeight 30– 40 m 30–40 m
Diameter
Diameter
1– 2 m 1–2 m
Operating Conditions
Operating Conditions
Gas oil inlet T 150– 300 °C
Gasinlet
Catalyst oil inlet
T T 675– 750 °C 150–300 ◦ C
Catalyst inlet T > 250 kg/m
Solid circulation 2s
675–750 ◦C

Catalyst to Oil
Solid (CTO) 4– 10 wt%>250 kg/m2 s
circulation
Dispersion steam 0– 5 wt%
Catalyst to Oil (CTO) 4–10 wt%
Pressure 150– 300 kPa
Dispersion steam 0–5 wt%
Solid Residence 3– 15 s
Pressure
Catalyst Properties 150–300 kPa
Average Residence 70 µ m
size
Solid 3–15 s
Density
Catalyst Properties1200 – 1700 kg/m2s
Typical Umf 0.001 m/s
Average size 70 µm
Density 1200–1700 kg/m2 s
Geldart Group
Typical Umf A
0.001 m/s
Geldart Group A

Berruti
Berruti etetal. al.[32]
[32]havehave produced
produced a comprehensive
a comprehensive reviewreview of gas–solid
of gas–solid flow in cir-
flow in circu-
culating fluidised
lating fluidised bedbed reactors
reactors such such
as FCC asrisers.
FCC risers. They explain
They explain and analyseandHarris
analyse and Harris
Da- and
Davidson
vidson [34]’s [34]’s
TypeType 1 models
1 models that predict
that predict onlyvariation
only axial axial variation
of the solidof the solid suspension,
suspension, and
and
TypeType 2 models
2 models that predict
that predict onlyvariation
only radial radial variation
by assuming by assuming multiple
multiple regions in regions
the flow in the
flow e.g., core-annulus
e.g., core-annulus or clustering
or clustering annular annular flow. Berruti
flow. Berruti et al.
et al. [32] [32]summarised
have have summarised Type 3 Type
3models
models in terms
in terms of CFD
of CFDmodelling of hydrodynamics
modelling of hydrodynamics of riser-type reactors. Godfroy
of riser-type reactors.etGodfroy
al.
[35]
et al.have compared
[35] have compareddifferent models
different with axial
models with and/or radial radial
axial and/or variation againstagainst
variation experi- exper-
mental data.
imental data.Gupta
Guptaetet al.al.
[36] reviewed
[36] reviewed literature on kinetic
literature on kineticmodels of theofFCC
models the riser.
FCC Theyriser. They
have also briefly reviewed hydrodynamics of the riser and
have also briefly reviewed hydrodynamics of the riser and paid particular attention paid particular attention to the to the
vaporisation section
vaporisation section of of the
theriser.
riser.Pinheiro
Pinheiroetetal. al.[13]
[13]produced
produceda acomprehensive
comprehensive review
review of of the
the entire FCC process, focusing on developments in modelling
entire FCC process, focusing on developments in modelling and optimisation. They have and optimisation. They
have also tackled the issue of steady state multiplicity in the FCC unit and its origin. Pin-
also tackled the issue of steady state multiplicity in the FCC unit and its origin. Pinheiro
heiro et al. [13] have explained in detail the literature and theory of process control of the
et al. [13] have explained in detail the literature and theory of process control of the FCC
FCC process. Vashisth et al. [37] have reviewed the computational fluid dynamics model-
process. Vashisth et al. [37] have reviewed the computational fluid dynamics modelling
ling of the bubbling bed behaviour of FCC catalyst particles in terms of two fluid models
of the bubbling bed behaviour of FCC catalyst particles in terms of two fluid models and
and particle-fluid models. However, their analysis was only concerned with the hydrody-
particle-fluid
namics of the models.flow andHowever, their analysis
not the reactive system. was Shahonly concerned
et al. with the
[7] summarised hydrodynamics
literature on
of the flow and not the reactive system. Shah et al. [7] summarised
gas-solid flow in riser type reactors and presented a critical review of CFD models of literature onFCCgas-solid
flow
risersinthenriser type reactors
compared the model andpredictions
presentedagainsta critical review ofresults
experimental CFD thatmodels of FCC risers
are available
then compared the model predictions against experimental
in the literature. They have reviewed the impact on model predictions of different closure results that are available in
the literature. They have reviewed the impact on model
models that are employed in CFD problems such as drag models, viscous models and predictions of different closure
models
kinetic models. They have also summarised the linking of riser hydrodynamic models and and
that are employed in CFD problems such as drag models, viscous models
kinetic
lumpedmodels.
crackingThey have
kinetics in also
CFDsummarised the linking ofRao
modelling. Madhusudana riser
et hydrodynamic
al. [38] have reviewed models and
lumped
the use of cracking kineticsmodelling
object-oriented in CFD modelling.
of the FCCMadhusudana
unit in industry.Rao Theyet describe
al. [38] have reviewed
in detail
the use
the concept and the methodology
of object-oriented modelling of object-oriented
of the FCC unitprocess modelling
in industry. Theyindescribe
the multipur-
in detail the
pose process
concept and simulator (MPROSIM)
the methodology framework and review
of object-oriented processthe application
modelling inof this
the frame-
multipurpose
work in simulator
process the case of(MPROSIM)
the FCC unit.framework
Tuning of the andmodel
review in the
MPROSIM and parameter
application opti-
of this framework in
misation
the case of arethe
also covered
FCC unit.inTuning
the paper. These
of the modelpapers are summarised
in MPROSIM in Table 2. optimisation
and parameter
are also covered in the paper. These papers are summarised in Table 2.
Energies 2022, 15, 308 4 of 50

Table 2. Summary of previous reviews in the literature relevant to FCC modelling.

Review Aspect of FCC Modelling Covered/Reviewed


1. Fluidised bed reactor models
Grace [31] 2. Gas mixing in dense or emulsion phase
3. Interphase gas transfer
1. Fluidised bed regime classification and applications
Berruti et al. [32] 2. Harris and Davidson’s Type I, II and III models for Circulating
Fluidised Bed (CFB) risers
1. Comparison of different models with axial and/or radial gradients
Godfroy et al. [35]
against experimental data
1. FCC Riser kinetic and hydrodynamic modelling
2. FCC Regenerator kinetic and hydrodynamic modelling
Gupta et al. [36]
3. Catalyst deactivation
4. FCC unit control models
1. History and evolution of the FCC unit
2. FCC modelling (both the riser and regenerator) in terms of kinetics
and hydrodynamics
Pinheiro et al. [13]
3. Riser-regenerator coupling
4. Steady state vs dynamic modelling
5. Control of the FCC unit
1. Experimental results literature from cold flow CFB risers
2. CFD models for cold flow gas-solid flow in CFB risers in terms of
Eulerian and Eulerian–Langrangian frameworks
Shah et al. [7] 3. Effect of constitutive models (i.e., drag models, gas viscous models,
KTGF closure models) on flow prediction
4. Reactive CFD models for FCC Risers
5. Modelling of the feed vaporisation section in FCC Risers
Madhusudana Rao 1. Object oriented modelling of the FCC unit in industry using the
et al. [38] multipurpose process simulator (MPROSIM) environment

The work described in Table 2 shows that various aspects of modelling the FCC unit
have already been thoroughly reviewed with such reviews available in the literature. How-
ever, a comprehensive review of traditional FCC unit modelling approaches is evidently
lacking in the literature. Of note is the lack of reviews that link experimental findings of
operations of FCC unit to the traditional modelling approaches and the simulation results
obtained from said models. The current study presents a review of the riser modelling and
analysis of the experimental data that have led to the assumptions made in modelling. The
modelling has been divided in terms of kinetics, where treatments of reaction rate equa-
tions, rate constants and orders of reactions are considered, and in terms of hydrodynamics,
where gas–solid mixing behaviour, flow, heat and mass transfer in the reactor are consid-
ered. The literature on the current understanding of carbenium ion reaction chemistry of
the riser is also reviewed, and this leads to the development of exciting mechanistic kinetic
models such as the single event kinetics that employ detailed explicit description of rection
pathways, in contrast to the simple lumped kinetics methodology. Additionally, treatment
of clusters and the subsequent flow heterogeneity in the riser reactor are also considered.
Constitutive units such as the disengaging and stripping section that are important to the
FCC reactor exit flows and dynamics are considered in terms of modelling. Finally, the
deficiencies of current traditional modelling approaches are identified for each unit and
recommendations for further work are made.

2. Fluid Catalytic Cracking (FCC) Riser


A review of the kinetic modelling of the FCC riser establishes three main aspects that
govern the modelling approach. Firstly, the description or characterisation of the reaction
mixture (feed and product). Kinetic modelling often requires an understanding of reacting
species and reactions occuring, however, because of the complexity of the reacting mixture
Energies 2022, 15, 308 5 of 50

in the FCC riser some considerations are needed to describe the mixture. Secondly, the
mechanism of reactions involved in the riser is generally well understood now, however
for kinetic studies this is usually linked to the level of detail used in describing the reaction
mixture. Lastly, the parameters resulting from the model, such as frequency factors and
activation energies, are also important considerations, especially for lumped models in
which these parameters are empirical and usually take values depending on the feed and
catalyst used. In this chapter, we give a brief overview of these aspects of kinetic modelling
and describe how they are relevant in different types of kinetic model.

2.1. Feed and Product Characterisation


Refineries process various types of crude oil depending on geological location. Addi-
tionally, market demands, and crude quality also fluctuates within a geological location
which means that the conventional gas oil feed to the FCC unit will continuously change
in feedstock quality [39]. Many sophisticated analytical techniques have been used by
researchers to determine the physical and chemical properties of the gas oil feed and subse-
quent products, which is quite relevant to the modelling of the FCC riser. Behera et al., and
Naik et al. [29,40] have reported 1 H and 13 C nuclear magnetic resonance (NMR) spectra of
vacuum gas oil FCC feed, in each case various gas oil feeds from different refineries were
investigated.
These spectra have given a closer look at the proton and carbon environments present
in the different molecular components of gas oil. According to Behera, Naik and their
teams [29,40] the peaks present in 1 H NMR spectra can be divided into three different
proton regions: aromatic hydrogens (chemical shift 6–9 ppm), aliphatic hydrogens (chem-
ical shift 0–4.5 ppm) and olefinic hydrogens (chemical shift 4.5–6 ppm). Additionally,
the 13 C NMR spectra can be divided into aliphatic (0–50 ppm), alcoholic (50–110 ppm),
aromatic carbons (110–150 ppm) and carboxylic carbons (150–200 ppm). The disadvantage
of analysis of 1 H and 13 C spectra is the overlap of resonances, particularly for the CH, CH2
and CH3 in the case of 13 C, and the minimal chemical shift dispersivity in 1 H, which makes
it different to distinguish certain groups. For example, Behera et al. [40] have found a
strong overlap of naphthenic carbons in the aliphatic region corresponding to 20–44 ppm
and 23–45 ppm). In order to circumvent this issue, researchers [40] have used distortionless
enhancement by polarisation transfer (DEPT) to improve spectral assignments by identify-
ing various substructures for more reliable quantitative interpretation. The result showed
that there are a variety of resonances present in the region 25–50 ppm corresponding to the
different CHn environments. Similar exercises could be undertaken for the proton regions
to distinguish between different proton environments. Naik et al. [29] have further subdi-
vided the hydrogen environments into Hα (2–4.5 ppm), Hβ (1–2 ppm) and Hγ (0.5–1 ppm).
Further chemical shift assignment of the gas oil can be found in the work of Behera and
colleagues [40] and references therein. This data from NMR studies paints a very clear
picture of the complexity of the structures of the different chemical components of the
FCC gas oil feed, and subsequently the product gas phase. Nevertheless, it is possible,
based on these structural variations, that the components of the feed and products can be
grouped into aliphatics, naphthenics and aromatics. Aliphatics are linear and/or open
branched hydrocarbons, which can be subdivided into olefins (unsaturated) and paraffins
(saturated). Naphthenes are ring-shaped alkanes while aromatics are based on unsaturated
benzene rings [39]. This is the so called ‘PONA’ classification that is commonly used
in petroleum refinery to characterise crude oil fractions [41]. Rodriguez et al. [42] have
reported GC × GC analysis of vacuum gas oil to study the relative amounts of each group
of compounds in typical FCC feed, they have found that the feed is 62.3 wt% aromatics,
29.2 wt% naphthenes, 8.5 wt% paraffins, and 5.5 wt% sulphurous compounds. But of
course, these values will be different for different crude oils. Although the above NMR
data give some information on the structure of the hydrocarbon components of feed and
products of the FCC riser, they give no information on the size of the hydrocarbon chains
and is limited by the sheer number of molecular components present in the mixture [29].
Energies 2022, 15, 308 6 of 50

The gas oil feed is made up of great number of hydrocarbon molecules, together 7 of 52 with
a relatively small quantity of impurities such as sulphur, nitrogen and metals, differing in
molecular weight, chemical composition and structure [43]. Due to the complexity in the
composition of the feed, it is impossible to analyse the individual components of the gas oil,
Sulphur containing hence often in refinery operations the composition of the mixture is characterised in terms
5.5
of large distillation cuts that are based on boiling point range and carbon number: gaso-
(wt%)
line (30–215 ◦ C, C5 –C12 ), kerosene (180–240 ◦ C, C11 –C15 ), diesel (240–360 ◦ C, C15 –C25 ),
Distillation boiling point cuts
vacuum gas oil (360–540 ◦ C, C25 –C45 ) and vacuum residuals (>540 ◦ C, >C45 ) [43]. Other
Gasoline (wt%) fractions may be 1.5defined depending on the existing refinery equipment capabilities (e.g.,
Diesel (wt%) rectifiers and separators).
3 Because the cuts are usually related to the final-end use, the
composition of the cuts in the gas oil is usually of importance to both marketing and the
>Diesel (wt%) 95.5
economics of the process. FCC feed typically has a higher composition of the higher boiling
Physical Properties
point range which are converted to products mixture with a higher composition of the
𝑘𝑔
Density ( at 15 °𝐶)more useful and 0.93valuable lower boiling point range cuts. Typical compositions of FCC
𝐿 feed and products are shown in Table 3. A broader characterisation of the feed may be
American Petroleum obtained from true boiling point (TBP) curves, which are a plot of the cumulative mass or
volume
Institute (API) gravity distillation
21.1 with increasing boiling point. The shape of the TBP curve depends on
the volatility of the components in a given gas oil, such that this curve may be unique to
(°API) the composition of a given FCC feed [44]. Figure 1 shows an example TBP plot for crude
Viscosity (cst at 100 oil, showing a possible link between distillation cut and chemical composition. The TBP
8.6
°C) distillation is one of the most commonly used experimental procedure employed to charac-
terise gas oil feeds [44], and researchers have used them as pseudo-component analysis
Mw (g/mol) 405.3
where there are specific cuts in generative product distributions from the FCC unit [45], and
HHV (Mj/kg) 42 the pricing of such feeds. From the TBP curve, information about the
therefore to influence
Mw—Number average relative volatility
molecular and, therefore,
weight, HHV—High size distribution, of the
heating value, components
determined byofdifferential
the gas oil mixture
scanning calorimetry. may be acquired. However, it is still impossible from this analysis to gain insight about
individual components in the mixture.

Figure 1. True boilingFigure


point 1.(TBP) of a naphthenic
True boiling crude
point (TBP) of oil (Redrawn
a naphthenic crude oilfrom [44]).from [44]).
(Redrawn

2.2. Reactions in the FCC Riser


Catalytic cracking of hydrocarbons in the gas oil feed is generally believed to be a
chain reaction that proceeds via carbenium ion theory [46,47] proposed by Whitmore et
al. [48]. The mechanism involves three steps: initiation, propagation and termination.
Energies 2022, 15, 308 7 of 50

Table 3. Feed and product characterisation from an FCC riser pilot plant using vacuum gas oil feed
and commercial equilibrium catalyst containing 15.5% zeolite [42].

Feed Products
Elemental analysis Distillation boiling point cuts
C (wt%) 86.60 Dry gas (C1 –C2 ) (wt%) 4
H (wt%) 12 LPG (C3 –C4 ) (wt%) 4
N (wt%) 0.23 Gasoline (C5 –C12 ) (wt%) 27
S (wt%) 1.17 Light cycle oil (C13 –C20 ) (wt%) 28
PONA Composition Heavy cycle oil (>C20 ) (wt%) 18
Paraffins (wt%) 62.3 Coke (wt%) 9
Naphthenes (wt%) 29.3
Aromatics (wt%) 8.5
Sulphur containing (wt%) 5.5
Distillation boiling point cuts
Gasoline (wt%) 1.5
Diesel (wt%) 3
>Diesel (wt%) 95.5
Physical Properties
kg
Density (L at 15 ◦ C) 0.93
American Petroleum Institute
21.1
(API) gravity (◦ API)
Viscosity (cst at 100 ◦ C) 8.6
Mw (g/mol) 405.3
HHV (Mj/kg) 42
Mw—Number average molecular weight, HHV—High heating value, determined by differential scanning
calorimetry.

2.2. Reactions in the FCC Riser


Catalytic cracking of hydrocarbons in the gas oil feed is generally believed to be a chain
reaction that proceeds via carbenium ion theory [46,47] proposed by Whitmore et al. [48].
The mechanism involves three steps: initiation, propagation and termination. These
reactions are thought to occur on the acid sites of the zeolite catalyst. The carbenium ions
produced on the active sight can undergo several reactions such as isomerism, alkylation,
β-scission and hydride transfer.
• Initiation: this is the first step that kicks off the cracking reactions. It is characterised
by attack of an active sight on a reacting molecule to produce an activated complex
(i.e., a carbocation). Initiation in the case of an olefin molecule is thought to proceed
via the direct attack of a Brönsted acid site on the reactant double bond to form a
carbenium ion [47]. This is the mechanism that would be expected for gaseous olefins,
such as the case in the FCC riser [49]. On the other hand, there is some debate over
the exact mechanism in which paraffins are activated, early work [50,51] suggested
that activation of paraffins required the existence, even in small quantities, of olefins.
In this mechanism, carbenium ion from the olefins (via the method described above
for olefins) would abstract a hydride ion on a nearby paraffin to generate an activated
paraffin carbenium ion. For a while, this mechanism was essentially correct [47], but
its insufficiencies in accounting for observed complex product distributions warranted
concern [52]. However, later Haag and Dessau [53,54] proposed a direct protonation
method of the paraffin (the so-called Haag-Dessau mechanism), that drew inspiration
from the 1994 Chemistry Nobel Prize winning work of George A. Olah (for contribu-
Energies 2022, 15, 308 8 of 50

tions to the chemistry of hydrocarbons in super-acidic solutions) and did not require
the presence of olefins in the activation of paraffins. Via this mechanism, a Brönsted
acid site directly protonates the paraffin reactant molecule forming a carbonium ion
complex which collapses by bond scission to give a carbenium ion. The protonation
produces a carbenium ion together with either an alkane (in the event of a C − C bond
scission) or hydrogen (in the event of a C − H bond scission) [53,54]. The Haag–Dessau
mechanism gained some acceptance as it was supported by some researchers [55–58].
Furthermore, activation of paraffins by Lewis sites has also been proposed [46,59]
involving the abstraction of a hydride from an alkane by a Lewis site resulting in a
carbenium ion formation. However, evidence from [60,61] has questioned the contri-
bution of Lewis sites. Nevertheless, more paraffin activation mechanisms have been
suggested [62–65]. Generally, all these mechanisms agree that the initiation step will
produce a carbenium ion, the subsequent steps after that become relatively routine
and generally agreed upon by most researchers. Initiation of both olefins and paraffins
is shown in Figure 2a,b.
• Propagation: this involves the transfer of a hydride ion from an adsorbed carbenium
ion to a reactant molecule, and always results in the formation of another carbenium
ion [47]. Simply put, a carbenium ion reacts with a paraffin or olefin so that another
carbenium ion is formed in the process. In the case of a paraffin molecule interacting
with a surface carbenium ion, it has been proposed that the chain propagation occurs
when the carbocation abstracts a hydride ion on the reactant paraffin. The former
carbocation then desorbs from the surface as a paraffin when the former paraffin
is left adsorbed as a carbenium ion. The resulting carbenium ion can then either
isomerise and/or crack further, which means the chain process continues. When the
cracking occurs via protolytic attack or β-scission, the resulting product is an olefin
or a paraffin, respectively. Additionally, it is also thought that adsorbed carbenium
ions may undergo 1,2 hydrogen shift so that the positive charge can migrate across the
molecule giving a possibility of branching [47]. See Figure 2c for possible propagation
mechanisms in the riser.
• Termination: this involves the desorption of the adsorbed carbenium ion to give an
Energies 2022, 15, 308 olefin, whilst restoring the active site to its initial9 state
of 52 [47]. This step results in the
termination of an active carbenium ion, thereby completing the chain.

Figure 2. (a) Haag–Dessau mechanism for initiation of paraffin, (Adapted from [66]) (b) initiation
Figure
mechanism2. (a) Haag–Dessau
of olefins, mechanism
(c) propagation mechanism for initiation
involving carbenium of paraffin,
ions (1: Hydride transfer, (Adapted from [66]) (b) initiation
2: β-scission) (Adapted from. [66]).
mechanism of olefins, (c) propagation mechanism involving carbenium ions (1: Hydride transfer,
2: β-scission)
Additional to (Adapted from
the main cracking [66]). discussed above, there are a number of
mechanism
acid catalysed reactions that are also occurring in the riser, and these include isomerisa-
tions, alkylations, branching, cyclisation and polymerisation [67]. By taking all these reac-
tions into account, Cortright, Yaluris and co [68–71] have managed to work out the cata-
lytic cycles for some alkanes taking place in the FCC riser. Figure 3 shows a simplified
catalytic reaction network, consisting of several reaction step cycles, that was shown to
account well for data representing the conversion of isobutane over zeolite Y catalyst.
Each cycle is derived by starting from a surface ion species, following reaction lines that
connect various surface ions until the reaction path returns to the original surface ion.
From the reaction network, olefins are produced via a series of oligomerisation and/or 𝛽-
Energies 2022, 15, 308 9 of 50

Additional to the main cracking mechanism discussed above, there are a number of
acid catalysed reactions that are also occurring in the riser, and these include isomerisations,
alkylations, branching, cyclisation and polymerisation [67]. By taking all these reactions
into account, Cortright, Yaluris and co [68–71] have managed to work out the catalytic cycles
for some alkanes taking place in the FCC riser. Figure 3 shows a simplified catalytic reaction
network, consisting of several reaction step cycles, that was shown to account well for data
representing the conversion of isobutane over zeolite Y catalyst. Each cycle is derived by
starting from a surface ion species, following reaction lines that connect various surface
ions until the reaction path returns to the original surface ion. From the reaction network,
olefins are produced via a series of oligomerisation and/or β-scission elementary steps
that produce carbenium ions, followed by deprotonation and desorption of the resulting
carbenium ion. Paraffins may be formed together with a surface isobutyl carbenium ion
from hydride transfer cycles. The various surface carbenium ions can undergo several
ergies 2022, 15, 308
other reactions such as isomerisation and β-scission to form other surface complexes and
gaseous products. From this analysis, Yaluris et al., suggest that any particular reaction
step in the cycle/network may be part of more than one catalytic cycle.

Figure
Figure 3. 3. Catalytic
Catalytic cycles cycles for of
for cracking cracking
isobutaneof isobutane
over zeolite USYover zeolite
at 773K USYplug
in assumed at 773K i
flow; solid lines denotes fast reactions, dotted lines denote slow reactions.Bold species are surface
solid lines denotes fast reactions, dotted lines denote slow reactions.Bold spec
complexes and species next to the reaction line are gaseous products or reactants. Reaction lines
plexesvarious
indicate and species
elementarynext
steps.to the reaction
(Redrawn line
from Yaluris are[69]).
et al. gaseous products or reactants.
cate various elementary steps. (Redrawn from Yaluris et al. [69]).
Evidently, once the reaction network is expanded to include cycles of all compounds
present in the FCC gas oil feed, it becomes very complicated or even impossible to compute.
Evidently,
However, the review ononce the reaction
characterisation abovenetwork is expanded
already showed that for suchto include
mixtures in cycle
the FCC riser, it is near impossible to account for every single compound present in the
present in the FCC gas oil feed, it becomes very complicated or even
mixture using currently available technology.
pute.TheHowever,
short overviewsthe review
about the feedon characterisation
characterisation above are
and FCC reactions already
presentedshowed
here
turesto highlight
in the FCCtwo important
riser, itfactors relevant
is near to the kineticto
impossible modelling
account of the
forFCC riser: single
every
in the mixture using currently available technology.
The short overviews about the feed characterisation and FCC reac
here to highlight two important factors relevant to the kinetic modelli
• Both the feed and the product mixture of the system are very co
Energies 2022, 15, 308 10 of 50

• Both the feed and the product mixture of the system are very complicated mixtures of
compounds, and their modelling is, to some extent, limited by the characterisation
techniques available to the process engineer,
• The reaction network for the FCC system, using conventional gas oil feed, is too
massive and too detailed to be modelled exactly. To some extent, modelling may be
limited by computation power.
In the regenerator, for example, it was easy to use simple chemical reaction kinetics
to describe the system mathematically because all the reacting species were known and
can easily be characterised (when an approximation for coke composition is made). Addi-
tionally, both the numbers of reacting species and of reactions are small, and therefore the
system/reaction network can be easily described mathematically. However, in the case of
the riser this is not the case, for example when considering only the C25 − C40 reacting
species in the gas oil feedstock, the number of isomers in this group exceeds 1014 , without
even accounting for heteroatoms such as N, O, S and metals [72], which are also present
in the feed molecules. Therefore, it is almost impossible to account for all the molecules
present in the reacting system. This problem of high dimensionality and increased coupling
in the reacting system is also encountered in systems such as polymerisation, hydrocracking,
isomerisation, aromatisation, reforming, Fischer–Tropsch synthesis and in some biological
processes [72]. Different kinetic models will be discussed in the next chapter that have been
used in the literature to describe FCC cracking in the riser, in each case, the modelling is
usually a compromise between the level of detail that is needed from the model by the
end user and the capabilities of the analytical techniques available to the simulator to
characterise and quantify the reaction mixture. In petroleum engineering, two classes of
analytical technique are used in kinetic modelling: chemical separation techniques such as
mass spectrometry (MS), liquid chromatography (LC), solvent extraction (SE), and physical
separation techniques such as distillation [72].

2.3. Kinetic Models of the Riser


2.3.1. Discrete Lumping
These are a priori lumping techniques that reduce the complex feedstock and large
reaction network of the FCC riser into a smaller network that groups reaction components
into groups, or lumps. In this technique, the lumped components, which may be grouped
based on chemical or physical similarity depending on the level of analytical characterisa-
tion available and the level of detail required, are a single reacting species. Consequently,
such an approach does not describe the system exactly but aims to capture essential ele-
ments of the real system with sufficient predictive power [73]. In the case of the FCC unit,
it is not usually important to obtain a complete description of the process, i.e., traditionally
only the yield of gasoline distillation cut in the product from the riser was of importance.
Once the lumps considered have been established, a simple reaction analysis may be used,
where a reaction network is made by conversions between different lumps. Each conver-
sion in the network has a reaction rate and, therefore, an associated reaction constant and
activation energy. In the general case, this may be described mathematically as:
N
dCi
Ri =
dt
= ∑ K ji ϕCjn − Kij ϕCim for i 6= j (1)
i

Eij
 
Kij = k ij exp − for i 6= j (2)
RT
where Ri is the rate of formation of lump i, C is the concentration, ϕ is a non-selective
deactivation constant, n & m are the reaction orders, N is the total number of lumps
and Kij is the reaction constant for the conversion between lumps i to j. The issue of
deactivation will be discussed in a later section. The parameters of the kinetic model are
then found in the Arrhenius form of the rate constant: k ij the pre-exponent factor (usually
defined as the frequency factor) and Eij the activation energy of the conversion i → j . The
Energies 2022, 15, 308 11 of 50

parameters are determined by fitting experimental data to the simple model described
above. Therefore, it follows that for the required regression task and optimisation of the
parameters, the characterisation methods available should be able to measure the evolution
of the concentration of the lumps. Another simplification of this method is that chemical
data of the lumps is not required, nor is the consideration of stoichiometry, conversions are
assumed to be one-to-one between the lumps [45]. In the traditional lumping technique, the
lumps are usually chosen according to the main refinery distillation cuts: gas oil, gasoline,
LPG and light gases. Alternatively, lumping may be undertaken according to the PONA
characterisation, depending on the analytical tools at the disposal of the process modeller.
One of the earliest applications of discrete lumping in the FCC process was by Week-
man and Nace [74]. Their kinetic model consisted of a reaction scheme consisting of lumps
for the gas oil feedstock, gasoline (C5 − BP of 220 ◦ C) and a lump consisting of coke and
light gases. Their reaction scheme posited that gas oil feed is converted to the two other
lumps in two parallel competing reactions, and that gasoline lump subsequently over-
cracks to the coke plus light gases lump. The reactions scheme assumes cracking of gas oil
to be second order and that of gasoline to be first order. Their model was validated using
experimental data from a laboratory moving bed reactor using a commercial grade catalyst.
It is fair to say their work pioneered the kinetic lumping methodology in FCC modelling.
In the years that followed, their model was expanded; Lee et al., and Yen et al. [75,76] have
split the coke plus light gases lump from Weekman and Nace’s model into two separate
lumps, creating a four-lump model, owing to the need to predict coke yields separately
from any other lump. In the scheme, direct cracking of both gas oil and gasoline to light
gases and to coke is added. Whereas the previous studies discussed have considered the
gas oil feedstock as a single lump, it may also be divided into lumps, again as long as the
analytic capabilities permit; Corella and Francés [77] have developed a model that splits the
gas oil lump into light and heavy fractions, Larocca et al. [78] split the gas oil lump into its
chemical compositions: aromatics, paraffins and naphthenics. Additionally, the rules for the
connections between the lumps are also flexible, for example, the previous model of Week-
man and the models derived from it assume that cracking of heavy lumps such as gasoline
and gas oil also directly results in coke formation, however, Ancheyta-Juárez et al. [79]
proposed a model including lumps for gas oil, gasoline, LPG, dry gas and coke in which
only the cracking of the gas oil lump could lead to coking. In this way, the rules for reaction
scheme are arbitrary and mean it is difficult to establish a universal lumping strategy for
the FCC cracking reactions. Consequently, the empirical kinetic constants determined from
kinetic studies are only comparable if the same reaction scheme is used to derive them. Re-
gardless, the simplicity of the methodology has seen a number of publications for different
number of lumped kinetic models for the FCC unit: four-lump models [79–82], five-lump
models [79,80,83,84], six-lump models [85–87], seven-lump models [88], eight-lump mod-
els [89,90], nine-lump models [30], 10-lump model [17,18], 11-lump models [91], 12-lump
models [92], and 14-lump models [93]. Figure 4 shows some lumped reaction schemes used
for the FCC riser. Table 4 shows some kinetic parameters of discrete lumped models from
various researchers. Table 4 also shows that the number of parameters increases with the
number of lumps in the model, and that the parameters such as the frequency factors vary
from model to model and for different feeds. The increase in the number of lumps in the
models means that more information about the product slate is produced from the model,
the obvious consequence to the regression exercise is that more model parameters must be
estimated simultaneously resulting in convergence problems [28].
Energies 2022, 15, 308 13 of 52
Energies 2022, 15, 308 12 of 50

Figure 4. Lumped kinetic models: (a) 3-lump model of Weekman and Nace [74], (b) 4-lump model of Lee et al. [75], (c) 5-lump model of Ancheyta-juarez et al. [80], and (d)
Figure 4. Lumped kinetic models: (a) 3-lump model of Weekman and Nace [74], (b) 4-lump model
10-lump model of Jacob et al. [17] (Ph , Nh , Ah , CAh for heavy fraction paraffins, naphthenics, aromatic substitutes and aromatic rings respectively, and Pl , Nl , Al , CAl for light
of Lee et
fraction paraffins, naphthenics, aromatic substitutes andal. [75], (c)
aromatic 5-lump
rings model
respectively, ofgasoline
G for Ancheyta-juarez
and C for lightetgases
al. [80], and (d) 10-lump model of Jacob
plus coke).
et al. [17] (Ph , Nh , Ah , CAh for heavy fraction paraffins, naphthenics, aromatic substitutes and
aromatic rings respectively, and Pl , Nl , Al , CAl for light fraction paraffins, naphthenics, aromatic
substitutes and aromatic rings respectively, G for gasoline and C for light gases plus coke).

Table 4. Some kinetic parameters of discrete lumping models.


4—Lump Parameters of Ancheyta-Juarez [80]
kij Eij
Transformation Units Value (kJmol−1 )
 
VGO → GS −1 −1 699 57.3
h wt%
 
VGO → LG − 1 − 1 125.28 52.7
h wt%
 
VGO → CK − 1 − 1 50.4 31.8
h wt%
 
GS → LG h−1 wt%−1 33.48 65.7
 
GS → CK h−1 wt%−1 72 × 10−6 66.6
 
Deactivation (ϕ = e−αt ) α h−1 315
6—Lump
 Parameters of Yakubu [85]
VGO → DZ h−1 wt%−1 2.86 × 107 53.9
 
VGO → GS h−1 wt%−1 5.20 × 107 57.1
 
VGO → CK h−1 wt%−1 1.45 × 105 32.4
 
VGO → LPG h−1 wt%−1 8.41 × 106 51.3
 
VGO → DG −1
h wt% −1 1.80 × 106 48.6
 
DZ → CK h−1 wt%−1 2.71 × 105 61.2
 
DZ → GS −1
h wt% −1 7.12 × 105 48.1
 
DZ → LPG h−1 wt%−1 1.26 × 104 67.8
 
DZ → DG h−1 wt%−1 1.22 × 104 64.3
 
GS → LPG h−1 wt%−1 7.88 × 103 56.2
 
GS → DG h−1 wt%−1 5.97 × 103 63.3
 
GS → CK −1
h wt% −1 7.31 × 103 61.8
 
LPG → DG h−1 wt%−1 1.23 × 104 55.5
 
LPG → CK h−1 wt%−1 2.16 × 103 52.5
 
DG → CK h−1 wt%−1 7.90 × 103 53.0
Lumps: VGO—Vacuum gas oil, GS—Gasoline, DZ—Diesel, LPG—Liquid petroleum gas, LG—Light gas, DG—Dry
gas, CK—Coke. Subscripts: ij representing transformation of some lump i to lump j ( i → j ).
Energies 2022, 15, 308 13 of 50

The models described above are plagued by the arbitrary nature of the reaction
rules (i.e., reactions are not in line with what is expected of cracking instead a one-to-one
conversion is assumed), and the non-specific nature of the determination of lumps. In
view of this, Gupta et al. [45] developed a new method of discrete lumping where the
lumps are characterised by TBP and in such a way that the lumps have a constant Watson
characterisation factor [94]. The Watson factor is essentially a measure of a crude fraction’s
aromaticity or paraffinicity. The difference from the previous discrete models is that the
physio-chemical properties of the lumps are known and, therefore, lumps can be treated as
pseudo-pure components. This model also has defined rules about conversions between the
lumps that are based on an intuitive understanding of cracking. Via their proposed scheme,
each lump on cracking gives two smaller lumps in one single reaction step, which is closer
to what would be expected of a cracking reaction. The stoichiometry is such that when one
mole of a given lump undergoes cracking, it gives one mole each of two smaller lumps
and the balance mass is formed as coke. From our understanding of the cracking reaction
mechanism from the previous section, a long chain paraffin for example, may underdo
direct protonation forming a smaller paraffin and a smaller carbenium ion, and both can go
on to undergo further reactions. Evidently, the present model by Gupta et al. [45] where
one pseudo-component cracks to two others is at least consistent with part of the theory
on cracking reaction mechanisms, unlike the first few models discussed here. Then the
chemical reaction for a pseudo-component is given by:

k i,m,n
PCi → PCm + PCn + αi,m,n (3)

where i, m and n are pseudo-components and α is the amount of coke formed per mole of
PCi that cracks. The mass balance for the coke formed per reaction:

αi,m,n = MWi − ( MWm + MWn ) for αi,m,n ≥ 0 (4)

In the scheme, PCi represents the several hundred compounds that have a molar mass
of MWi . To calculate the molecular weights, the authors used the correlation by [95]:

MWi = 204.38Tb0.118 S1.88



g exp(0.00218Tb ) × exp −3.07Sg (5)

1
 
Tb3
Sg = 1.21644  (6)
Kw

1
MeABP 3
Kw = (7)
Sg
where Sg is the specific gravity for each pseudo component, Tb is the normal boiling point
of each component and Kw is the Watson characterisation factor. MeABP is the mean
average boiling point of the components. Therefore, to find these characteristics for each
component, the TBP distillation curve for the feed and the entire fraction density is required.
The next step is to determine a reaction rate for the cracking of the ith pseudo-component.
For this, researchers [45] have assumed first order kinetics for each component, so that the
2
rate equations for cracking in a small element of volume πD 4 × ∆z is given by:

ri,m,n = ϕ × k i,m,n Ci × ( Mcat × ∆t) (8)

where the ri,m,n is the rate of cracking of the ith component, k i,m,n is the rate constant of
the cracking of i to give m and n pseudo-components, Ci is the concentration of i in the
control volume, Mcat is the mass flow rate of catalyst into the control volume and ∆t is the
residence time of the catalyst in the control volume (which is found from ∆z uc where uc is
Energies 2022, 15, 308 14 of 50

the catalyst velocity). To estimate the kinetic constants in order to complete the model, the
authors have produced a semi-empirical equation, given below:
   
α

E0 MWiv
 exp − i,m,n
τ − exp(− MWi )
k i,m,n = k0 exp − × (9)
RT 1 − exp(− MWi )

where k0 , E0 , v and τ are model parameters that are correlated to experimental data. The
authors [45] state that these four parameters are constant for any feed and catalyst pair.
Parameter τ is said to correlate to coke forming tendency while parameter v is said to
correlate to activation energy of individual pseudo component in terms of its molecular
weight. Gupta et al. [45] found that when the above kinetic model is combined with a
suitable hydrodynamic model, they were able to predict well the overall gas oil conversion,
product yields and temperature along the riser height. This model avoids the arbitrary
nature of the reaction rules between the lumps and allows for the extension of lumps
beyond just the wide boiling ranges of distillation, in their paper, the authors used 50 lumps
to describe the system. The model is appropriate to use as long as the parameters for that
particular feed and catalyst pair are known.
The simplicity of the lumping strategy discussed here makes it an attractive option
for the kinetic modelling of the FCC riser, hence the vast amount of literature on the
subject. Workers using this approach hardly have to think about such things as the reaction
mechanisms and mass transfer limitations, and the system’s simplicity also requires very
little computational power. However, despite these advantages, some disadvantages of the
methodology should also be considered by the simulation engineer. Namely, the reactivity
of the lumps may not be similar which means no global reaction rate constant may be
established for any one lump [72], they are highly feed- and catalyst-specific which means
they extrapolate very poorly to different pairs of feed and catalyst [87], and the description
of the composition of the feed by boiling points and in some instances by a broad range of
chemical similarities is too simplistic for any reliable model to be built with any kind of
generalisation [23]. The use of a large number of lumps solves partially the problem of poor
extrapolation of the models to different catalyst-feed pairs [87], and the pseudo-component
model by Gupta solves partially the problem of stoichiometry and arbitrary reaction rules
between the lumps; however, this model still suffers from reliance of model parameters on
the feed–catalyst combination. More detailed methods are required in order to eliminate
these problems.

2.3.2. Continuous Lumping


It was seen that in discrete lumping, the reacting mixture in the FCC riser was treated
as a conventional mixture where it was possible to distinguish between a finite number
of components or lumps. If the lumping assumption is again relaxed, then the gas oil
system is again thought to be a complex mixture of an astronomically large number of
chemical species. The foundations of the continuous lumping methodology come from
the work of Aris and Gavalas [96] who postulate that such mixtures may be thought of
as having an infinite number of chemical species. It is then no longer possible [96] or
worthwhile [97] to speak of properties like mass, concentration and reactivity for individual
species or lumps. Instead, only the distribution of such properties is considered. M(x)
is then the mass distribution function such that M(x)dx is the mass of a cut of species
A(x)dx for range (x, x + dx). x is then identified as a characteristic index (such as boiling
point temperature or chromatographic retention time in the case of the reactive FCC riser
mixture such that A(x)dx is a sub-mixture whose index lies in the range (x, x + dx). Similar
definitions can be established for other properties such as concentrations for the cut of
species. Any characteristic may be used as an index, the only requirement is that the
characteristic unequivocally identifies the reactants [98]. In order to describe the kinetic
model for a continuous system, as we did for the case of the discrete model, we follow
the procedure by Laxminarasimhan et al. [97]. They use the TBP curve to characterise
Energies 2022, 15, 308 15 of 50

the cracking reaction mixture by converting the TBP curve into a continuous distribution
such that for the components in the mixture the weight fraction is normalised in TBP, (θ).
In this case C (θ, t) for any given time instance is the distribution function of component
concentrations corresponding to normalised TBP. By definition, C (θ, t)dθ then corresponds
to the fraction or proportion of species with a boiling point in the range θ − (θ + dθ ). Using
i to index the reacting mixture for any component, then it follows that Ci (t)di = C (θ, t)
for the general case of i. Usually the model equations for the reactor are derived in the
c vs. k reference frame [99] where k is the reaction rate constant, and then subsequently
converted back to the C vs. θ coordinates. It is commonplace that the cracking rate constant
is treated as being a monotonic function of normalised TBP such that high boiling point
molecules crack faster than low boiling point molecules [100]. Therefore, accordingly from
the continuous description of the components in terms of θ, it is reasonable to assume
that the rate constants can also be treated as a continuous function of θ as well. Evidently,
it is possible to extend this approach to develop a model with multiple characterisation
variables, rather than just θ. However, for the present case, at any time instance:

C (θ, t) × dθ = c(k, t) × D (k) × dk (10)

where the normalised temperature is given by:

TBP − TBP(l )
θ= (11)
TBP(h) − TBP(l )

Such that TBP(h) and TBP(l ) are the highest and lowest possible boiling point temper-
atures of the reactive mixture, corresponding to the highest molecular weight component
and lowest molecular weight component in the continuous mixture, respectively.
D (k) is the species-type distribution function such that D (k)dk is the number of species
with a reactivity in the range (k, k + dk). This approach is such that c(k, t) corresponds to
the concentration of species with a reactivity k. The monotonic function of k vs. θ allows
that an independent function of the two variables can be developed; for example, it may be
taken as a simple power law type equation [97]:

k 1
= θα (12)
k max

α is a model parameter, k max is the rate constant of the highest boiling point species,
corresponding to θ = 1. This power law equation makes it explicit that the rate constant
for the lowest boiling point species is zero, that is k = 0 at θ = 0, which is a reasonable
assumption in the FCC riser cracking mixture where the smallest chain molecules such
as methane are not thought to crack any further. Of course, other forms of the k vs. θ
relationships may be used if they are representative of results from experiments. The D (k)
can be regarded as a Jacobian of i − k coordinate transformation, so that by a chain rule:
  
di di dθ
D (k) = = (13)
dk dθ dk

If the total number of species in the reaction mixture is N (for N → ∞ in a complex


di
mixture), then dθ ≈ N, so that:
 
α
D (k) = N k α −1 (14)
kαmax

As reported by Laxminarasimhan et al. [97], D (k) must satisfy:


Z k max
1
D (k)dk = 1 (15)
N 0
Energies 2022, 15, 308 16 of 50

With the established background, the mass balance for the species with reactivity k
can be formulated in the k-plane as follows:
Z k max
dc(k, t)
= −kc(k, t) + p(k, K )Kc(K, t) D (K )dK (16)
dt k

The function p(k, K ) is the yield function, which basically means it determines how
much of a species with reactivity k is produced from species with reactivity K. The species-
type distribution function (i.e., the Jacobian mapping), D (K ), in the integral accounts for
the cracking of species with reactivity K, since the mass balance is carried out in the k space.
The integral then sums over the cracking of species up to species with the maximum
reactivity k max , assumed here to be the heaviest hydrocarbon. It is worth noting that the
form of the mass balance equation has previously been derived by McCoy and Wang [101]
using particle breakage and molecular fragmentation theory where two or more fragments
result from each fission or fragmentation event. This approach requires the consideration
of stoichiometric coefficients of the scission events and the incorporation of linear or non-
linear kinetics into the model. For the present case, the only term left for consideration
is the yield function, which may take different forms depending on the specific reaction
mixture. Laxminarasimhan et al. [97] have thoroughly discussed the form of the yield
function and have noted that a skewed Gaussian-type distribution function is the best
candidate for the yield function for cracking reactions. They explained that the skew of
the distribution is due to the maximum yield of products from cracking of components
with reactivity K occurs just below components with reactivity K, so that the peak of the
distribution is biased towards K. Laxminarasimhan et al. [97] give the following function
as the yield function:
    α 2  
0
k
1  K − 0.5  
p(k, K ) = √ exp−   − A + B (17)
  
S0 2π α1

where:  !
0.5 2

A = exp − (18)
a1
  
k
B = δ 1− (19)
K
S0 can be solved by substituting the yield function into the following mass balance cri-
terion: Z K
p(k, K ) D (k)dk = 1 (20)
0
    α 2  
0
k
Z K
1 K − 0.5  
∴ √ exp−   − A + B D (k)dk = 1 (21)
   
0 S0 2π α1

From the kinetic model described above, the model parameters that must be estimated
empirically from experimental data are k ma , α0 , α1 , and δ. For the parameter estimation
exercise, a suitable optimisation function must be established, but first the main mass
balance equation must be discretised in order to solve numerically as analytical solutions
are not possible. Laxminarasimhan et al. [97] recommend discretisation by dividing the
K-plane into N nodes, so that the main balance equation may be written as a small difference
equation for any ith node and small-time step ∆t. ∆t and N may be varied depending on
the required accuracy and available computation power.
The analysis presented above uses simple first order cracking reactions, however,
it is possible to incorporate higher order reactions. Cicarelli et al. [102] have analysed a
Energies 2022, 15, 308 17 of 50

system of cracking type reactions that were governed by non-linear cracking reactions
using the continuous approximation, by combining basic continuum ideas from Aris and
co-workers [96,103,104]. Their approach follows the fragmentation methodology assuming
equal stoichiometric distribution of all crackates, and they also include considerations
for the adsorption of species using the Langmuir-Hinselwood kinetics. The approach
described here is formulated based on a single characterisation variable, but although
often not addressed or attempted, an extension to multiple characterisation variables is
also possible. Peixoto et al. [105,106] have already demonstrated this for the catalytic
cracking process by having aromatic, paraffinic and naphthenic structures be continuously
characterised by carbon number.
The continuous lumping methodology that was derived above can be seen to be an
extension of the work of pseudo-component lumping by Gupta et al. [45]. However, it
still suffers from the same limitations as those discussed in the discrete lumping section,
model parameters are still dependent on feedstock composition and/or catalyst, therefore
requiring refitting of parameters for different feedstock-catalyst pairs. From the model
described above, it is also evident that the continuous model description is less trivial
than in the discrete case, and therefore will be more computationally demanding. Another
limitation of continuous approximation that is worth noting for the reader is that discussed
by Ho and White [107] for long-time (or high-conversion) validity of the continuous
approximation. At longer times or high conversion, the reaction mixture gradually becomes
less crowded as great proportions of higher boiling point components are lost, and the
validity of the continuum assumption begins to break down, especially for low order
reactions in plug flow. Generally, the validity of the continuum approximation is dependent
upon the kinetics, feed properties, reactor type and pore diffusion, and this is described in
depth in the work of Ho and White [107].
For the mathematical description of continuous mixture approximation in the general
case, the reader is directed to see the literature [14,96,99,103,104,107–110]. For continuous
lumping in catalytic cracking the literature [102,105,111] is worth exploring.

2.3.3. Single Event Kinetics and Structure-Oriented Lumping


The single event kinetics approach to the modelling of acid-catalysed reactions in
complex mixtures was developed through a series of publications [22,23,112–115]. The
concept establishes the link between the reactivity of a chemical species in the mixture to its
geometry and a limited number of intrinsic kinetic parameters, which are only dependent
on the type of molecule and the type of reaction involved [72,116]. The methodology
behind this modelling approach is in describing the kinetics of the elementary steps that
occur in the complex reactive mixture, but it has already been established that the reaction
network is massive and therefore the number of elementary steps occurring is also gigantic.
However, what can be established from the carbenium mechanism described in the previous
section, and what was noticed [23] is that although the number of steps is large, they belong
to a limited number of types (i.e., exocyclic protolytic scissions, hydride abstractions,
methyl shifts, β-scissions, aromatisation etc.). Therefore, it becomes feasible to model
the reactions in such complex mixtures (such as the FCC riser) by decomposing them
into their elementary forms, in a similar fashion to the elementary radical chemistry
(initiation, propagation, chain transfer and radical termination) kinetic modelling used in
polymerisation [117,118]). In such a methodology, an understanding of the elementary steps
occurring in the reactor is required, Figure 5 shows some elementary reactions occurring
during cracking in the riser.
them into their elementary forms, in a similar fashion to the elementary radical chemistry
(initiation, propagation, chain transfer and radical termination) kinetic modelling used in
polymerisation [117,118]). In such a methodology, an understanding of the elementary
steps occurring in the reactor is required, Figure 5 shows some elementary reactions oc-
Energies 2022, 15, 308 18 of 50
curring during cracking in the riser.

Figure 5. Some elementary steps of paraffins, olefins, naphthenes and aromatics in the FCC riser.
(Adapted from [113]).

In order to determine the rate of any given elementary state, the transition state theory
Figure 5. Some elementary steps of paraffins, olefins, naphthenes and !aromatics in the FCC riser.
is used [113]:
(Adapted from [113]). 0 kB T ∆S06= ∆H 06=
k = exp − (22)
h R RT
In order to determine the
where k0 is the rate
rate of any
constant given
of the elementary
elementary state,
step, k B and h are the transition
the Boltzmann andstate the-
Planck
6 6
ory is used [113]: constants respectively, ∆S and ∆H are the standard change in entropy and change in
0 = 0 =

enthalpy associated with the reaction of the reactant to form the transition/intermediate
state. The change in standard entropy consists of contributions from translational, vibra-
tional and rotational. If the rotational contribution is taken as representative of the structure
and considering the possibility for chirality, then this contribution can be explained via
global symmetry numbers. It can be written that:
.. 0
S0 = S − Rln(σ ) (23)
.. 0
where S is the intrinsic entropy term and σ is the symmetry number of the species. The
entropy change of reaction for the formation of the activated complex or transition state is
then given by: σ 
06 = r
∆Ssym = Rln 6= (24)
σ
To account for chirality, a global symmetry number is introduced using the number of
chiral centres in the species, n:
σ
σgl = n (25)
2
Analysing for ∆S06= leads to:
! !
k T σgl,r ∆S06= ∆H 06=
k = B
0
exp − = ne k (26)
h σgl,6= R RT
Energies 2022, 15, 308 19 of 50

σgl,r
where ne = σgl,6= , ne is the number of the so-called single events. The rate of the elementary
k0
step is therefore a multiple of that of a single event, k. In this way the number of single
events has been separated from the rate coefficient for the elementary step and, therefore,
the effect of the structure of the carbenium ion on the change of entropy can be accounted
for. By way of the Polanyi equation [113], the effect of structure and carbon number on the
∆H 06= can be calculated for an exothermic step as follows:

Ea = E0 − α|∆Hr | (27)

and for the endothermic step:

Ea = E0 − (1 − α)|∆Hr | (28)

Ea is the activation energy, and is related to the rate constant of a single event through
an Arrhenius type expression for temperature dependency:
 
Ea
k = A exp − (29)
RT

Since the structure dependency has been taken out using the number of single events,
the Polanyi parameters (Ea , α) have unique values for any given type of elementary step.
The rate constants are, therefore, invariant with respect to the structure of the carbenium ion
for any single event or elementary step [22]. This is the fundamental advantage of this type
of model, that the kinetic rate constants are effectively invariant for that elementary step
for any complex mixture or feed. ∆Hr for a single event can be calculated from enthalpies
of formation for both the reactant and the product.
In order to carry out this analysis, a reaction network for the system must be estab-
lished so that all the elementary steps and single events involved may be determined. It has
already been established that the networks involved, especially for a mixture as complex
as that of FCC gas oil feed, is gigantic; therefore, it can only be sufficiently generated by
computer algorithms. This falls under the now well-established field of computer-assisted
chemistry or computational chemistry. Generally, species involved in the reaction network
are stored in the computer in the form of standardised labels, which can be broken down
into three vectors: one vector of dimension 1 and two vectors of dimension C (where C is
the number of carbon atoms in the molecule). The first vector contains the index of the
carbon atom carrying the charge, the first C-dimension vector carries information about
whether a carbon atom is primary (p), secondary (s), tertiary (t) or quaternary (q), and
the last vector indicates the type of carbon atoms present in the species (i.e., saturated,
unsaturated, cyclic etc). The labelling is such that the encoding is unique for each species,
and generally in a way that is consistent with the International Union of Pure and Applied
Chemistry (IUPAC) naming rules. In order to numerically represent the elementary steps
in the computer simulation, the vector representation of a reacting species is usually con-
verted to a Boolean connectivity matrix form. In this way, Boolean matrix elements for a
species mij and m ji are 1 if there exists a bond between atoms i and j. If no bond exists, then
the elements are 0. Such a storage system allows the structure of a molecule to be easily
computed. An auxiliary vector is used to store information about double bonds, and a
scalar is used to tract the position of the charged carbon in the species. Figure 6 shows how
Boolean connectivity matrices are used to compute elementary steps, by way of example, a
methyl shift elementary step is shown. This representation and computation of reaction
networks has been thoroughly reviewed elsewhere [23,112,115,116,119].
ecule to be easily computed. An auxiliary vector is used to store information about double
bonds, and a scalar is used to tract the position of the charged carbon in the species. Figure
6 shows how Boolean connectivity matrices are used to compute elementary steps, by way
of example, a methyl shift elementary step is shown. This representation and computation
Energies 2022, 15, 308 20 of 50
of reaction networks has been thoroughly reviewed elsewhere [23,112,115,116,119].

Figure 6. Representation Figure 6. Representation


of a methyl of a methylstep
shift elementary shift of
elementary step of aion:
a carbenium carbenium ion: (a)form,
(a) skeletal skeletal(b)
form,
(b) Boolean matrix representation, (c) auxiliary vectors for the transformation (7 = unsaturated carbon,
Boolean matrix representation, (c) auxiliary vectors for the transformation (7 = unsaturated carbon,
8 = saturated carbon for vector 3), (Adapted from [23]).
8 = saturated carbon for vector 3), (Adapted from [23]).
From the system described above, it should be clear that the generation and subse-
quent description of reaction networks is possible and can be automated via specialised
From the system described above, it should be clear that the generation and subse-
computer programs, or algorithms could be written for unique cases. In catalytic cracking
quent description of reaction
simulations,networks
specialisedis possible
automated and can
reaction be generation
network automated via specialised
algorithms have been used
computer programs, or algorithms could be written for unique cases. In catalytic cracking
by various researchers [112–116,120–128]. Once the network has been generated, it follows
that rate equations need to be determined so that they may be incorporated into a suitable
simulations, specialised automated reaction network generation algorithms have been
hydrodynamic model for the reactor.
used by various researchers [112–116,120–128].
To determine Once
the rate equation, the network
the method of Fromenthas
[23] been generated,
is followed it
here. Consider,
by way of example, the rate equation for the β-scission elementary step of a paraffinic
secondary carbenium ion R1+ into an acyclic olefin and an acyclic paraffinic carbenium ion
R2+ , the rate equation for the step is written as:

r β = ne k β (s; s)CR+ (30)


1

The inaccessible concentration of R1+ is then written in terms of measurable partial


pressures of paraffins and olefins, by assuming equilibrium or pseudo steady state balance
between formation and disappearance. Hence all reactions resulting in the formation of
the carbenium ion R1+ are considered together with those resulting in the removal of R+
from the reaction mixture. In the present example, the formation occurs by protonation
of olefins the same length as R+ , direct protonation of paraffins resulting in scission to
give R1+ and some other paraffin and hydride transfer between paraffin P1 and adsorbed
Energies 2022, 15, 308 21 of 50

carbenium ions, and the disappearance occurs by deprotonation and hydride transfer from
paraffins, therefore:
!
1
k pr (s)(1 − CB+ ) ∑(ne ) pr K 0 O j ↔ Or pOj + k protol (s) (1 − CB+ ) ∑(ne ) protol p Pj

CR + = N
1 j J (31)

+k ht CB+ (ne )ht p P1
where N = (ne ) av.De k De (s) = k ht ∑(ne ) protol p Pj .
j
K 0 is the equilibrium constant for isomerisation between O j and Or , (ne ) av.De is the
average number of deprotonation events of R1+ , and CB+ is the relative concentration of
the Brönsted acid sites on the catalyst that the carbenium ions are adsorbed to, so that at
equilibrium (steady-state):
CB+ = ∑ CR1+ (32)

Therefore, elimination of CR+ leads to the rate equation for the β-scission events in terms
1
of the partial pressures of species in the gas phase, which importantly, can be measured.
The network, which is undeniably too large, can be reduced by disregarding the
involvement of methyl and p type carbenium ions as they are less stable compared to
those of p, t and q- ions [23]. Furthermore, the resulting network can be reduced further by
introduction of some a posteriori lumping technique to group isomer species according
to their carbon number and their degrees of branching without losing much information
from the original network [72]. According to Froment [23], the lumps should be chosen in
terms of the analytical capabilities that are available to the study, so that initial values of
partial pressures required in the model described above are available or measurable. In
this way, the rate equation for the transformation between some lump L1 to another lump
L2 can be derived by summing the rates of all the elementary steps which convert all the
species making up of L1 (carbenuim ions and molecules) into species making up L2 . As an
example, the rate of isomerisation by protonated cyclopropane (PCP)-branching of some
paraffinic lump L1 to form a lump L2 [23]:
   
rPCP ( L1 → L2 ) = ∑ CR+ (s)ne kPCP (s; s) D (s) + ∑ CR+ (s)ne kPCP (s; t) D (s)
i i
s,s  s,t  (33)
+ ∑ CR+i (t)ne kPCP (t; s) D (t) + ∑ CR+i (t)ne kPCP (t; t) D (t)
t,s t,t

where ne is the number of PCP single events for (s; s), (s; t), (t; s) and (t; t) transformations
and D (s) and D (t) represents the fraction of s or t carbenium ions which desorb by hydride
transfer. Using the quasi-steady state approximation for the carbenium ion concentration
and accounting for the total site balance:

rPCP ( L1 → L2 ) = (LC)PCP (s; s) FPCP (s; s)kPCP (s; s) p L1 + (LC)PCP (s; t) FPCP (s; t)kPCP (s; t) p L1
(34)
+(LC)PCP (t; s) FPCP (t; s)kPCP (t; s) p L1 + (LC)PCP (t; t) FPCP (t; t)kPCP (t; t) p L1
The FPCP (i; j) contain the different single event rate coefficients, and (LC)PCP (i; j) are
the lumping coefficients that depend only on the network of elementary steps and choice
of lumps but not on the single event rate constants [23]. The lumping coefficients, for
global rate equations of the lumped model, can be calculated and stored for any given
reaction network [113]. Expressions like those above have been shown for cracking [124]
and for endocyclic β-scission [113]. Extension to other elementary reaction types is trivial
and can be found in the work of Surla et al. [116]. In order to estimate rate constants, the
kinetic model above is incorporated to a suitable transport model for the reactor, by way of
example, for simple plug flow model is shown below:

dFi (z)
= ρbulk Aϕri (35)
dz
Energies 2022, 15, 308 22 of 50

where Fi (z) is the molar flow rate of some component or lump i at some axial coordinate z of
the reactor, A is the reactor cross-sectional area, ϕ is the non-selective deactivation function,
ρbulk is the bulk density of the reactor bed and ri is the rate equation for the lump i as
defined in the kinetic model. From this point, the analysis is like that described for previous
kinetic models, where the experimental data are fitted by regression to some optimisation
function in order to find the kinetic parameters. The main advantage of the single event ki-
netic model is that the models produce kinetic parameters that extrapolate well, as the rates
are independent of feedstock, an assumption that was proven by Standl et al. [126] who
produced kinetic constants for the cracking of 1-pentene over ZSM-5 catalyst, and managed
to use those parameters to reproduce the experimental data for the cracking of different
olefins with impressive accuracy. This demonstrated one powerful observation, because the
rates are only concerned with the type of elementary step and structure of the molecule, the
rates can be calculated for cracking of simple molecules, which produce smaller reaction
networks such as the case of Standl et al. [126], and extrapolated for models with more
complex reaction schemes (such as that of gas oil cracking in FCC risers) without the need
to recalculate the rate constants. Beirnaert et al. [129] have calculated rate constants for
cracking of 2,24-trimethylpentane over USY zeolite, Quintana-Solórzano et al. [124] for
cracking of hydrotreated gas oil over RE-USY equilibrium catalyst, Quintana-Solórzano
et al. [123] for cracking of mixtures of 1-octene and cycloalkanes (cycloalkanes used: methyl-
cyclohexane and n-butylcyclohexane) over RE-USY catalyst, Quintana-Solórzano et al. [130]
for partially hydrotreated vacuum gas oil over RE-USY and Feng et al. [115] for cracking of
n-decane over RE-Y catalyst (shown in Table 5).

Table 5. Single-event rate kinetic parameters for cracking of paraffins over RE-Y catalyst by Feng
et al. [115].

−1 −1 EA
Parameter Value at 450 ◦ C (mol.kgcat .s )
(kJ.kmol−1 )
kHA (s) 8459 97, 870
kHA (t) 10.59 × 103 98, 570
kHD (s) 3.766 × 10−2 34, 520
kHD (t) 3.72 × 10−4 102, 500
kPr (s) 1.008 × 107 43, 920
kPr (t) 2.715 × 107 36, 360
kDe (s) 1.767 × 10−2 32, 980
kDe (t) 1.492 × 10−4 55, 680
kHS (s : s) 1.292 × 1012 117, 800
kHS (s : t) 9.212 × 108 27, 690
kHS (t : t) 1.331 × 1012 59, 720
kMS (s : s) 2.887 × 1013 34, 550
kMS (s : t) 2.204 × 108 45, 960
kMS (t : t) 5.107 × 1015 56, 480
kPCP (s : s) 1.985 × 1010 10, 600
kPCP (s : t) 2.902 × 1012 27, 270
kPCP (t : t) 1.283 × 109 88, 180
kLCr (s : s) 1.093 23, 860
kLCr (s : t) 3.984 51, 890
kLCr (t : s) 8.188 × 10−1 51, 030
kLCr (t : t) 1.702 × 101 64, 330
kBCr (s : s) 7.163 × 103 56, 820
kBCr (s : t) 6.573 × 102 21, 830
kBCr (t : s) 7.383 × 102 13, 450
kHT (s) 2.091 × 102 60, 120
kHT (t) 3.557 × 102 25, 570
Subscript: HA—Hydride abstraction, HD—Hydride donation, Pr—Protonation, De—Deprotonation, HS—
Hydride shift, MS—Methyl shift, PCP—Protonated cyclopropane, Cr—Cracking (i.e., β-scission), HT—Hydride
transfer. Superscripts: L—Lewis site, B—Bronsted site. Brackets: Type of carbenium ion: s—secondary, t—tertiary.
Energies 2022, 15, 308 23 of 50

Hopefully, what is evident from the above analysis of the single event kinetics is
that this approach solves the two fundamental problems that were identified in the two
previous classes of models; the transformations are no longer arbitrarily determined, and
the reaction rate constants are no longer feed-dependent and therefore extrapolate well
for other feeds. However, as with the other approaches, it carries its own disadvantages:
the model is highly computationally demanding, and the level of feed characterisation
required is more detailed than the previous models.
A short summary of the different modelling approaches is presented in Table 6. One
other issue of note that has not been discussed here is the catalyst. The work described so
far establishes that different catalysts have been used by various studies over the years as
the FCC catalyst technology has evolved and developed. Each of the methods discussed
here may be used with various catalysts, however, the values of the kinetic parameters
may differ owing to the catalyst’s selectivity and activity. On the other hand, because of the
fundamental nature of the single event model, kinetic parameters from this method have
been shown to extrapolate well for various feeds as long as the catalyst used remains the
same. This attribute is an indicator that these kinetic parameters are intrinsic reaction rate
parameters that are catalyst specific but not feed specific for any given elementary step,
as opposed to the empirical parameters founding from lumping. Therefore, for a lumped
methodology, it is important to consider both the catalyst and feed characteristics before
kinetic parameters are used for predictive purposes.

Table 6. Comparitive summary of the different kinetic modelling approaches in FCC riser unit.

Feed and Product


Model Reaction Chemistry Main Features
Characterisation
Reactions considered as one-to-one
Model parameters are the frequency
transformations between lumps, usually
factors and activation energies of
Components of the reaction higher boiling point lumps to lower boiling
the lump-to-lump transformations.
mixture grouped into lumps point lumps, although the connections
Discrete These models are simple and less
based on boiling point range between the lumps have usually varied
lumping computationally demanding
or carbon number and between different researchers. This approach
because the large complex reaction
molecular structure. is empirical and not representative of the
mixture has been reduced to a small
carbenium ion chemistry that is known to
number of lumps.
govern cracking reactions.
Properties of individual components in the
mixture such as reactivity can be indexed via
normalised boiling point. The individual The model parameters are the
Reaction mixture thought of
components in the mixture are assumed to parameters associated with the
as a continuous mixture of an
Continuous all be involved in the same type of reaction, relationship between reactivity and
infinite number of molecules,
Lumping i.e., all assumed to be undergoing normalised TBP (e.g., k max & α) and
which are characterised by
fragmentation or cracking; however, it is parameters associated with the
normalised TBP.
known that other types of reactions are also yield function.
taking place in the riser due to carbenium ion
chemistry (e.g., isomerisation, alkylation etc.)
The parameters are the frequency
factors and activation energies of
A computer algorithm creates a reaction the specific elementary steps.
network of elementary steps between the Very detailed feed characterisation
molecules in the reaction mixture based on is required to determine the values
Single
Molecular reconstruction carbenium ion theory. The algorithm uses the of the kinetic parameters which
event
methods numerical representation of the reacting makes this model very time
kinetics
species together with the reaction rules to consuming. However, because of
create a large network consisting of all the fundamental nature of the
intermediate carbenium ions. reaction chemistry, the parameters
determined are intrinsic and will
extrapolate well for changing feeds.
Energies 2022, 15, 308 24 of 50

2.3.4. Catalyst Deactivation Function


Cerqueira et al. [131] have reviewed the phenomenon of catalyst deactivation of FCC
catalysts and have divided deactivation in such heterogenous catalysts as being caused by
physical or chemical means. The former includes phenomena such as sintering and occlu-
sion, while the latter includes chemical degradation by materials such as alkaline metals,
irreversible poisoning by permanent adsorption of impurities on the active sites and fouling
by coke deposition. The work only focuses on deactivation by deposition of coke as this is
the deactivation that is reversed in the regenerator and usually the primary source of deacti-
vation [132,133]. Coke deposition occurs very rapidly in the FCC riser [134] and, therefore,
the activity of the catalyst is quickly reduced in the lower height of the riser. Deactivation
by coke is usually either by poisoning of acid sites where coke molecules block the acid sites
directly, affecting the activity linearly and also possibly affecting the selectivity [135] or by
pore blockage where the coke impedes reactants’ access to the reaction site [131,136]. The
effect of pore blockage is much more significant [134,137]. Evidence of deactivation by coke
has been showing by [42,138] by comparing the N2 absorption-desorption data of clean and
spent catalyst. They have found that spent catalysts underwent physical deterioration of
physical properties associated with catalytic activity (e.g., Brunauer-Emmett-Teller (BET),
meso- and micropore areas), with the deterioration being feed-type dependent. Changes in
average pore diameter after coke deposition also confirms the suggestion that it influences
selectivity. In terms of modelling this phenomenon, a difficulty arises in the mathematical
description of the rate of deactivation [139]. In kinetic modelling, the effect of deactivation
is incorporated by including the so-called deactivation function (ϕ above), which is defined
as follows [139]:
r
ϕi = 0i (36)
ri
rCK
ϕCK = 0
(37)
rCK
where subscript 0 corresponds to the rate of reaction on clean catalyst, so that the presence
of coke affects both the rate of production of reation i and the further production of
coke itself. The above accommodate the possibility that the deactivation function may
be different for each reactant or product, which is called selective deactivation, or that
the deactivation function is the same for all the reactions (non-selective deactivation).
Appropriate mathematical functions for ϕ are such that they can describe the initially high
catalyst activity that diminishes rapidly followed by a levelling off (see Table 7) e.g., the
exponential function, A exp(− Bt) where A and B are parameters to be simultaneously
estimated with the kinetic rate parameters, and t may be residence time or coke on catalyst.
For the former, deactivation function is referred to as a time-on-stream (TOS) model and the
latter are referred to as coke-on-catalyst (COC) models. In the case of lumping kinetics, the
deactivation model parameters (i.e., A & B) add to the number of kinetic parameters being
estimated which can result in over-parameterization [83]. Bollas et al. [83] have compared
different deactivation scenarios, non-selective deactivation, reactant-oriented deactivation
and product-oriented deactivation, incorporated into a five-lump kinetic model and found
out that non-selective deactivation was sufficient in describing the deactivation phenomena.
This is consistent with what was reported by Corella [140]. In fact, the only case that Bollas
found where substantial advantage was gained by using selective deactivation was in the
case of product-oriented deactivation. However, the fact that non-selective deactivation
is proven sufficient and also reduces model parameters has resulted in most deactivation
models found in the literature being non-selective. Table 7 summarises some of the widely
used types of deactivation functions in the literature.
Energies 2022, 15, 308 25 of 50

Table 7. Summary of types of deactivation catalysts used in the literature (Redrawn from [7]).

Weekman [15] Jacob et al. [17]


α
ϕ = exp(−αtc ) ϕ=
TOS ( pGO )(1+ βtγc )
ϕ = t−
c
n
α, β, γ are model
α, n model parameters parameters
Farag et al. [141] Krambeck [142] Pitault et al. [143]
1−( 1 ) ϕ = B+exp B +1
COC ϕ = exp(−Kc CCK ) ϕ= CCK b ( A·CCK )
Kc is a model parameter b is a model parameter B, A are model parameters
tc = residence time (s). CCK = coke on catalyst concentration (kg coke/ kg catalyst). pGO = partial pressure of oil at
riser bottom.

Because of the empirical nature of the deactivation model parameters, these models
suffer from being feedstock-catalyst pair dependent, which is expected according to data
from Rodriguez et al. [42]. However, the extent of this dependence has yet to be fully
studied in the literature to the same extent as that of the dependency of lumping kinetic
parameters to the feedstock–catalyst combination. Therefore, the extrapolation of any
model parameters to different feeds may be poor especially the differences in the rates of
decay of activity between the two approaches to modelling deactivation in Table 7 [77]. To
account for some of the variations in decay rates, Fernandes et al. [144] recommend the use

of the first differential form of the deactivation function, dx = −α x ϕd , where x is either
residence time or content of coke on catalyst (wt%). They have compared the model in
both TOS and COC modes and have found similar profiles for open loop responses, for
both approaches, with some minor differences. However, they have not compared their
models against different experimental data to assess how well both models extrapolate to
different feeds. Hence, the level to which the first order differential proposed accounts for
the variations across different feeds is not yet known. Nevertheless, deactivation models in
Table 7 have continued to produce satisfactory results.

2.4. Riser Hydrodynamics


The flow in the riser involves the gas and the solids moving up the riser height and
exiting at the top where they are separated via the disengaging section. The solids are then
sent to the regenerator for regeneration of deactivated catalyst and subsequently recycled
back to the riser at the bottom. This arrangement is sometimes referred to as a circulating
fluidised bed (CFB) arrangement. The riser is then operated in the fast fluidisation regime.
Unlike in the bubbling/turbulent bed in the regenerator, the upper surface of fast fluid
beds is more diffuse because of greater freeboard activity [145]. At higher gas velocities,
there exists a transport velocity, Vt , beyond which the rate of particle carry-over in the
riser increases sharply so that if particles are not recycled at the bottom, then the bed
would run empty very quickly [33,145]. Hence, the catalyst circulation rate is important
in regulating the dilution of the flow inside the riser. In fact, Berruti et al. [32] and Kunii
and Levenspiel [1] have shown that beyond the transition to turbulent flow, two modes
of operation are present: (i) fixed inventory system (FIS) where the solid mass flux is a
dependent variable, and (ii) variable inventory system (VIS) where the solid mass flux is an
independent variable. In the FIS system, there is no controllable solid storage inventory
in the riser; the solid inventory is established by setting the gas velocity and solid charge;
whereas VIS systems vary the gas and the solid mass flux to achieve a certain solid inventory
in the riser. As will also be seen in the regenerator, the understanding of the flow and
gas–solid mixing characteristics is important in order to model the system. The gas–solid
flow behaviour inside the FCC riser has been well studied using cold flow CFB setups at
pilot scale, using a variety of measurement techniques such as pressure taps, high frame
rate cameras, optical probes, radioactive tracers, laser Doppler anemometry (LDR) and
positron emission particle tracking (PEPT). These studies have already been reviewed
elsewhere [7,32,146,147] and will not be re-evaluated in this present work, only some of
their conclusions will be discussed here.
Energies 2022, 15, 308 26 of 50

2.4.1. Axial and Radial Profiles


Pressure, velocity and axial profiles have been measured at different locations along
the riser height and radial positions. Figure 7 shows the variation of solids volume fraction
in the axial direction of the riser reported by researchers [148–152] and collated by Shah
et al. [7]. Shah and colleagues [7] describe this axial profiles as having a characteristic ‘S-
Energies 2022, 15, 308 28 of 52
shape’. This sort of profile in solids fraction is discussed in greater detail in the regenerator;
however, the profiles in Figure 7 show a smoother transition between the dense bottom
and the lean freeboard region. In fact, from Figure 7 it can be concluded that the boundary
boundary
between thebetween the dense
dense bottom andbottom and theisfreeboard
the freeboard more diffuse, is more
unlikediffuse,
in theunlike in the
profiles for pro-
the
regenerator, and may not be discernible by the naked eye. Although the experimentsthe
files for the regenerator, and may not be discernible by the naked eye. Although fromex-
periments from which the axial profiles are measured are carried
which the axial profiles are measured are carried out at conditions dissimilar from those out at conditions dis-
ofsimilar from those
an industrial FCCof an industrial
riser in normal FCC riser in
operation, thenormal
resultsoperation, the results
are nevertheless are never-
important for
theless important
understanding for inside
of flow understanding of flow
the FCC riser. Whatinside theofFCC
is also noteriser.
is thatWhat is also of
the collated note
data areis
thatcompiled
also the collated
from data are also
a variety of compiled from aand
flow conditions variety of flowagreement
its general conditions(i.e.,
andinitsterms
general
of
agreement
the ‘s-shape’ (i.e., in terms
profile) does giveof the ‘s-shape’inprofile)
confidence does give
this observed confidence
result. From the in datathis
in observed
Figure 7
weresult. From the
can make data inobservation
a general Figure 7 wethat can make a general
the bottom dense observation
region solids that volume
the bottom dense
fraction
region solids volume fraction lies in the range 0.2 − 0.4, while the top
lies in the range 0.2–0.4, while the top dilute (‘lean’) region solids volume fraction is in the dilute (‘lean’) region
solids0–0.1.
range volumeThefraction
transition is in
region can be0seen
the range − 0.1.toThe
occupytransition region
the region can be0.2–0.5
between seen to inoccupy
terms
ofthe
theregion between riser,
dimensionless 0.2 − which
0.5 in terms of the to
corresponds dimensionless
about 30% ofriser, which
the riser corresponds
height. However, to
these 30% ofwill
aboutresults thebe riser height. However,
influenced these and
by gas velocity results will be
catalyst influenced
circulation by gas
rates. velocity
Pärssinen
andZhu
and catalyst
[152]circulation
have shown rates.
thatPärssinen
at constantandcatalyst
Zhu [152] have shown
circulation rate,that
the at constant cata-
decreasing the
gas
lystvelocity resulted
circulation rate, intheadecreasing
clearer distinction betweenresulted
the gas velocity the bottom in a dense
clearerregion and the
distinction be-
top freeboard. This is expected as lowering the gas velocity approaches
tween the bottom dense region and the top freeboard. This is expected as lowering the gas the regime of
regenerator operation.the
velocity approaches Higher
regime velocity would then
of regenerator clearly lead
operation. Higherto a smoother transition
velocity would then
between the top
clearly lead to aand bottomtransition
smoother regions. between the top and bottom regions.

Figure 7. Axial solid volume profiles along the riser height (Redrawn from Shah et al. [7] for vari-
Figure 7. Axial solid volume profiles along the riser height (Redrawn from Shah et al. [7] for various
ous gas and solid flux).
gas and solid flux).

Derouinetetal.
Derouin al.[153]
[153]have
havereported
reportedradial
radialvariation
variationofofgas
gasvelocity
velocityand
andcatalyst
catalystflux
flux
from cold flow of FCC catalyst in a CFB riser (Figure 8a,b respectively), and
from cold flow of FCC catalyst in a CFB riser (Figure 8a,b respectively), and Miller and Miller and
Gidaspow [154] have reported radial variation of solids volume fraction (Figure
Gidaspow [154] have reported radial variation of solids volume fraction (Figure 8c). The 8c). The
radial solids profiles show a solid lean region in the centre of the riser and a solid rich
region near the walls of the riser. This is the so-called core–annulus radial profile, which
shows a solid lean core moving upward with the gas surrounded by a solid rich annus
layer at the walls of the riser. These observations have also been made by various research-
ers [155–158]. The mean solids concentration is observed to decrease from bottom to the
Energies 2022, 15, 308 27 of 50

radial solids profiles show a solid lean region in the centre of the riser and a solid rich region
near the walls of the riser. This is the so-called core–annulus radial profile, which shows a
solid lean core moving upward with the gas surrounded by a solid rich annus layer at the
walls of the riser. These observations have also been made by various researchers [155–158].
The mean solids concentration is observed to decrease from bottom to the top of the
riser [153], and higher gas velocities increase solids volume fraction at the same radial
position and increases the solids concentration in the annulus [154] (see Figure 8c). The
gas and solids in the core are seen to move faster than in the regions near the wall. The
radial profiles are often described as being quadratic about the centreline [32], however,
the observed symmetry is not always perfect [153]. Two reasons have been suggested for
the two separate solids flows in the core and annulus region: (i) energy dissipation due
to friction at the wall and particle collisions [7], (ii) and formations of clusters inside the
suspension [154,159,160]. Yerushalmi and Cankurt [161] are some of the initial reported
observations of clusters in CFB risers, where solids are seen moving in groups (or ‘strands’
and ‘ribbons’) within the rising gas. Recent studies [162,163] have also captured movement
Energies 2022, 15, 308 29 of 52
of clusters along the riser wall. They have found that clusters can vary dramatically in terms
of size, from a few particles to length scales in the order of a few riser diameters. Horio
The radial profiles are often described as being quadratic about the centreline [32], how-
and Kuroki [164,165]ever, have the also
observedreported
symmetry thatis not clusters
always perfect are[153].
notTwo only confined
reasons have been tosug-
the riser wall
but can be found in the gestedcore region.
for the two separateThese
solidsclusters
flows in thethen
core andpresent a flow
annulus region: inhomogeneity
(i) energy dissi- that
pation due to friction at the wall and particle collisions [7], (ii) and formations of clusters
is different from thatinside
observed in the regenerator where the voids are
the suspension [154,159,160]. Yerushalmi and Cankurt [161] are some of the initial seen rising through
the emulsion and, therefore, presents
reported observations a different
of clusters in CFB risers,challenge
where solids with regards
are seen moving into modelling the
groups
(or ‘strands’ and ‘ribbons’) within the rising gas. Recent studies [162,163] have also cap-
flow. Figure 8b also tured
shows that ofsolids
movement clusters flow
along the inriser
thewall.annulus
They havecanfound bethatnegative
clusters can i.e.,
vary downflow
of solids instead of upward
dramatically co-current
in terms of size, flow
from a with gas.toThis
few particles length downward
scales in the ordersolid
of a fewvelocity has
riser diameters. Horio and Kuroki [164,165] have also reported that clusters are not only
also been observed by various researchers [166,167]. Of course, downward
confined to the riser wall but can be found in the core region. These clusters then present flow of solids
at the wall was alsoa flow observed
inhomogeneityin solids back-mixing
that is different in bubbling/turbulent
from that observed in the regenerator where the beds. The
voids are seen rising through the emulsion and, therefore, presents a different challenge
definition of the diameter
with regardsof to
the core is
modelling thesomewhat
flow. Figure 8b alsodebatable, the flow
shows that solids core–annulus
in the annulus boundary
can be defined as the can beposition
negative i.e.,where
downflowthe radial
of solids insteadvelocity profileflow
of upward co-current of with
thegas.
solids
This becomes
downward solid velocity has also been observed by various researchers [166,167]. Of
negative or it can becourse,
defined as the position where the time-averaged
downward flow of solids at the wall was also observed in solids back-mixing in particle velocity is
zero, with the two different assumptions giving different results especially in terms of
bubbling/turbulent beds. The definition of the diameter of the core is somewhat debatable,
the core–annulus boundary can be defined as the position where the radial velocity profile
modelling [145]. Experiments have,
of the solids becomes however,
negative or it can beshown
defined asthat the core
the position where occupies about 60–80%
the time-averaged
of the riser diameter particle
whilevelocity
the annulus
is zero, withoccupies
the two different about 20–40%
assumptions ofdifferent
giving the riser cross-section
results espe- [7].
cially in terms of modelling [145]. Experiments have, however, shown that the core occu-
It is reasonable to assume that the ranges given are large enough
pies about 60 − 80% of the riser diameter while the annulus occupies about 20 − 40% ofthat the choice of core
diameter does not affect the
the riser result.[7]. It is reasonable to assume that the ranges given are large enough
cross-section
that the choice of core diameter does not affect the result.

Figure 8. (a) Radial gas velocity profiles, (redrawn with data from [153] for Ug = 5.2 m/s) (b) radial
profiles of catalyst flux, (redrawn with data from [153] for Ug = 5.2 m/s) and (c) Radial solid volume
profile (redrawn with data from [154] for Ug = 2.89 m/s), for different solid flux.
Energies 2022, 15, 308 28 of 50

2.4.2. Gas and Solid Mixing


For gas mixing, a distinction can be made between mixing of gas at the macro and
micro scales [145]. The former relates to gas-solid mixing patterns inside the riser unit at
reactor scale, which greatly affect the vertical mixing phenomena, and to the turbulent
eddy currents at the small scale [168]. On the other hand, micro-mixing is related to
molecular transport phenomena such as diffusion in porous particles and is usually only of
significance in cases of fast reactions where the reactions are mass transfer limited [169]. In
cases of chemical reaction-limited processes, the concentration of gas at the reaction site in
heterogenous systems is assumed to be the same as that of the bulk flow, so that only meso
and macro mixing phenomena is important. At the meso-scale, gas mixing is influenced by
various flow structures such as clusters, and by the existence of different regions (such as
dense bottom and particle lean freeboard) at the macro-scale [33]. Therefore, it is clear that
the gas mixing occurring in the riser is different in specific regions of the riser [170,171].
Because of variations in process conditions and their influence on gas mixing, it is expected
that the concepts of mixing will be different depending on the process [172]. In processes
such as the catalytic cracking in FCC risers, the product of the process is the gas therefore
the back-mixing of gas is undesirable [33]. In these cases, process variables such as gas flow
rate and catalyst circulation rate are optimised to minimise the level of gas back-mixing in
the riser. Quantitative study of the mixing behaviour of gases in the CFB riser is usually
carried out using gas tracer experiments. Experimental data (such as concentration of
tracer gas at various positions in the riser) is plotted against a suitable mixing model (such
as dispersion or residence time distribution (RTD). Schugerl [173] suggested that three
dispersion coefficients are required to quantitatively describe mixing of a gas in gas-solid
fluidised beds: Dga , which characterises the extent of axial back-mixing of gas, Dgr , which
characterises the extent of radial dispersion, and lastly Dg which is the effective dispersion
coefficient. Schugerl [173] reports that these coefficients are related by:

D2
 
Dg = Dga + βv2 (38)
Dgr

where v is the gas velocity, D is the vessel diameter and β is a dimensionless measure of the
1
radial flow profile (i.e., parabolic flows have β = 192 ). Dga can be obtained from RTD data
of the tracer gas, and Dgr can be evaluated from tracer gas concentration measurements
obtained at different lateral positions in the riser that are downstream of the tracer injection
position. Grace et al. [33] have collated data from various researchers [171,174–177] and
Dga
plotted the variation of the inverse Peclet number ( Dv ) with superficial gas velocity. Their
plot showed that axial gas dispersion increases as solids mass flux increases (while gas flow
rate is unchanged) and it increases as gas velocity decreases (while solid flux is unchanged).
As gas velocity increases towards pneumatic transport regime, Li and Wu [171] reported
that the spread of the RTD decreased showing that the amount of gas back-mixing also
decreased. On the other hand, gas back-mixing increased when the solid density of the
suspension increased, mainly as a result of the increased solid downflow that is observed
at the wall wall region [176,177]. RTD comparison of solids and gases by Van Zoonen [175]
revealed that axial dispersion of gas is lower compared to that of solids, but both RTD data
show flow patterns that deviate from what is expected for plug flow. Bai [177] concluded
that while the degree of axial dispersion of the gas is small compared to that of the solids,
it still is significant enough that it should not be ignored. On the other hand, early data
from researchers [161,167,178] suggested that back-mixing of gas decreased substantially
beyond the turbulent regime towards fast and pneumatic fluidisation regimes, so that
gas may be assumed to have substantial plug flow in that regime. However, that claim
was questioned by Li and Weinstein [170] who reported considerable gas back-mixing in
fast fluidisation regime, clearly in contrast to the earlier work. The results from Werther
et al. [179] suggested that regions of the bed with low solid concentrations (such as the
dilute upward moving core) may be sufficiently described using plug flow for gas mixing,
Energies 2022, 15, 308 29 of 50

which suggests the gas behaviour in the fast velocity risers is much better accounted for
with a two-region model that have separate descriptions for gas mixing in each region (i.e.,
the so-called core-annular models). In this case, the plug flow of gas would be assumed
in the lean core and dispersion accounted for in the solid rich annular region. According
to Li and Weinstein [170], the complicated axial gas mixing characteristics are due to the
presence of macroscopic flow structures (i.e., clusters) that form particle groups of various
sizes, in both the vertical and lateral directions of the CFB risers. The result is both spatial
and temporal distributed heterogeneity in the riser suspension. As a result, time-averaged
measurements at some fixed position in the riser may not average over this observed
heterogeneity [170]. Observations of clusters in CFB risers have already been described in
the previous section.
On the other hand, some research groups [180,181] have found evidence of poor
lateral mixing in industrial CFB risers. They reported three crucial observations about
radial dispersion: (a) dispersion is maximum in the absence of solids, (b) dilute transport in
the riser reduces Dgr , and (c) increased solid density of the suspension increases turbulence
and therefore an increase in gas mixing. These results show that solids can enhance or
dampen gas turbulence in the riser depending on their concentration. Therefore, which
effect solids have on the gas should be considered in modelling gas mixing. It is usually
difficult to compare the results of gas dispersion coefficients from different researchers
because of a wide variation of results, stemming from differences in such things as injection
locations of tracer material, injection technique and models applied [146]. However, from
the results discussed here it is evident that the gas mixing (both radial and axial) is highly
influenced by the concentration of solids in the bed. The downflow of solids, which is
usually observed at the walls of the bed, results in increased gas back-mixing, and this is
affected by both the gas velocity and the solids flux. The poor lateral mixing of the gas is
clearly a significant downside of the CFB risers especially in areas where a high level of
gas-solid mixing is required such as the FCC riser. It is therefore important that models
reflect this, hence the need for more 2/3D models of the FCC riser.
Generally, studies that investigate mixing of particles in CFB risers have been carried
out in a similar manner to that used for gases: fitting experimental data to dispersion or
RTD models. By studying pressure changes at the bottom dense zone of a cold flow CFB
riser, Svensson et al. [182,183] found that this region is hydrodynamically quite similar to
that of bubbling bed regime, like behaviour seen in the FCC regenerator. Most of the gas is
seen passing through this region in the form of bubbles, or empty voids. Wei et al. [184]
using FCC particles in a commercial scale 5.8 m diameter riser, reported data that suggest
that the solid axial dispersion coefficient (Dsa ) was somewhat dependent on the riser
diameter. This result suggests that wider riser reactors experience more solids back mixing
compared to narrow beds. If only the riser diameter is changed without increasing the gas
flow rate, then it is easy to see why this result would hold, as wider risers would have
lower superficial gas velocities. Lee and Kim [185] have studied axial solid mixing in a
turbulent bed using quasi steady state heat transport in the axial direction. They have
correlated the axial solids Peclet number to the Archimedes number:

Pe = 4.22 × 10−3 Ar (39)

In the turbulent regime it was shown that fine particles exhibited higher dispersion
coefficients. The authors attribute this to differences in wake fractions, influence of turbu-
lence and motion of solid clusters. Increase in gas velocity was shown to result in increase
in both axial and radial solid dispersion coefficients. This may reasonably be explained by
the fact that higher gas velocities increase the turbulence in the bed resulting in better heat
and mass transfer.
This section has discussed some of the experimental studies into the mixing behaviour
of gas and solid in CFB risers. It is evident that the mixing characteristic of the reacting
suspension in the FCC riser is dependent upon solid concentration and gas–solid flow rates.
There is some ambiguity in the literature about whether plug flow assumption is valid to
Energies 2022, 15, 308 30 of 50

describe the mixing of gas and solids, although researchers seem to lean towards plug flow
only being suitable for dilute beds, or regimes of the bed. Solid mixing is more complex
due to the observed core-annular profile and downflow near the wall, hence, to model the
system more realistically these features must be considered. The next section reviews some
of the common modelling approaches of the FCC riser, based on the lessons described here
from the literature.

2.5. Hydrodynamic Models of Circulating Fluidised Bed (CFB) Risers


The distribution of gas and solids in the CFB riser has already been discussed, and most
experimental data indicate that there is an uneven distribution of both the gas and the solids
across the riser cross section. This observation is no doubt important to the performance of
such riser reactors and, therefore, important to its modelling. Three macroscopic balance
equations are fundamental to the modelling of the FCC riser; material balance of transport
equation for species to describe the evolution of species concentration across the reactor,
momentum balances for solids and gas to describe their velocities in the riser, and an energy
balance to describe the evolution of temperature across the reactor. For material balances, a
review of the literature on mixing in the riser reveals two types of models used to describe
the system: (i) pseudo-homogenous models, and (ii) two-region or core–annulus models.
The type (i) models are usually simple models that assume a single-phase suspension
flowing up the riser and so they take no regard of the heterogeneity present in the riser.
On the other hand, type (ii) models make a distinction between the annulus and flow
in the core to account for the observed difference between these two regions in the riser.
The modelling assumed here is that of the fully developed region of the riser which is
downstream of the gas entry point. Of course, the bottom region of the FCC riser (the
so-called vaporisation region) is more complicated as it is the region where gas oil feed and
lifting steam meets the hot catalyst, and the feed is subsequently vaporised. There is then a
simultaneous existence of solids, liquid and gas, which makes it more complex to describe
in mathematical models. As a result, this region will be treated separately.
• Pseudo-homogeneous models These assume that flow is turbulent, but non-isotropic
so that variations present in the vertical directions may be different from those occur-
ring horizontally [168], but no macroscopic flow heterogeneity in different regions
of the bed is made, that is, at the macro scale flow is essentially similar through the
bed. The mass balance equation will usually include dispersion coefficients for both
axial and radial positions to account for mixing. Grace et al. [33] present the following
general equation for simple homogeneous dispersion models:

∂2 C
 
∂C U ∂C ∂ ∂C
=− + Dga + Dgr r + S + RC (40)
∂t ε ∂z ∂z ∂r ∂r

where S and RC are the source term and the rate equation for the gas component as
defined by the kinetic model. Note that here only the gas equation is shown as this is
usually the important mass balance for the process with gaseous products. In the case
of one-dimensional models [161,171,184,186,187] the radial term may be dropped so
that the dispersion equation becomes:

∂C U ∂C ∂2 C
=− + Dg + S + RC (41)
∂t ε ∂z ∂z
Other possible simplifications include the assumption of steady state, which is quite
common in FCC riser modelling due to the perceived insignificance of the riser dy-
namics compared to that of the regenerator [5]. If plug flow is assumed for the gas,
Dg = 0, and if the gas is assumed to be well mixed, Dg = ∞.
• Core–annulus models These models are based on the core–annular profile that was
observed in fully developed CFB risers by various researchers. The generalised core
Energies 2022, 15, 308 31 of 50

equations are shown below, contributions for core-annulus interchange from [32] and
contributions for dispersion from [179]:

∂2 C
 
∂C ∂C ∂ ∂C 2k
= −U + Dga + Dgr r − (Cc − Ca ) (42)
∂t ∂z ∂z ∂r ∂r rc

Both axial and radial dispersion have been included, but for 1D models either one
can be ignored. the transient term may also be ignored for steady-state simulation,
which is a common assumption for modelling FCC risers. The gas in the core is usually
assumed to be stagnant or be descending at relatively low velocity, hence convective
flow terms may be ignored, the material balance equations for gas in this region are
given by [32]:
∂C 2krc
− 2 (Cc − Ca ) = 0 (43)
∂t R − rc2
rc , R are the core radius and the riser radius respectively, and subscripts a, c are the
annulus and the core regions respectively. The convective flow terms may be included
for the solid phase in order to capture the effect of the downward flow of solids in the
annular region that is shown in Figure 8b.
The material balance equations are coupled to the momentum equations via velocities
of both solids and gases. Sometimes, especially in cases where velocity is considered
constant or given empirically [188] researchers have found it sufficient to only consider
the material balance equations. But in systems such as the FCC riser where considerable
expansions are expected during vaporisation and, therefore, acceleration is expected,
evolution of velocity is found via momentum equations.
For momentum balances, Tsuo and Gidaspow [189] have collapsed the Navier–Stokes
equations to provide one-dimensional momentum balance equations for solids and gases
along a CFB riser reactor. The 1-D equations are derived by averaging over the pipe radius,
and the viscosity terms are replaced by Fanning friction-type coefficients ( f i for i = s, g) in
order to account for the energy loses within the flowing phase. All pressure drop is assumed
to be on the gas phase, and a solid compressive stress modulus (G) term is introduced on
the solids phase [189]:
For gases:

 2 f g ε g ρ g v2g
 
∂ ρg ε g vg ∂ ρg ε g vg vg ∂P
+ =− + β vs − v g + + ε g ρg g (44)
∂t ∂z ∂z D
For solids:

∂(ρs ε s vs ) ∂(ρs ε s vs vs ) ∂ε s  2 f s ε s ρs v2s


+ = −G + β v g − vs + + ε s ρs g (45)
∂t ∂z ∂z D
Above, the term with the β accounts for the friction between the two phases, and the
bracketed term corresponds to the slip velocity between the phases. This slip velocity is
important because the particles in the riser will generally have a longer residence time than
the gases. Admittedly, Tsuo and Gidaspow [189] have reported that the 1-D momentum
balance model is poor at predicting the cluster formation or solid reverse flow in CFB risers
and, therefore, often poorly predicts the velocity and volume fractions. However, because
reaction engineering models of the FCC risers are largely 1-D models, this equation (or its
variants) is usually used [5,45]. Short reviews for the correlations for β and f i are available
in the literature [7,190]. The above description uses what is commonly called two-fluid
modelling or Eulerian-Eulerian (E-E) phase description where the continuum assumption is
made [7,13,190–192]). Alternatively, the Eulerian-Langrangian (E-L) approach may be used
where the continuum assumptions for the gas still holds but the particle are considered
as a discrete phase [7]. For a cluster of particles, whose diameter dcl is known, a force
Energies 2022, 15, 308 32 of 50

balance is carried out around the cluster to determine the evolution of particle velocity in
the riser [193–195]:
 2 
C v − v
!
3
πdcl ∂vs D ρ g g c πd2cl ρ g πd3cl g ρc πd3cl g
ρ = + − (46)
6 ∂t 2 4 6 6

where CD is the gas–solid interphase coefficient or drag coefficient similar to β above in the
E-E momentum equations. Corelations for dcl relative to the particle diameter exist in the
literature [196]. Both E-E and E-L modelling approaches are now well established and have
been thoroughly reviewed elsewhere [7,192].
A review of the literature on FCC riser models (Table 8) show that there are generally
two approaches to describing the temperature of the riser: (i) thermal equilibrium being
catalyst and gas phase, and (ii) considering the thermal resistance between the catalyst and
the gas phase. In type (i) models, only one energy balance is made for the riser where it is
assumed that both phases have the same temperature at any axial and/or radial position.
The steady-state energy balance in this case for plug flow of gas and solids can be written
as [197]:
∂T A RS ε c ∑ ∆Hi ri
= (47)
∂z Fc Cpc + ∑ Fj Cp j
where i is the lump-to-lump reaction iterator, and j is the lump iterator, ∆Hi is the enthalpy
change of reaction of lump-to-lump reaction i and ri is the rate of reaction i. For type (ii)
models, separate equations for both catalyst and gas phase are developed so that exchange
of energy between the two phases is considered:

∂Tg A RS ka gs Tc − Tg
= (48)
∂z Fg Cp g

A RS ka gs Tg − Tc
 
∂Tc
= + ρc ε c ∑ ∆Hi ri (49)
∂z Fg Cp g
a gs is the effective heat transfer area per unit volume at the gas–solid interphase and
h gs is the heat transfer coefficient between the gas and the catalyst, which is usually taken
from Nusselt number correlations for the flow. Note that here the heat required for the
cracking reactions is assumed to be from the catalyst side as reactions are assumed to be
occurring on the catalyst surface; however, other researchers have used the gas side for the
heat of cracking [5]. The boundary equations for the differential equations discussed here
are such that conditions at z = 0 are equal to the conditions at the feed vaporisation section.

2.6. Feed Vaporisation


The vaporisation process occurs at the bottom of riser, where preheated atomised
gas oil feed, atomising steam and the hot regenerated catalyst from the regenerator come
into contact and mix. During this contact, the atomised feed droplets undergo intense
mixing and vaporise almost instantly [198]. The heat and mass transfer processes occurring
during this stage of the riser are very complex, which introduces a great challenge to the
modelling of FCC risers. The heat from the catalyst provides both the sensible heat, the heat
required for vaporisation of the feedstock and the heat to power the highly endothermic
cracking reactions occurring in the FCC riser, hence the regenerated catalyst temperature is
a crucial parameter in the FCC riser system. Because of a high degree of turbulence, large
thermal gradients and high flow inhomogeneity in the vaporisation section, it makes it the
most complex region to model in the entirety of the riser reactor [199]. Studies have also
shown that the diameters of the atomised feed [200,201], the nozzle and particle wetting
characteristics of particles [202–205] also play crucial roles in feed vaporisation, proving the
complexity of the vaporisation section to modelling. However, according to plant data from
Ali et al. [199], the entirety of the vaporisation process only occurs in 0.1 s, corresponding to
Energies 2022, 15, 308 33 of 50

only about 3% of the residence time of catalyst in the riser. For this reason, a large majority
of researchers have assumed instant vaporisation of the feed gas oil; therefore, this section
of the riser can be treated as a heat-transfer section (corresponding to z = 0 for the riser
axial position) in which catalyst, steam and feed meet. This is described mathematically by
for catalyst and gas respectively [5]:
!
FGO  Fs C ps
TgFS − Ts + ∆HvGO

Tc = TcRG − C pGO TGO − TgFS + (50)
Fc C pc FGO

B
TgFS = −C (51)
A − log( P × yGO )
where A, B and C are Antoine coefficients for the gas oil feed and P × yGO is the partial
pressure of the gas oil after vaporisation. T, F, C are the temperature, mass flow rate and
heat capacities respectively, with subscripts c, GO, s corresponding to catalyst, gas oil and
dispersion steam and FS, RG corresponding to feed section and regenerator respectively.
∆HvGO is the enthalpy change of vaporisation of gas oil, correlations for it are available in
the literature. To account for vapor phase expansion after vaporisation, a simple gas law
equation with compressibility factor is used:

Fs + FGO
vg = (52)
ρ g (1 − ε cRG ) A RG

Mw g
ρg = P × (53)
RTgFS Zg

where A RG , Mw g , Zg are the riser cross sectional area, average molar mass of the gas
phase and the compressibility factor of the gas phase. The initial catalyst velocity may be
calculated from the circulation rate:
FcRG
Fc = (54)
ρc ε cRG A RG

The above describes the simple vaporisation section model (for z = 0) where instant
vaporisation is assumed. Contrary to this, the vaporisation section may be described using
more detailed models that account for the penetration length of the droplet phase through
the bed. For this, momentum and material balance equations for the droplet phase are
included so that they are coupled with those for the catalyst and the gaseous phase, using
either E-E or E-L approaches. These models have been thoroughly reviewed elsewhere in
the literature [7,198]. Table 8 summarizes the FCC riser models in the literature.
Energies 2022, 15, 308 34 of 50

Table 8. Summary of FCC riser models in the literature.

Momentum Balance
Authors 1/2D Kinetic Approach Kinetic Model Deactivation Function Vaporisation Material Balance Energy Balance
Approach
10-lump model Parameters from [18] Pseudo-homogenous phase,
[2] 1D COC by [142] No slip Thermal equilibrium
of [18] Coking tendency by [18,142,206] plug flow of gas and solids.
Pseudo-homogenous phase, N/A (gas-solid slip
646 partially lumped components
COC by [18], N2 plug flow of gas. Empirical accounted for but
[113] 1D Single event kinetics and reaction network with Instant Thermal equilibrium
deactivation considered. coke formation from not shown in
44169 global reactions.
Jacob [17] equations)
GS, GS, LG, CK lumps Two-phase gas-solid flow, Interphase resistance
[5,6] 1D 4-lump model COC Instant E-E
considered, parameters from [80] plug flow of gas and solids. considered
50 pseudo-components Compartments in series for
Pseudo-component Instant
[45] 1D considered, rates characterised by COC by [143] gas and solids E-E Thermal equilibrium
model IG molar expansion
normal boiling temp and MR. (pseudo-homogeneous).
6-lump model of VGO, DO, LCO, GS, LPG, FG Pseudo-homogenous phase,
[207] 1D COC Instant E-L Thermal equilibrium
Takatsuka [208] lumps considered plug flow of gas.
VGO, LCO, GS, LPG, FG, CK
Pseudo-homogenous phase,
6-lump model lumps considered, inverse Interphase resistance
[209,210] 2D COC N/A plug flow of gas and solids E-E
of [211] parameter estimation used for considered.
with axial and radial axis.
parameters
Empirical Slip factor
GS, LG, GS, CK lumps Plug flow for gas and solid
[188] 1D 4-lump model of [80] TOS by [14] Instant correlation by Thermal equilibrium
considered, parameters from [80] (pseudo-homogeneous).
Patience [212]
Kinetic parameters expressed
Semi-continuous using the beta distribution [214] Plug flow for gas and solid
[213] 1D COC by [137] Instant N/A Thermal equilibrium
model as a function of boiling point (pseudo-homogeneous).
temp
GS, GS, LG, CK lumps
Pseudo-homogenous phase,
[88] 1D 4-lump model considered, parameters TOS by [14] Instant N/A Thermal equilibrium
plug flow of gas and solids.
form [188]
6-lump model of VGO, LCO, GS, LPG, FG, CK Plug flow for gas and solid
[197] 1D COC instant E-E Thermal equilibrium
Takatsuka et al. [208] lumps considered (pseudo-homogeneous).
Energies 2022, 15, 308 35 of 50

Table 8. Cont.

Momentum Balance
Authors 1/2D Kinetic Approach Kinetic Model Deactivation Function Vaporisation Material Balance Energy Balance
Approach
42 pseudo components
Pseudo-component considered, rates characterised by Pseudo-homogenous phase,
[215] 1D COC from [45] Instant E-E Thermal equilibrium
model normal boiling temp (∆T = 30 ◦ C) plug flow of gas and solids.
and densities.
VGO, GS, LG, CK lumps Compartments or CSTR in
[216] 1D 4-Lump model considered, kinetic parameters COC by [143] instant series for gas and solids E-L Thermal equilibrium
of [75,143] (pseudo-homogeneous).
GO, GS, CK + LG lumps
Pseudo-homogenous phase,
[217] 1D 3-lump model of [74] considered, parameters from [74] COC Instant No slip Thermal equilibrium
plug flow of gas and solids
Coking rate by Lee [218]
Pseudo components considered;
Pseudo-component Pseudo-homogenous phase,
[219] 1D rates characterised by normal COC by [143] Instant E-E Thermal equilibrium
model plug flow of gas and solids
boiling temp (∆T = 15–30 ◦ C).
GO, DZ, GS, LPG, DG, CK lumps
Two-phase gas-solid flow, Interphase resistance
[85] 1D 6-lump model considered, parameter estimation COC as defined by [6] Instant E-E
plug flow of gas and solids. considered
for parameters.
Feed molecular reconstruction
Auto reaction generation similar
Hybrid COC from [86]
to [221,222] Reaction network
structure-oriented Deactivation by N2 and Pseudo-homogenous phase,
[220] 1D with 3827 molecules and Instant N/A Isothermal reactor
model and heavy aromatics plug flow of gas and solids
7572 reactions.
bond-electron matrix from [86]
Adsorption and diffusion
considered.
Lumps: GO—Gas oil, VGO—Vacuum gas oil, LCO—Light cycle oil, GS—Gasoline, LG—Light gases, DZ—Diesel, CK—Coke, LPG—Liquid petroleum gas. Deactivation: COC—Coke-
on-catalyst, TOS—Time-on-stream. Momentum equations phase description: E-E—Eulerian, E-L—Eulerian–Lagrangian. Pseudo-component modelling: ∆T—Temperature difference for
the narrow cuts in True boiling point curve corresponding to a pseudo-component. N/A—Equation not included in the model, or not shown on the journal article. Material balance:
Pseudo-homogeneous implies solid–gas suspension is treated as a single-phase ignoring core–annulus flow structure.
2.7. Riser Performance Prediction
The prediction of the models, employing different kinetic approaches, is com
against known experimental data available in the literature.
Energies 2022, 15, 308 Figure 9 shows a comparison of model predicted profiles of gas 36oilof 50 conversi
ported by various researchers [45,188,197,223] to the commercial data reported by
ouin et al.Performance
2.7. Riser [153] andPrediction
pilot scale plant data reported by Shah et al. [224]. All the simu
gas oilTheconversion
prediction profiles showemploying
of the models, similar trends to kinetic
different the measured
approaches, data along the riser h
is compared
Both setsknown
against of data, model predictions
experimental data availableand measured
in the literature. data, show a steep rise in the co
sion within the first 20% of the riser height, thenprofiles
Figure 9 shows a comparison of model predicted followed of gas
byoil conversion re-
a plateauing effect th
ported by various researchers [45,188,197,223] to the commercial data reported by Derouin
to the rest of the riser. This is because the bottom of the riser is where the catalyst is
et al. [153] and pilot scale plant data reported by Shah et al. [224]. All the simulation gas oil
active, but profiles
conversion it is quickly deactivated
show similar trends toby the the formation
measured of coke,
data along thereby
the riser height. resulting
Both in
reactions formodel
sets of data, the remainder
predictionsof and the reactor.data,
measured Thisshowis also evident
a steep rise ininthe
the temperature p
conversion
inwithin
Figure the10 first 20% of the
reported byriser height,
[6,219] then followed
showing steep by a plateauing gradient
temperature effect through to first 2
in the
the rest of the riser. This is because the bottom of the riser is where
the riser followed by a levelling off. Figure 9 shows that typical gas oil conversion is the catalyst is most
active, but it is quickly deactivated by the formation of coke, thereby resulting in slow
range 45 − 80%. Shah et al. [7] have attributed the discrepancy in the final gas oil co
reactions for the remainder of the reactor. This is also evident in the temperature profiles in
sion of 10
Figure [153,224]
reportedto by the flow
[6,219] conditions.
showing The feedgradient
steep temperature and theincatalyst flow
the first 20% rates are 5
of the
times
riser higher
followedinbythe case of Derouin
a levelling off. Figurethan thatthat
9 shows of Shah,
typicalresulting in higher
gas oil conversion gas
is in theoil conv
inrange 45–80%.
the case of Shah et al. [7]
Derouin. have attributed
Therefore, the discrepancy
the wide range ininthe the final
gas gasoil oil conversion predict
conversion
of [153,224] to the flow conditions. The feed and the catalyst flow rates are 5–10 times
various researchers in Figure 9 can be attributed to differences in flow condition
higher in the case of Derouin than that of Shah, resulting in higher gas oil conversion in the
impact
case of of the catalyst
Derouin. Therefore,to the
thewide
finalrange
conversion
in the gas isoilalso noteworthy
conversion predicted asbydata of Deroui
various
Shah are conducted
researchers in Figure 9two can bedecades
attributed apart, and in that
to differences in flowsame periodThe
conditions. there wasofa switch
impact
the catalyst to the final conversion is also noteworthy
use of zeolite Y to ZSM-5 in the FCC unit which resulted in higher conversionsas data of Derouin and Shah are [225
conducted two decades apart, and in that same period
early 2000s saw another shift to hierarchical pore architecture zeolite catalystthere was a switch from use of (e.g.
zeolite Y to ZSM-5 in the FCC unit which resulted in higher conversions [225]. The early
Meso
2000sY) sawthat further
another shiftbumped
to hierarchicalup theporeconversion; therefore,
architecture zeolite catalystit(e.g.,
is reasonable
NaY, Meso Y) to assum
some of the bumped
that further discrepancies in the gas therefore,
up the conversion; oil conversion reported
it is reasonable to is attributable
assume that someto the ca
used,
of theespecially
discrepancies because
in the gaslump kinetics are
oil conversion reliant
reported of the catalyst
is attributable to theand feed
catalyst used.
used,
especially because lump kinetics are reliant of the catalyst and feed used.

90

80

70
Gas oil conversion (wt%)

60

50

40 Zhu 2011 (4-lump)


Derouin 1997 real plant data
30
Shah 1977 pilot plant data
20 Gupta 2007 (pseudocomponent

10 Fernandes 2003 (6-lump)


Ahari 2008 (4-lump)
0
0 0.2 0.4 0.6 0.8 1
Dimensionless riser height (-)
Figure 9. Comparison of gas oil conversion predicted profiles of FCC riser with measured data of
Shah 1977 [224] and Derouin 1997 [153].
Figure 9. Comparison of gas oil conversion predicted profiles of FCC riser with measured d
Shah 1977 [224] and Derouin 1997 [153].

In Figure 10 in the case where heat transfer resistance is considered between th


alyst and the gas, the vaporisation at the bottom of the riser is assumed to only he
20% of the riser height for gas to approach the catalyst temperature which casts dou
the assumption of other cases where this is also assumed to occur instantaneously. F
10 shows that the exit temperature of the catalyst from the riser is in the range 780 −
Energies 2022, 15, 308 or 507 − 527 °C which is consistent with data on the spent catalyst entry37temperatur
of 50
the regenerator which we have shown in Part II of this paper.

950

900

850
Temperature (K)

800

750
catalyst (In-Su Han 2001)

gas (In-Su Han 2001)


700
Catalyst + gas (Yong Li et al 2017)

650
0 0.2 0.4 0.6 0.8 1
Dimensionless riser height (-)
Figure
Figure10.
10.Comparison
Comparison ofof predicted
predictedtemperature
temperature profiles
profiles of models
of models with gas–catalyst
with gas–catalyst thermal therma
librium and with
equilibrium catalyst–gas
and with catalyst–gasinterphase resistance.
interphase resistance.

In Figure 10 in the case where heat transfer resistance is considered between the
2.8. Shortcomings
catalyst andthe
and the gas, Future Recommendations
vaporisation at the bottom of the riser is assumed to only heat
up gas
Theoilfollowing
to the temperature
shortcomingsat whichhave
it all turns
beentoidentified
vapor; therefore,
from thethegas continues
review to riser
of FCC
rise
els: in temperature up the riser. In this period, the hot catalyst provides both this sensible
heat and that heat required for cracking, hence catalyst temperature goes down while gas
(a) temperature
Vaporisation:increases.theOnvaporisation
the other hand,has in themostly been assumed
case of thermal equilibriumtobetween
be instanta
(Table 8) to avoid complicating the model. The recent review by Nguyen
catalyst and gas, the gas is vaporised and then heated to the same temperature as the et al.
catalyst
showed in thethat
vaporisation
models that section, and then
describe the catalyst–gas
vaporisation suspension
of atomised providesin
droplets thefluidised
necessary heat for the cracking reactions hence its temperature decreases along the height
show wide variations in vaporisation times depending on flow
of the riser. It may be reasonable to assume that vaporisation occurs so rapidly that it is
conditions and,
fore, even(as
instantaneous if the assumption
discussed of in
earlier), as rapid
both vaporisation
cases discussedmay
here, be
butvalid in some cases, i
the subsequent
not be
sensible intemperature
heat others. Hence, gain of this
thesection
gas is shownneedsintothebe considered
case of Han andmoreChung rigorously
[6]
thatmodels
it takes about 20% of the riser height for gas to approach the catalyst
to distinguish between the two cases, as ignoring it may lead to inacc temperature
which casts doubt on the assumption of other cases where this is also assumed to occur
predictions in certain cases.
instantaneously. Figure 10 shows that the exit temperature of the catalyst from the riser is
in the range 780–800 K or 507–527 ◦ C which is consistent with data on the spent catalyst
entry temperature into the regenerator which we have shown in Part II of this paper.

2.8. Shortcomings and Future Recommendations


The following shortcomings have been identified from the review of FCC riser models:
(a) Vaporisation: the vaporisation has mostly been assumed to be instantaneous (Table 8)
to avoid complicating the model. The recent review by Nguyen et al. [198] showed
that models that describe vaporisation of atomised droplets in fluidised beds show
wide variations in vaporisation times depending on flow conditions and, therefore,
Energies 2022, 15, 308 38 of 50

even if the assumption of rapid vaporisation may be valid in some cases, it may not
be in others. Hence, this section needs to be considered more rigorously in the models
to distinguish between the two cases, as ignoring it may lead to inaccurate predictions
in certain cases.
(b) Gas and solid dispersion: experimental data from tracer studies considered in this
work showed that plug flow of gas may only be assumed in the case where the con-
centration of solids in the riser is low. However, Table 8 shows that researchers almost
always assumed plug flow of gas, which may cause some inaccurate predictions in
some cases where the plug flow assumption is inadequate. Hence, it is recommended
that dispersion be considered, especially since the literature on dispersion coefficients
of both solids and gases in CFB risers is vast.
(c) Cluster formation and core–annular phase inhomogeneity: the literature about the
flow structure of CFB risers is now generally in agreement about the core–annulus
pattern in the fully developed region of the riser. However, all the models discussed
in Table 8 have ignored this phenomenon and assumed that no flow separation is
observed (i.e., pseudo-homogeneous) and, therefore, cannot predict the core–annulus
profile or its effect on riser performance. 1D momentum balance equations for the
riser have also been shown by Tsuo and Gidaspow [189] to be poor at describing the
effect of particle clusters on flow, hence more 2D models are required to better predict
the riser performance.
(d) Kinetic modelling: Table 8 shows that most workers have overwhelmingly used
discrete lumping methodology for describing the reactions in the FCC riser. Several
workers have reported the values for the kinetic parameters for the conversion be-
tween different lumps which are then used by secondary workers in their models
without any recalibration. However, as was discussed before, the lump parameters are
very feedstock and catalyst dependent, and as shown by Shah et al. [7], these lumped
kinetic parameters failed to predict well the experimental data of Nace et al. [16]. This
is a clear indication that such parameters are highly influenced by reaction condi-
tions, and therefore are more error prone when extrapolated to different operating
conditions. The world has now come to a point where the computational power is
great enough that molecular and structure-oriented lumping techniques such as single
event kinetics which have been proved to produce parameters that are feedstock
independent, are now worth pursuing. The literature on such parameters has also
grown; therefore, such models can be incorporated in FCC models without the need
for new parameter estimation exercises.
(e) Thermal balances: the majority of models in Table 8 assume that heat transfer re-
sistance between the two phases may be ignored; however, as was shown by the
model from [5], thermal equilibrium between the phases is not reached in the most
significant region of the riser (first 20%) and, therefore, this would be expected to
affect the model predictions. There is also a need to explore how the various heat
transfer coefficients at the gas–catalyst interphase affect the model predictions as this
is yet to be investigated.
(f) Catalyst deactivation: various models for catalyst deactivation are available in the
literature and have been used by resarchers for FCC riser simulations. However, the
effect of these different deactivation models on the riser predictions is not yet fully
understood, hence a comparative study of the effect off different deactivation models
is required.

3. Modelling FCC Unit Constitutive Components


Disengager and Stripper
This disengaging and the stripping sections are riser termination devices forming
the top part of the FCC reactor vessel; these sections, often modelled together as one
unit [5,6,207], are where the catalyst is separated from the product gases. Because of this
separation of gas and solid phases, it is commonplace to assume that no further cracking
Energies 2022, 15, 308 39 of 50

reactions occur in this unit [5]. Stripping steam is used to remove the product gas entrained
with the catalyst and remove some unreacted hydrocarbons that may be adsorbed to the
catalyst surface or within the catalyst pores. These hydrocarbons in/on the catalyst serve
as coke precursors that eventually form the so-called catalyst-to-oil (CTO) coke [5,226].
From the regenerator point of view, the unreacted hydrocarbons behave just like coke in
that their combustion releases heat that can be used to power the riser reactions; however,
these hydrocarbons produce up to 3.7 times more heat compared to regular coke which
can result in temperature swings in the regenerator [226], and therefore their removal from
spent catalyst is of importance to the controlled operation of the FCC unit. The stripping
steam displaces these unreacted hydrocarbons from the catalyst and pushes them up to be
removed with the other product gases, so that CTO coke is removed from spent catalyst.
The major important variables in the disengaging/stripping section are the holdups
of catalyst and gas, the flow rates and concentrations of exit flows (i.e., flows to the stack
and deactivated catalyst flow to the regenerator) and the pressure drop. Hence, because
the flow patterns in the unit are less important, only global conservation balances around
the unit are required to calculate these important process variables. Before balances can be
made for this section, the CTO coke needs to be considered together with the catalytic coke
from the riser reactions. Han and Chung [5,6] have developed the following expression for
unstripped CTO coke concentration (wt% of catalyst), depending on the stripping steam
rate and the CTO ratio:

CckCTO = CckCTO0 + k ss exp(−( Ess (CTO) Fss ) (55)

where Fss is the stripping steam rate, Ess and k ss are parameters of the stripping function,
CTO is the ratio of catalyst flow rate to feed oil flow rate, CckCTO0 is the minimum coke
content attainable by stripping and CckCTO is the concentration of CTO coke unstripped
from the catalyst. Alternatively, Fernandes et al. [207] developed an empirical function of
stripper temperature based on pilot plant data to calculate the contribution of CTO coke to
coke-on-catalyst concentration in the stripper:

CckCTO = exp(5.2113 − 0.0144TST ) (56)

Both expressions of CckCTO given above consist of a constant followed by an additional


term which is exponential in some property of the flow (e.g., temperature or flow rate). In
contrast, Arbel et al. [2] developed a correlation for CckCTO where the additional term was
a linear function of stripping steam flow rate:

CckCTO = 0.0002 + 0.0018(1 − kFss ) (57)

where k is the stripping coefficient (k = 0.08) and Fss is the stripping steam flow rate.
Ultimately, the correlations produced for CckCTO are predicting the concentration of un-
stripped CTO coke on the catalyst surface, which is a quantity related to the efficiency of
the stripper. The stripper efficiency is a measure of how well the stripper removes CTO
residue from the spent catalyst surface. Garcia-Dopico and Garcia [226] have reviewed the
literature data on stripper efficiency, their analysis found that the efficiency is dependent
on temperature, pressure, steam flow rate and catalyst residence time. The correlations
given above for CckCTO , which can be treated here as a proxy for stripper efficiency, show
the dependency on temperature and steam flow rate, while the pressure dependency is
not included. Researchers [226] have explained that efficiency decreases at higher pressure
because partial pressure for vaporising hydrocarbons is greater making stripping difficult.
However, because of the pressure balance in the FCC unit, they explain that pressure cannot
be changed in the stripper without affecting the entire unit operation, therefore the effect
Energies 2022, 15, 308 40 of 50

of pressure in the variation of efficiency may be assumed constant. Be that as it may, they
have developed the following stripper efficiency function using data from the literature:
   τST 
( FCTORS − FCTOST )
 
9445 0.225PST 
ηST = = 1001 − 1.6 exp − 12.5   (58)
FCTORS TST M steam
catalyst

(F −F )
where CTORS CTOST
FCTORS translates to the mass of CTO removed by stripping steam as a
fraction of the CTO residue entering the stripper unit, M steam is the mass flow rate of
catalyst
steam per flow rate of catalyst through the stripper, PST is the pressure inside the stripper
and τST is the catalyst residence time in the stripper. The stripper efficiency correlation
can be used to determine the CckCTO using a simple balance equation for the CTO residue.
However, this pathway requires knowledge of CTO residue concentration at the riser outlet,
hence the kinetic model used in this analysis should be able to make a distinction between
different types of coke produced in the riser so that CTO residue concentration at the riser
outlet can be calculated. Such kinetic models have been produced and published in the
literature [227]. In the absence of such models, the correlations for CckCTO described above
should be sufficient.
For the total concentration of coke-on-catalyst in the stripper to be a sum of the catalytic
coke from the riser outlet and the CTO coke, the following calculation is used:

CckST = CckCTO + CckRS (59)

Following the analysis of Han and Chung [5,6], the mass balances for the gas and
catalyst phases over the disengage/stripping unit are as follows:

dw g
= FgRS + Fss − FgRT − FcRS CckST (60)
dt
dwc
= FcRS − FcSV − FcMF (61)
dt
In the gas equation, FgRS is the gas flow rate from the riser termination, Fss is the steam
flow rate into the stripper, FgRT is the exit flow rate of gas flowing from the reactor unit to
the main fractionator and the last term is the flow rate of CTO coke precursor gas entrained
in the steam via the stripping process. In the catalyst equation, FcRS is the catalyst flow
rate terminating from the riser, FcSV is the flow rate of spent catalyst through the slide
valve in the transport line to the regenerator and the last term represents the flow rate of
unseparated catalyst entrained with gaseous components to the main fractionating vessel.
It is important to note at this point that most researchers [2,5,6,207,217] are in agreement
that, contrary to the riser where a steady state balance was sufficient, the holdup and
residence times of the stripper section are important to its operation and the FCC unit
as whole, hence dynamic balances are required in this section. The coke balance over
the disengage/stripper unit allows the determination of coke-on-catalyst concentration
entering the regenerator. The balance, using the CST assumption and considering both
catalytic and CTO coke is as follows [5]:

dCckRT F
= cRS (CckRS + CckCTO − CckRT ) (62)
dt wc

where CckRT is the total coke-on-catalyst concentration in the disengager/stripper which


is equal to the concentration of coke on spent catalyst leaving the reactor vessel to the
regenerator and CckCTO is the unstripped CTO coke concentration. Arbel et al. [2] have also
included a term in their coke balance to account for coke that is a result of Conradson Carbon
Residue (CCR). CCR coke is seldom explicitly included in many FCC models [5,6,207].
This type of residue is different from the coke described above and is formed when heavy
hydrocarbons in the feed are thermally decomposed on the surface of the catalyst. Because
Energies 2022, 15, 308 41 of 50

of the non-catalytic nature of the residue, it will normally occur on non-active sites of the
catalyst [2] and, therefore, has a reduced effect on catalyst deactivation. This, together with
the limited data on differentiation of different types of coke produced in the FCC riser unit,
could be the reason why it is not explicitly accounted for in most FCC models. Nevertheless,
a lot of useful results have been produced even with its exclusion from modelling analysis.
Gas component balances over the disengage/stripper using the assumption of CST
can be used to predict the evolution of product gas concentrations in the wet gas flowing to
the main fractionating column:
!
dyig yig dw g
 
y F C
= RSi FgRS − cRS ckRS − FgRT + ; for i, j = gas lumps considered (63)
dt wg ∑ j y jRS wg dt

Temperature in the stripper can be predicted using a simple energy balance around
the disengager/stripper unit, under the CST and gas–solid thermodynamic equilibrium
assumptions [5,6,144,195]:

 dTST out
   
wc C pc + w g C pg = Fss Hss − Fss HwRT + FcRS Hcin − Hcout + FgRS Hgin − Hgout − Fss ∆HST + Qloss (64)
dt
j
where Hi is the enthalpy of i (= catalyst or gas phase from riser outlet) and j indicates the
stream going in or out of the stripper. H can be calculated from:
 
Hi = C pi TiRS − Tre f (65)

where C pi is the average specific heat catalyst for i phase and Tre f is some reference tem-
perature from which enthalpy is measured. ∆HST is the change in enthalpy associated
with the stripping process. Qloss is the heat lost from the system through the walls of the
unit to the surroundings, the simple heat exchanger equation can be used to estimate this
heat loss (i.e., Qloss = U.Ae f f ( TST − Tair ) where symbols have their usual definitions). The
adiabatic assumption, where Qloss is assumed to be zero, is also a common simplification
made for this system by various workers [2,217] with great utility. However, to the best
of our knowledge, no paper has provided evidence that the adiabatic assumption affects
the predictions in any significant way, and heat loss to the environment is still an impor-
tant aspect of the energy considerations of the process, and therefore we recommend its
inclusion in the models for completeness. Another common simplification is to assume
that the stripping steam does not affect the energy balance in a significant way so that the
Fss Hss term disappears. However, this is usually only valid if the stripping steam flow rate
is small compared to that of the catalyst and the riser outlet gases.
Pressure and pressure drop in the disengaging-stripper are the other important process
variables which should be included in the model. This is because the difference in pressure
between the stripping section and the regenerator dense bed determines the flow rate
of spent catalyst (i.e., catalyst circulation rate) between the two units and, therefore, the
interaction and coupling between the regenerator and the reactor system [5] Pressure along
the disengaging-stripper section is estimated using the ideal gas equation, with deviation
from ideal gas behaviour accounted for via the compressibility factor [5,6,144,196,197]:

w g RTST Zg
PST = (66)
Mwg VgST

with Zg as the compressibility factor of the gas phase, Mwg is the average molar mass of the
gas phase and VgST is the volume of the disengaging-stripping vessel occupied by gas. VgST
is calculated from subtracting the volume of solid holdup in the vessel from the overall
Energies 2022, 15, 308 42 of 50

vessel volume. The pressure at the base of the stripper, however, is higher than PST due to
the static pressure exerted by the catalyst bed [5]. This pressure can be estimated from:

wc ghST
PSTbottom = PST + (67)
VST

The equations given in this section are the model for the disengaging-stripping section,
predicting exit concentrations from the reactor system, coke-on-catalyst concentration
on spent catalyst and the pressure at the bottom of the stripper which controls catalyst
circulation rate. Although it was not explicitly stated, the boundary conditions of these
equations are such that the entry conditions (concentrations, temperature etc.) into the
disengaging-stripping unit are the same as the riser termination conditions. Therefore,
in an integrated model of the FCC unit, these boundary conditions are constrained by
the riser.

4. Conclusions
Overall, modelling of the FCC unit using traditional reaction engineering approaches
has certainly come a long way. This is especially evident in the riser kinetics through the
introduction and development of approaches such as continuous lumping and single event
kinetics. Single-event kinetic models establish a more fundamental way to understand
the reactions occurring in the riser using current knowledge of carbenium chemistry.
Additionally, analytical technology has come a long way since the early 1960s when FCC
modelling first began, with tools now able to provide more detailed characterisation of
FCC reaction mixture, which has greatly advanced the fundamental approach to kinetic
modelling. However, lumped modelling remains the most common methodology used by
researchers because of their relative ease and low computational cost, which make them
ideal for optimisation and control studies of the FCC unit. In terms of hydrodynamics, the
regime of fluidisation of the FCC riser unit means there is significant flow heterogeneity
during its normal operation. Clusters of catalyst particles are observed in the suspension
and a core–annular profile consisting of a solid dense annulus and a lean core is observed
in the fully developed region of the riser. In contrast, we demonstrate in Part II of this paper
that the heterogeneity in the regenerator arises, instead, from bubbles rising through an
emulsion phase and with radial profile showing a solid lean phase in areas of high bubble
density. This heterogeneity is an important factor in the conversion in the vessel and has
been modelled to varying degrees by researchers. In the riser, more work is still needed
to incorporate the core–annular structure that has been observed experimentally, into the
riser models. This requires that more models be developed in 2D so that the impacts of this
observed flow structure to FCC yields can be better understood. Lastly, FCC models have
mostly simplified the vaporisation section by assuming that this process is instantaneous.
More detailed models of vaporisation have been developed in the last two decades that
deal with the process in time scales that are comparable to those in the riser; therefore, we
recommend that more work is done to incorporate these new vaporisation models into the
riser so that more reliable and realistic models are developed.

Author Contributions: Conceptualization, T.W.S., R.P., I.M.M. and Y.M.J.; methodology, T.W.S. and
Y.M.J.; software, T.W.S.; validation, T.W.S. and Y.M.J.; formal analysis, T.W.S., R.P., I.M.M. and Y.M.J.;
investi-gation, T.W.S. and Y.M.J.; resources, T.W.S., R.P., I.M.M. and Y.M.J.; data curation, T.W.S.;
writing—original draft preparation, T.W.S.; writing—review and editing, T.W.S., R.P., I.M.M. and
Y.M.J.; visualization, T.W.S.; supervision, R.P., I.M.M. and Y.M.J. All authors have read and agreed to
the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Energies 2022, 15, 308 43 of 50

Acknowledgments: The main author (T.W.S.) would like to acknowledge the Government of
Botswana for funding his PhD research studies.
Conflicts of Interest: There is no conflict of interest regarding the publication of this research.

References
1. Kunii, D.; Levenspiel, O. Fluidization Engineering; Butterworth-Heinemann: Oxford, UK, 1991.
2. Arbel, A.; Huang, Z.; Rinard, I.H.; Shinnar, R.; Sapre, A.V. Dynamic and Control of Fluidized Catalytic Crackers. 1. Modeling of
the Current Generation of FCC’s. Ind. Eng. Chem. Res. 1995, 34, 1228–1243. [CrossRef]
3. Kunii, D.; Levenspiel, O. Entrainment of solids from fluidized beds I. Hold-up of solids in the freeboard II. Operation of fast
fluidized beds. Powder Technol. 1990, 61, 193–206. [CrossRef]
4. Kunii, D.; Levenspiel, O. Fluidized reactor models. 1. For bubbling beds of fine, intermediate, and large particles. 2. For the lean
phase: Freeboard and fast fluidization. Ind. Eng. Chem. Res. 1990, 29, 1226–1234. [CrossRef]
5. Han, I.-S.; Chung, C.-B. Dynamic modeling and simulation of a fluidized catalytic cracking process. Part I: Process modeling.
Chem. Eng. Sci. 2001, 56, 1951–1971. [CrossRef]
6. Han, I.-S.; Chung, C.-B. Dynamic modeling and simulation of a fluidized catalytic cracking process. Part II: Property estimation
and simulation. Chem. Eng. Sci. 2001, 56, 1973–1990. [CrossRef]
7. Shah, M.; Utikar, R.P.; Pareek, V.K.; Evans, G.M.; Joshi, J.B. Computational fluid dynamic modelling of FCC riser: A review. Chem.
Eng. Res. Des. 2016, 111, 403–448. [CrossRef]
8. Zhang, X.; Zhou, Y.; Zhu, J. Enhanced fluidization of group A particles modulated by group C powder. Powder Technol. 2020, 377,
684–692. [CrossRef]
9. Shaul, S.; Rabinovich, E.; Kalman, H. Typical Fluidization Characteristics for Geldart’s Classification Groups. Part. Sci. Technol.
2014, 32, 197–205. [CrossRef]
10. Zhou, Y.; Xu, J.; Zhu, J. Different bubble behaviors in gas-solid fluidized bed of Geldart group A and group C+ particles. Powder
Technol. 2021, 384, 431–441. [CrossRef]
11. Kong, W.; Tan, T.; Baeyens, J.; Flamant, G.; Zhang, H. Bubbling and Slugging of Geldart Group A Powders in Small Diameter
Columns. Ind. Eng. Chem. Res. 2017, 56, 4136–4144. [CrossRef]
12. Lee, J.-R.; Lee, K.-S.; Hasolli, N.; Park, Y.-O.; Lee, K.-Y.; Kim, Y.-H. Fluidization and mixing behaviors of Geldart groups A, B and
C particles assisted by vertical vibration in fluidized bed. Chem. Eng. Process. Process. Intensif. 2020, 149, 107856. [CrossRef]
13. Pinheiro, C.I.C.; Fernandes, J.L.; Domingues, L.; Chambel, A.J.S.; Graça, I.; Oliveira, N.M.C.; Cerqueira, H.S.; Ribeiro, F.R. Fluid
Catalytic Cracking (FCC) Process Modeling, Simulation, and Control. Ind. Eng. Chem. Res. 2011, 51, 1–29. [CrossRef]
14. Weekman, V. Lumps, Models, and Kinetics in Practice; A.I.Ch.E. Monograph Series: New York, NY, USA, 1979; Volume 75, pp. 1–29.
15. Weekman, V.W., Jr. Kinetics and Dynamics of Catalytic Cracking Selectivity in Fixed-Bed Reactors. Ind. Eng. Chem. Process. Des.
Dev. 1969, 8, 385–391. [CrossRef]
16. Nace, D.M.; Voltz, S.E.; Weekman, V.W., Jr. Application of a Kinetic Model for Catalytic Cracking. Effects of Charge Stocks. Ind.
Eng. Chem. Process. Des. Dev. 1971, 10, 530–538. [CrossRef]
17. Jacob, S.M.; Gross, B.; Voltz, S.E.; Weekman, V.W. A lumping and reaction scheme for catalytic cracking. AIChE J. 1976, 22, 701–713.
[CrossRef]
18. Gross, B.; Jacob, S.M.; Nace, D.M.; Voltz, S.E. Simulation of Catalytic Cracking Process. U.S. Patent No. US3960707A, 1 June 1976.
19. Arbel, A.; Rinard, I.H.; Shinnar, R.; Sapre, A.V. Dynamics and Control of Fluidized Catalytic Crackers. 2. Multiple Steady States
and Instabilities. Ind. Eng. Chem. Res. 1995, 34, 3014–3026. [CrossRef]
20. Elshishini, S.; Elnashaie, S. Digital simulation of industrial fluid catalytic cracking units: Bifurcation and its implications. Chem.
Eng. Sci. 1990, 45, 553–559. [CrossRef]
21. Quann, R.J.; Jaffe, S.B. Structure-oriented lumping: Describing the chemistry of complex hydrocarbon mixtures. Ind. Eng. Chem.
Res. 1992, 31, 2483–2497. [CrossRef]
22. Vynckier, E.; Froment, G.F. Modeling of the kinetics of complex processes based upon elementary steps. In Kinetic and Thermody-
namic Lumping of Multicomponent Mixtures; Elsevier: Amsterdam, The Netherlands, 1991; pp. 131–161.
23. Froment, G. Kinetic modeling of acid-catalyzed oil refining processes. Catal. Today 1999, 52, 153–163. [CrossRef]
24. Dave, D.; Zhang, N. Multiobjective optimisation of fluid catalytic cracker unit using genetic algorithms. Comput. Aided Chem. Eng.
2003, 14, 623–628. [CrossRef]
25. Mora, R.L.; Pascual, M.P.; Miranda, M. Maximixing Propylene Production in a FCC Unit. In Proceedings of the 19th World
Petroleum Congress, Madrid, Spain, 29 June–3 July 2008.
26. Akah, A.; Al-Ghrami, M. Maximizing propylene production via FCC technology. Appl. Petrochem. Res. 2015, 5, 377–392. [CrossRef]
27. John, Y.M.; Patel, R.; Mujtaba, I.M. Maximization of propylene in an industrial FCC unit. Appl. Petrochem. Res. 2018, 8, 79–95.
[CrossRef]
28. Naik, D.V.; Karthik, V.; Kumar, V.; Prasad, B.; Garg, M.O. Kinetic modeling for catalytic cracking of pyrolysis oils with VGO in a
FCC unit. Chem. Eng. Sci. 2017, 170, 790–798. [CrossRef]
Energies 2022, 15, 308 44 of 50

29. Naik, D.V.; Kumar, V.; Prasad, B.; Behera, B.; Atheya, N.; Singh, K.K.; Adhikari, D.K.; Garg, M.O. Catalytic cracking of pyrolysis
oil oxygenates (aliphatic and aromatic) with vacuum gas oil and their characterization. Chem. Eng. Res. Des. 2014, 92, 1579–1590.
[CrossRef]
30. Zhang, R.; Li, L.; Liu, Z.; Meng, X. Nine-Lump Kinetic Study of Catalytic Pyrolysis of Gas Oils Derived from Canadian Synthetic
Crude Oil. Int. J. Chem. Eng. 2016, 2016, 9148925. [CrossRef]
31. Grace, J.R. Fluidized Bed Reactor Modeling. In Chemical Reactors; ACS: Washington, DC, USA, 1981; Volume 22, pp. 3–18.
[CrossRef]
32. Berruti, F.; Pugsley, T.S.; Godfroy, L.; Chaouki, J.; Patience, G.S. Hydrodynamics of circulating fluidized bed risers: A review. Can.
J. Chem. Eng. 1995, 73, 579–602. [CrossRef]
33. Grace, J.R.; Avidan, A.A.; Knowlton, T.M. Circulating Fluidized Beds; Blackie Academic & Professional: Glasgow, UK, 1997.
34. Harris, B.J. Modeling options for circulating fluidized beds: A core/annulus deposition model. In Circulating Fluidized Bed
Technology IV; Elsevier: Amsterdam, The Netherlands, 1993; pp. 32–39.
35. Godfroy, L.; Patience, G.S.; Chaouki, J. Radial Hydrodynamics in Risers. Ind. Eng. Chem. Res 1999, 38, 81–89. [CrossRef]
36. Gupta, R.; Kumar, V.; Srivastava, V.K. Modeling and simulation of fluid catalytic cracking unit. Rev. Chem. Eng. 2005, 21, 95–131.
[CrossRef]
37. Vashisth, S.; Motlagh, A.H.A.; Tebianian, S.; Salcudean, M.; Grace, J.R. Comparison of numerical approaches to model FCC
particles in gas–solid bubbling fluidized bed. Chem. Eng. Sci. 2015, 134, 269–286. [CrossRef]
38. Rao, R.M.; Rengaswamy, R.; Suresh, A.; Balaraman, K. Industrial Experience with Object-Oriented Modelling: FCC case study.
Chem. Eng. Res. Des. 2004, 82, 527–552. [CrossRef]
39. Sadeghbeigi, R. Fluid Catalytic Cracking Handbook; Elsevier: Amsterdam, The Netherlands, 2012. [CrossRef]
40. Behera, B.; Ray, S.S.; Singh, I. Structural characterization of FCC feeds from Indian refineries by NMR spectroscopy. Fuel 2008, 87,
2322–2333. [CrossRef]
41. Tissot, B.P.; Welte, D.H. Kerogen: Composition and Classification. In Petroleum Formation and Occurrence; Springer:
Berlin/Heidelberg, Germany, 1984; pp. 131–159. [CrossRef]
42. Rodríguez, E.; Izaddoust, S.; Valecillos, J.; Bilbao, J.; Arandes, J.M.; Castaño, P.; Epelde, E.; Elordi, G. Lessening coke formation
and boosting gasoline yield by incorporating scrap tire pyrolysis oil in the cracking conditions of an FCC unit. Energy Convers.
Manag. 2020, 224, 113327. [CrossRef]
43. Stratiev, D.S.; Marinov, I.; Nedelchev, A.; Velkov, I.; Stratiev, D.D.; Veli, A.; Mitkova, M.; Stanulov, K. Evaluation of approaches for
conversion of ASTM into TBP distillation data of oil fractions. Oil Gas Eur. Mag. 2014, 4, 216–221.
44. Behrenbruch, P.; Dedigama, T. Classification and characterisation of crude oils based on distillation properties. J. Pet. Sci. Eng.
2007, 57, 166–180. [CrossRef]
45. Gupta, R.K.; Kumar, V.; Srivastava, V.K. A new generic approach for the modeling of fluid catalytic cracking (FCC) riser reactor.
Chem. Eng. Sci. 2007, 62, 4510–4528. [CrossRef]
46. Dwyer, J.; Rawlence, D.J. Fluid catalytic cracking: Chemistry. Catal. Today 1993, 18, 487–507. [CrossRef]
47. Corma, A.; Orchillés, A. Current views on the mechanism of catalytic cracking. Microporous Mesoporous Mater. 2000, 35–36, 21–30.
[CrossRef]
48. Whitmore, F.C.; Schmerling, L.; Ipatieff, V.N. Advances in Catalysis; Academic Press: Cambridge, MA, USA, 1934; Volume 26.
49. Hiraoka, K.; Kebarle, P. Stabilities and energetics of pentacoordinated carbonium ions. The isomeric protonated ethane ions and
some higher analogs: Protonated propane and protonated butane. J. Am. Chem. Soc. 1976, 98, 6119–6125. [CrossRef]
50. Pansing, W.F. The Catalytic Cracking of Hexadecane—Effects of Impurities, Olefins, and Steam. J. Phys. Chem. 1965, 69, 392–399.
[CrossRef]
51. Greensfelder, B.S.; Voge, H.H.; Good, G.M. Catalytic Cracking of Pure Hydrocarbons. Ind. Eng. Chem. 1945, 37, 1168–1176.
[CrossRef]
52. Kotrel, S.; Knözinger, H.; Gates, B. The Haag–Dessau mechanism of protolytic cracking of alkanes. Microporous Mesoporous Mater.
2000, 35-36, 11–20. [CrossRef]
53. Haag, W.O.; Dessau, R.M. Duality of mechanism for acid-catalyzed paraffin cracking. In Proceedings of the 8th International
Congress on Catalysis, Berlin, Germany, 2–6 July 1984; Volume 2, pp. 305–316.
54. Haag, W.O.; Dessau, R.M.; Lago, R.M. Chemistry of Microporous Crystals. Stud. Surf. Sci. Catal. 1991, 60, 255–265.
55. Narbeshuber, T.F.; Vinek, H.; Lercher, J.A. Monomolecular Conversion of Light Alkanes over H-ZSM- 5. J. Catal. 1995, 157,
388–395. [CrossRef]
56. Lercher, J.A.; van Santen, R.R.; Vinek, H. Carbonium ion formation in zeolite catalysis. Catal. Lett. 1994, 27, 91–96. [CrossRef]
57. Kwak, B.; Sachtler, W. Effect of Ga/Proton Balance in Ga/HZSM-5 Catalysts on C3 Conversion to Aromatics. J. Catal. 1994, 145,
456–463. [CrossRef]
58. Krannila, H.; Haag, W.O.; Gates, B.C. Monomolecular and bimolecular mechanisms of paraffin cracking: N-butane cracking
catalyzed by HZSM-5. J. Catal. 1992, 135, 115–124. [CrossRef]
59. Tung, S.; Tung, S.E.; McIninch, E. Zeolitic aluminosilicate II. Surface oxide diffusion, dynamic (time variant) Lewis acids and
catalytic activity with hexane on decationized Y. J. Catal. 1968, 10, 175–182. [CrossRef]
60. Farneth, W.E.; Gorte, R.J. Methods for Characterizing Zeolite Acidity. Chem. Rev. 1995, 95, 615–635. [CrossRef]
Energies 2022, 15, 308 45 of 50

61. Abbot, J.; Guerzoni, F.N. Roles of Brønsted and Lewis sites during cracking of n-octane on H-mordenite. Appl. Catal. A Gen. 1992,
85, 173–188. [CrossRef]
62. Kissin, Y.V. Primary Products in Hydrocarbon Cracking over Solid Acidic Catalysts under Very Mild Conditions: Relation to
Cracking Mechanism. J. Catal. 1998, 180, 101–105. [CrossRef]
63. Kissin, Y.V. Chemical Mechanism of Hydrocarbon Cracking over Solid Acidic Catalysts. J. Catal. 1996, 163, 50–62. [CrossRef]
64. Kazansky, V.B.; Senchenya, I.N. Quantum chemical study of the electronic structure and geometry of surface alkoxy groups
as probable active intermediates of heterogeneous acidic catalysts: What are the adsorbed carbenium ions? J. Catal. 1989, 119,
108–120. [CrossRef]
65. Frash, M.V.; Solkan, V.N.; Kazansky, V.B. A quantum-chemical study of hydride transfer from alkanes to carbenium cations in the
gas phase: A comparison with liquid- and solid-acid catalytic systems. J. Chem. Soc. Faraday Trans. 1997, 93, 515–520. [CrossRef]
66. Caeiro, G.; Carvalho, R.; Wang, X.; Lemos, M.A.; Lemos, F.; Guisnet, M.; Ribeiro, F.R. Activation of C2–C4 alkanes over acid and
bifunctional zeolite catalysts. J. Mol. Catal. A Chem. 2006, 255, 131–158. [CrossRef]
67. Bartholomew, C.H.; Farrauto, R.J. Fundamentals of Industrial Catalytic Processes; John Wiley & Sons: Hoboken, NJ, USA, 2011.
68. Yaluris, G.; Rekoske, J.E.; Aparicio, L.M.; Madon, R.J.; Dumesic, J.A. Isobutane Cracking over Y-Zeolites: I. Development of a
Kinetic-Model. J. Catal. 1995, 153, 54–64. [CrossRef]
69. Yaluris, G.; Rekoske, J.E.; Aparicio, L.M.; Madon, R.J.; Dumesic, J.A. Isobutane Cracking over Y-Zeolites: II. Catalytic Cycles and
Reaction Selectivity. J. Catal. 1995, 153, 65–75. [CrossRef]
70. Yaluris, G.; Madon, R.J.; Dumesic, J.A. 2-Methylhexane Cracking on Y Zeolites: Catalytic Cycles and Reaction Selectivity. J. Catal.
1997, 165, 205–220. [CrossRef]
71. Cortright, R.D.; Dumesic, J.A.; Madon, R.J. Catalytic cycles for hydrocarbon cracking on zeolites. Top. Catal. 1997, 4, 15–26.
[CrossRef]
72. De Oliveira, L.P.; Hudebine, D.; Guillaume, D.; Verstraete, J.J. A Review of Kinetic Modeling Methodologies for Complex
Processes. Oil Gas Sci. Technol. 2016, 71, 45. [CrossRef]
73. Ho, T.C. Modeling of Reaction Kinetics for Petroleum Fractions. In Practical Advances in Petroleum Processing; Springer: New York,
NY, USA, 2006; pp. 653–694. [CrossRef]
74. Weekman, V.W.; Nace, D.M. Kinetics of catalytic cracking selectivity in fixed, moving, and fluid bed reactors. AIChE J. 1970, 16,
397–404. [CrossRef]
75. Lee, L.-S.; Chen, Y.-W.; Huang, T.-N.; Pan, W.-Y. Four-lump kinetic model for fluid catalytic cracking process. Can. J. Chem. Eng.
1989, 67, 615–619. [CrossRef]
76. Yen, L.C.; Wrench, R.E.; Ong, A.S. Reaction kinetic correlation equation predicts fluid catalytic cracking coke yields. Oil Gas J.
1988, 86, 2.
77. Corella, J.; Francés, E. Analysis of the Riser Reactor of a Fluid Cracking Unit. ACS Symp. Ser. 1991, 452, 165–182. [CrossRef]
78. Larocca, M.; Ng, S.; de Lasa, H. Fast catalytic cracking of heavy gas oils: Modeling coke deactivation. Ind. Eng. Chem. Res. 1990,
29, 171–180. [CrossRef]
79. Ancheyta-Juárez, J.; López-Isunza, F.; Aguilar-Rodríguez, E. 5-Lump kinetic model for gas oil catalytic cracking. Appl. Catal. A
Gen. 1999, 177, 227–235. [CrossRef]
80. Ancheyta-Juárez, J.; López-Isunza, F.; Aguilar-Rodríguez, E.; Moreno-Mayorga, J.C. A Strategy for Kinetic Parameter Estimation
in the Fluid Catalytic Cracking Process. Ind. Eng. Chem. Res. 1997, 36, 5170–5174. [CrossRef]
81. Ancheyta-Juárez, J.; Murillo-Hernández, J.A. A Simple Method for Estimating Gasoline, Gas, and Coke Yields in FCC Processes.
Energy Fuels 2000, 14, 373–379. [CrossRef]
82. 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 reactorI. Fuel 2003, 82, 1559–1569. [CrossRef]
83. Bollas, G.M.; Lappas, A.A.; Iatridis, D.K.; Vasalos, I.A. Five-lump kinetic model with selective catalyst deactivation for the
prediction of the product selectivity in the fluid catalytic cracking process. Catal. Today 2007, 127, 31–43. [CrossRef]
84. Ancheyta-Juárez, J.; Sotelo-Boyás, R. Estimation of Kinetic Constants of a Five-Lump Model for Fluid Catalytic Cracking Process
Using Simpler Sub-models. Energy Fuels 2000, 14, 1226–1231. [CrossRef]
85. John, Y.M.; Mustafa, M.A.; Patel, R.; Mujtaba, I.M. Parameter estimation of a six-lump kinetic model of an industrial fluid catalytic
cracking unit. Fuel 2019, 235, 1436–1454. [CrossRef]
86. Xiong, K.; Lu, C.; Wang, Z.; Gao, X. Kinetic study of catalytic cracking of heavy oil over an in-situ crystallized FCC catalyst. Fuel
2015, 142, 65–72. [CrossRef]
87. Souza, J.A.; Vargas, J.V.C.; von Meien, O.F.; Martignoni, W.P.; Ordonez, J.C. The inverse methodology of parameter estimation
for model adjustment, design, simulation, control and optimization of fluid catalytic cracking (FCC) risers. J. Chem. Technol.
Biotechnol. 2009, 84, 343–355. [CrossRef]
88. Heydari, M.; Ebrahim, H.A.; Dabir, B. Modeling of an Industrial Riser in the Fluid Catalytic Cracking Unit. Am. J. Appl. Sci. 2010,
7, 221–226. [CrossRef]
89. Wang, X.S.; Weng, H.X.; Mao, X.J.; Liu, F.Y.; Sha, Y.X.; Den, X.L. Investigation of the lumped kinetic model for catalytic cracking
IV determination of kinetic parameters in heavy gas oil network. Acta Sin. 1988, 4, 11–17.
90. Gao, H.; Wang, G.; Xu, C.; Gao, J. Eight-Lump Kinetic Modeling of Vacuum Residue Catalytic Cracking in an Independent Fluid
Bed Reactor. Energy Fuels 2014, 28, 6554–6562. [CrossRef]
Energies 2022, 15, 308 46 of 50

91. Mao, X.; Weng, H.; Zhu, Z.; Wang, S.; Zhu, K. Investigation of the lumped kinetic model for catalytic cracking: III. Analyzing
light oil feed and products and measurement of kinetic constants. Acta. Pet. Sin. 1985, 1, 11–18.
92. Oliveira, L.L.; Biscaia, E. Catalytic cracking kinetic models. Parameter estimation and model evaluation. Ind. Eng. Chem. Res.
1989, 28, 264–271. [CrossRef]
93. Sun, T.D.; Zhong, X.; Chen, Y.; Yu, B.T. Study and application of a lumping FCC kinetics model for vacuum residue. Petrol. Process.
Petrochem. 2001, 32, 41–44.
94. Watson, K.M.; Nelson, E.F. Improved Methods for Approximating Critical and Thermal Properties of Petroleum Fractions. Ind.
Eng. Chem. 1933, 25, 880–887. [CrossRef]
95. Edmister, W.C.; Lee, B.K. Applied Hydrocarbon Thermodynamics, 2nd ed.; Gulf Professional Publishing: Houston, TX, USA, 1984;
Volume 1.
96. Aris, R.; Gavalas, G.R. On the theory of reactions in continuous mixtures. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Sci. 1966,
260, 351–393.
97. Laxminarasimhan, C.S.; Verma, R.P.; Ramachandran, P.A. Continuous lumping model for simulation of hydrocracking. AIChE J.
1996, 42, 2645–2653. [CrossRef]
98. Amhammed, M.M. Applications of Lumping Kinetics Methodology to Complex Reactive Mixtures. Ph.D. Thesis, Heriot-Watt
University, Edinburgh, UK, 2013.
99. Chou, M.Y.; Ho, T.C. Continuum theory for lumping nonlinear reactions. AIChE J. 1988, 34, 1519–1527. [CrossRef]
100. Lababidi, H.M.S.; Alhumaidan, F.S. Modeling the Hydrocracking Kinetics of Atmospheric Residue in Hydrotreating Processes by
the Continuous Lumping Approach. Energy Fuels 2011, 25, 1939–1949. [CrossRef]
101. McCoy, B.J.; Wang, M. Continuous-mixture fragmentation kinetics: Particle size reduction and molecular cracking. Chem. Eng.
Sci. 1994, 49, 3773–3785. [CrossRef]
102. Cicarelli, P.; Astarita, G.; Gallifuoco, A. Continuous kinetic lumping of catalytic cracking processes. AIChE J. 1992, 38, 1038–1044.
[CrossRef]
103. Aris, R.; Astarita, G. Continuous lumping of nonlinear chemical kinetics. Chem. Eng. Process. Process. Intensif. 1989, 26, 63–69.
[CrossRef]
104. Astarita, G.; Ocone, R. Lumping nonlinear kinetics. AIChE J. 1988, 34, 1299–1309. [CrossRef]
105. Peixoto, F.C.; de Medeiros, J.L. Reactions in multiindexed continuous mixtures: Catalytic cracking of petroleum fractions. AIChE
J. 2001, 47, 935–947. [CrossRef]
106. Peixoto, F.C.; Platt, G.M.; Pessoa, F.L.P. Vapor–liquid equilibria of multi-indexed continuous mixtures using an equation of state
and group contribution methods. Chem. Eng. J. 2000, 77, 179–187. [CrossRef]
107. Ho, T.C.; White, B.S. On the continuum approximation of large reaction mixtures. AIChE J. 2010, 56, 1894–1906. [CrossRef]
108. Aris, R. Prolegomena to the rational analysis of systems of chemical reactions II. Some addenda. Arch. Ration. Mech. Anal. 1968,
27, 356–364. [CrossRef]
109. Krambeck, F.J. Computers and modern analysis in reactor design. Inst. Chem. Eng. Symp. Ser. 1984, 87, 733.
110. Chou, M.Y.; Ho, T.C. Lumping coupled nonlinear reactions in continuous mixtures. AIChE J. 1989, 35, 533–538. [CrossRef]
111. McCoy, B.J. Continuous kinetics of cracking reactions: Thermolysis and pyrolysis. Chem. Eng. Sci. 1996, 51, 2903–2908. [CrossRef]
112. Froment, G.F. Single Event Kinetic Modeling of Complex Catalytic Processes. Catal. Rev. 2005, 47, 83–124. [CrossRef]
113. Dewachtere, N.V.; Santaella, F.; Froment, G.F. Application of a single-event kinetic model in the simulation of an industrial riser
reactor for the catalytic cracking of vacuum gas oil. Chem. Eng. Sci. 1999, 54, 3653–3660. [CrossRef]
114. Moustafa, T.M.; Froment, G.F. Kinetic Modeling of Coke Formation and Deactivation in the Catalytic Cracking of Vacuum Gas
Oil. Ind. Eng. Chem. Res. 2003, 42, 14–25. [CrossRef]
115. Feng, W.; Vynckier, E.; Froment, G.F. Single event kinetics of catalytic cracking. Ind. Eng. Chem. Res. 1993, 32, 2997–3005.
[CrossRef]
116. Surla, K.; Guillaume, D.; Verstraete, J.J.; Galtier, P. Kinetic Modeling using the Single-Event Methodology: Application to the
Isomerization of Light Paraffins. Oil Gas Sci. Technol. 2011, 66, 343–365. [CrossRef]
117. D’Hooge, D.R.; Reyniers, M.-F.; Marin, G.B. Methodology for Kinetic Modeling of Atom Transfer Radical Polymerization.
Macromol. React. Eng. 2009, 3, 185–209. [CrossRef]
118. De Rybel, N.; van Steenberge, P.H.M.; Reyniers, M.-F.; Barner-Kowollik, C.; D’Hooge, D.R.; Marin, G.B. An Update on the Pivotal
Role of Kinetic Modeling for the Mechanistic Understanding and Design of Bulk and Solution RAFT Polymerization. Macromol.
Theory Simul. 2017, 26, 1600048. [CrossRef]
119. Clymans, P.J.; Froment, G.F. Computer-generation of reaction paths and rate equations in the thermal cracking of normal and
branched paraffins. Comput. Chem. Eng. 1984, 8, 137–142. [CrossRef]
120. Prickett, S.E.; Mavrovouniotis, M.L. Construction of complex reaction systems—I. Reaction description language. Comput. Chem.
Eng. 1997, 21, 1219–1235. [CrossRef]
121. Joshi, P.V. Molecular and Mechanistic Modeling of Complex Process Chemistries: A Generic Approach to Automated Model
Building. Ph.D. Thesis, University of Delaware, Newark, DE, USA, 1998.
122. Christensen, G.; Apelian, M.R.; Hickey, K.J.; Jaffe, S.B. Future directions in modeling the FCC process: An emphasis on product
quality. Chem. Eng. Sci. 1999, 54, 2753–2764. [CrossRef]
Energies 2022, 15, 308 47 of 50

123. Solorzano, R.Q.; Thybaut, J.W.; Marin, G.B. A single-event microkinetic analysis of the catalytic cracking of (cyclo)alkanes on an
equilibrium catalyst in the absence of coke formation. Chem. Eng. Sci. 2007, 62, 5033–5038. [CrossRef]
124. Solorzano, R.Q.; Thybaut, J.W.; Galtier, P.; Marin, G.B. Simulation of an industrial riser for catalytic cracking in the presence of
coking using Single-Event MicroKinetics. Catal. Today 2010, 150, 319–331. [CrossRef]
125. Xue, G.; Huixin, W.; Thybaut, J.; Marin, G. Catalytic cracking of cycloparaffins admixed with olefins: 2. Single-event microkinetic
(semk) assessment. China Pet. Process. Petrochem. Technol. 2014, 16, 84–90.
126. Standl, S.; Tonigold, M.; Hinrichsen, O. Single-Event Kinetic Modeling of Olefin Cracking on ZSM-5: Proof of Feed Independence.
Ind. Eng. Chem. Res. 2017, 56, 13096–13108. [CrossRef]
127. von Aretin, T.; Standl, S.; Tonigold, M.; Hinrichsen, O. Optimization of the product spectrum for 1-pentene cracking on ZSM-5
using single-event methodology. Part 2: Recycle reactor. Chem. Eng. J. 2017, 309, 873–885. [CrossRef]
128. Standl, S.; Kirchberger, F.M.; Kühlewind, T.; Tonigold, M.; Sanchez-Sanchez, M.; Lercher, J.A.; Hinrichsen, O. Single-event kinetic
model for methanol-to-olefins (MTO) over ZSM-5: Fundamental kinetics for the olefin co-feed reactivity. Chem. Eng. J. 2020, 402,
126023. [CrossRef]
129. Beirnaert, H.C.; Alleman, J.R.; Marin, G.B. A fundamental kinetic model for the catalytic cracking of alkanes on a USY zeolite in
the presence of coke formation. Ind. Eng. Chem. Res. 2001, 40, 1337–1347. [CrossRef]
130. Quintana-Solórzano, R.; Thybaut, J.W.; Marin, G.B.; Lødeng, R.; Holmen, A. Single-Event MicroKinetics for coke formation in
catalytic cracking. Catal. Today 2005, 107–108, 619–629. [CrossRef]
131. Cerqueira, H.S.; Caeiro, G.; Costa, L.; Ribeiro, F.R. Deactivation of FCC catalysts. J. Mol. Catal. A Chem. 2008, 292, 1–13. [CrossRef]
132. Guisnet, M.; Magnoux, P. Deactivation by coking of zeolite catalysts. Prevention of deactivation. Optimal conditions for
regeneration. Catal. Today 1997, 36, 477–483. [CrossRef]
133. Guisnet, M.; Magnoux, P. Organic chemistry of coke formation. Appl. Catal. A Gen. 2001, 212, 83–96. [CrossRef]
134. Reyes, S.C.; Scriven, L.E. Analysis of zeolite catalyst deactivation during catalytic cracking reactions. Ind. Eng. Chem. Res. 1991,
30, 71–82. [CrossRef]
135. Froment, G.F. A Quantitative Approach of Catalyst Deactivation by Coke Formation; Elsevier: Amsterdam, The Netherlands, 1980; pp.
1–19. [CrossRef]
136. Forzatti, P.; Lietti, L. Catalyst deactivation. Catal. Today 1999, 52, 165–181. [CrossRef]
137. Froment, G.F. The Modeling of Catalyst Deactivation by Coke Formation; Elsevier: Amsterdam, The Netherlands, 1991; pp. 53–83.
[CrossRef]
138. Rodríguez, E.; Elordi, G.; Valecillos, J.; Izaddoust, S.; Bilbao, J.; Arandes, J.M.; Castaño, P. Coke deposition and product distribution
in the co-cracking of waste polyolefin derived streams and vacuum gas oil under FCC unit conditions. Fuel Process. Technol. 2019,
192, 130–139. [CrossRef]
139. Froment, G. Modeling of catalyst deactivation. Appl. Catal. A Gen. 2001, 212, 117–128. [CrossRef]
140. Corella, J. On the Modeling of the Kinetics of the Selective Deactivation of Catalysts. Application to the Fluidized Catalytic
Cracking Process. Ind. Eng. Chem. Res. 2004, 43, 4080–4086. [CrossRef]
141. Farag, H.; Ng, S.; de Lasa, H. Kinetic modeling of catalytic cracking of gas oils using in situ traps (FCCT) to prevent metal
contaminant effects. Ind. Eng. Chem. Res. 1993, 32, 1071–1080. [CrossRef]
142. Krambeck, F.J. Continuous mixtures in fluid catalytic cracking and extensions. In Mobil Workshop on Chemical Reaction in Complex
Mixtures; van Nostrand Reinhold: New York, NY, USA, 1991; pp. 42–59.
143. Pitault, I.; Forissier, M.; Bernard, J.R. Determination of Kinetic Constants of Catalytic Cracking in Microactivity Test Modeling.
Can. J. Chem. Eng. 1995, 73, 498–504. [CrossRef]
144. Fernandes, J.L.; Domingues, L.H.; Pinheiro, C.I.C.; Oliveira, N.M.C.; Ribeiro, F.R. Influence of different catalyst deactivation
models in a validated simulator of an industrial UOP FCC unit with high-efficiency regenerator. Fuel 2012, 97, 97–108. [CrossRef]
145. Davidson, J.F.; Clift, R.; Harrison, D. Fluidization, 2nd ed.; Academic Press: Cambridge, MA, USA, 1985.
146. Bi, H.T.; Ellis, N.; Abba, I.A.; Grace, J.R. A state-of-the-art review of gas–solid turbulent fluidization. Chem. Eng. Sci. 2000, 55,
4789–4825. [CrossRef]
147. Lim, K.S.; Zhu, J.X.; Grace, J.R. Hydrodynamics of gas-solid fluidization. Int. J. Multiph. Flow 1995, 21, 141–193. [CrossRef]
148. Rhodes, M.J.; Geldart, D. The hydrodynamics of re-circulating fluidized beds. In Circulating Fluidized Bed Technology, Proceedings of
the First International Conference on Circulating Fluidized Beds, Halifax, NS, Canada, 18–20 November 1985; Elsevier: Amsterdam, The
Netherlands, 1985; pp. 193–200. [CrossRef]
149. Louge, M.; Chang, H. Pressure and voidage gradients in vertical gas-solid risers. Powder Technol. 1990, 60, 197–201. [CrossRef]
150. Li, I.; Kwauk, M. Two-Phase Flow: The Energy-Minimization Multi-Scale Method; Metallurgy Industry Press: Beijing, China, 1994.
151. Wei, F.; Lin, H.; Cheng, Y.; Wang, Z.; Jin, Y. Profiles of particle velocity and solids fraction in a high-density riser. Powder Technol.
1998, 100, 183–189. [CrossRef]
152. Pärssinen, J.H.; Zhu, J.-X. Particle velocity and flow development in a long and high-flux circulating fluidized bed riser. Chem.
Eng. Sci. 2001, 56, 5295–5303. [CrossRef]
153. Derouin, C.; Nevicato, D.; Forissier, M.; Wild, G.; Bernard, J.-R. Hydrodynamics of Riser Units and Their Impact on FCC Operation.
Ind. Eng. Chem. Res. 1997, 36, 4504–4515. [CrossRef]
154. Miller, A.; Gidaspow, D. Dense, vertical gas-solid flow in a pipe. AIChE J. 1992, 38, 1801–1815. [CrossRef]
Energies 2022, 15, 308 48 of 50

155. Martin, M.P.; Turlier, P.; Bernard, J.R.; Wild, G. Gas and solid behavior in cracking circulating fluidized beds. Powder Technol. 1992,
70, 249–258. [CrossRef]
156. Martin, M.P.; Derouin, C.; Turlier, P.; Forissier, M.; Wild, G.; Bernard, J.R. Catalytic cracking in riser reactors: Core-annulus and
elbow effects. Chem. Eng. Sci. 1992, 47, 2319–2324. [CrossRef]
157. Senior, R.C.; Brereton, C. Modelling of circulating fluidised-bed solids flow and distribution. Chem. Eng. Sci. 1992, 47, 281–296.
[CrossRef]
158. Wong, R.; Pugsley, T.; Berrut, F. Modelling the axial voidage profile and flow structure in risers of circulating fluidized beds.
Chem. Eng. Sci. 1992, 47, 2301–2306. [CrossRef]
159. Noymer, P.D.; Glicksman, L.R. Descent velocities of particle clusters at the wall of a circulating fluidized bed. Chem. Eng. Sci.
2000, 55, 5283–5289. [CrossRef]
160. Pandey, P.; Turton, R.; Yue, P.; Shadle, L. Nonintrusive Particle Motion Studies in the Near-Wall Region of a Pilot-Scale Circulating
Fluidized Bed. Ind. Eng. Chem. Res. 2004, 43, 5582–5592. [CrossRef]
161. Cankurt, N.T.; Yerushalmi, J. Gas Backmixing in High Velocity Fluidized Beds. In Fluidization; Cambridge University Press:
Cambridge, MA, USA, 1978; p. 387.
162. Cocco, R.; Shaffer, F.; Hays, R.; Karri, S.B.R.; Knowlton, T. Particle clusters in and above fluidized beds. Powder Technol. 2010, 203,
3–11. [CrossRef]
163. McMillan, J.; Shaffer, F.; Gopalan, B.; Chew, J.W.; Hrenya, C.; Hays, R.; Karri, S.R.; Cocco, R. Particle cluster dynamics during
fluidization. Chem. Eng. Sci. 2013, 100, 39–51. [CrossRef]
164. Horio, M.; Kuroki, H. Three-dimensional flow visualization of dilutely dispersed solids in bubbling and circulating fluidized
beds. Chem. Eng. Sci. 1994, 49, 2413–2421. [CrossRef]
165. Horio, M.; Nonaka, A. A generalized bubble diameter correlation for gas-solid fluidized beds. AIChE J. 1987, 33, 1865–1872.
[CrossRef]
166. Grace, J. High-velocity fluidized bed reactors. Chem. Eng. Sci. 1990, 45, 1953–1966. [CrossRef]
167. Bader, R.; Findlay, J.; Knowlton, T.M. Gas/solids flow patterns in a 30.5-cm-diameter circulating fluidized bed. In Circulating
Fluidized Bed Technology, Proceedings of the Second International Conference on Circulating Fluidized Beds, Compiégne, France, 14–18
March 1988; Elsevier: Amsterdam, The Netherlands, 1988; pp. 123–137. [CrossRef]
168. Van Deemter, J.J. Mixing. Fluidization; Academic Press: New York, NY, USA, 1985; pp. 331–355.
169. Stubington, J.F.; Chan, S.W. On the phase location and rate of volatiles combustion in bubbling fluidized bed combustors. Chem.
Eng. Res. Des. 1990, 68, 195–201.
170. Li, J.; Weinstein, H. An experimental comparison of gas backmixing in fluidized beds across the regime spectrum. Chem. Eng. Sci.
1989, 44, 1697–1705. [CrossRef]
171. Li, Y.; Wu, P. A study on axial gas mixing in a fast fluidized bed. In Circulating Fluidized Bed Technology III; Basu, P., Ed.; Pergamon
Press: Toronto, ON, Canada, 1991; pp. 581–586.
172. Zhu, J.X.; Bi, H.T. Development of CFB and High Density Circulating Fluidized Beds. In Circulating Fluidized Bed Technology IV;
Avidan, A.A., Ed.; AIChE: New York, NY, USA, 1993.
173. Schugerl, K. Experimental Comparison of Mixing Processes in Two and Three Phase Fluidized Beds. In Proceedings of the
International Symposium on Fluidization, Eindhoven, The Netherlands, 6–9 June 1967; pp. 782–794.
174. Dry, R.J.; White, C.C. Gas residence-time characteristics in a high-velocity circulating fluidised bed of FCC catalyst. Powder Technol.
1989, 58, 17–23. [CrossRef]
175. Van Zoonen, D. Measurements of diffusional phenomena and velocity profiles in a vertical riser. In Proc. Symp. on The Interaction
between Fluids and Particles, 1962; Instn, Chem. Engrs: Rugby, UK, 1962; pp. 64–71.
176. Brereton, C.M.H.; Grace, J.R.; Yu, J. Axial gas mixing in a circulating fluidized bed. In Circulating Fluidized Bed Technology,
Proceedings of the Second International Conference on Circulating Fluidized Beds, Compiégne, France, 14–18 March 1988; Elsevier:
Amsterdam, The Netherlands, 1988; pp. 307–314. [CrossRef]
177. Bai, D.R. Residence time distributions of gas and solids in a circulating fluidized bed. In Fluidization VII; Engineering Foundation:
New York, NY, USA, 1992; pp. 195–202.
178. Helmrich, H.; Schugerl, K.; Janssen, K. Decomposition of NaHCO3 in laboratory and bench scale circulating fluidized bed reactors.
In Circulating Fluidized Bed Technology, Proceedings of the First International Conference on Circulating Fluidized Beds, Halifax, NS,
Canada, 18–20 November 1985; Elsevier: Amsterdam, The Netherlands, 1986; pp. 161–166. [CrossRef]
179. Werther, J.; Hartge, E.U.; Kruse, M.; Potter, O.E.; Nicklin, D.J. Gas mixing and interphase mass transfer in the circulating fluidized
bed. In Fluidization VII; Engineering Foundation: New York, NY, USA, 1992; pp. 257–264.
180. Couturier, M.; Doucette, B.; Stevens, D.; Poolpol, S.; Razbin, V. Temperature, gas concentration and solid mass flux profiles
within a large circulating fluidized bed combustor. In Proceedings of the Eleventh International Conference on Fluidized Bed
Combustion, Montreal, QC, Canada, 21–24 April 1991; pp. 21–24.
181. Zheng, Q.; Xing, W.; Lou, F. Experimental study on radial gas dispersion and its enhancement in circulating fluidized beds. In
Fluidization VII; Engineering Foundation: New York, NY, USA, 1992; pp. 285–293.
182. Svensson, A. Fluid Dynamics of the Bottom Bed of Circulating Fluidized Bed Boilers. Ph.D. Thesis, Chalmers University of
Technology, Gothenburg, Sweden, 1995.
Energies 2022, 15, 308 49 of 50

183. Svensson, A.; Johnsson, F.; Leckner, B. Bottom bed regimes in a circulating fluidized bed boiler. Int. J. Multiph. Flow 1996, 22,
1187–1204. [CrossRef]
184. Wei, F.; Lin, S.; Yang, G. Gas and solids mixing in a commercial FCC regenerator. Chem. Eng. Technol. 1993, 16, 109–113. [CrossRef]
185. Lee, G.S.; Kim, S.D. Axial mixing of solids in turbulent fluidized beds. Chem. Eng. J. 1990, 44, 1–9. [CrossRef]
186. Edwards, M.; Avidan, A. Conversion model aids scale-up of mobil’s fluid-bed MTG process. Chem. Eng. Sci. 1986, 41, 829–835.
[CrossRef]
187. Foka, M.; Chaouki, J.; Guy, C.; Klvana, D. Natural gas combustion in a catalytic turbulent fluidized bed. Chem. Eng. Sci. 1994, 49,
4269–4276. [CrossRef]
188. Ahari, J.S.; Farshi, A.; Forsat, K. A mathematical modeling of the riser reactor in industrial FCC unit. Pet. Coal 2008, 50, 15–24.
189. Tsuo, Y.P.; Gidaspow, D. Computation of flow patterns in circulating fluidized beds. AIChE J. 1990, 36, 885–896. [CrossRef]
190. Gidaspow, D. Multiphase Flow and Fluidization: Continuum and Kinetic Theory Descriptions; Academic Press: Cambridge, MA, USA,
1994.
191. Drew, D.A. Mathematical Modeling of Two-Phase Flow. Annu. Rev. Fluid Mech. 1983, 15, 261–291. [CrossRef]
192. Lu, H.; Gidaspow, D.; Wang, S. Computational Fluid Dynamics and the Theory of Fluidization; Springer: Singapore, 2021.
193. Pugsley, T.S.; Berruti, F. A predictive hydrodynamic model for circulating fluidized bed risers. Powder Technol. 1996, 89, 57–69.
[CrossRef]
194. Sabbaghan, H.; Sotudeh-Gharebagh, R.; Mostoufi, N. Modeling the acceleration zone in the riser of circulating fluidized beds.
Powder Technol. 2004, 142, 129–135. [CrossRef]
195. Fernandes, J.L.; Pinheiro, C.I.C.; Oliveira, N.M.C.; Inverno, J.; Ribeiro, F.R. Model Development and Validation of an Industrial
UOP Fluid Catalytic Cracking Unit with a High-Efficiency Regenerator. Ind. Eng. Chem. Res. 2008, 47, 850–866. [CrossRef]
196. Xu, G.; Kato, K. Hydrodynamic equivalent diameter for clusters in heterogeneous gas–solid flow. Chem. Eng. Sci. 1999, 54,
1837–1847. [CrossRef]
197. Fernandes, J.L.; Pinheiro, C.I.C.; Oliveira, N.; Ribeiro, F.R. Modeling and simulation of an operating industrial fluidized catalytic
cracking (FCC) riser. In Proceedings of the 2nd Mercosur Congress on Chemical Engineering and 4th Mercosur Congress on
Process Systems Engineering, Rio de Janeiro, Brazil, 14–18 August 2005; pp. 1–4.
198. Nguyen, T.T.B.; Mitra, S.; Pareek, V.; Joshi, J.B.; Evans, G. Comparison of vaporization models for feed droplet in fluid catalytic
cracking risers. Chem. Eng. Res. Des. 2015, 101, 82–97. [CrossRef]
199. Ali, H.; Rohani, S.; Corriou, J.-P. Modelling and Control of a Riser Type Fluid Catalytic Cracking (FCC) Unit. Chem. Eng. Res. Des.
1997, 75, 401–412. [CrossRef]
200. Du, B.; Warsito, W.; Fan, L.-S. Flow dynamics of gas-solid fluidized beds with evaporative liquid injection. China Particuology
2006, 4, 1–8. [CrossRef]
201. Gehrke, S.; Wirth, K.E. Evaporating liquid in a high-density riser. In Proceedings of the International Congress on Particle
Technology, Nuremberg, Germany, 27–29 March 2007.
202. Newton, D.; Fiorentino, M.; Smith, G.B. The application of X-ray imaging to the developments of fluidized bed processes. Powder
Technol. 2001, 120, 70–75. [CrossRef]
203. Bruhns, S.; Werther, J. An investigation of the mechanism of liquid injection into fluidized beds. AIChE J. 2005, 51, 766–775.
[CrossRef]
204. Ahmadi Motlagh, A.H.; Pougatch, K.; Maturi, A.; Salcudean, M.; Grace, J.R.; Grecov, D.; McMillan, J. CFD Study of Wet
Agglomerate Growth and Breakage in a Fluidized Bed Containing Hot Silica Sand Particles with Evaporative Liquid Injection.
Ind. Eng. Chem. Res. 2019, 58, 4396–4411. [CrossRef]
205. Ahmadi Motlagh, A.H.; Grace, J.R.; Briens, C.; Berruti, F.; Farkhondehkavaki, M.; Hamidi, M. Experimental analysis of volatile
liquid injection into a fluidized bed. Particuology 2017, 34, 39–47. [CrossRef]
206. Voorhies, A. Carbon Formation in Catalytic Cracking. Ind. Eng. Chem. 1945, 37, 318–322. [CrossRef]
207. Fernandes, J.L.; Verstraete, J.J.; Pinheiro, C.I.C.; Oliveira, N.M.C.; Ribeiro, F.R. Dynamic modelling of an industrial R2R FCC unit.
Chem. Eng. Sci. 2007, 62, 1184–1198. [CrossRef]
208. Takatsuka, T.; Sato, S.; Morimoto, Y.; Hashimoto, H. A reaction model for fluidized-bed catalytic cracking of residual oil. Int.
Chem. Eng. 1987, 27, 107–116.
209. Souza, J.A.; Vargas, J.V.C.; Von Meien, O.F.; Martignoni, W.; Amico, S.C. A two-dimensional model for simulation, control, and
optimization of FCC risers. AIChE J. 2006, 52, 1895–1905. [CrossRef]
210. Souza, J.A.; Vargas, J.V.C.; Von Meien, O.F.; Martignoni, W.P. Modeling and simulation of industrial fcc risers. Therm. Eng. 2007, 6,
19–25. [CrossRef]
211. Petrobras Six. Riser Improvement Project; Internal Report; Springer: São Mateus do Sul, Brazil, 2001. (In Portuguese)
212. Patience, G.S.; Chaouki, J.; Berruti, F.; Wong, R. Scaling considerations for circulating fluidized bed risers. Powder Technol. 1992,
72, 31–37. [CrossRef]
213. Hernández-Barajas, J.R.; Vázquez-Román, R.; Félix-Flores, M.G. A comprehensive estimation of kinetic parameters in lumped
catalytic cracking reaction models. Fuel 2009, 88, 169–178. [CrossRef]
214. Gupta, P.L.; Gupta, R.C. The monotonicity of the reliability measures of the beta distribution. Appl. Math. Lett. 2000, 13, 5–9.
[CrossRef]
Energies 2022, 15, 308 50 of 50

215. Zhang, J.; Wang, Z.; Jiang, H.; Chu, J.; Zhou, J.; Shao, S. Modeling fluid catalytic cracking risers with special pseudo-components.
Chem. Eng. Sci. 2013, 102, 87–98. [CrossRef]
216. Arya, D.; Gupta, R.K. Simulation Studies on Variable Cluster Size in FCC Riser Reactor. Part. Sci. Technol. 2014, 32, 144–150.
[CrossRef]
217. Boum, A.; Corriou, J.P.; Latifi, A. Comparison of model predictive control strategies for a fluidized catalytic cracker. Int. J. Eng.
Technol. 2017, 6, 181–190. [CrossRef]
218. Lee, E.; Groves, F.R., Jr. Mathematical model of the fluidized bed catalytic cracking plant. Trans. Soc. Comput. Simul. Int. 1985, 2,
219–236.
219. Li, Y.; Chu, J.; Zhang, J. A new generic reaction for modeling fluid catalytic cracking risers. Chin. J. Chem. Eng. 2017, 25, 1449–1460.
[CrossRef]
220. Chen, Z.; Feng, S.; Zhang, L.; Wang, G.; Shi, Q.; Xu, Z.; Zhao, S.; Xu, C. Molecular-level kinetic modeling of heavy oil fluid
catalytic cracking process based on hybrid structural unit and bond-electron matrix. AIChE J. 2020, 67, e17027. [CrossRef]
221. He, S.; Lucio-Vega, J.; Zhang, L.; Shi, Q.; Horton, S.R.; Al-Khattaf, S.; Al-Jundi, I.; Elmutasim, O.; Zhao, S.; Klein, M.T. Integrating
Visualization Methods with Chemical Kinetics Model Solution and Editing Tools. Energy Fuels 2017, 31, 9881–9889. [CrossRef]
222. Chen, Z.; Feng, S.; Zhang, L.; Shi, Q.; Xu, Z.; Zhao, S.; Xu, C. Molecular-level kinetic modelling of fluid catalytic cracking slurry
oil hydrotreating. Chem. Eng. Sci. 2019, 195, 619–630. [CrossRef]
223. Zhu, C.; Jun, Y.; Patel, R.; Wang, D.; Ho, T.C. Interactions of flow and reaction in fluid catalytic cracking risers. AIChE J. 2010, 57,
3122–3131. [CrossRef]
224. Shah, Y.T.; Huling, G.P.; Paraskos, J.A.; McKinney, J.D. A Kinematic Model for an Adiabatic Transfer Line Catalytic Cracking
Reactor. Ind. Eng. Chem. Process. Des. Dev. 1977, 16, 89–94. [CrossRef]
225. Shi, J.; Wang, Y.; Yang, W.; Tang, Y.; Xie, Z. Recent advances of pore system construction in zeolite-catalyzed chemical industry
processes. Chem. Soc. Rev. 2015, 44, 8877–8903. [CrossRef] [PubMed]
226. García-Dopico, M.; García, A.; Santos Garcíac, A. Modelling fluidized catalytic cracking unit stripper efficiency. Chem. Ind. Chem.
Eng. Q. 2015, 21, 95–105. [CrossRef]
227. García-Dopico, M.; García, A.; Garcíac, A.S. Modelling coke formation and deactivation in a FCCU. Appl. Catal. A Gen. 2006, 303,
245–250. [CrossRef]

You might also like