You are on page 1of 21

Computers and Chemical Engineering 100 (2017) 198–218

Contents lists available at ScienceDirect

Computers and Chemical Engineering


journal homepage: www.elsevier.com/locate/compchemeng

Modeling of the maleic anhydride circulating fluidized bed reactor


Pranava Chaudhari, Sanjeev Garg ∗
Department of Chemical Engineering, Indian Institute of Technology, Kanpur 208016, India

a r t i c l e i n f o a b s t r a c t

Article history: Maleic Anhydride (MA) production by selective oxidation of n-butane in a Multi-Tubular Fixed Bed reac-
Received 20 May 2016 tor is constrained by the flammability limits. The use of Fluidized Bed circumvents this constraint but
Received in revised form 1 November 2016 suffers from high back mixing. The Circulating Fluidized Bed (CFB) reduces the back mixing of the cata-
Accepted 6 February 2017
lyst and enhances the catalyst activity by reactivating the catalyst. A simple reactor model with suitable
Available online 14 February 2017
hydrodynamic and kinetic models at the operating conditions of both the pilot and commercial MA CFB
reactors is proposed. A dispersion model with variable gas density is used for modeling the hydrodynam-
Keywords:
ics of these reactors. The reactors’ configurational complexities are also accounted for in the model. The
Circulating fluidized bed
Maleic anhydride
data fitting exercise of the proposed reactor model is performed. The proposed method simultaneously
Reactor modeling minimizes the total bias based on the individual biases for each component of the multi component
Variable density reactor reactor system and the total error in a multi-objective framework.
Data fitting exercise © 2017 Elsevier Ltd. All rights reserved.
Multi-objective optimization

1. Introduction compared to the MTFBR (∼1.8%) (Contractor, 1999). The lower con-
centration of Bu in the MTFBR is controlled by its flammability limit.
The various valuable chemicals and intermediates are produced In the open literature, three major problems associated with the
in the process industries by selective oxidation of the small alka- scale-up of a fluidized bed are reported (Contractor and Sleight,
nes in the presence of specific catalysts (Schlogl, 2009). The Maleic 1987). Firstly, the “gulf streaming” that creates a wide gas residence
Anhydride (MA) production by the selective oxidation of n-butane time distribution that may lead to reduced conversions; secondly,
(Bu) in the presence of Vanadium-Phosphorus-Oxide (VPO) catalyst the “significant solid and gas back-mixing” that may lead to the
is one of the most successful example (Ballarini et al., 2006) (Global loss in the selectivity of the desired product and thirdly the “lim-
market of $2.8 billion with projected 7.2% increase for 2014–2019; ited oxygen availability” because of high Bu fraction in the feed that
Source: report by research and market ID: 3292451, June 2015). may lead to the over reduction of the catalyst.
The Multi-Tubular Fixed Bed Reactor (MTFBR) is predominantly The use of a MA Circulating Fluidized Bed (CFB) reactor has been
used in the industrial processes (Contractor, 1999). The limited heat proposed (Contractor and Sleight, 1987) by DuPont that not only
removal in a MTFBR generally leads to hot spots within the reactor retains all the advantages of a fluidized bed reactor over a MTFBR
in case of exothermic reactions. A previous study (Chaudhari and but also overcomes the three problems associated with the flu-
Gupta, 2012) has shown that the optimal production of the MA in idized bed reactors mentioned above. In a CFB reactor, the extra
a MTFBR is controlled by constraints of the flammability limit and oxygen is fed separately through the spargers that results in less
the auto ignition temperature (AIT). over reduction of the catalyst. The separate oxygen feed in the MA
The fluidized bed reactors have much better heat transfer char- CFB reactor allows the reactor to operate with higher Bu concen-
acteristics and uniform temperature control that also reduces the tration (25%) (Contractor, 1999) as compared to the fixed bed and
chances of hot spots in the fluidized reactor. Moreover, in the flu- MTFBRs. The catalyst particles in a CFB reactor are cycled between
idized bed reactors the catalyst particles also act as flame arrestors the fuel enriched and the oxygen enriched environments. The lat-
by quenching the free radicals of the homogeneous combustion tice oxygen present on the VPO catalyst contributes to the reaction
reactions (Contractor et al., 1987). The fluidized bed reactors can, in the fuel enriched environment and is previously reported as
thus, be operated with much higher Bu (∼4%) in the feed as the Lattice Oxygen Contribution (LOC) of the catalyst (Patience
and Bockrath, 2010). The Langmuir-Hinshelwood reaction kinetics
reported in the reactor model of a MTFBR (Chaudhari and Gupta,
∗ Corresponding author. 2012) is, thus, not applicable for the operating conditions of a MA
E-mail addresses: pranavch@iitk.ac.in (P. Chaudhari), sgarg@iitk.ac.in, CFB reactor. Several other kinetic models are reported in the open
s12garg@gmail.com (S. Garg). literature for the CFB reactors (Gascon et al., 2006; Lorences et al.,

http://dx.doi.org/10.1016/j.compchemeng.2017.02.012
0098-1354/© 2017 Elsevier Ltd. All rights reserved.
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 199

takes place. The correlations for criteria of regime transition along


Nomenclature with the corresponding references are outlined in Table 1. The cor-
  relations given in Table 1 use U0 measured at the outlet i.e., U0OUT ,
ACS,i Cross-sectional area of the reactor section i m2 as reported in the open literature (Patience and Bockrath, 2009).
CT Total concentration
 in the gaseous phase Pugsley et al. (1992) have proposed a MA CFB reactor model
kmol/m3 without the data fitting exercise with the pilot or commercial MA
DT,i Diameter of the reactor section i (m) CFB reactor data. The hydrodynamic model reported with this reac-
dp Mean particle diameter (m)   tor model requires experimental pressure drop profiles. Pugsley
Fj Molar flow of the jth -component kmol/s
  and Berruti, (1996) have suggested a predictive core annular hydro-
fconv Conversion factor kg/ms2 bar dynamic model that did not require the experimental pressure drop
HT Total height of the reactor profiles. The reported core annular hydrodynamic model considers
HT,i Height of the reactor section i (m) a downward solid movement in the annular region near the reactor
NL  capacity of the lattice site in the cata-
Total oxygen wall. This downward solid movement disappears in the CFB reac-
lyst bulk mmolO2 /kg of cat tors operating in the DSU regime with high Gpz values (Grace et al.,
PT Total pressure in the gaseous phase (bar) 1999). Thus, the reported core annular hydrodynamic model is suit-
pj Partial pressure of the jth -component (bar) able for the combustion reactors that operate at lower Gpz values
Rj Reaction rate jth -component and is not suitable for the catalytic reactors that operate at much
  of the
higher Gpz values (Berruti et al., 1995). The absence of a downward
kmol/kg of cat − s

  solid movement in the annular region of the catalytic reactors limits
Rg Gas constant kJ/kmol − K
  the applicability of many other core annular hydrodynamic mod-
Rg Gas constant m3 bar/K − kmol els reported in the open literature (Bai et al., 1995; Gupta, 1998;
 
rk’ Rate of the kth catalytic reaction kmol/kgofcat − s Puchyr et al., 1997) for modelling the hydrodynamics of a MA CFB
T Temperature (K) reactor. A generic fluidized-bed reactor hydrodynamic model is also
reported using a dispersion model with dense and lean sections in
vpz,t Terminal
 settling velocity of the single particle
m/s terms of the solid hold up (Abba et al., 2003). The model is generic
Yj Mole fraction of the gaseous jth -component (–) as it is applicable to bubbling fluidized bed (BFB), turbulent flu-
idized bed (TFB) and fast fluidized bed (FF) regimes. The limitation
of this generic model is that most of the catalytic reactors operate in
Greek symbols/greek letters   the DSU regime and this generic model does not consider the DSU
g Viscosity of the gaseous mixture kg/ms
  regime. Thus, a reactor model is desirable with suitable hydrody-
g Density of the gaseous mixture kg/m3
 3
 namic and kinetic models valid at the operating conditions present
cat Density of the catalyst kg/m in a MA CFB reactor. Moreover, the desired reactor model should
ε Void fraction of the reactor be simple and accurate and the same is proposed in this work.
bed m3 ofvoid/m3 ofreactor The motivation of the proposed work is described in Section
OL Fraction of occupied lattice oxygen sites in the cat- 1. A brief description of a Pilot Plant (PP) and a Commercial Plant
alyst bulk (–) (CP) MA CFB reactor is given in Section 2. The development of the
proposed reactor model is discussed in Section 3. A detailed data
Superscript/subscript fitting exercise using the PP and CP reactors is discussed in Section
i Section of the reactor (1: transport bed; 2: riser) 4. The results obtained from the data fitting exercise are discussed
k’ Specific catalytic reaction step in Section 5. The conclusions of this work are given in Section 6.
k" Adsorption desorption step (1: O2 ; 2: bu; 3: MA)

2. Brief description of the MA CFB reactors

2004, 2003; Patience and Lorences, 2006; Shekari and Patience, The MA CFB PP reactor (Patience and Bockrath, 2010) comprises
2013; Wang and Barteau, 2003, 2002, 2001; Xiao-Feng et al., 2002). of a transport bed, a riser, a riser stripper, a regenerator and a regen-
Ballarini et al. (2006), in their review, conclude that despite being erator stripper. The MA CFB CP reactor (Patience and Bockrath,
one of the most studied reactions in the selective oxidation of the 2010) has no regenerator stripper but instead a cyclone separa-
small alkanes the final picture for MA production over VPO catalyst tor is present. The details of these two reactors are not given here
is still not clear. for brevity and may be found elsewhere (Hutchenson et al., 2010;
Several studies (Avidan and Yerushalmi, 1982; Bi and Grace, Patience and Bockrath, 2010). Here, the schematic of a MA CFB reac-
1995; Chan et al., 2010; Grace et al., 1999; Gupta, 1998; Yerushalmi tor is shown in Fig. 2. Fresh Bu and nitrogen (N2 ) along with the
and Cankurt, 1979; Zhang et al., 2015) report the different oper- recycle gas, which contains Bu, oxygen (O2 ), carbon dioxide (CO2 )
ating hydrodynamic regimes of the different gas-solid systems. and carbon mono-oxide (CO) (Patience and Bockrath, 2010) are fed
The different hydrodynamic regimes present in the CFB reactors to the bottom of the transport bed. The velocity of the gas mix-
with solid recirculation are, namely, Dense Suspension Upflow ture is sufficient enough to fluidize and lift the catalyst particles
(DSU), Fast Fluidized Bed (FFB), Core Annular Flow Bed (CAFB) coming from the regenerator in the upward direction. Some fresh
and Pneumatic Conveying (PC). Recently, Zhu and Zhu (2008) have O2 is fed separately through the spargers to avoid the flammability
suggested another hydrodynamic regime present in the CFB reac- limit at the inlet gas mixture. It is noted that the fresh O2 is fed
tors as Circulating-Turbulent Fluidized Bed (C-TFB). The superficial at three different locations at three different heights in the trans-
gas velocity, U0 , and the solid mass flux, Gpz , are the two main port bed to avoid the over reduction of the catalyst. The oxidation
parameters for defining these different hydrodynamic regimes. reactions take place on the catalyst particles as the gas and solid
A schematic representation of these regimes and the criteria are move through the transport bed and the riser section of the reac-
shown in Fig. 1. These criteria are the critical values of U0 (Uff , UCA , tor. The catalyst is reduced by the consumption of its lattice oxygen
UC-TFB and UPC ) and Gpz (Gpz SCC ; SCC: saturation carrying capacity) (Patience and Bockrath, 2010). The reduced catalyst is separated
across which the transition from one regime to another regime from the gas mixture at the outlet of the riser and send to the regen-
200 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Fig. 1. Regime map for gas-solid system with and without solid circulation.

Table 1
Correlations for the terms involved in the criteria for regime transition as regime map given in Fig. 1.

References Correlations for the criteria of the regime transition mentioned in Fig. 1
 0.55  0.36
g Uff dp Gpz g (cat − g ) gdp3
Uff Gupta (1998) = 12.55
g 2g
vpz, t cat
1.192 −1.064 −0.064
Grace et al. (1999) Uff = 0.0113Gpz g [g g (cat − g )]
⎛ ⎞0.59
 (cat − g ) gdp3
−0.2
= 22.8⎝ ⎠
Uff /ε Gpz
Kim et al. (2004)   g

vpz, t 2g
cat 1−ε vpz, t
 GSCC 0.542   (cat − g ) gdp3
0.105
U
0
SCC pz g
Gpz Bi and Fan (1991) = 21.6
g U0 2g
gdp
1.85
SCC  (cat − g ) gdp3
0.63  
Gpz dp U0 cat − g

g
Bai and Kato (1995) = 0.125
g 2g g
gdp
 (cat − g ) gdp3
0.105  GSCC 0.542
g pz
UCA Bi and Fan (1991) UCA = 21.6 gdp
2g g U0
UC−TFB Qi (2012) UC−TFB = 0.0041Gpz + 1.78
 0.31  −0.139   (cat − g ) gdp3
−0.021
0.347 Gpz dp g
UPC Bi and Fan (1991) UPC = 10.1(gdp )
g DT 2g

erator. In the regenerator, the reduced catalyst is fluidized with the model. The various reactor and catalyst parameter values used in
fresh oxygen to replenish the consumed lattice oxygen (Patience the modeling are given in Table 2.
and Bockrath, 2010). In the MTFBR and fluidized bed reactors, the
oxidation reactions over the catalyst particles and the oxidation of
the reduced catalyst particles by the gaseous oxygen occur simul- 3. Reactor modeling
taneously, while in the MA CFB reactor these two steps occur in two
different zones. The desired product, MA, is separated from the out- 3.1. Kinetic model
let gas mixture in the form of maleic acid and a significant portion
of the remaining gas is recycled back to the reactor (Patience and The heterogeneous catalytic reaction mechanisms may broadly
Bockrath, 2010). be divided into two categories based on the role of the catalyst.
Thus, as discussed above, the MA CFB reactor is a multi- In the first case, the catalytic surface acts only as an adsorption
component reactor system (MCRS) involving 8 components. 7 (NGC ) platform (e.g., Langmuir–Hinshelwood, Eley–Rideal, etc.) and in the
of these are present in the gas phase (Bu, MA, O2 , H2 O, CO2 , CO and second case, the catalyst plays an active role in the reaction through
N2 ) and 1 in the solid phase (the lattice oxygen as the fraction of the redox reaction (Mars–Van Krevelen). In the kinetic model devel-
occupied lattice oxygen sites in the catalyst bulk, OL ). The regener- opment for these cases quasi steady state assumption (QSSA) for the
ator is assumed to be ideal and at the steady state with complete catalyst sites is generally assumed. Several reported kinetic models
regeneration of the lattice oxygen in each cycle. In the proposed applying QSSA on all the catalyst sites at the steady state conditions,
work, the modeling is done only for the transport bed and the riser applicable for the MA MTFBR, are summarized in the open literature
of the MA CFB reactor. The riser stripper, cyclone separator, regen- (Xiao-Feng et al., 2002). These kinetic models are not applicable for
erator and regenerator stripper are not modeled. Thus, the inlet of the MA CFB reactors because even for the steady state operation of
the transport bed is considered as the inlet and the outlet of the these reactors the catalyst particles are circulated between a fuel
riser is considered as the outlet in the proposed MA CFB reactor enriched environment and an oxygen enriched environment. Thus,
the catalyst is never truly in a steady state condition.
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 201

Table 2
Reactor and catalyst parameter values.

Reactor properties PP reactor CP reactor Reference

HT, 1 (m) , DT, 1 (m) 11.15, 4.2 6, 0.3 Patience and Bockrath (2010)
HT, 2 (m) , DT, 2 (m) 28.5, 1.8 24, 0.15 Patience and Bockrath (2010)
zOIN, 1 (m) , zOIN, 2 (m) , zOIN, 3 (m)
2  2 2
0.45, 1.9, 5.5 0.9, 2.1, 3.7 Shekari et al. (2010)
kg
Cat 1573 1573 Patience and Bockrath (2009)
m3

dp (␮m)  70 70 Shekari et al. (2010)
NL mmolO2 /kg Cat 310 310 Patience and Bockrath (2010)

experimental feed (Wang and Barteau, 2001) was reported whereas


the CFB reactor is known to have 3.6–12.9% O2 (Patience and
Bockrath, 2010) in the feed. Shekari and Patience (2013) reported a
kinetic model for the MA CFB reactors operating under similar con-
ditions and in terms of the OL . In their kinetic model the selectivity
of the MA is a ratio of the kinetic parameters which is inconsis-
tent with the observed experimental data for the MA CFB reactors
(Patience and Bockrath, 2010). Among the other transient kinetic
models, the kinetic model proposed by Gascon et al. (2006) involves
the least number of variable catalyst sites (O2 and OL ). The reaction
mechanism (Gascon et al., 2006) involves other catalyst sites (Bu ,
MA and OS ) and QSSA is made for these sites.
A modification of the Gascon’s kinetic model (Gascon et al.,
2006) is proposed for the MA CFB reactors in this study. The pro-
posed modified kinetic model uses a single variable catalyst site.
Along with the reported QSSA on Bu , MA and OS sites, QSSA on O2
is assumed. A fundamental approach is used to develop the mod-
ified kinetic model equations using the above assumption on O2
site and the reaction mechanism proposed by Gascon et al. (2006).
The proposed kinetic model as given in Table 3 has the following
three differences with the Gascon’s kinetic model
I. Instead of considering O2 as an independent variable, it is
calculated from Eq. (1).

Keq, 1 pO2
O2 = (1)
1 + Keq, 1 pO2 + Keq, 2 pBu + Keq, 3 pMA
Fig. 2. A schematic of MA CFB reactor.
II. The extra term for the MA adsorption, Keq,3 pMA , is present in
the denominator of the expressions for r4 to r8 as given in Table 3
Several reaction mechanisms and transient kinetic models are and represents the adsorption-desorption equilibrium of MA.
reported (Gascon et al., 2006; Lorences et al., 2004, 2003; Patience III. The rate expression for r8 in Table 3 is proportional to the
and Lorences, 2006; Shekari and Patience, 2013; Wang and Barteau, adsorbed MA and OS as reported in the reaction mechanism but
2002, 2001; Xiao-Feng et al., 2002) by simulating the experimental not present in the Gascon’s kinetic model (Gascon et al., 2006).
conditions similar to that of the MA CFB reactor. This is achieved by rate constants,
operating the experimental reactors in cycles with the fuel enriched
The  (kk ) and the equilibrium
constants, Keq, k" as defined in the reactions
and the oxygen enriched feeds. The transient state of the catalyst (k = 1, 2, ..., Nrxn ; k" = 1 : O2 ; 2 : Bu; 3 : MA) given in Table 3
particles under such operating conditions is modeled in terms of are expressed in the form of Arrhenius equation and Van’t Hoff
variable catalyst sites. These variable catalyst sites arereported equation given by Eqs. (2) and (3), respectively
as the fraction of adsorption sites occupied by O2 O2 and the
     
fraction of occupied lattice oxygen sites in the catalyst bulk OL EA,k 673 
kk (T ) = kk0 exp 1− ; k = 1, 2, ..., Nrxn (2)
by Gascon et al. (2006); as O2 , OL , the fraction of lattice surface 
T
  673Rg
sites occupied by oxygen and MA OS and MA
S by Xiao-Feng et al.
 5+ 4+ C4

(2002); the Vanadium sites V ,V and V by Lorences et al.   
(2004, 2003) and Patience and Lorences (2006) and as OL by Shekari 0
Heq,k"e; 673
Keq,k"e; (T ) = Keq,k"e; exp  1− ; (k"e; = 1 :
and Patience (2013) and Wang and Barteau (2002, 2001). It is noted 673Rg T
that other catalyst sites are also involved in the reported reaction
O2 ; 2 : Bu; 3 : MA) (3)
mechanisms but the QSSA is made for these catalyst sites in their
respective kinetic models.
The LOC information for both the MA CFB reactors (Patience and
The reported values of the parameters used in these equations
Bockrath, 2010) is given as the variable catalyst site, OL . Though 0
are given in Table 4. The two extra parameters, Keq, and Heq,3 ,
the kinetic models proposed by Wang and Barteau (2002, 2001) 3
are based on OL but the models are not applicable for the MA corresponding to the MA adsorption are introduced in the pro-
CFB reactors. In their model, either too much oxygen (21%) in the posed kinetic model. The reaction rates, Rj , for each of the NGC
experimental feed (Wang and Barteau, 2002) or no oxygen in the gaseous components and OL are written in terms of the intrinsic
202 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Table 3
Modified Gascon’s kinetic model.

Oxidation of surface lattice site from adsorbed oxygen


k1  
O2 + 2VS →2OS + V r1 = k1 O2 1 − OS
Diffusion of lattice oxygen from subsurface lattice sites to surface lattice sites:
k2  
OL →OS r2 = k2 OL − OS
Reactions happening at the adsorbed oxygen site:
k3
6.5O2 + C4 H10 (g) →4CO2 (g) + 5H2 O + 6.5V r3 = k3 pBu O2
k4
3.5O2 + C4 H10 (g) →C4 H2 O3 (g) + 4H2 O + 3.5V r4 = k4 pBu 2O
2
Reactions happening at the lattice oxygen site with adsorbed Bu and MA:
 2
k5 pBu OS
 
k5
7OS + C4 H10 →C4 H2 O3 + 4H2 O + 7VS r5 =
1 + Keq, 1 pO2 + Keq, 2 pBu + +Keq, 3 pMA
k6 pBu OS
 
k6
9OS + C4 H10 →4CO + 5H2 O + 9VS + V r6 =
1 + Keq, 1 pO2 + Keq, 2 pBu + +Keq, 3 pMA
k7 pBu OS
 
k7
13OS + C4 H10 →4CO2 + 5H2 O + 13VS + V r7 =
1 + Keq, 1 pO2 + Keq, 2 pBu + Keq, 3 pMA
k8 pBu OS
 
k8
6OS + C4 H2 O3 →4CO2 + H2 O + 6VS + V r8 =
1 + Keq, 1 pO2 + Keq, 2 pBu + Keq, 3 pMA

Table 4
Reported and tuned values (simultaneous data fitting exercise) of the kinetic parameters.

Kinetic parametera Reported values Final Tuned Values (Simultaneous Kinetic parameterb Reported values
(Gascon et al., 2006) data exercise fitting) (Gascon et al., 2006)
   
k10 10−6 × kmol/kg − s 5.52 8.81 EA,1 104 × kJ/kmol 3.09
   4 
k20 10−6 × kmol/kg − s 3.65 13.44 EA,2 10 × kJ/kmol 13.40
   4 
k30 10−6 × kmol/kg − bar − s 2.33 0.57 EA,3 10 × kJ/kmol 13.81
   4 
k40 10−6 × kmol/kg − bar − s 14.3 1.41 EA,4 10 × kJ/kmol 5.03
   4 
k50 10−6 × kmol/kg − bar − s 5.82 19.93 EA,5 10 × kJ/kmol 3.24
   4 
k60 10−6 × kmol/kg − bar − s 8.15 6.51 EA,6 10 × kJ/kmol 10.61
   4 
k70 10−6 × kmol/kg − bar − s 2.77 2.55 EA,7 10 × kJ/kmol 8.20
   4 
k80 10−6 × kmol/kg − bar − s 1.45 6.69 EA,8 10 × kJ/kmol 13.79
0
   4 
Keq,1 H1 10 × kJ/kmol −3.06
0
1/bar 55.85 34.52
 4 
Keq,2 H2 10 × kJ/kmol −11.43
0
 1/bar 94.213 5.50
Keq,3
1/bar
4
 – 7.97
H3 10 × kJ/kmol – −2.42

a
Kinetic parameters which are used as tuning parameters in data fitting exercise.
b
Kinetic parameters which are used as reported and not used as tuning parameters in data fitting exercise.

Table 5 3.2. Hydrodynamic modeling


Rate expressions for different components.


NRxn The identification of the operating regimes for the transport bed
Rj k j rk and the riser of the MA CFB PP and CP reactors is necessary for the
k =1
development of a suitable hydrodynamic model. The experimental
OUT , G
values of the variables U0, i pz, i and εi , (i = 1: transport bed; i = 2:
RC4 H10 −r3 − r4 − r5 − r6 − r7
riser) are required to determine the operating regimes of the given
RC4 H2 O3 r4 + r5 − r8 OUT and G
section i. The experimental values of U0, 2 pz,2 for the riser are
RO2 −r1 − 6.5r3 − 3.5r4 calculated from the reported data (Patience and Bockrath, 2010) for
RH2 O 5r3 + 4r4 + 4r5 + 5r6 + 5r7 + r8 both the MA CFB PP and CP reactors. The analysis based on these
RCO2 4r3 + 4r7 + 4r8 values, along with the correlations given in Table 1, conclude that
RCO 4r6 the risers of both the MA CFB PP and CP reactors operate in the
DSU regime. The experimental value of U0, OUT for the transport bed
RN2 0 1
cannot be calculated directly from the reported data (Patience and
R L −r2 OUT is calculated by
Bockrath, 2010). An approximate value for U0,
O

OUT with A
 1 
multiplying U0, 2 CS, 2 /ACS, 1 assuming negligible variation
in U0 in the riser. The analysis based on these values, along with
reactions rates with the stoichiometric coefficients, k’j in Eq. (4). the correlations given in Table 1, conclude that the transport beds
The stoichiometric coefficients are given in Table 5. of both the MA CFB PP and CP reactors operate in the C-TFB regime.
Another hydrodynamic study is reported (Patience and Bockrath,

NRxn
2009) on the PP reactor only (not on the CP reactor) with the com-
Rj = k j rk ; j = 1, 2, ..., NGC and OL (4)
OUT , G
plete information for U0, pz,i and εi , (i = 1: transport bed; i = 2:
k =1 i
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 203

   
riser). The analysis based on this experimental data also results 1 d dYj (x) d CT (x) Yj (x) U0 (x)
in the same conclusion for the MA CFB PP reactor, i.e., the trans- Dge (x) CT (x) −
HT dx dx dx
port bed and the riser of PP reactor operates in the C-TFB and DSU
 
regimes, respectively.
+ cat 1 − ε (x) HT Rj (x) = 0 (7)
The hydrodynamic modeling of the gas and solid flow behav-
iors in a CFB reactor can be done using any of the following three
approaches – Approach 1: Assuming ideal flow reactors, either as
Dge d2 Cj (x) dCj (x)
 
PFR or CSTR; Approach 2: Using the first principle models with − U0 + cat 1 − ε (x) HT Rj (x) = 0 (8)
Computation Fluid Dynamics (CFD); and Approach 3: Empirical HT dx2 dx
modeling with non-ideal flows. Approach 1 is the simplest of the The overall continuity equation (Eq. (9)) used in the proposed
three approaches but lacks good prediction capabilities. Approach model considers the effect of the volumetric changes due to both
2 is good in visualizing the flow behavior. Nonetheless, the predic- the density variation (as a result of pressure drop) and the reactions.
tion capabilities of the resulting reactor model may not be good Note that Eq. (9) in the model is an overall material balance and
until most of the fluid flow phenomena are appropriately modeled. hence only NGC − 1 gaseous species (Bu, MA, O2 , H2 O, CO2 and CO)
Incorporating this constraint makes the model computationally balances are independent and considered in the individual species
intensive. Approach 3 is further classified in two sub-approaches balances given by Eq. (7). The mass transfer resistance over the
– Approach 3a: Modeling the flow complexities by approximating catalyst particles and the diffusional resistance within the catalyst
them to the assumed flow behaviors (e.g., the core-annular model particles are neglected. The effective dispersion coefficient because
or the cluster flow model) based on the experimental observations of the non-ideal flow is assumed to be the same   each species
for
and Approach 3b: PFR with dispersion or tank in series models (i.e., and constant throughout the transport bed Dge, 1 and the riser
“One-parameter models”). The single parameter in the PFR with
 
Dge, 2 .
dispersion model is the dispersion coefficient while it is the num-
 
ber of tanks for the tanks in series model. The experimental studies
cat 1 − ε (x) HT
on the residence time distributions of the pulse inputs in both the dU0 (x) dCT (x)
CT (x) + U0 (x) − RT (x) = 0 (9)
gas (Liu, 2001) and solid phases (Andreux et al., 2008) of the CFB dx dx CT (x)
systems have shown significant breakthrough time which repre-
Abba et al. (2002) in their variable-gas-density fluidized bed
sents a behavior close to the PFR than the CSTR. The tail behavior
reactor model suggested the superficial gas velocity be defined
in the output response, E(t), also shows deviation from the ideal
as given in Eq. (10). Further, the species continuity equation and
plug flow behavior. In the reviews reported in the open literature
the reactions were proposed to be written in terms of Fj (x) for the
by Bi (2004) and Breault (2006), the gas and solid hydrodynamics
MCRS. The expression for Fj (x) reported in their model considered
for CFB reactors are modeled as a PFR with dispersion. The same
only the convective molar flow and neglected the dispersive molar
approach is used in this study. The radial and axial dispersions are
flow. A correct expression for Fj (x), as given in Eq. (5), considers
modeled with only the effective
  coefficients
dispersion  (Schugerl,
both the convective and dispersive molar flows. The use of this
1967) for both the gas Dge and the solid Dpe phases. A uniform
expression, though, makes it quite difficult to express the reaction
radial distribution is assumed for both the gas velocity and the solid
rates in terms of Fj (x). The use of Eqs. (7) and (9) in terms of Yj , CT
hold up as suggested by Liu (2001).
and U0 is proposed as a much easier way of developing the reactor
model for the variable gas density systems.

3.2.1. Gas phase modeling FT (x) T (x) PTIN


U0 (x) = U0IN (10)
Choong et al. (1998) have suggested that the dispersive molar F IN T IN PT (x)
T T
Disp
flow rate of a gaseous species, Fj (x), in a binary mixture is pro- The pressure drops in the two sections (i = 1: transport bed;
portional to its mole fraction gradient analogous to the Fick’s law i = 2: riser) of the CFB reactor are calculated using Eq. (11). The
of diffusion. The justification for the suggestion was that the sum of first two terms represent the contributions of the gas and solid
dispersion flow rates for both the components must be zero. This weights, respectively. The third term represents the contribution
justification was made on the fact that the density variation or the of the acceleration effects and is present only in the transport bed.
total concentration variation across the length of flow is counter This term is absent in the riser as the solid flow is developed in the
balanced with the variation in the gas velocity by the continuity transport bed itself. The last two terms represent the contributions
equation in an open system. of the gas-wall and the solid-wall frictions, respectively. The fric-
The molar flow rate of a gaseous species, Fj (x), at a given dimen-
  tion factors for the gas, fg and the solid, fs can be calculated from
sionless height, x x = z/HT , with the dispersion in the MCRS the correlations given in Bai et al. (1997). The effect of the various
(similar to the binary case of Choong et al., 1998) is given by Eq. terms in Eq. (11) are analyzed on the reported hydrodynamic data
(5). The total molar flow rate is only convective and is given by Eq. (Patience and Bockrath, 2009) for both the sections of the MA CFB
(6). PP reactor, which are operating in the C-TFB and the DSU regimes,
respectively. This analysis concludes that solid weight is the sig-
nificant contributor (∼98%) for the pressure drop. Eq. (12) is, thus,
ACS Dge (x) CT (x) dYj (x)
Disp
Fj (x) = FjConv (x) + Fj (x) = ACS U0 (x) CT (x) Yj (x) − (5) used in the proposed model for calculating the pressure drop. The
HT dx
void fraction (ε1 (z)) for the C-TFB regime is reported to be axially
uniform (Zhu and Zhu, 2008) and, thus, a constant value, ε1 (z) = ε1 ,
FT (x) = ACS U0 (x) CT (x) (6)
is considered. The DSU operating regime for the riser is character-
ized with a significant axial variation only at the bottom section
The continuity equation for a given species in the MCRS with because of the acceleration effects, followed by a near uniform void
dispersion, variable gas density and reactions is given by Eq. (7). fraction in the upper section of the riser work (Issangya, 1998). In
This equation in the limiting case of constant gas density simplifies MA CFB reactors, the bottom section of the riser operates in the C-
to the expression given in Eq. (8) and is similar to the expression TFB regime and hence the void fraction of the riser is characterized
reported by Danckwerts (1953). only by the upper section of the DSU regime and is considered con-
204 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Table 6
Model equations as simultaneous 1st order ODEs.

dT (x) HT  
= TiOUT − TiIN
dx Hi  
fconv cat 1 − εi gHT
dCT (x) PT (x) dT (x)
=− −
dx Rg T (x)  Rg T
2 (x) dx
cat 1 − εi HT
dU0 (x) U0 (x) dCT
=− (x) + RT
dx CT (x) dx CT (x)
⎡   ⎤
  HT cat 1 − εi  
d dYj
(x) =−
1 dYj (x) dCT (x)
+
U0 (x) HT
⎣ dYj (x) + Yj RT (x) − Rj (x) ⎦ ; (j = 1, 2, ...NGC − 1)
dx dx CT (x) dx dx Dge, i dx CT (x) U0 (x)

⎡   ⎤
 dL  dOL
cat 1 − εi HT
d O
(x) = 
Gpz, i HT
 ⎣ (x) − RL (x)⎦
dx dx dx Gpz, i NL O
cat 1 − εi Dpe, i

d dYj
Yj (x) = (x) ; (j = 1, 2, ...NGC )
dx dx
d L dOL
 (x) = (x)
dx O dx

stant ε2 (z) = ε2 . This is also in agreement with a recently reported persive molar flow rates is given by Eq. (15) and is similar to Eq.
experimental work (Wang et al., 2014). It is noted that this unique (5).
set up of the C-TFB operating regime at the bottom and the DSU
 
CL (x) = NL cat 1 − ε OL (x) (14)
regime at the top ensures the assumption of two different but con- O
stant void fractions (constant solid hold up) in the transport bed ⎡   ⎤
and the riser. It is noted that if a reactor is operating in the DSU ACS NL cat 1−ε Dpe
dOL
= ACS Gpz NL OL (x) + ⎣− (x)⎦
v Disp
regime only, then the axial variation in the void fraction will be FL (x) = F Con
L +F
L
(15)
O 
O O
HT dx
required to be considered in the hydrodynamic model.
The compressibility factor at the reported experimental con- The constant solid phase 
ditions (Grace et al., 1999; Issangya et al., 1996) is almost one.
 coefficients in the
effective dispersion
transport bed and the riser Dpe, 1 and Dpe, 2 are considered similar
Thus, the gas phase is considered an ideal gas mixture and the total to the gas phase effective dispersion coefficients. The molar balance
concentration is given by Eq. (13).
⎡ ⎤
HT
2
Gpz,1 2
2fg εi g U0,i 2fs HT
2
Gpz,i
dPT ⎢ g gε HT +  g 1 − ε (x) HT +   + HT +   ⎥
= fconv ⎢ ⎥(11)
i Cat i
− HT,1 DT,i DT,i
dx ⎣ Cat 1 − ε1 Cat 1 − εi ⎦

dPT (x)
  for the lattice oxygen across a differential segment of the reactor
= −fconv cat 1 − εi gHT i (12) between heights x and x + dx is given by Eq. (16). The balance for OL
dx
is given by Eq. (17) which is obtained by combining Eqs. (15) and
(14).
PT (x)
CT (x) =
Rg T (x)
(13) dFL (x)  
O
= ACS cat 1 − ε HT RL (x) (16)
dx O
The MA CFB reactors have excellent heat removal characteris-  
tics. The temperature rise in the transport bed and the riser of a MA d L
  d dOL
CFB PP reactor are reported as 7 K and 4 K, respectively (Patience HT Gpz NL  (x) − cat 1 − ε Dpe NL (x)
dx O dx dx
and Bockrath, 2010) irrespective of the very high exothermic reac-  
tions involved in the selective oxidation of butane. The equations − cat 1 − ε HT2 RL (x) = 0 (17)
in form of the constant temperature gradients, as given in Table 6, O

for both sections of the MA CFB PP reactor are used in the proposed
model with the reported values for the temperature rise in each
section. In the MA CFB CP reactor the temperature rise values are 3.2.3. Summary of the basic model for the MA CFB reactor
considered to be zero since the CP reactor operation is reported at The complete set of model equations in terms of 2NGC + 3 simul-
isothermal conditions (Patience and Bockrath, 2010). taneous 1st order ODEs are summarized in Table 6. The first three
ODEs are for the temperature, the total concentration and the con-
tinuity equation, respectively. The NGC 2nd order ODEs represent
3.2.2. Solid phase dispersion model the dispersion model for NGC − 1 gaseous species and OL as given
The deactivation of the catalyst is modeled in terms of the by Eq. (7) and Eq. (17), respectively. These NGC 2nd order ODEs are
available lattice oxygen. The concentration of the available lattice converted into a set of 2NGC 1st order ODEs. 2NGC + 3 variables (T,
oxygen at the dimensionless height, x, is given by Eq. (14). The CT , U0 , OL , dOL /dx, Yj , and dYj /dx; j = 1, 2 . . ., NGC −1) involved in the
lattice oxygen capacity of the catalyst (NL ) is reported (Wang and 2NGC + 3 1st order ODEs are termed as the model variables. Thus, a
Barteau, 2001) as 310 mmol O2 /kg and used in the proposed work. total of 2NGC + 3 specified values are required to solve these model
The molar flow for the lattice oxygen with both convective and dis- equations.
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 205

Table 7 and solid mixing in the C-TFB regime, the values of the Peclet num-
Parameter values used in Eq. (18).
bers for the transport bed section (Pege,1 and Pepe,1 ) of both the PP
ai bi ci di and CP reactors are used as tuning parameters. Both the solid and
i = 1: transport bed 2.3 0 2.7 1 gas effective dispersion coefficients are related to the correspond-
i = 2: riser 1 0 4.5 −1.7 ing Peclet numbers with the superficial gas velocities as reported
by Bi (2004). It is againnoted
that U0,i used in Eqs. (18)–(20) is mea-
f
sured at the outlet i.e., U0, i , as reported in the literature (Patience
3.2.4. Hydrodynamic parameters of the model

The hydrodynamic parameters involved in the reactor model and Bockrath, 2009).
are the two void fractions ε1 , ε2 and the four effective disper- U0, i HT
  Dge, i = (19)
Pege, i
sion coefficients Dge, 1 , Dge, 2 , Dpe, 1 and Dpe, 2 . A correlation for εi
(i = 1: transport bed; i = 2: riser) of the MA CFB PP reactor is reported U0, i HT
Dpe, i = (20)
in the open literature (Patience and Bockrath, 2009) as given by Eq. Pepe, i
(18). The correlation involves 4 parameters (ai , bi , ci and di ) in both
the sections. The reported d2 value is used with a sign change from 3.3. Conditions for the reactor configuration complexities
positive to negative. This is required to fit the trend in the reported
graph between C0,2 and U0 (Wang and Barteau, 2001). The same The reactor model in the form of 2NGC + 3 1st order ODEs, given
sign change is also reported in the literature (Patience and Bockrath, in Table 6, remains same in both the transport bed as well as the
2009). In the absence of the “fines” data for the MA CFB PP and CP riser. The parameters and the model variables values change from
reactors (Patience and Bockrath, 2010), b1 is used as 0 in the pro- the transport bed to the riser. The sudden change in the variable
posed model. It is noted that for the riser section (i = 2) b2 is already values also happen at the mixing points of the oxygen spargers in
reported as 0 (Keraron, 2009) and the same is used in the proposed the transport bed. The model variable values also change due to the
model. The values of the parameters (ai , bi , ci and di ) used for the sudden change in the cross-sectional area between the transport
 are given in Table 7. The same correlation given
MA CFB PP reactor bed and the riser. The relationships between the model variables at
by Eq. (18) for εi is used for the MA CFB CP reactor but the param- the interface of each section using the fundamental balances and
the basic assumptions are discussed next.
eters are retuned. It is noted that the correlation given by Eq. (18)
use U0 measured at the outlet of the transport bed and the riser i.e.,
OUT , as reported in the literature (Patience and Bockrath, 2009). 3.3.1. Conditions at the mixing point of oxygen spargers
U0, i The location of the fresh oxygen mixing point through kth (k = 1,
2 and 3) sparger is represented by x = xIN, O2 . The temperature, pres-
εi − εmf U0, i − Umf k
=   (18a) sure, molar balances, etc., are conserved at each of the mixing points
1 − εmf Gpz O2 O2
C0, i U0, i − Umf + + Vg, i (i.e., at xIN, k
(−) and xIN, k
(+)) for the steady state operation of the
cat (1−εmf )
reactor. The fresh feed of oxygen is assumed to be at the same tem-
where, perature as the reaction mixture before and after the mixing point.
Gpz The pressure at a given height, for assumed constant temperature,
di
C0, i = 1 + ci U0, i
  (18b) represents the constant total concentration at that height. The side
cat 1 − εmf
additions of the fresh oxygen result in change of the relative con-
 2   g 1 centration values of the different gaseous species so as to keep the
Vg, i = ai dp, 50 Cat − g (18c) total pressure/concentration constant before and after the mixing
18g (fines)bi
point.
The four effective dispersion coefficients which represent the The molar balances for the lattice oxygen, (NGC − 1) gaseous
measure of gas and solid mixing are the other set of hydrodynamic species along with the total molar balance in the gas phase are
parameters. Breault (2006), based on the reported correlations in given by Eqs. (21)–(23).
the literature, concluded that the different correlations applicable  O2
  O2

for different operating regimes result in different order of values FL xIN, k (−)
= FL xIN, k (+)
(21)
O O
for the gas and solid dispersion coefficients. Bi (2004), based on    
the experimental data from the open literature, reported that the   Fj
O 2
xIN,k (+) (for j = 1, 2, ..., NGC − 1; j =
/ O2 )
Fj O2
xIN,k (−) =   (22)
Peclet numbers based on the effective  diffusivity decrease
 with
 Fj
O2
xIN,k (+) − FOIN,k
2
(for j = O2 )
an increase in Gpz both for the gas Pege and solid phases Pepe .
   
Moreover, after the onset of the DSU regime Pege and Pepe become FT O2
xIN, k (−)
= FT O2
xIN, k (+)
− FOIN, k (23)
2
constant (Bi, 2004). The constant values are attributed to the dis-
appearance of the downward flow in the annular region in the DSU It is assumed that the fresh oxygen feed through the spargers only
regime. Wei et al. (1993) reported that the dispersion coefficient affect the gaseous phase and not the lattice oxygen of the catalyst as
represented in terms of the Peclet numbers is proportional to the evident from Eqs. (21) and (22). This is valid if and only if QSSA is not
assumed on OL . The assumption makes both the convective F Con v
square root of the diameter of the reactor. Hence, the constant val- L 
O
ues of the Peclet numbers (Pege and Pepe ) could be different for the Disp
and dispersive components F of FL to be continuous before and
PP and CP reactors. The range of diameters of the experimental reac- L
O O
tors used in the study by Bi (2004) is 0.03 m–0.94 m. The diameter of after the mixing point, which are represented by Eqs. (21) and (24).
the PP reactor riser (0.15 m) lies well within this range, whereas the The convective molar flow rates of the gaseous species other than
diameter of the CP reactor riser (1.8 m) lies outside this range. Thus, the oxygen are not affected by the fresh oxygen feed as given in Eq.
the constant value of the Peclet numbers (Pege and Pepe ) for the PP (25). The fresh O2 feed is assumed to be added only in the molar
reactor model is used as 10 as suggested by Bi (2004) whereas for convective flow of O2 as given in Eq. (26a). It is noted that the kinetic
the CP reactor model this value is tuned. parameters of the reactor model are further tuned in the proposed
The transport bed of both the MA CFB PP and CP reactors operate work. The reaction extent and thus the molar flow rate values pre-
in the C-TFB regime. In the absence of a proper analysis of the gas dicted by the model depend on the values of the kinetic parameters
206 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

in the tuning exercise. The values of the tuning parameters are 3.4. Boundary conditions
varied in the tuning exercise. The specific values of these tuning
parameters may lead to specific situations. The conditions applied (Abba et al., 2002) used the Danckwerts boundary conditions, as
in the model for the reactor configuration complexities should be given in Eq. (32), in the dispersion model considering the variable
applicable for such specific situations. For example, one such spe- gas density.
cific situation at the oxygen mixing points could be that the gaseous  
dCj (x) HT
O2 is completely consumed before any of the oxygen mixing points. = Cj (x) − CjIN at x = 0
dx U0IN Dge,1
 O is achieved
The specific situation of complete O2 consumption  in Eq. (32)
(26a) when the model predicted value of FOConv xIN,2 (+) becomes
k dCj (x)
2
=0 at x = 1
equal to FOIN, k as compared to being greater than it. After this situ- dx
2
ation Eq. (26a) is not applicable as the convective molar flow just The modified Danckwerts boundary conditions for a variable
before the mixing point cannot be negative and should be zero. density binary gas mixture were proposed by Choong et al. (1998).
Thus, for this case the condition given by Eq. (26b) is used in the The modified conditions consider the dispersive molar flow rates of
model in place of Eq. (26a). different species to be proportional to the gradients of their mole
    factions Choong et al. (1998). It is noted that the present reactor
F Conv xO2 = F Conv xO2 (24) system is a MCRS with significant pressure drop. Thus, the modified
 L IN, k (−) L  IN, k (+)
O O Danckwerts boundary conditions given by Eq. (33) are used in this
study. It is noted that for the limiting case of constant density, Eq.
    (32) are obtained from Eq. (33) but the reverse is not possible. A
FjConv xIN,k
O2
(−) = FjConv xIN,k
O2
(−) similar set of modified Danckwerts boundary conditions given by
Eq. (34) for the solid phase component, OL , is used.
(forj = 1, 2, ..., NGC − 1; j =
/ O2 ) (25)  
dYj (x) HT FTIN FjIN
= Yj (x) − at x = 0
dx ACS,1 CTIN Dge,1 FTIN
(forj = 1, 2, ..., NGC − 1) (33)

Conv
  Conv
  IN,k
 Conv
  IN,k
 dYj (x)
=0 at x = 1
FO O2
xIN,k (−) = FO O2
xIN,k (+) − FO ifFO O2
xIN,k (+) > FO (26a) dx
 
2 2 2 2 2

  F INL
dOL (x) Gpz,1 HT 
FOConv O2
xIN, k (−) =0 (26b) =   OL (x) − O
at x = 0
2 dx ACS,1 Gpz,1 NL
Dpe,1 cat 1−ε (34)
The temperature and pressure continuity along with Eqs. (21)–(26) dOL (x)
are the basis for 2NGC + 3 relationships between 2NGC + 3 model =0 at x = 1
dx
variable values just before and after the mixing points, given in
The modified Danckwerts boundary conditions have been used only
Table 8.
for the NGC − 1 components in the gaseous phase (Eq. (33)), thus, CT
and U0 at the inlet and outlet are related by Eq. (35).
3.3.2. Conditions at the interface of the transport bed and the riser ACS,1 U0 (x) CT (x) = FTIN at x = 0
The interface between the transport bed and the riser is rep- (35)
resented by x = xINT . The temperature and pressure are constant ACS,2 U0 (x) CT (x) = FTOUT at x = 1
immediately before and after the interface (i.e., at xINT (−) and The species molar flow rate of the NGC − 1 gaseous species and the
xINT (+)), resulting in a constant total concentration at the inter- molar flow rate of OL are given by Eqs. (36) and (37) at the outlet.
face according to Eq. (13). The molar flow rates are also conserved It is noted that the dispersive flow rates according to the modified
at the interface for the steady state operation of the reactor. The Danckwerts boundary conditions are zero at the outlet. The pres-
conservation of species molar balances for the lattice oxygen, the sure at the outlet of the MA CFB PP and CP reactors is reported
NGC − 1 gaseous species along with the total molar balance in the (Patience and Bockrath, 2010). The temperature is calculated at the
gaseous phase are given by Eqs. (27)–(29). There is no addition from outlet of both the reactors from the reported information (Patience
the outside of reactor at the interface, and hence the mole frac- and Bockrath, 2010). The total concentration at the outlet is then
tion Yj of each of the (NGC − 1) gaseous species and OL is assumed calculated by Eq. (5).
to be continuous as given by Eqs. (30) and (32). The temperature
and pressure continuity along with Eqs. (27)–(31) are the basis for FjOUT = ACS,1 U0 (x) CT (x) Yj (x) at x = 1 (forj = 1, 2, ..., NGC − 1) (36)
the 2NGC + 3 relationships, given in Table 8 between 2NGC + 3 model
variable values at the outlet of the transport bed and at the inlet of F OUT
L = ACS,2 Gpz,2 NL OL (x) at x = 1 (37)

O
the riser.
    4. Data fitting exercise
FL xINT (−) = FL xINT (+) (27)
O O
The data fitting exercise is done in two steps, namely, model
    tuning and model validation. The available experimental data of
Fj xINT (−) = Fj xINT (+) ; (j = 1, 2, ..., NGC − 1) (28)
both the MA CFB PP and CP reactors are divided into the training
    and validation sets. The model tuning is done by error minimiza-
FT xINT (−) = FT xINT (+) (29) tion between the experimental data and the model predicted values
using the training set. The error minimization results in the tuned
   
Yj xINT (−) = Yj xINT (+) ; (j = 1, 2, ..., NGC − 1) (30) values of the kinetic and hydrodynamic model parameters. The
model validation is done by analyzing the accuracy of the reactor
    model predictions with the tuned parameters using the validation
OL xINT (−) = OL xINT (+) (31) set.
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 207

Table 8
Conditions at the oxygen mixing points and at the interface of the transport bed and the riser.
O2
Conditions at x = xIN, k
Conditions at x = xINT

T (−) = T (+) T (−) = T (+)


 = PT (+)
PT (−)
O2 −
PT (−) = PT (+) 
U0 xIN, = U0 (−) = ACS, 2 /ACS, 1 U0 (+)
k
 
FOIN, k
2
U0 (+) 1 −
ACS, 1 U0 (+) CT (+)
OL (−) = OL (+) OL (−) = OL (+)  
ACS, 1 1 − ε1 Dpe, 1
dOL dOL dOL (+) dOL (−)
dx
(−) =
dx
(+)
dx
=   dx
ACS, 2 1 − ε2 Dpe, 2
FT (+)
Yj (−) = Yj (+) ; (j = 1, 2, ..., NGC ; j =
/ O2 ) Yj (+) = Yj (−) ; (j = 1, 2, ..., NGC )
FT (+) − FOIN, k
2
FT (+) YO2 (+) − FOIN, k
2
if FOConv (+) ≥ FOIN, k
YO2 (−) = FT (+) − FOIN, k 2 2
2

0 if FOConv (+) < FOIN, k


2 2
dYj dYj dYj (+) ACS, 1 Dge, 1 dYj (−)
(−) = (+) ; (j = 1, 2, ..., NGC ; j =
/ O2 ) = ; (j = 1, 2, ..., NGC )
dx dx dx ACS, 2 Dge, 2 dx
dYO2
(+) if FOConv (+) ≥ FOIN, k
dx 2 2
dYO2
(−) =  
dx dYO2 (+) U0 (+) HT FOIN, k
− YO2 (+) − 2
if FOConv (+) < FOIN, k
dx Dgz, 1 FT (+) 2 2

4.1. Experimental data 4.2. Model simulation

The experimental data for both the MA CFB PP and CP reactors 4.2.1. Numerical methods for the model simulation
are required for the data fitting exercise of the respective reactor A total of 2NGC + 3 conditions are required to simulate the MA
models. The experimental data for both the reactors are reported in CFB reactor model consisting of 2NGC + 3 simultaneous 1st order
terms of the conversion of Bu (XBu ), the selectivity of MA (SMA ), the ODEs given in Table 6. Many relations as given by Eqs. (33)–(37)
mole fractions of some components at the inlet and/or of other com- and discussed in the Section 3.4 may act as the boundary condi-
ponents at the outlet, the lattice oxygen contribution (LOC) and the tions. These conditions require the information of the molar flow
IN, Exp OUT, Exp
catalyst mass flow rate (Patience and Bockrath, 2010). The exper- rates both at the inlet (Fj ) and at the outlet (Fj ) of the
imental data for both the reactors are divided in two parts: the reactor. The pressure and temperature are also available at the out-
training and validation sets. The ratio of these sets is kept at 2:1. let of the reactor. It is noted that only some part of the experimental
The training and validation sets are made to have similar range of data, discussed in Section 4.1, of each experimental case can used
values for each of the reported information. as the required conditions for the model simulation because the
To simulate a reactor-model, the information for all the NGC + 1 remaining information is required for the error minimization in
components are required at either the inlet of the transport bed the model tuning step. It is also observed that the complete set of
and/or at the outlet of the riser. Thus, the reported experimental required 2NGC + 3 conditions at one point can only be obtained at
data of both the reactors (PP and CP) are converted into the indi- the outlet of the reactor based on the available information and
vidual molar flow rates of each component at the inlet (FjIN ) and the use of Eqs. (33)–(37). The same is not possible at the inlet of
at the outlet (FjOUT ). This is achieved by applying the material bal- the reactor. Hence, the reactor model can be simulated as an ini-
ances on a process flow sheet as shown in Fig. 2. It is again stressed tial value problem (IVP) only in the backward direction from the
that the modeling is done only for the two sections of the reac- outlet to the inlet of the reactor with the initial conditions given
tors, namely, the transport bed and the riser. Out of the reported in the first column of Table 9. It is evident, from Table 9, that the
information one information was relaxed to get a unique solution model simulation as an IVP uses the information of the molar flow
OUT, Exp
of the mass balances for the designed flow sheet. The LOC value rates only at the outlet (Fj ) of the reactor. The reactor model
was relaxed based on the assumption that the LOC measurements can also be simulated as a boundary value problem (BVP) with the
depend on the residence time distribution of the solid catalyst par- boundary conditions given in the second column of Table 9. It is evi-
ticles which is not reported in the experimental data (Patience and dent from the second column of Table 9 that the model simulation
Bockrath, 2010). The reported ratio of the oxygen feed rates through as a BVP uses the information of the molar flow rates only at the
IN, Exp
the three spargers is 4:2:3 for the MA CFB PP reactor based on the intel (Fj ) of the reactor. Thus, different information in the form
reported number of nozzles at the each sparger (Hutchenson et al., of molar flow rates, Fj
OUT, Exp
and Fj
IN, Exp
are required for the model
2010). The ratio of the oxygen feed rates used for the MA CFB CP
simulation as an IVP and a BVP, respectively. Thus, the remaining
reactor is 1:1:0.325 based on the reported information that the feed IN, Exp OUT, Exp
information Fj and Fj used in defining the error formu-
rates at the lower and the middle spargers of the CP reactor were
lation are also different for the model simulations as an IVP and a
of the order of 1000 kg/h, whereas at the upper sparger it varied
BVP, respectively.
between 250 and 500 kg/h (Hutchenson et al., 2010).
The reactor model includes the interface between the transport
bed and the riser along with the three oxygen mixing points as
discussed above. The conditions at these points make the ODEs
unstable at these points and thus a stiff ode solver ode15s of
208 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Table 9
Boundary and Initial conditions for simulating the model as a BVP and an IVP, respectively.

Initial Conditions for IVP (In terms of output molar flow rates) Boundary Conditions for BVP (In terms of input molar flow rates)
  PTOUT  −  PTOUT
CT 1− = CT 1 =
 −  OUT
Rg T OUT  −  OUT
Rg T OUT
T 1 =T T 1 =T
 − FTOUT  + FTIN
U0 1 = U0 0 =
ACS, 2 CT (1− ) ACS, 1 CT (0+
) 
 −  FjOUT dYj (0+ ) U0 (0+ ) HT  + FjIN
Yj 1 = OUT j = 1, 2, ..., 6 = Yj 0 − j = 1, 2, ..., 6
FT dx Dge, 1 FTIN
 
 − F OUT
L dOL (0+ ) Gpz, 1 HT  + F INL

L
O 1 = O
ACS, 2 Gpz, 2 NL dx
=   L
O 0 − O
ACS, 1 Gpz, 1 NL
Dpe, 1 cat 1 − ε1
dYj (1− ) dYj (1− )
=0 (j = 1, 2, ..., 6) =0 (j = 1, 2, ..., 6)
dx dx
dOL (1− ) dOL (1− )
=0 =0
dx dx

et al., 2000). To solve these nonlinear algebraic equations bvp4c


algorithm requires an initial guess. The solution of the reactor
model depends on the various model parameter values. The values
of these parameters change at each iteration of the model tuning
step while searching for the optimal values and, thus, at each itera-
tion a different initial guess for the solution is required. Moreover,
it is observed that bvp4c is not able to solve the reactor model with
randomly chosen initial guess values. Thus, the model simulation
as a BVP is found to be inappropriate in the iterative model tuning
step of the data fitting exercise. The IVP code does not need any
initial guess of the model solution and is also found to be almost 50
times faster as compared to the BVP code in terms of the CPU time
for a given tolerance. Thus, the model simulation as an IVP is used
in the iterative model tuning step of the data fitting exercise. This
approach also has its own limitations. Some approximations are
required because of the backward simulation which are discussed
in the next section.

Fig. 3. Profile of FBu (x) obtained from the model simulation as an IVP and a BVP.
4.2.2. Extrapolation required in the model simulation as an IVP in
® model tuning exercise
MATLAB is used to simulate the model as an IVP. The shoot-
The backward simulation of the model as an IVP leads to an
ing method is inappropriate for simulating such unstable ODEs
opposite trend in the variation of the mole fractions of the differ-
as BVP (Shampine et al., 2000). Nonetheless, the algorithm bvp4c
® ent components compared to the real trends observed with the
of MATLAB , a collocation based method, is used to simulate the
forward flow of the gas and solid phases. The mole fractions of
model as a BVP. The elemental profiles for all the four elements
the components acting as the reactants (Bu, O2 and OL ) increase
involved in the system, C, H, O and N, are observed to be constant
with the backward simulation as an IVP. On the contrary the mole
in the model simulation both as an IVP and as a BVP. The constant
fractions of the components acting as the products (MA, H2 O, CO2
elemental profiles validate the model equations and the conditions
and CO) decrease. The extent of the decrease depends on the val-
applied at the different interfaces given in Table 8. The simulated
ues of the model parameters. In model tuning step these parameter
BVP and IVP results in the form of profiles for FjModel (x) are observed
values keep changing. The reactor model with a given set of param-
to be consistent in the presence of the appropriate conditions. The eter values sometimes over predict and sometimes under predict
appropriate conditions, here, are the required boundary conditions the extent of different reactions. The over prediction of the reac-
for the BVP simulations obtained as a result from the IVP simula- tion extents leads to a faster decrease in the mole fractions of the
tion and vice versa. The same consistent profiles for FjModel (x) are products in the backward simulation of the model as an IVP in com-
not observed when two different sets of 2NGC + 3 conditions, given parison to the experimental data. This leads to zero values for the
in the two columns of Table 9, are used for the model simulation mole fractions of some products before reaching the inlet of the
as an IVP and as a BVP as shown in Fig. 3 for j = Bu. This is because reactor model. It is noted that if the model is simulated in the for-
OUT, Exp IN, Exp
different sets of information (Fj and Fj ) are used for sim- ward direction and that results in complete consumption of any
ulating the model as an IVP and as a BVP. It is also observed that the of the reactants, the corresponding reaction rates automatically
profiles obtained from both the simulations are under predicting becomes zero and the concentration of the given reactant remains
the corresponding experimental values but to different extents. zero thereafter. In the backward simulation of the model if the mole
The model simulation both as a BVP and an IVP have their own fraction of any of the product becomes zero, it will not remain zero
limitations, which makes both the methods suitable for different automatically for the remaining part of the reactor. This is because
steps of the data fitting exercise. In the model simulation as a BVP the reactants are still going to be present in the simulation of the
®
the collocation based bvp4c algorithm of MATLAB converts the reactor model. The specific rate expressions based on these reac-
system of ODEs into a system of nonlinear algebraic equations at tant concentration values forces the simulation of reactor model to
different mesh points between the two boundary points (Shampine further decrease the amount of such products. The mole fraction
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 209

the initial guess of the model solution for the BVP simulation. The
model validation exercise is done by simulating the model as a BVP
only with the obtained optimal model parameter values.

4.3. Model parameters for the model tuning exercise

The model parameters in the MA CFB reactor model involve sev-


eral kinetic and hydrodynamic parameters. The kinetic parameters
are the pre-exponential factors (kk0 and Keq, 0
k"
; k = 1, 2,. . ., NRxn and
k¨ = 1: Bu, 2: O2 , 3: MA), the activation energies (EA, k ; k = 1, 2,. . .,
NRxn ) and the reaction enthalpies (Heq,k ; k = 1: Bu, 2: O2 , 3: MA)
as given in Eqs. (2) and (3). The reported values of these parameters
(Gascon et al., 2006) are given in Table 4. It is reported (Gascon et al.,
2006) that these values are obtained using commercial VPO catalyst
produced by Amoco in a fluidized bed reactor. On the other hand the
experimental data used for the data fitting exercise are generated
using the VPO catalyst produced by DuPont (commercially named
Fig. 4. Profile of FMA (x)simulating the same model as an IVP and a BVP. The extrap-
as VPP catalyst) in both the PP and CP MA CFB reactors (Patience and
olation needed in IVP simulation when the model is over predicting the formation Bockrath, 2010). Thus, the number of catalyst sites (per weight of
of MA. the catalyst) on which the specific reactions are taking place might
be different that may lead to differences in the pre-exponential fac-
tors. Thus, kk0 (k = 1, 2,. . ., NRxn ) and Keq,
0
k"
(k = 1: O2 , 2: Bu) are used
of the products can be set to zero in the reactor model once they
as the tuning parameters while the values of EA, k’ (k = 1, 2,. . ., NRxn )
reach to zero value, but the model will not be able to penalize the
and Heq, k" (k = 1: O2 , 2: Bu) are used as reported (Gascon et al.,
over predictability of the reaction rates properly as it will give a fix 0
2006) in the proposed reactor model. Keq, and Heq, k = 3:MA
value of the error irrespective of the extent of the over predictability k"=3:MA
of the reaction rates. Thus, the tuned model obtained from such a are also used as two more tuning parameters in the absence of any
model tuning exercise will have an inherent bias for the given prod- reported values.
uct. Thus, to penalize the over predictability similar to penalizing The hydrodynamic parameters include a total of five parameters
the under predictability of the reaction rates, the mole fractions each for both the transport bed and the riser. Three of these param-
of the products are kept decreasing in the model, even after they eters, ai , ci and di are used in the correlations for the void fractions
reach to zero value. This is just a mathematical exercise used in as given in Eq. (18) and the rest two are the Peclet numbers (Pge,i
the model tuning step to achieve a reactor model with unbiased and Pepe,i ) representing the effective dispersion coefficients in the
error in the prediction of its products and does not represent the gas and solid phases. The hydrodynamic parameters of the MA CFB
real concentration values inside the reactor. These negative values PP reactor, except the two Peclet numbers (Pge,1 and Pepe,1 ) for the
for the products, which are not involved in rate expressions (H2 O, transport bed, are already reported and discussed in Section 3.2.2.
CO2 and CO) as given in Table 3, does not lead to any infeasible
reaction rates. On the contrary, the negative mole fraction of the 4.4. Objective formulation for the model tuning exercise
product MA which also takes part in the series reaction leads to
infeasible negative reaction rates. Thus, the reactor simulation as The kinetic and hydrodynamic parameters in the MA CFB reac-
an IVP is stopped as soon as the mole fraction of the MA reaches to tor model control the predicted reaction extent, k (k = 1, 2, . . .,
zero value at x = xconstraint (YMA (xconstraint ) = 0). As discussed earlier NRxn ), of each reaction. In a MCRS with multiple reactions, k can
both over and under predictions are desired and, thus, calculated to be related to the change in the molar flow rates for each compo-
guide the model tuning exercise towards the optimal solutions. To nents across the reactor Fj ( FjIN − FjIN ; j = 1, 2, . . ., NGC and OL )
achieve this the molar flow rates are extrapolated at the inlet of the and the stoichiometry of the reactions. If the number of com-
reactor by using the slope of all the molar flow rates Fj (xconstraint ) at ponents in a MCRS are more than the number of the reactions
the reactor height xconstraint as shown in Fig. 4 for FMA . It is observed involved, then each Fj can be expressed in terms of independent
from Fig. 4 that FMA achieves the zero value before xconstraint when k (which are less than the number of components). If the number
YMA (xconstraint )becomes zero. This is because of the dispersive term of reactions becomes more than the number of the components, as
of FMA being considered along with the convective term. observed in the proposed kinetic model (Table 3), then each Fj
This constraint in the model simulation as an IVP is even possible can be independently  expressed in terms
 of the overall conversion
with the converged tuned model parameter values. This is because of the reactants, Xj j = Bu, O2 and OL , and the overall selectivity
the error is randomly distributed with positive and negative val- of the products, Sj (j = MA, H2 O, CO2 and CO), as given by Eqs. (38)
ues for different experimental cases for obtaining a statistically and (39), respectively. It is noted that the different definitions of
better prediction. This constraint is not present in the simulation selectivity and yields have been used interchangeably in the open
of the model as a BVP as shown in Fig. 4. This is because the literature (Carberry, 2001). The definition of Sj used in this work is
molar flow rates in the BVP simulation are fixed at the inlet using as defined by Coker, (2001) with Bu as the limiting reactant. The
the boundary conditions. This forces Fj (x = 0) for all the products expressions for the experimental and the model predicted (both as
(MA, H2 O, CO2 and CO) to have either positive or zero values. This an IVP and a BVP) Xj and Sj are given in Table 10.
ensures that in the BVP simulation the Fj (x) for products cannot
achieve negative values as the reactions lead to further increase in FjIN − FjOUT  
Fj (x > 0; j : products). Thus, even though the model tuning exercise Xj = j = Bu, O2 andOL (38)
FjIN
is performed by the IVP simulations, as it does not need any initial
guess of the model solution and also takes significantly less time FjOUT − FjIN
than the BVP simulations, the final analysis of the tuned model is Sj = IN − F OUT
(j = MA, H2 O, CO2 andCO) (39)
done using the BVP simulations. Here, the IVP simulation provides FBu Bu
210 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Table 10
Expressions for experimental and model (for simulation as IVP and BVP) predicted Xj and Sj .

Experimental Model (IVP) Model (BVP)


IN, Exp OUT, Exp OUT, Exp IN, Exp
Fj − Fj FjIN, Model − Fj Fj − FjOUT, Model
Exp
Xj = IN, Exp
XjModel = XjModel = IN, Exp
Fj FjIN, Model Fj
OUT, Exp IN, Exp OUT, Exp IN, Exp
Fj − Fj Fj − FjIN, Model FjOUT, Model − Fj
Exp
Sj = IN, Exp OUT, Exp
SjModel = OUT, Exp
SjModel = IN, Exp
IN, Model OUT, Model
FBu − FBu FBu − FBu FBu − FBu

It is noted that the expressions for experimental values of Xj and of 300 generations and subsequent runs of 150 generations till
Exp Exp the convergence is reached with less than 1% improvement in the
Sj (Xj and Sj ) are the same for IVP or BVP model simulations
objective values. In every GA run of 150 generations 35% of the
but the expressions for the model values of Xj and Sj (XjModel and
population obtained from the previous GA run is kept and the rest
SjModel ) are different. This is depicted by the two different profiles 65% is randomly generated. The lower and upper bounds for each
for Fj (x) obtained from the IVP and BVP model simulations as shown of the 12 kinetic parameters are kept as 0.1 times to 10 times of
in Fig. 3. The input of the model simulations as an IVP and a BVP are the reported values for the first 300 generations. The bounds are
OUT, Exp IN, Exp
Fj and Fj , whereas the output of the model simulations as relaxed in every 150 generations by dividing and multiplying by
an IVP and a BVP are FjIN, Model and FjOUT, Model , respectively. Thus, the 3 in the minimum and maximum values of each kinetic param-
eter values obtained from the previous GA run, respectively. The
calculated model values of Xj,u and Sj,u (XjModel and SjModel ) given in
lower and upper bounds for the two hydrodynamic parameters,
Table 10 are different in the model simulations as an IVP and a BVP. Pge and Pepe , for the transport bed of the PP reactor operating in the
The percentage error, Ej,u , between the experimental value and the C-TFB regime are kept as 5 and 50. This is based on the reported
model value of Xj,u and Sj,u for a given component, j, and experiment, values (Bi, 2004) for the risers operating with the Gpz values less
u, is calculated using Eq. (40). It is noted that Ej,u is not defined than 200 kg/m2 s. The lower bound for the Peclet numbers for the
for the inert component (N2 ) as it will always remain zero. This CP reactor is decreased to 1 from 5 based on the square root ratio
makes the total number of responses, NResp , to be equal to NGC with of the diameters of the PP and CP reactors (transport bed as well
NGC − 1 responses in the gas phase and one response in the solid as riser) as discussed in Section 3.2.4. This makes the lower and
phase. The objective for the model tuning, IError , is defined as the upper bounds for Pge and Pepe for the transport bed and the riser
average of the absolute values of Ej,u for all the reactive components of the CP reactor as 1 and 50. The lower and upper bounds for the
and for all the experimental cases of the training set, as given by other hydrodynamic parameters (ai , ci and di ; i = 1: transport bed,
Eq. (41). It is noted that in a given experiment, u, Ej,u value for one i = 2: riser) are appropriately relaxed with each GA run based on the
component (say, j = J) is usually linked with Ej,u values of the other minimum and maximum values obtained in the previous run.
components (j = / J). Nonetheless, for any component Ej,u value of
one experiment (say, u = U) is completely independent to that of the
other experiment (u = / U). This leads to a possibility that the model 4.5. Procedure for the data fitting exercise
tuning done with the objective, IError , may result in Ej,u values for a
given component averaged over all the experiments predominantly The kinetic parameters are considered same for the PP and CP
positive or negative. This average value for a given component, j, reactors whereas the hydrodynamic parameters are assumed to be
is proposed and defined as Biasj , given by Eq. (42). The target Biasj different. Thus, a step by step data fitting exercise is proposed. In
value for any component is zero for a good model prediction. Hence, the Step 1 data fitting exercise, the PP data is used for the data
to achieve this target, a second objective, IBias , is defined in terms fitting exercise of the PP reactor model. In the Step 2 data fitting
of the Biasj as given in Eq. (43). Thus, the model tuning exercise exercise, the CP data is used for the data fitting exercise of the
is done with these two objectives using a multi-objective genetic CP reactor model. Most of the kinetic (Gascon et al., 2006) and
®
algorithm (GA) implemented in MATLAB (gamultiobj). hydrodynamic parameters (Bi, 2004; Patience and Bockrath, 2009)
⎡ ⎤ for the PP reactor are reported except the two kinetic parameters
Exp Model
Xj,u − Xj,u   0
(Keq, k"=3:MA
and Heq,k¨ = 3:MA ) and the two hydrodynamic param-
⎢ 100 × j = Bu, O2 andOL ⎥
⎢ Xj
Exp
⎥ eters (Pge,1 and Pepe,1 ). Two data fitting exercises as the Step 1a
Ej,u (%) = ⎢
⎢ Exp
⎥ (40)

and the Step 1b are performed. The Step 1a is performed on the
Model
Sj,u − Sj,u
⎣ ⎦ PP data by relaxing the 12 reported kinetic parameters (as men-
100 × Exp
(j = MA, H2 O, CO2 andCO) tioned in Section 4.3) along with the two unknown hydrodynamic
Sj,u
parameters as the tuning parameters. The Step 1b is performed on
⎡ ⎤ the PP data by relaxing all the 10 reported hydrodynamic param-
1 
NResp NExp
IError = Min ⎣ |Ej, u (%) |⎦ (41)
eters (as mentioned in Section 4.3) along with the two unknown
NResp × NExp kinetic parameters as the tuning parameters. The tuned values of
j=1 u=1
the kinetic parameters obtained from the Step 1 data fitting exer-
cise on the PP data are then used without any further tuning in
1 
NExp
the Step 2 data fitting exercise on the CP data. The hydrodynamic
Biasj = Ej, u (%) (42)
NExp parameters used in the Step 2 data fitting exercise on the CP data
u=1
⎡ ⎤ are all the hydrodynamic parameters (ai , ci , di , Pge,i and Pepe,i ; i = 1:

NResp transport bed, i = 2: riser). Thus, in summary, in the proposed step
1
IBias = Min ⎣ |Biasj |⎦ (43) by step data fitting exercise, the kinetic parameters of the CP reactor
NResp model are tuned from the PP reactor data on the PP reactor model.
j=1
It is observed that the reported conversion and selectivity data
The default values of the GA parameters are used in gamulti- trends in the PP and CP reactors are different. Hence, it is envis-
obj. The population size of 10× number of tuning parameters is aged that the step by step data fitting exercise though logical may
used. Several GA optimization runs are made with the first run not be appropriate to fit all the data. Thus, another fitting exercise
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 211

Fig. 5. Residual plots of the model values as compared to the experimental values for the commercial plant; 䊐: XBu ; O:XO2 ; : XL ; y: SMA ; : SCO2 ; : SCO ; : SH2 O (blank:
O
training set; filled: validation set) (a) Model values of PP reactor without any data fitting exercise on PP reactor data (b) Model values obtained from step by step data fitting
exercise on PP reactor data (c) Model values obtained from simultaneous data fitting exercise on PP and CP reactor data combined.

is performed as simultaneous data fitting exercise. This is done by ues of the Biasj for all the reactive components (j = Bu, O2 , OL , MA,
fitting both the PP and CP data simultaneously on the PP and CP CO2 , CO and H2 O), IError and IBias without any data fitting exercise
reactor models. The kinetic parameters are kept the same for the are shown in Figs. 6a and 7a, respectively.
PP and CP reactor models whereas the hydrodynamic parameters The first observation made from Figs. 5a, 6a and 7a is that the
are kept different for the two reactor models. In the end, the pre- model is behaving similar on the training and validation sets. This
diction capability of CP reactor models obtained from both the step validates the division of the experimental data in the two sets and,
by step and the simultaneous data fitting exercise are compare using thus, the model validation exercise based on such a partitioning
the validation set of experimental data. can be accepted. The second observation made from these figures
is that the model values are quite different from the experimental
values without any data fitting exercise. The Ej, u values are quite
5. Results and discussion high (Fig. 5a), which makes the IError values of the order of 50. The
third observation made from Fig. 5a is that the Ej,u values for any
The MA CFB PP reactor model is first simulated as a BVP that given component are either dominantly positive or negative for all
does not involve any approximations using the reported kinetic the experimental cases. This observation results in significant Biasj
and hydrodynamic parameters values both on the training and val- values as shown in Fig. 6a. Positive Biasj values for all the reactants
idation sets to get reference values for Ej,u , Biasj , IError and IBias . The j = Bu and O2 and OL in Fig. 6a represent that the model values of Xj
values of the two unknown hydrodynamic parameters, Pge,1 and for all the reactants are being under predicted in comparison to the
Pepe,1 are taken as the reported values of Pge,2 and Pepe,2 . The values experimental values. Similarly, it can be said that model values of
for the two unknown kinetic parameters, Keq, 673 and Heq,k¨ = 3 are
k"=3 Sj for j = MA and CO are being under predicted while for j = CO2 are
taken as the reported values of Keq,673 and Heq,k¨ = 2 . Thus, the Ej,u being over predicted in comparison to the experimental values. It is
k"=2
values obtained without any data fitting exercise are plotted against noted that Ej,u values for j = H2 O are quite low. This can be attributed
Exp Exp to the fact that the amount of H2 O produced in the conversion of
the experimental values of Xj and Sj in Fig. 5a. This figure rep-
resents a comparison between the model predicted values and the Bu to MA, CO2 or CO is almost same. The maximum value of the
reported experimental values. It is noted that multiplication factors IBias is equal to IError . This equality of objectives happens when Ej,u
(0.2 for SH2 O and 0.66 for SCO2 and SCO ) are used in the experimental for any component j remain positively or negatively biased for all
values plotted on the horizontal axis to shift the points horizontally. experimental cases (u = 1, 2, ..., NExp ) as observed in Fig. 5a. Here, the
This exercise is done for a better visual analysis. The reference val- significant Biasj values makes IBias very close to the value of IError .
212 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Fig. 6. Biasj for different component and for both training and validation sets of MA CFB PP reactor. (a) obtained from PP reactor model without any data fitting exercise (b)
obtained from PP reactor model using step by step data fitting exercise (c) obtained from PP reactor model using simultaneous data fitting exercise.

Fig. 7. Results of the data fitting exercise in terms of the objective values (IError and IBias ) calculated on both the training set (T. S.) and the validation set (V. S.) and by simulating
the model as a BVP. (a) Results for the PP data (b) Results for the CP data. First two bars in both the figures represent objective values without any data fitting exercise. Next
four bars for the PP data in (a) and two bars for the CP data in (b) represent objective values for the PP and CP reactor models obtained from the step by step data fitting
exercise. Last two bars in both figures represent objective values for the PP and CP reactor models obtained from the simultaneous data fitting exercise.

These observations suggest the need for the data fitting exercise shown as the first bars of Fig. 8a and b, respectively. It is observed
of the model. Since the model tuning exercise is performed by simu- by comparing the values from Figs. 7a and 8a that the objective val-
lating the model as an IVP, the reference values of the two objectives ues for the BVP and IVP simulations are different. This is attributed
(IError and IBias ) are also calculated by simulating the model as an IVP to the previous observation of different profiles Fj (x) for the BVP
using the base values of the model tuning parameters. The values and IVP simulations shown in Fig. 3 and as discussed in Section 4.4.
of the two objectives for both the training and validation sets are
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 213

Fig. 8. Evolution of tuning exercise as IError and IBias values calculated between the experimental data (training and validation) and the model simulation values (IVP and
BVP) with the tuned parameters corresponding to the selected point from different Pareto sets converged after each GA run. Here, the number of generations represent the
cumulative sum after each GA run.

which to make this decision, the solution closest to the origin in the
Pareto set is chosen as the final solution of the model tuning exer-
cise in this work. The normalized objective values between 0 and 1
are used in the final selection. The decision variables corresponding
to the chosen solution represent the tuned model parameters. The
model obtained with these model parameters is called the tuned
reactor model. The tuned reactor model after each GA run are sim-
ulated both as an IVP and a BVP to calculate the values of the IError
and IBias both on the training and validation datasets.
The first observation with the convergence of the model tun-
ing exercise is made based on the different Pareto sets shown in
Fig. 9 and the objective values given in Fig. 8a corresponding to the
selected solutions from the Pareto sets. It is observed that both the
objective values decrease till 900 generations (5th iteration of the
GA runs) and after that they are almost converged. The tuned model
parameters are also observed to be converged after 900 genera-
tions with less than 1% variation after these generations. The second
observation is made on the guidance capability of the model tun-
Fig. 9. Evolution of tuning exercise as the converging Pareto sets with each iteration
of optimization runs. Here, the number of generations represent the cumulative sum ing exercise. It is clear from Fig. 8a and b that the improvement in
after each GA run. terms of the two objective values on the training set with the num-
ber of generations is properly reflected on the validation set too.
This observation is further verified by comparing Fig. 8a–d for the
The evolution of the model tuning exercise with the number of simulation of the model as an IVP and a BVP where similar trends
GA runs is shown in Fig. 9 for the Step 1a data fitting exercise. The are also observed. The use of an IVP simulation in the tuning exer-
number of generations shown in Fig. 9 represent the cumulative cise is validated based on these observations and the fact that an
number of the generations after each GA run. The results are shown IVP simulation takes significantly less computational time than the
in the form of different converged Pareto sets after each GA run. A BVP simulation. Though, it is noted that the BVP simulation is more
Pareto set represents a set of non-dominated solutions. Each solu- accurate as it does not require any extrapolation while calculating
tion of the Pareto set is non-dominated to each other because the IError and IBias . Thus, the two objective values using BVP simulations
improvement in any one of the objective (say, IError ) is only possi- are used to further analyze the final converged results. The final
ble at the cost of worsening the other objective (say, IBias ). Usually a converged results of Step 1a and Step 1b of the data fitting exercise
unique solution is required to obtain a final set of the tuned model on the PP data both for the training and validation data sets are
parameter values which represent the minimum IError as well as shown in Fig. 7a. The IBias values of the order of IError signifies that
minimum IBias . In the absence of any process design constraints on
214 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Fig. 10. Residual plots of the model values as compared to the experimental values for the commercial plant; 䊐: XBu ; O: XO2 ; : XL ; y: SMA ; : SCO2 ; : SCO ; :SH2 O (blank:
O
training set; filled: validation set) (a) Ej, u (%) values without any data fitting exercise on CP reactor data (b) Converged Ej, u (%) values obtained from step by step data fitting
exercise on CP reactor data (c) Converged Ej, u (%) values obtained from simultaneous data fitting exercise on PP and CP reactor data combined.

the model with tuned parameters from Step 1b data fitting exercise tors (0.2 for SH2 O and 0.66 for SCO2 and SCO ). The simulated values
still gives biased predications for most of the components. The Step of the Biasj values and the two objectives (IBias and IError ) without
1a of PP data fitting exercise with twelve kinetic parameters and any data fitting exercise on the CP data are given in Figs. 11a and
two unknown hydrodynamic parameters, reduces both the objec- 7b, respectively, for both the training and validation sets.
tive values significantly. This justifies the use of kinetic parameters It is observed in Fig. 10a that the Ej,u values are significantly
for the model tuning in the Step 1a data fitting exercise on PP data. high and unevenly distributed with either positive or negative val-
The converged values of Ej,u and Biasj are also given in Figs. 5b and ues for different components. This is also clear in Fig. 11a as the
6b. The comparison of Fig. 5a and b shows that the values of Ej,u are Biasj values of all the three reactants (j = Bu, O2 and OL ) are negative
significantly decreased and are also evenly distributed with posi- while for products it is negative for MA and positive for CO2 and
tive and negative values after the (Step 1a) data fitting exercise. The CO. The opposite trends in the Biasj values are observed in Figs. 6a
same can be observed by comparing Fig. 6a and b as the Biasj val- and 10a. The PP reactor model with the use of the reported kinetic
ues of each component significantly decreased after the data fitting parameters for the PP data was under predicting the reactant con-
exercise. versions which is addressed with the use of the tuned parameters.
A similar exercise is done on the CP data for getting the refer- It is observed that with the tuned parameter values the CP reac-
ence values before performing Step 2 data fitting exercise, as done tor model now over predicts the reactant conversions as shown
on the PP data before the Step 1 data fitting exercise. This is done by in Fig. 9. It is noted that the hydrodynamic parameters of the PP
simulating the CP reactor model as a BVP on the training and val- reactor are used for the CP reactor and could be the reason for this
idation sets with the tuned kinetic parameters obtained from the observation. This validates the Step 2 data fitting exercise on the CP
Step 1a data fitting exercise. Since the analysis of the tuned model data by tuning all the 10 hydrodynamic parameters as mentioned
is done by simulating it as a BVP, the reference values are obtained in Section 4.5.
by simulating the model as BVP only. The hydrodynamic param- The results of the step by step data fitting exercise (Step 2) for the
eters are also kept the same (as obtained for PP reactor model). CP data in terms of Ej,u , Biasj and the two objective values calculated
Ej,u values for the CP reactor without any data fitting exercise are by the BVP simulation of the converged CP reactor model are shown
Exp Exp in Figs. 10b, 11b and 7b, respectively. It is observed from Fig. 10b
plotted against the experimental values of Xj and Sj in Fig. 10a
for both the training and validation sets. A similar exercise of hori- that even though the Ej,u values decreased after Step 2 data fitting
zontally shifting the points for better visual analysis is implemented exercise but the values are either dominantly positive or negative
in this figure as done in Fig. 5 by using different multiplication fac- for any given component. This is shown in Fig. 11b with significant
positive and negative Biasj values. This makes the IBias values of the
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 215

Fig. 11. Biasj for different components and for both the training and validation sets of the MA CFB CP reactor. (a) obtained from the CP reactor model without any data fitting
exercise (b) obtained from the CP reactor model using the step by step data fitting exercise (c) obtained from the CP reactor model using the simultaneous data fitting exercise.

order of the IError values as shown by the third and fourth bars in in Fig. 10c in comparison to the Biasj values shown in Fig. 10b. A
Fig. 7b. Thus, the reactor model for the CP reactor obtained from similar observation is made in terms of the two objective values
the step by step data fitting exercise shows significant bias in the (IError and IBias ). The comparison between the simultaneous and the
form of over prediction for MA and under prediction for CO2 and step by step data fitting exercise in Fig. 7 shows significant improve-
CO. The CP reactor tuned model with such inherent biases is not ment in the objective value, IBias , for the commercial plant (Fig. 7b)
appropriate for optimization studies. at the cost of not so significant worsening of the objective values
The kinetic parameters tuned from the PP data fitting exercise for the pilot plant (Fig. 7a). It is also noted that the model obtained
are unable to fit the CP reactor data as evident from the previous dis- from the simultaneous data fitting exercise shows better IBias values
cussion. This validates the simultaneous data fitting exercise using for the validation set of the PP reactor data. Thus, the simultane-
the PP and CP reactor data simultaneously on the respective mod- ous data fitting exercise results in a CP reactor model with better
els as discussed in Section 4.5. The Ej,u , Biasj and the two objective prediction capability in terms of any inherent bias for any compo-
(IError and IBias ) values for the PP data calculated using the PP reac- nent. In summary, the tuned PP and CP reactor models from the
tor model obtained from the simultaneous data fitting exercise are simultaneous data fitting exercise with lower IBias values represent
shown in Figs. 5c, 6c and the last two bars of Fig. 7a, respectively. better prediction capability models than the models obtained from
Similarly, the Ej,u , Biasj and the two objective (IError and IBias ) values step by step data fitting exercise.The tuned kinetic parameter values
for the CP data calculated using CP reactor model obtained from from the simultaneous data fitting exercise are given in Table 4 and
the simultaneous data fitting exercise are shown in Figs. 10c, 11c the hydrodynamic parameters are given in Table 11. It is observed
and the last two bars of Fig. 7b, respectively. It is observed that the that the tuned kinetic parameters are different than the reported
PP reactor models obtained from the step by step and the simulta- values. This is attributed to the different catalyst used in the study
neous data fitting exercise show similar behavior in terms of the for the reported kinetic model development than that used in the
Ej,u and Biasj values. The CP reactor models obtained from the two PP and CP MA CFB reactors as discussed in Section 4.3. Ballarini
different data fitting exercises show quite different behavior. It is et al. (2006) in their review reported that the activity of the VPO
observed by comparing Fig. 10b and c that the CP reactor model catalyst is affected by the preparation method. It is observed from
obtained from the simultaneous data fitting exercise gives the Ej,u Table 4 that the rate constants for the reactions involving the for-
values evenly distributed with both positive and negative values, mation and consumption of OS sites are predicted to be higher than
unlike the CP reactor model obtained from the step by step data fit- the reported values. On the other hand, the rate constants for the
ting exercise. This results in much lower Biasj values as observed reactions happening on the O2 sites are predicted to be less than
216 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Table 11
Tuned values (simultaneous data fitting) of the hydrodynamic parameters of the PP and CP MA CFB reactor model.

Hydrodynamic Parameter Pege, 1 Pepe, 1 Pege, 2 Pepe, 2 a1 c1 d1 a2 c2 d2

MA CFB reactor PP CP PP CP CP CP CP CP CP CP CP CP

Final tuned values 47.4 1.06 20.7 6.09 1.52 10.73 1.86 0.77 1.16 0.69 0.22 −1.79

the reported values. Moreover, the tuned values of Keq, 0 and Hk" Several assumptions are made during the model development
k"

values for k = 3: MA adsorption – desorption equilibrium are of the to keep it simple. The simplification of the expression for pressure
order of the tuned values for k = 1: O2 and 2: Bu adsorption – des- drop from Eqs. (11) to (12) is one of the such assumptions. The
0
orption equilibriums. This justifies the use of Keq, and Hk" terms final tuned model is simulated with the pressure drop including
k"
in the modified kinetic model. all the frictional and acceleration term as given in Eq. (11). It is
The tuned hydrodynamic parameters are given in Table 11. It is observed that the change in the two objective values (IError and IBias )
again noted that only two hydrodynamic parameters are tuned for are less than 1% in the training and validation sets. This validates the
the PP reactor whereas ten hydrodynamic parameters are tuned for simplification for the pressure drop expression. Negligible external
the CP reactor. It is observed that the tuned value of Pege,1 for the and internal mass transfer resistances in the gas film and within
PP reactor is near to its upper bound on the other hand the same the microspores of catalyst particles, respectively, are the other
for the CP reactor is on its lower bound. The tuned value of Pepe,1 assumptions made in the model. Mears criterion (Mears, 1971) and
for the PP reactor is also greater than the same obtained for the CP Weisz-Prater criterion (Weisz and Prater, 1954) are generally used
reactor, but both the values are well within the bounds (1–50: CP for checking the validity of these two assumptions, respectively.
reactor, as discussed in Section 4.4). This difference is attributed These criteria need the experimental rate data. The validity of these
to the fact that the data fitting exercise tries to fit the differences two assumptions in absence of such rate data is made by simulat-
observed in the PP and CP reactor data with the same kinetic and ing the final tuned model with and without these assumptions. The
different hydrodynamic parameters as discussed before. The values Rj (x) for different components considering the two mass transfer
of the Peclet numbers represent high gas and solid mixing in the resistances are calculated from Eq. (44) in the place of Eq. (4).
transport bed of the CP reactor than the PP reactor. The high gas
mixing in the riser of the CP reactor is also observed from the tuned 1 
2
value of Pege,2 for the CP reactor. The solid mixing in the riser of the Rj (x) = 3 rcat Rj, rcat (x) drcat (44)
CP reactor is observed to be similar to that of the PP reactor as the
rcat =0
tuned value of Pepe,2 for the CP reactor is near to the reported value
(Bi, 2004) for the PP reactor. The tuned values of Pege,1 and Pege,2 for
Rj, rcat (x) in Eq. (44) represents the rate for component j within the
the CP reactor are at their lower bound. The tuned values of Pege,1
catalyst (present at reactor height, x) at the radius rcat . A similar
and Pege,2 being at their lower bounds represent a significant gas
mixing in the CP reactor. A sensitivity analysis is done by further approach is followed to calculate Rj, rcat (x) implementing the exter-
decreasing these values (one by one and together). It is observed nal and internal mass transfer resistances, as reported (Chaudhari
that 50% reduction of these values resulted in less than 1% change and Gupta, 2012) in the model of the MTFBR. The correlation pro-
in the two objective values. The lower tuned values of a1 and a2 for posed by Gunn (1978) is used for the mass transfer coefficient in the
the CP reactor in comparison to the reported values for PP reactor CFB reactor as suggested by Breault (2006). No significant change
signify lower drift velocity Vg,i (Eq. (18c)) in the CP reactor than is observed in the value of Rj (x) calculated from Eq. (4) or Eq. (44).
that in the PP reactor. The lower tuned values of c1 and c2 along On the other hand, the simulation of the model considering these
with the tuned values of d1 and d2 for the CP reactor in comparison mass transfer resistances leads to significant increase (more than
to the reported values for the PP reactor signify lower distribution 100 times) in the computational time. This validates the model
coefficient C0,i (Eq. (18)) in the CP reactor than in the PP reactor. assumption of negligible external or internal mass transfer resis-
This means the radial distribution of the solid void fraction and the tances for the catalyst particles. This further validates the use of
axial gas velocity is more uniform in the CP reactor than that in the kinetic parameter tuning exercise of the PP and CP reactor data.
PP reactor. One of the salient feature of the proposed model is the consider-
ation of variable gas density in the form of axial variation of CT (x)

Fig. 12. Axial profile of (a) CT (x) and (b) U0 (x) simulated using the tuned PP and CP MA CFB reactor models.
P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218 217

and U0 (x). The axial profiles of CT (x) and U0 (x) predicted by the for the model tuning exercise than the BVP in absence of the proper
tuned MA CFB PP and CP reactor models are shown in Fig. 12 for initial guess for the model solution. The final analysis of the tuned
one of the given experimental conditions of the PP and CP each. It model is done using the model simulation as a BVP. The use of the
is noted that the experimental conditions of the PP and CP reac- two objectives, IError and IBias , is proposed for the model tuning exer-
tors are different, which is also observed by two different profiles cise of a MCRS. The kinetic parameters are assumed to be same for
of CT (x) and U0 (x) predicted by the tuned PP and CP reactor mod- both the PP and CP reactors whereas the hydrodynamic parameters
els. It is observed from Fig. 12a that the decrease in CT (x) is of the are assumed to be different for the two reactors. Two data fitting
order of 10% in the transport bed and the riser of the PP and CP exercises are performed, namely, the step by step and the simultane-
reactors. Similarly, it is observed from Fig. 12b that the increase in ous data fitting exercises. It is observed from the model validation
U0 (x) is of the order of 30% in the transport bed and 15% in the exercise that the tuned PP and CP reactor models obtained from
riser of the PP and CP reactors. The sudden increase in U0 (x) at the simultaneous data fitting exercise show better prediction capa-
xINT is because of the sudden decrease in the cross-sectional area bility in comparison to the models obtained from the step by step
between the transport bed and the riser. The linear variation of data fitting exercise. Similar results obtained in all the model tun-
CT (x) shows that the first term in the expression for dCT (x) /dx ing and validation exercises also validate the use of the proposed
in Table 6 (due to pressure drop) is the dominant term. The con- two-objectives (IError and IBias ) minimization for modeling the MA
stant bed void fraction assumption results in a constant pressure CFB reactors.
drop and a constant dCT (x) /dx ignoring the non-significant second
term in the expression for dCT (x) /dx in Table 6 (due to temperature
Funding
drop). Two terms lead to the axial variation in the U0 (x), the pres-
sure drop and the overall reaction rate RT (x), as discussed in Section
This research did not receive any specific grant from funding
3.2.1. It is observed from the simulations of the tuned model that
agencies in the public, commercial, or not-for-profit sectors.
the contribution of the reaction rate in the axial variation of U0 (x)
is of the order of 10% and the remaining contribution is due to the
pressure drop. This is the reason why the profile of U0 (x) is observed References
to be almost linear in Fig. 12b. Thus, consideration of axial varia-
Abba, I.A., Grace, J.R., Bi, H.T., 2002. Variable-gas-density fluidized bed reactor
tions in CT (x) and U0 (x) in the model is validated from the tuned model for catalytic processes. Chem. Eng. Sci. 57, 4797–4807.
model simulations. The assumption of the uniform radial distribu- Abba, I.A., Grace, J.R., Bi, H.T., Thompson, M.L., 2003. Spanning the flow regimes:
tion both for the gas velocity and the solid hold up can be relaxed generic fluidized-bed reactor model. AIChE J. 49, 1838–1848.
Andreux, R., Petit, G., Hemati, M., Simonin, O., 2008. Hydrodynamic and solid
by considering radial dispersion along with the axial dispersion as
residence time distribution in a circulating fluidized bed: experimental and 3D
suggested in the literature (Liu, 2001). The radial and axial disper- computational study. Chem. Eng. Process. Process Intensif. 47, 463–473.
sions in the reactor model result in partial differential equations Avidan, A.A., Yerushalmi, J., 1982. Bed expansion in high velocity fluidization.
(PDE). The same is attempted by Abba et al. (2003). A PDE model Powder Technol. 32, 223–232.
Bai, D., Kato, K., 1995. Saturation carrying capacity of gas and flow regimes in CFB.
may represent the real reactor better than the proposed model but J. Chem. Eng. Japan 28, 179–185.
with significant increase in the computational time. Thus, with the Bai, D., Zhu, J.-X., Jin, Y., Yu, Z., 1995. Internal recirculation flow structure in vertical
given prediction capability of IError under 17% and IBias under 6%, the upflow gas-solids suspensions Part I. A core-annulus model. Powder Technol.
85, 171–177.
use of proposed simple MA CFB reactor model is recommended. Bai, D., Issangya, A.S., Zhu, J.X., Grace, J.R., 1997. Analysis of the overall pressure
balance around a high-density circulating fluidized bed. Ind. Eng. Chem. Res.
36, 3898–3903.
Ballarini, N., Cavani, F., Cortelli, C., Ligi, S., Pierelli, F., Trifirò, F., Fumagalli, C.,
6. Conclusions Mazzoni, G., Monti, T., 2006. VPO catalyst for n-butane oxidation to maleic
anhydride: a goal achieved, or a still open challenge? Top. Catal. 38, 147–156.
A simple MA CFB reactor model is proposed in this work. The Berruti, F., Chaouki, J., Godfroy, L., Pugsley, T.S., Patience, G.S., 1995. Hydrodynamics
of circulating fluidized bed risers: a review. Can. J. Chem. Eng. 73, 579–602.
transport bed and the riser of MA CFB reactor are modeled assum-
Bi, H.T., Fan, J., 1991. Regime transitions in gas-solid circulating fluidized beds. In:
ing ideal operation of the regenerator at the steady state operation AIChE Annual Meeting, Los Angeles, California, pp. 17–22.
with the complete regeneration of the lattice oxygen in each cycle. Bi, H.T., Grace, J.R., 1995. Flow regime diagrams for gas-solid fluidization and
upward transport. Int. J. Multiphase Flow 21 (1229), 1236.
The non-ideal plug flow, both for the gas and solid phases, is mod-
Bi, X.T., 2004. Gas and solid mixing in high-density CFB risers. Int. J. Chem. React.
eled with the help of dispersion coefficients (Dge and Dpe ) for both Eng. 2, 1542–6580.
the phases in the proposed model. The reported kinetics Gascon Breault, R.W., 2006. A review of gas–solid dispersion and mass transfer coefficient
et al. (2006) appropriate for the operating conditions of the MA correlations in circulating fluidized beds. Powder Technol. 163, 9–17.
Carberry, J.J., 2001. Chemical and Catalytic Reaction Engineering. Dover
CFB reactor is suitably modified. The transport bed and the riser of Publications, New York.
both the MA CFB PP and CP reactors are found to be operating in the Chan, C.W., Seville, J.P., Parker, D.J., Baeyens, J., 2010. Particle velocities and their
C-TFB and the DSU regimes, respectively. The appropriate hydro- residence time distribution in the riser of a CFB. Powder Technol. 203, 187–197.
Chaudhari, P., Gupta, S.K., 2012. Multiobjective optimization of a fixed bed maleic
dynamic correlations for the identified operating regimes of the anhydride reactor using an improved biomimetic adaptation of NSGA-II. Ind.
MA CFB reactor are used in the proposed model. The density vari- Eng. Chem. Res. 51, 3279–3294.
ation because of the significant pressure drop in the DSU and the Choong, T.S.Y., Paterson, W.R., Scott, D.M., 1998. Axial dispersion in rich, binary gas
mixtures: model form and boundary conditions. Chem. Eng. Sci. 53,
C-TFB regime is implemented in the dispersion model with a very 4147–4149.
simple approach. The proper conditions are implemented in the Coker, A.K., 2001. Modeling of Chemical Kinetics and Reactor Design, 1st ed. Gulf
proposed reactor model for the configurational complexities of the Professional Publishing, Houston, Texas.
Contractor, R.M., Sleight, A.W., 1987. Maleic anhydride from C-4 feedstocks using
reactor in the form of variation in its cross – section area between
fluidized bed reactors. Catal. Today 1, 587–607.
the transport bed and the riser of the MA CFB reactor. The appropri- Contractor, R.M., Bergna, H.E., Horowitz, H.S., Blackstone, C.M., Malone, B., Torardi,
ate conditions for the addition oxygen fed at the different heights C.C., Griffiths, B., Chowdhry, U., Sleight, A.W., 1987. Butane oxidation to maleic
anhydride over vanadium phosphate catalysts. Catal. Today 1, 49–58.
of the transport bed are also implemented in the proposed model.
Contractor, R.M., 1999. Dupont’s CFB technology for maleic anhydride. Chem. Eng.
The modified Danckwerts boundary conditions for the variable gas Sci. 54, 5627–5632.
density are used in the proposed model. Danckwerts, P.V., 1953. Continuous flow systems: distribution of residence times.
A detailed data fitting exercise is performed on the proposed MA Chem. Eng. Sci. 2, 1–13.
Gascon, J., Valenciano, R., Tellez, C., Herguido, J., Menendez, M., 2006. A generalized
CFB PP and CP reactor models using the reported PP and CP reactor kinetic model for the partial oxidation of n-butane to maleic anhydride under
data. The model simulation as an IVP is found to be more suitable aerobic and anaerobic conditions. Chem. Eng. Sci. 61, 6385–6394.
218 P. Chaudhari, S. Garg / Computers and Chemical Engineering 100 (2017) 198–218

Grace, J.R., Issangya, A.S., Bai, D., Bi, H.T., Zhu, J., 1999. Situating the high-density Qi, M., 2012. Hydrodynamics and Micro Flow Structure of Gas–Solid Circulating
circulating fluidized bed. AIChE J. 45, 2108–2116. Turbulent Fluidized Beds. Doctoral Dissertation. University of Western Ontario.
Gunn, D.J., 1978. Transfer of heat or mass to particles in fixed and fluidised beds. Schlogl, R., 2009. Concepts in selective oxidation of small alkane molecules. In:
Int. J. Heat Mass Transf. 21, 467–476. Mizuno, N. (Ed.), Modern Heterogeneous Oxidation Catalysis: Design,
Gupta, S.K., 1998. Modelling the Hydrodynamics of Large Scale Circulating Reactions and Characterization. Wiley-VCH Verlag GmbH & Co. KGaA,
Fluidized Bed Risers. PhD Thesis. The University of Calgary. Weinheim, Germany, pp. 1–42.
Hutchenson, K.W., La Marca, C., Patience, G.S., Laviolette, J.P., Bockrath, R.E., 2010. Schugerl, K., 1967. Experimental comparison of mixing processes in two- and
Parametric study of n-butane oxidation in a circulating fluidized bed reactor. three-phase fluidized beds. In: Drinkenburg, A. (Ed.), Proceedings of the
Appl. Catal. A: Gen. 376, 91–103. International Symposium on Fluidization. Netherlands University, Amsterdam,
Issangya, A.S., Bai, D., Bi, H.T., Lim, K.S., Zhu, J., Grace, J.R., 1996. Axial solids holdup pp. 782–794.
profiles in a high-density circulating fluidized bed riser. In: CFB5, 5th Shampine, L.F., Kierzenka, J., Reichelt, M.W., 2000. Solving boundary value
International Conferences on Circulating Fluidized Beds, Beijing, pp. 1–7. problems for ordinary differential equations in MATLAB with bvp4c. Tutorial
Issangya, A.S., 1998. Flow Dynamics in High Density Circulaing Fluidized Bed. The Notes, 437–438.
University Of British Columbia. Shekari, A., Patience, G.S., 2013. Transient kinetics of n-butane partial oxidation at
Keraron, M., 2009. Conversion du méthanol en oléfines sur SAPO-34: modélisation elevated pressure. Can. J. Chem. Eng. 91, 291–301.
cinétique et design de réacteur. Doctoral dissertation. École Polytechnique de Shekari, A., Patience, G.S., Bockrath, R.E., 2010. Effect of feed nozzle configuration
Montréal. on n-butane to maleic anhydride yield: from lab scale to commercial. Appl.
Kim, S.W., Kirbas, G., Bi, H., Lim, C.J., Grace, J.R., 2004. Flow behavior and regime Catal. A: Gen. 376, 83–90.
transition in a high-density circulating fluidized bed riser. Chem. Eng. Sci. 59, Wang, D., Barteau, M.A., 2001. Kinetics of butane oxidation by a vanadyl
3955–3963. pyrophosphate catalyst. J. Catal. 197, 17–25.
Liu, J., 2001. Particle and Gas Dynamics of High Density Circulating Fluidized Beds. Wang, D., Barteau, M.A., 2002. Oxidation kinetics of partially reduced vanadyl
Doctoral Dissertation. The University of British Columbia. pyrophosphate catalyst. Appl. Catal. A: Gen. 223, 205–214.
Lorences, M.J., Patience, G.S., Díez, F.V., Coca, J., 2003. Butane oxidation to maleic Wang, D., Barteau, M.A., 2003. Differentiation of active oxygen species for butane
anhydride: kinetic modeling and byproducts. Ind. Eng. Chem. Res. 42, oxidation on vanadyl pyrophosphate. Catal. Lett. 90, 7–11.
6730–6742. Wang, C., Zhu, J., Barghi, S., Li, C., 2014. Axial and radial development of solids
Lorences, M.J., Patience, G.S., Díez, F.V., Coca, J., 2004. Transient n-butane partial holdup in a high flux/density gas–solids circulating fluidized bed. Chem. Eng.
oxidation kinetics over VPO. Appl. Catal. A: Gen. 263, 193–202. Sci. 108, 233–243.
Mears, D.E., 1971. Tests for transport limitations in experimental catalytic reactors. Wei, F., Lin, S., Yang, G., 1993. Gas and solids mixing in a commercial FCC
Ind. Eng. Chem. Process Des. Dev. 10, 541–547. regenerator. Chem. Eng. Technol. 16, 109–113.
Patience, G.S., Bockrath, R.E., 2009. Drift flux modelling of entrained gas–solids Weisz, P.B., Prater, C.D., 1954. Interpretation of measurements in experimental
suspensions. Powder Technol. 190, 415–425. catalysis. Adv. Catal. 6, 143–196.
Patience, G.S., Bockrath, R.E., 2010. Butane oxidation process development in a Xiao-Feng, H., Cheng-Yue, L., Chen, B.H., Silveston, P.L., 2002. Transient kinetics of
circulating fluidized bed. Appl. Catal. A: Gen. 376, 4–12. n-butane oxidation to maleic anhydride over a VPO catalyst. AIChE J. 48,
Patience, G.S., Lorences, M.J., 2006. VPO transient oxidation kinetics. Int. J. Chem. 846–855.
React. Eng. 4, 1542–6580. Yerushalmi, J., Cankurt, N.T., 1979. Further studies of the regimes of fluidization.
Puchyr, D.M.J., Mehrotra, A.K., Behie, L.A., Kalogerakis, N.E., 1997. Modelling a Powder Technol. 24, 187–205.
circulating fluidized bed riser reactor with gas–solids downflow at the wall. Zhang, H.L., Degrève, J., Dewil, R., Baeyens, J., 2015. Operation diagram of
Can. J. Chem. Eng. 75, 317–326. circulating fluidized beds (CFBs). Procedia Eng. 102, 1092–1103.
Pugsley, T.S., Berruti, F., 1996. A predictive hydrodynamic model for circulating Zhu, H., Zhu, J., 2008. Gas-solids flow structures in a novel circulating-turbulent
fluidized bed risers. Powder Technol. 89, 57–69. fluidized bed. AIChE J. 54, 1213–1223.
Pugsley, T.S., Patience, G.S., Berruti, F., Chaouki, J., 1992. Modeling the catalytic
oxidation of n-butane to maleic anhydride in a circulating fluidized bed
reactor. Ind. Eng. Chem. Res. 31, 2652–2660.

You might also like