You are on page 1of 16

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Contents lists available at ScienceDirect

Chemical Engineering Research and Design


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

CFD study of diesel oil hydrotreating process in the


non-isothermal trickle bed reactor
Amir Heidari, Seyed Hassan Hashemabadi
Computational Fluid Dynamics (CFD) Research Laboratory, School of Chemical Engineering, Iran University of
Science and Technology (IUST), Narmak, Tehran 16846-13114, Iran

a b s t r a c t
In the present study, the EulerianEulerian multiphase approach was implemented to simulate the hydrotreating
processes (hydrodesulfurization (HDS) and hydrodearomatization (HDA)) in the trickle bed reactor (TBR) by means of
computational uid dynamics (CFD) technique. A new gasliquid interphase heat transfer coefcient was used in the
CFD model to predict the reactor performance at non-isothermal conditions. The effects of feed inlet temperature,
gas and liquid velocities, operational pressure and hydrogen sulde concentration of the gas phase were investigated
to calculate the reactions conversions and the bed temperature distribution. Also, the inuence of bed porosity and
adiabatic operational conditions on the reactor temperature and HDS reaction conversion was discussed. The results
showed at adiabatic reactor, HDS reaction conversion increased about 9% compared to constant wall temperature
condition. Furthermore, it was found that neglecting the gasliquid interphase heat transfer effect in the CFD model
increases the average relative error between numerical and the experimental data at prediction of HDS conversion
about 5% and also affects the bed temperature pattern.
2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords:

Computational uid dynamics (CFD); Hydrotreating; Non-isothermal condition; Trickle bed reactor;

Gasliquid heat transfer

1.

Introduction

Hydrodesulfurization (HDS) and hydrodearomatization (HDA)


are the well-known hydrotreatment (HDT) processes of fossil fuels impurities reduction in the oil industries. The HDT
processes are generally performed in the packed bed reactors known as trickle bed reactor (TBR) (Ranade et al., 2011). In
majority of the TBRs, gas and liquid phases move continuously
in the co-current downow mode through the high pressure
bed (Mederos et al., 2009b; Ranade et al., 2011; Wang et al.,
2013). During the HDT process, hydrogen transfers from the
gas to liquid phase and reacts with the target components to
reduce oil impurities. TBRs are generally characterized to have
low pressure drop and low catalyst loss. There are no moving
parts and TBRs have comparatively lower maintenance costs.
On the other hand, issues such as catalyst deactivation and

cementations are also TBRs main characteristics (Mederos


and Ancheyta, 2007; Sattereld, 1975).
In the past, many attempts were made to nd out about different aspects of TBR performance using numerical methods.
Reviewing published studies shows that most of the research
works were focused on development of appropriate models to
predict bed hydrodynamic properties by means of CFD technique. Investigating the effect of pressure drop (Atta et al.,
2010; Attou and Ferschneider, 2000; Aydin and Larachi, 2005,
2008; Bazmi et al., 2011, 2012; Boyer et al., 2007; Brkljac et al.,
2007; Iliuta et al., 2002; Iliuta and Larachi, 2002, 2005; Iliuta
et al., 2000a,b; Lopes and Quinta-Ferreira, 2008, 2009b; Propp
et al., 2000; Salimi et al., 2013; van der Merwe et al., 2007),
mal-distribution and liquid holdup (Bazmi et al., 2012; Gunjal
et al., 2003; Lappalainen et al., 2009; Lopes and Quinta-Ferreira,
2009a; Martnez et al., 2012; Mederos et al., 2009a; Strasser,

Corresponding author. Tel.: +98 21 77240376; fax: +98 21 77240495.


E-mail address: hashemabadi@iust.ac.ir (S.H. Hashemabadi).
Received 8 April 2014; Received in revised form 16 August 2014; Accepted 22 September 2014
Available online 30 September 2014
http://dx.doi.org/10.1016/j.cherd.2014.09.016
0263-8762/ 2014 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

550

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Nomenclature
a, b and C constants (dimensionless)
specic gasliquid interface (m1 )
aGL
as
specic surface of a particle (m1 )
concentration of ith species at liquid phase
CL,i
(mol/m3 )
heat capacity (J/kg K)
Cp
diffusion coefcient (m2 /s)
D
dp , dh , dr particle, hydraulic and reactor diameter (m)
Ergun constants
E1 , E2
2
Etvs number Eo = L g(dh ) / in Eq. (14)
Eo
(dimensionless)
Eo
modied Etvs number (dimensionless)
interphase momentum exchange coefcient
F
(kg/m3 s)
fe
wetting efciency (dimensionless)
Froude number (dimensionless)
Fr
gravity (m/s2 )
g
Henrys coefcient (MPa cm3 /mol)
H
specic enthalpy (W/m3 )
h
hG,L
gasliquid interphase heat transfer (W/m2 K)
modied Galileo number (dimensionless)
Ga
K
mass transfer coefcient (m/s)
KPoly/Di/Mono dynamic equilibrium constant (dimensionless)
Kw
Watson characterization factor (dimensionless)
thermal conductivity (W/m K)
k
ks
HDS reaction constant (((m3 )2.16 )/kg(kmol)1.16 s)
k,k
forward and backward reaction rate constants
(m3 /kg s)
kad
HDS reaction adsorption coefcient (m3 /kmol)
liquid hourly supercial velocity (h1 )
LHSV

m
interphase mass transfer rate (kg/m3 s)
Nu
Nusselt number (dimensionless)
Prandtl number (dimensionless)
Pr
Reynolds number, Re =  U dh / in Eq. (14)
Re
(dimensionless)
Re
modied Reynolds number (dimensionless)
pressure (Pa)
P
q
heat ux (W/m2 )
R
universal gas constant (J/mol K)
reaction rate of ith species (kg/m3 s)
Ri
r
radius (m)
source term (kg/m3 s)
S
SG
specic gravity (dimensionless)
temperature (K)
T
cubic average boiling point (K)
TCABP
TMABP
molal average boiling point (K)
TMeABP mean average boiling point (K)
t
time (s)
velocity (m/s)
u
v
molar volume (m3 /mol)
mass fraction (dimensionless)
x
molar fraction (dimensionless)
xmi
Greek symbols
, and  constants (dimensionless)
 ,  and   constants (dimensionless)
HR
heat of reaction (J/mol)

volume fraction (dimensionless)


average bed porosity (dimensionless)







dynamic viscosity (N s/m2 )


solubility coefcient (N l/kg MPa)
ratio of catalyst volume to summation of inert
and catalyst volume (dimensionless)
density (kg/m3 )
surface tension (N/m)

Subscripts
b
boiling point or bulk porosity
gas
G
ith species
i
int
interaction
kth phase
k
liquid phase
L
pressure
P
r
rth phase
solid phase
S
T
temperature

2010; Zeiser et al., 2001) and wetting efciency (Al-Dahhan and


1995; Baussaron et al., 2007a,b; Cheng et al., 2012;
Dudukovic,
Kundu et al., 2003; Lappalainen et al., 2008; Lopes and QuintaFerreira, 2010c; Mills and Dudukovic, 1981; Pironti et al., 1999;
Ring and Missen, 1991; van Houwelingen et al., 2006) were
the main subjects on the numerical study of trickle bed
reactors. The investigations of TBR performance at different
operational conditions (Gunjal and Ranade, 2007; Heidari and
Hashemabadi, 2014; Jarullah et al., 2011; Lopes and QuintaFerreira, 2007, 2010a,b,c; Lopes et al., 2007; Mapiour et al.,
2010), scale up challenges of TBRs (Hickman et al., 2013; Sie
and Krishna, 1998), reactor performance in the industrial scale
(Alvarez and Ancheyta, 2008; Alvarez et al., 2007; Jarullah

et al., 2012a,b; Mederos and Ancheyta, 2007; Munoz


et al., 2005;
Murali et al., 2007) and study of catalysts deactivation problems (Kallinikos et al., 2008; Kam et al., 2005; Pacheco et al.,
2011; Skala et al., 1991) were also other interesting subjects for
many researchers. Despite the fact that many researches were
published on hydrodynamics and yield of TBRs, studies about
heat transfer in the trickle ow regime have not been widely
reported. The available researches about heat transfer at TBRs
can be summarized in four categories; wall to bed heat transfer (Flvio Pinto Moreira et al., 2006; Habtu et al., 2011; Larachi
et al., 2003; Mariani et al., 2003; Mousazadeh et al., 2012), particles to uids heat transfer (Borremans et al., 2004; Guardo
et al., 2006, 2007; Larachi et al., 2003), gasliquid interphase
heat transfer (Heidari and Hashemabadi, 2013) and the effective axial (Flvio Pinto Moreira et al., 2006) and radial thermal
conductivity (Babu et al., 2007; Flvio Pinto Moreira et al., 2006;
Lamine et al., 1996; Larachi et al., 2003).
In this contribution, the non-isothermal trickle bed reactor
was simulated by CFD method and implementation of a new
gasliquid interphase heat transfer model that was developed
by Heidari and Hashemabadi (2013). The appropriate interphase momentum, mass and heat interaction models were
implemented in EulerianEulerian multiphase approach. The
reported HDS and HDA reactions kinetics were optimized to
nd the best values that present intrinsic chemical behaviors of reactions mechanisms based on the reactor real model.
Extensive simulation works were conducted to examine the
effects of gas and liquid velocities, operational pressure, feed
temperature and gas phase hydrogen sulde concentration on

551

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

the reactor performance. The CFD results were compared with


reported experimental data (Chowdhury et al., 2002). Also, the
importance of gasliquid interphase heat transfer effect on the
accuracy of CFD model at prediction of reactions conversions
and bed temperature distribution were studied.

2.

Governing equations

2.1.

Modeling assumption

forms:
G
=
G + L

FGL

(1)

i=r

(k k uk )
+ (k k uk 2 ) = k Pk + (k uk ) + k k g
t
+

k


rk ukr m
kr ) + Fint,k (uk ur )
(urk m

G 2 dp 2

S
(1 G )

G
G + L

E1 G (1 G )
G 2 dp 2

1
G + L

E1 G (S )
L 2 dp 2

0.333 

S
(1 G )

FLS =

0.667

(5)

0.667

0.333 

E2 L uL (S )
L dp

(6)


(7)

Table 1 presents supplementary correlations to estimate oil


properties, wetting efciency and porosity distribution.

2.3.

Species conservation equations

In the EulerianEulerian multiphase approach, species equation can be written as follows:


k k xk,i
+ (k k xk,i Uk ) = (k k Di,k xk,i )
t
k


rk,i m
kr,i ) + fe k Rk,i
(m

(8)

i=r

In the present study, an EulerianEulerian multiphase


approach was used to present gas (continuous phase) and liquid (discrete phase) continuity and momentum equations as
follows:


(k k )
rk m
kr ) + Sq
(m
+ (k k uk ) =
t

E2 G uG (1 G )
S
+
G dp
(1 G )

Hydrodynamics equations

To develop the momentum, mass and heat equations the following assumptions were regarded as:

2.2.

E1 G (1 G )

E2 G (uG uL )(1 G )
S
+
G dp
(1 G )

FGS =

1. Catalysts deactivation was ignored in the bed.


2. Vaporizations of diesel oil components were neglected due
to their high molecular weights.
3. Two-lm theory was used to dene interphase mass transfer between gas and liquid phases.
4. Mass transfer resistance is controlled by the gas and liquid interface and consequently it was assumed that no
resistance exists between other phases.
5. The solid phase was taken as third phase with known
porosity distribution and zero velocity in the domain.
6. According to Lopes and Quinta-Ferreira (2007) and Atta
et al. (2007a,b), due to low gasliquid interaction at trickle
ow regime, capillary pressure force can be neglected in
the simulation of TBRs. Accordingly, in this work, capillary
force was ignored in the CFD model.

(2)

i=r

The last term in Eq. (2), Fint,k , shows momentum exchange


between phases and with considering effect of wetting efciency, fe , is evaluated by (Lappalainen et al., 2009):
Fint,G = fe FGL (1 fe )FGS

(3)

Fint,L = fe (FLG FLS )

(4)

In order to develop interphase momentum exchanges, Fkr ,


Attou and Ferschneider (1999) model was applied in the CFD
model. In this model, effective diameter of particles was used
in the derivation of interphase interaction terms as a correction to account for the presence of the liquid lm. Their model
in the three phase system can be expressed by the following

The correlations required for calculation of species molecular diffusion, Di , and interphase mass transfer are presented
in Table 1. Also, diesel oil component concentration and distillation information are tabulated in Table 2.
In the current work, molecular diffusion was used to
describe mass diffusion instead of packed beds axial and
radial dispersion coefcients, Eq. (8). According to Ranade
et al. (2011) main factors contributing to mass and liquid
dispersion in the packed beds are non-uniform porosity distribution, partial wetting, dead-zones and channeling. These
factors have the vital effects on dispersion in the trickle bed
reactors. The main idea to develop axial and radial mass dispersion (diffusive term) models in the packed beds is the
investigation of hydrodynamics and porous media effects on
species governing equations (Jarullah et al., 2012b; Mederos
and Ancheyta, 2007; Mederos et al., 2006, 2009a, 2012). Therefore, in the models in which hydrodynamic and bed properties
(porosity distribution, partial wetting, dead-zones and channeling) are regarded, using axial or radial dispersion coefcient
causes over prediction of species diffusion terms and reactor
performance. Here, it should be noticed that in the packed
beds with large particles diameter, mechanical dispersion cannot be fully explained by porosity distribution functions and
appropriate models need to be used to improve the mathematical description of the reactor. In this work, due to small
particles size, the effect of mechanical dispersion was ignored
in the CFD model.
The ability of EulerianEulerian multiphase approach on
evaluation of axial and radial dispersion was studied by different researchers (Atta et al., 2007a; Bazmi et al., 2012;
Gunjal et al., 2003). They found a good agreement between

552

Table 1 Correlations to estimate diesel oil properties, bed characteristics and interphase mass transfer correlations.
Parameter

Correlation
L = 0 + P + T

Liquid density (Ahmed, 2007)

Liquid viscosity (Ahmed, 2007)

DL,i = 8.93 108


vi = 0.285v1.048
c

i or L

Diffusion coefcient (Ahmed, 2007; Perry


and Green, 2008; Reid et al., 1987; Riazi,
2005)

TMeABP =

0.2896
vcL = 0.5567TMeABP
SG0.7666
15.6

TMABP + TCABP
,
2


TCABP =

v0.267
T
L
v0.433 L

1
1.8

n
 


TMABP =

n


xmi Tbi

i=1

i (1.8Tbi 459.67)

i=1

1/3

PG,i
LG,i = m
GL,i i : H2 or H2 S
CL,i , m
i
 H
0.5
2
KL,i aL
L uL
L
= 1.8
i : H2 or H2 S
DL,i
L
L DL,i
vi
Hi =
i : H2 or H2 S

i L

2 
(T 273.15)
1
2
H2 = 0.5597 0.4294 103 (T 273.15) + 3.0753 109
+ 1.944 106 (T 273.15) + 0.835 103
106
20 C
20 C

GL,i = KL,i aGL


m

Interphase mass transfer (Goto and


Smith, 1975; Korsten and Hoffmann,
1996)

Liquid thermal conductivity (Riazi and


Faghri, 1985)

Liquid heat capacity (Kesler and Lee, 1976)

Bed porosity
function (Bazmi
et al., 2011)
(constants were
selected for sock
bed)

H2 S = exp(3.3670 0.00847(T 273.15)) 103


k = 1.7307(1.8 TMeABP ) SG15.5 C
= exp(21.78 8.07986t + 1.12981t2 0.05309t3 )
= 4.13948 + 1.29924t 0.17813t2 + 0.00833t3
 = 0.19876 0.0312t 0.00567t2
t = (1.8T 460)/100
Cp =  ( +   T)
 = 1.4651 + 0.2302 Kw
 = 0.306469 0.16734 SG15.5 C
  = 0.001467 0.000551 SG15.5 C
1/3
Kw = ((1.8 TMeABP ) )/SG
15.5 C
= (b + D) + (1 (b + D))

exp

C dr
p

3 
2 

ai (r/dp )2
(r/dp )(3+2(i1)) +bi

i=1

i
1
2
3

a
1.803
1.185
0.02649

b
0.0479
0.3566
0.001925

C
0.1252

D
0.045

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Surface tension (Tsonopoulos et al., 1986)


Wetting efciency (Lappalainen et al.,
2008)

P = ([0.167 + 16.181 100.04250 ][P 1.450 107 ] 0.01 [0.2999 + 263 100.06030 ] [P 1.450 107 ] ) 16.018
2.54
2
T = ([0.0133 + 152.4(0 + P )
](T 1.8 520) [8.1 106 0.0622 100.764(0 +P ) ](T 1.8 520) ) 16.018
3.444
a
L = 3.141 107 (T 1.8 460)
[log 10 (API)]
a = 10.313[log 10 (T 1.8 460)] 36.447
141.5
API =
131.5
SG15.6
4
 = (0.0017237 Tb0.05873 SG0.64927 (L G )) /1000
0.014
 0.185
 0.025
fe = 0.335Re L Eo 0.188Ga G (1 + FrG )

553

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Table 2 Diesel oil (liquid phase) specications


(Chowdhury et al., 2002).
Distillation curve
(ASTM D2887)

IBP (0.5 vol.%): 180 C


50 vol.%: 337 C
FBP (99.5 vol.%):
434 C
Sulfur: 16 298 ppm
Nitride: 218 ppm

Components

xwt 100 (wt%)

Monoaromatics
Diaromatics
Triaromatics

17.96
8.77
1.64

Tetraaromatics
Naphthenes
Parafn

0.95
19.25
49.78

k


rk hrk m
kr hkr ) + Sk
(Qrk + m

(13)

i=r

where , , Qrk and S are thermal conductivity, viscosity, interphase heat transfer according to Eqs. (14) and (18) and the
source term due to reactions heat, respectively.

CFD and experimental data for prediction of axial and


radial dispersion. Gunjal and Ranade (2007) indicated that
using molecular diffusion coefcient with appropriate hydrodynamics model (EulerianEulerian multiphase model and
considering porosity distribution) can properly describe reactive TBR performance. The results of their CFD model showed
good agreement with experimental data. In the current work,
the same model proposed by Gunjal and Ranade (2007) was
used in the CFD model. Furthermore, wetting effects were
considered on the CFD model to improve model accuracy.

2.3.1.

P
(  h ) + (k k Uk hk ) = ( k Tk ) + k uk k k
t k k k
t

2.4.1.

Gasliquid interphase heat transfer

A new correlation was used to account interphase heat transfer between gas and liquid phases. The correlation covers the
wide range of operational conditions and uids physical properties at trickle ow regime with the following form (Heidari
and Hashemabadi, 2013):

 Re 0.378  Pr 0.499

NuGL = 2.185

NuGL =

ReL

PrL

Eo0.627

hG,L dh
kG

(14)

(15)

Reaction kinetics

The last term in Eq. (8), Rk,i , stands for consumption or production of ith species due to chemical reactions. The following
reaction rate expressions were used to describe the HDS and
HDA reactions (Chowdhury et al., 2002):

where dh is the packed bed hydraulic diameter and dened as

Ar S + 2H2 Aromat + H2 S

2.4.2.

(9)

Poly-aromatics + H2 Di-aromatics

(10)

Di-aromatics + 2H2 Mono-aromatics

(11)

Mono-aromatics + 3H2 Naphthene

(12)

The reactions rate equations are tabulated in Table 3. It


should be noticed that HDA reactions are reversible and the
rate constants for the backward reactions (kPoly , kDi and
kMono ) can be estimated by vant Hoff equation (Chowdhury
et al., 2002).

2.4.

Energy conservation equation

In the current work, the non-isothermal conditions were evaluated for the reactor CFD model. The criteria for study a
packed bed at isothermal condition were reported by different researchers. Doraiswamy and Tajbl (1974) showed that the
practical reactor diameter-to-particles diameter ratio (dR /dp )
should be less than 4 for ignoring radial temperature. Carberry
(2001) stated that the maximum value for assumption of
isothermal temperature in the radial direction is dR /dp < 6.
In this work, the ratio of dR /dp based on catalysts equivalent diameter is about 13.6 (dR = 19 mm and dp = 1.4 mm) that
does not satisfy criteria of isothermal operation for the studied TBR. Furthermore, according to Carberry and White (1969)
the product yield in the packed bed reactors using 2D models is quite sensitive to temperature effects. Therefore, in the
current study, evaluation of temperature distribution in the
bed is necessary to achieve the best results for prediction of
reactor performance. The energy conservation equation for
each phase (gas, liquid and solid) based on EulerianEulerian
multiphase approach can be developed in the following form:

dh =

4
as (1 )

(16)

Liquidsolid and gassolid interphase heat transfer

Boelhouwer et al. (2001) derived a correlation to evaluate rate


of solid to liquid heat transfer at fully wetted trickle bed
reactor. In the current study, due to catalysts partial wetting condition, their correlation was modied to account for
the gas and liquid phases heat transfer with solid phase as
follows:
1/3

NuGS = (1 fe ) 0.111Re0.8
G PrG
1/3

NuLS = fe 0.111Re0.8
L PrL

2.4.3.

(17)
(18)

Wall to bed heat transfer

Heat transfer from the reactor wall to the bed at TBRs is significantly sensitive to the liquid ow rate. Based on experimental
data, Mariani et al. (2003) proposed the following correlation
to account wall to bed Nusselt number, Nuwall , for different
bed to particle diameter aspect ratios (4.7, 8.2, 17.2 and 34.3)
as
1/3

Nuwall = Re0.76
Lsu PrL
ReLsu =

ULc dp
L

(19)

(20)

where ULc and dp are liquid supercial velocity at core of the


bed and particle equivalent diameter, respectively.

3.

CFD model and numerical algorithms

In this study, as a benchmark, Chowdhury et al. (2002)


experimental work was used to consider the effect of
developed interphase heat transfer correlations in the prediction of hydrotreating reactions conversions. They studied
hydrodesulfurization and hydrodearomatization of diesel oil

554

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Table 3 Hydrodesulfurization and hydrodearomatization reactions kinetics.


Reactions (Chowdhury et al., 2002)

Heat of reactions (HR ) (Stanislaus and


Cooper, 1994; Chowdhury et al., 2002)
ks C1.6 C0.56
Ar S H2

= 1+50,000C

HDS

RAr

HDA (polyaromatics)
HDA (diaromatics)
HDA (monoaromatics)

RPoly = kPoly CPoly + kPoly CDi


RDi = kDi CDi + kDi CMono
RMono = kMono CMono + kMono CNaph

H2 S

in a tubular reactor with 500 mm length and 19 mm diameter, Fig. 1. The reactor consisted of two non-reactive zones at
two ends lled by inert particles with 0.2 mm average diameter. In the reactive zone, the trilobe catalysts (1.6 mm average
diameter and 3.5 mm average length) and inert spherical particles were mixed with ratio of 1:1.25 (vol/vol). In this study, the
equivalent diameter of particles was evaluated about 1.4 mm
on the basis of Sauter Mean Diameter (SMD) of trilobe catalysts. The more detail about effect of particles different size on
CFD model results of Chowdhury et al. (2002) reactor was presented by Gunjal and Ranade (2007). The thermal conductivity
of the solid phase on the basis of volumetric average of catalysts and inert particles thermal conductivity was evaluated
about 58 W/m K. The reactor wall temperature was controlled
by thermal system to prevent high increase in reactor temperature. Table 4 shows 18 different operational conditions used
at CFD simulations. The boundary conditions at the inlet zone
are equal with operational conditions, Table 4; furthermore,

67 000
66 462
133 024
199 203

the wall temperature is equal with the bed inlet temperature.


Fig. 1 shows other boundary conditions of the CFD model. The
initial conditions of the simulations were adjusted based on
Table 4 data. In order to reduce divergence at the beginning of
the simulation, the initial hydrogen concentration in the gas
phase was dened to zero.
Due to reactor shape and uniform ow through the bed,
a two dimensional (2D) axisymmetric computational domain
was used for CFD simulations, Fig. 1. Governing equations
(Eqs. (1), (2), (8) and (13)) were solved by nite volume method
(Patankar, 1980). Semi-Implicit Method for Pressure Linked
Equations (SIMPLE) was implemented for evaluating pressure
and velocity coupling in the computational domain. In order
to achieve more accurate simulations, all of the convective
terms in the transport equations were discretized by QUICK
(Quadratic Upwind Interpolation for Convective Kinematics)
scheme (Patankar, 1980). The simulations were performed at
unsteady state conditions with time step of 0.01 s. The energy

Fig. 1 Reactor dimensions, computational domain and boundary conditions.

555

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Table 4 Operational and boundary conditions in the CFD simulation.


Parameter

LHSV (h1 )

Tinlet , Twall (K)

P (MPa)

Temperature

sim-T1
sim-T2
sim-T3
sim-T4

573
593
613
653

4
4
4
4

2
2
2
2

200
200
200
200

0.014
0.014
0.014
0.014

Pressure

sim-P1
sim-P2
sim-P3

593
593
593

2
4
8

2
2
2

200
200
200

0.014
0.014
0.014

LHSV

sim-L1
sim-L2
sim-L3

593
593
593

4
4
4

1
2
4

200
200
200

0.014
0.014
0.014

sim-Q1
sim-Q2
sim-Q3
sim-Q4

593
593
593
593

4
4
4
4

2
2
2
2

100
200
300
500

0.014
0.014
0.014
0.014

sim-H2 S1
sim-H2 S2
sim-H2 S3
sim-H2 S4

593
593
593
593

4
4
4
4

2
2
2
2

200
200
200
200

0
0.014
0.03
0.08

Gas ow rate

Hydrogen
sulde
concentration

Cases

equations relaxation factors were dened small enough to


prevent numerical instability during the solution. The simulations were terminated when the bed average temperature
and average mass fraction of sulfur components at the entire
of solution domain have no signicant change with the time,
simultaneously. The average time that was required to reach
steady state conditions was about 10 000 s. All of the simulations were done on a PC with 8 GB RAM and Intel CoreTM i7
960 (8M Cache, 3.20 GHz) processor.

4.

Results and discussion

4.1.

CFD model validation

The hydrodynamic correlations of CFD model were validated


in the authors previous work (Salimi et al., 2013) and for
conciseness, validation procedure was not presented here.
In order to validation of the CFD model based on the conditions that reactions kinetics by Chowdhury et al. (2002) were
developed, the CFD simulations at different inlet temperatures (sim-T1, sim-T2, sim-T3 and sim-T4, Table 4) were done
in a model with uniform porosity, fully wetted particles and

QG,NTP /QL

xvH

2S

isothermal bed. Fig. 2 depicts CFD simulations results in contrast with Chowdhury et al. (2002) experimental data. The
mean relative error of 5.5% between CFD and experimental
results (Chowdhury et al., 2002) was obtained which shows
that the developed CFD model can predict HDS reactor performance with appropriate accuracy.

4.2.

Mesh independency

Three different mesh sizes were examined to nd optimum


mesh for CFD simulations. In this sensitivity analysis, achieving the nal HDS reaction conversion at the minimum run
time and no signicant change on its value were used as mesh
independency criteria. As shown in Table 5, the second mesh
type satises the mentioned criteria and was selected as the
optimum mesh for numerical simulations. Additionally, the
mesh density was rened near the bed wall to account wall
effect on the radial porosity distribution, Fig. 1.

4.3.

Modied reaction kinetics

In this investigation, the inuence of porosity distribution,


bed non-isothermal conditions and also particles partial wetting effect were accounted in the reactions kinetics constants
(ks,Poly,Di,Mono Table 3) to nd intrinsic HDS and HDA reactions
kinetics. According to Mederos et al. (2009a) if temperature
effects are neglected in the reactor model, it leads to the reaction rates with several orders of magnitude greater than those
calculated at non-isothermal conditions. Also, Ranade et al.
(2011) pointed out that for development of rate equations and
evaluation of kinetic parameters, all aspect of reactor conditions should be considered in the mathematical model. To

Table 5 Mesh independency results at sim-T1


operational condition of Table 4.
Mesh type

Fig. 2 Validation of HDS reactor CFD model.

1
2
3

Cell
numbers

HDS reaction
conversion (%)

Run time
(h)

9000
22 000
88 000

40
42.6
42.7

10
45
115

556

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Table 6 New values of reactions kinetic constants for


non-isothermal conditions and partially wetted
catalysts.
Reaction

HDS (ks ), Eq. (9)


HDA (polyaromatics
kPoly ), Eq. (10)
HDA (diaromatics kDi ),
Eq. (11)
HDA (monoaromatics
kMono ), Eq. (12)

Isothermal and
fully wetted
catalysts
(Chowdhury et
al., 2002)

Non-isothermal
and partially
wetted catalysts

2.5 1012
2.66 105

7.166 1012
4.23 105

8.5 102

1.33 103

6.04 102

9.62 102

this aim, new values of reactions constants were estimated


by comparing of HDS conversion obtained from numerical
(this work) and experimental data (Chowdhury et al., 2002)
at different operational temperatures (sim-T1, sim-T2, sim-T3
and sim-T4, Table 4). The best values were evaluated by using
bisection method. Table 6 presents the new values of reactions
constants for the HDS and HDA reactions.

4.4.

Fig. 4 Axial temperature distribution (at r = 0) and heat of


the total reactions through the bed at sim-T3 conditions,
Table 4.

Feed inlet temperature effect

Fig. 3 shows HDS reaction conversion at different inlet temperatures (sim-T1, sim-T2, sim-T3 and sim-T4, Table 4). As it
can be inferred, the CFD simulation results show proper agreements with experimental data (Chowdhury et al., 2002). The
relative error between the experimental and optimized CFD
model with new reactions constants (Section 4.3) was found
about 1.8%. Therefore, it can be concluded that the modied
reactions kinetics can be used to study the effects of different
operational parameters on the performance and local properties of the HDS reactor, more accurately.

4.4.1.

Temperature distribution

The axial temperature distribution and total heat of HDS and


HDA reactions at sim-T3 operational condition (Table 4) are
presented in Fig. 4. As shown in this gure, in the reactive
zone, the bed temperature increases rapidly due to generated
heat by the hydrotreating reactions (Eqs. (9)(12)). The temperature increases to the highest value (a hot spot with 623 K)

Fig. 5 HDS reaction conversion at different operational


pressures (sim-P1, sim-P2 and sim-P3, Table 4).
approximately at z = 0.31 m and then begins to decrease
through the bed. The reason of reduction in the bed temperature after z = 0.31 m can be found at reactor constant wall
temperature which was xed at 613 K. When the reactants
enter the reactive zone, great amounts of thermal energy
release due to high reaction rates at the beginning of reactive
zone. At the same time, increase in the bed temperature accelerates reactions rates and the bed temperature rises much
faster. On the other hand, the bed constant wall temperature
which was set at 613 K eliminates generated heat in the bed. As
shown in Fig. 4, the total heat of reactions begins to decrease
after the inlet part of reactive zone which is due to reduction
in the reactants concentrations. On the other hand, removing
the heat through the reactors wall nally causes the bed temperature decreases through the bed at z > 0.31 m. Lastly, the
rate of reduction in the temperature occurs more rapidly at
z > 0.4 m, where the reactive zone reaches to its end and the
energy equation (Eq. (13)) is solved with no heat source terms.

4.5.

Fig. 3 HDS reaction conversion at different inlet


temperatures (sim-T1, sim-T2, sim-T3 and sim-T4, Table 4).

The bed operational pressure effect

Fig. 5 depicts the effect of the bed operational pressure (sim-P1,


sim-P2 and sim-P3, Table 4) on the conversion of HDS reaction. The results show a good agreement between reported

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

557

Fig. 7 HDS reaction conversion at different LHSV values


(sim-L1, sim-L2 and sim-L3, Table 4).

Fig. 6 (A) Axial temperature distribution (at r = 0) and (B)


mass fraction of desulfurized components through the bed
at constant wall temperature and isolated wall in the
sim-P1 conditions, Table 4.

experimental data (Chowdhury et al., 2002) and CFD simulation results with 4.9% relative error. As it can be observed,
at constant gas and liquid ow rates, the reaction conversion
improves with an increase in the bed pressure. Enhancement
in the conversion rate can be caused by two reasons; rst,
when the operational pressure increases, the gas phase compresses and thus its velocity decreases through the bed. At
the same time, the momentum interaction between gas and
liquid phases causes the liquid velocity to reduce through the
bed, which, in turn, brings about increase in the liquid phase
residence time. Second, in the high reactor pressures, the
hydrogen concentration increases in the gas phase and therefore the hydrogen interphase mass transfer enhances which
gives rise to more conversion in the liquid phase reactants. As
a result, growth in the HDS conversion rate can be observed at
high operational pressures.

4.5.1.

isolated condition compared to constant wall temperature,


accelerates the reactions rates and reactants conversion. Consequently, as it can be seen in Fig. 6B, the mass fraction of
sulfur components in the adiabatic reactor decreases more
than a reactor with constant wall temperature. The nal conversion for HDS reaction in the adiabatic condition was about
78% while it is obtained about 69% at constant wall temperature condition.
As Fig. 6A illustrates, the liquid phase temperature
decreases about 0.5 K after the reactive zone and remains
0.41 m. The reason of this small reduction
constants after z =
in the liquid temperature can be found in the liquidgas and
liquidsolid interphase heat transfer rates. In effect, the interphase heat transfer is not a phenomenon with innite rate and
specic time is required that phases reach the thermal equilibrium. As a result, after the reactive zone a small decrease is
observed in the liquid phase temperature.

4.6.

Effect of LHSV on HDS reaction conversion

Fig. 7 shows HDS reaction conversion versus Liquid Hourly


Supercial Velocity (LHSV LHSV is ratio of liquid volumetric ow rate per reactor volume) at sim-L1, sim-L2 and sim-L3

Performance of adiabatic reactor

Fig. 6 compares the effect of isolated reactor wall and constant


wall temperature on axial distribution of liquid temperature and sulfur components concentration at sim-P1 (Table 4)
condition. According to Fig. 6A, an ascending difference is
observed between temperature values at constant wall temperature and isolated conditions with maximum difference
about 16 K. The more increase in the bed temperature at

Fig. 8 Axial temperature distribution at r = 0 in the sim-L1


and sim-L3 conditions, Table 4.

558

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Fig. 9 Total HDA reaction conversion at different LHSV


values (sim-L1, sim-L2 and sim-L3, Table 4).
conditions, Table 4. The relative error between CFD results
and experimental data (Chowdhury et al., 2002) is about 3.8%
which shows good agreement between numerical and experimental results. As it can be inferred from Fig. 7 when LHSV
increases, conversion of HDS reaction decreases. Actually, at
higher values of LHSV the liquid residence time decrease
and accordingly, reaction conversion reduces in the reactor.
The difference between CFD and experimental data at high
LHSV values can be explained by effect of hydrogen interphase
mass transfer rate on HDS reaction conversion. According to
interphase mass transfer coefcient term, KL,i aL (Table 1), an
increase in the liquid velocity enhances the rate of interphase
mass transfer. Therefore, at high LHSVs the hydrogen concentration in the liquid phase increases. Based on CFD results, it
seems that hydrogen concentration is predicted more than
real conditions at higher LHSVs and consequently, more prediction in HDS conversion occurs at numerical results respect
to experimental data.
In Fig. 8, axial distribution of temperature in the TBR
is depicted at two different conditions (sim-L1 and sim-L3,
Table 4). As shown, the temperature of the bed at sim-L1
shows higher values in comparison with sim-L3. The reason
of such behavior can be found in more conversion at sim-L1
and subsequently, more heat generation compared to simL3. According to Fig. 8, the bed maximum temperature in
the sim-L1 condition occurs at z = 0.35 m while in the other
case, sim-L3, the highest temperature happens at the outlet

Fig. 10 HDS reaction conversion at different gas ow rate


values (sim-Q1, sim-Q2, sim-Q3 and sim-Q4, Table 4).

of reactive zone, z = 0.4 m. The cause of this difference in the


hot spot positions can be found in the liquid phase residence
time. In the sim-L1, the liquid phase moves with lower velocity through the bed than that of sim-L3. As a result, in sim-L1
most of conversion happens at the entrance of the reactive
zone and the hot spot occurs closer to the inlet compared with
the sim-L3. In contrast, the higher velocity at sim-L3 condition
moves the hot spot to the end of the reactive zone due to low
residence time of reactants.

4.6.1.

HDA reactions conversion

Fig. 9 illustrates the effect of liquid velocity in terms of LHSV


parameter on the total HDA conversion. As the gure shows,
increase in the LHSV values reduces reaction conversion due
to decrease in reactants residence time. Moreover, the deviation between numerical and experimental data (Chowdhury
et al., 2002) is high and CFD results show about 45% average relative error with respect to the experimental data. The
reason of deviation between CFD and experimental results
can be due to inuence of temperature on the rate of HDA
backward reactions which damps forward reactions rates and
reduces overall conversion of aromatic components. According to bed non-isothermal condition and vant Hoff equation,
at high temperatures HDA backward reactions become as fast
as forward reactions. Therefore, reduction in the total HDA
conversion can be observed in the numerical results.

Fig. 11 (A) Radial temperature distribution at z = 0.4 m and (B) contour of temperature in the sim-Q1 condition (Table 4).

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

4.7.

Gas ow rate effect

In Fig. 10, the HDS reaction conversion is plotted against


gas ow rate (sim-Q1, sim-Q2, sim-Q3 and sim-Q4, Table 4).
According to the gure, increase in the gas ow rate improves
HDS reaction conversion, although no remarkable enhancement can be observed. The numerical results show about 6.4%
relative error versus available experimental data (Chowdhury
et al., 2002).
Increase in the gas velocity affects directly on the liquid phase velocity due to interphase momentum coupling;
subsequently, reactants residence time reduces in the bed.
Therefore, it is expected that reaction conversion decreases at
higher gas ow rates while the reverse behavior is observed.
Such inconsistency can be explained by inuence of higher gas
velocities on increase of hydrogen interphase mass transfer.
In comparison with low velocity condition, at higher velocities the bed hydrogen concentration remains close to the inlet
condition. Consequently, the rate of interphase mass transfer
enhances between gas and liquid phases which causes more
hydrogen concentration in the liquid phase. Hence, as shown
in Fig. 10, more conversion in the HDS reaction occurs at higher
gas ow rates.
The radial temperature distribution at the end of reactive
zone (z = 0.4 m) and the contour of temperature on longitudinal cross section are shown in Fig. 11 for sim-Q1 condition
(Table 4). As Fig. 11A suggests, the temperature prole is
affected by the bed porosity distribution so that wherever
the porosity is high, the bed temperature increases except
in the vicinity of the wall. In the locations with high local
porosity, due to more velocity of liquid phase, hydrogen interphase mass transfer increases about 4% respect to neighboring
locations according to interphase mass transfer correlation of
Table 1. Consequently, enhancement in the hydrogen concentration in the liquid phase increases rates and heat of the
reactions. On the other hand, it is expected that decrease
in the catalysts volume fraction at high local porosity positions leads to reduction of the reactions rates and the bed
temperature. As a result, it can be concluded that formation
of the local hot spots due to small changes in the catalysts volume fraction is affected by interphase mass transfer
and the bed temperatures increases in the high local porosity positions far from the wall. In the vicinity of the wall,
where the bed porosity is at the highest value, the bed temperature sharply decreases. In this region, the maximum
heat transfer from the bed occurs because of the proximity to the reactor wall which is set at constant temperature
(Twall = 593 K).
In Fig. 11B, the contour of bed temperature is shown at simQ1 condition. According to this gure, the hottest region is
approximately formed at the end of the reactive zone. The
position of hot spot depends on velocity of the phases and
the bed porosity distribution. It forms in the location near
the center of the bed axis located in the range of r > 2 mm
and r < 4 mm where the local porosity is more than neighboring locations. The estimation of hot spot position helps
to reduce thermal runaway problems in the bed and enables
reactor engineers to have an appropriate control over the process.

4.8.

559

Fig. 12 HDS reaction conversion at different hydrogen


sulde concentrations (sim-H2S1, sim-H2S2, sim-H2S3 and
sim-H2S4, Table 4).

sim-H2S3 and sim-H2S4, Table 4). The inuence of increase


in the gas phase hydrogen sulde can be seen in the form
of reduction of hydrogen concentration in the gas phase
and correspondingly, in the liquid phase. Actually, at lower

Effect of gas phase H2 S concentration

Fig. 12 illustrates the HDS conversion at different hydrogen


sulde concentrations of the gas phase (sim-H2S1, sim-H2S2,

Fig. 13 (A) Reactive zone average temperature and (B) total


heat generated by HDS and HDA reactions (sim-H2S1,
sim-H2S2, sim-H2S3 and sim-H2S4, Table 4).

560

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Fig. 14 Mass fraction of Poly-, Di- and Mono-aromatics


along the bed at sim-H2S1 conditions, Table 4.

values of hydrogen mole fraction in the gas phase, rate


of interphase mass transfer reduces which causes smaller
hydrogen concentration in the liquid phase and therefore
reduction in the conversion of reactions. As it is depicted
in Fig. 12, the current CFD model which considers for the

effect of interphase heat transfer, provides a better match


to the experimental data in comparison with other CFD
results (Gunjal and Ranade, 2007) at isothermal conditions.
The average of relative error between CFD and experimental data (Chowdhury et al., 2002) is about 13% for the current
study and 21% for Gunjal and Ranade (2007). The deviation
between numerical and experimental results at prediction
of HDS conversion might be attributed to independency
of adsorption coefcient of hydrogen sulde to the temperature in the HDS reaction kinetics and also developed
reaction kinetics constants. However, it seems that due to
small changes in the bed temperature, adsorption coefcient has no signicant effect on the errors. Therefore, the
main reason of deviation between experimental and numerical results might be related to the undesired dependency of
HDS reaction kinetic from hydrogen and hydrogen sulde concentrations.
Fig. 13 depicts the average temperature of the reactive
zone and heat of HDS and total HDA reactions. According
to Fig. 13A, the temperature shows no signicant changes
with increase in the hydrogen sulde concentration while
due to decrease in HDS conversion (Fig. 12) and produced
heat by the HDS reaction, it seems that the bed temperature
should be decreased. To explain constant temperature trend,
the released heat by HDS and total HDA reactions should be
computed, simultaneously. In Fig. 13B, values of generated
heat by the reactions are shown at individual bars. The gure shows that with increase in the hydrogen sulde volume
fraction in the gas phase, the produced heat by HDS reaction decreases. In contrast, due to reduction in HDS reaction
rate, the released heat by total HDA reactions takes ascending
trend at higher hydrogen sulde concentrations. At the high
concentration of hydrogen in the bed, the rate of hydrogen
consumption by the HDS reaction is considerably more than
HDA reactions. On the other hand, at low hydrogen concentration, the rate of hydrogen consumption by HDS reaction
declines which causes growth in the rate of hydrogen consumption by the HDA reactions. With more decrease in the
hydrogen concentration (sim-H2 S1 to sim-H2 S4, Table 4), the
contribution of HDA reactions to consumption of hydrogen
increases. Therefore, as seen in Fig. 13B, due to higher generated heat by HDA reactions in comparison with HDS reaction
and also approximately constant overall produced heat, the
bed average temperature shows no signicant changes at different hydrogen sulde concentration, Fig. 13A.

4.8.1. Distribution of aromatic components axial mass


fraction

Fig. 15 (A) Effect of gasliquid interphase heat transfer in


the CFD prediction of HDS reaction conversion (sim-P1,
sim-P2 and sim-P3, Table 4) and (B) axial (r = 0) bed
temperature at sim-P3 conditions.

In Fig. 14, distribution of average mass fraction of aromatic components is shown along the reactor length at
sim-H2S1 conditions. As observed, mass fractions of polyand di-aromatics components continuously decrease while
mono-aromatics mass fraction shows a maximum at the
inlet of the reactive zone. According to Eqs. (10)(12), HDAs
reactions kinetics are presented by a chain mechanism. In
the case of mono-aromatics reaction, due to high conversion of di-aromatics species respect to mono-aromatics, an
increase in the concentration of mono-aromatics components
occurs along the bed with a maximum point at z = 0.17 m.
After that, due to reduction in the di-aromatics concentration and consumption of mono-aromatics components, mass
fraction of mono-aromatics begins to decrease through the
reactor.

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

561

Fig. 16 Contour of the bed temperature: (A) with gasliquid interphase heat transfer and (B) without gasliquid interphase
heat transfer at sim-P3 conditions, Table 4.

4.9.
Gasliquid interphase heat transfer effect on
simulation results
In order to demonstrate gasliquid interphase coefcient
effect in the prediction of HDS reaction conversion and bed
temperature distribution, the CFD simulations at sim-P1,
sim-P2 and sim-P3 (Table 4) were repeated with eliminating
gasliquid interfacial heat transfer. As Fig. 15 shows, in the
case of no gasliquid heat transfer, Fig. 15A, the HDS conversion prediction is about 5% higher than non-isothermal
conditions. In effect, neglecting the heat transfer between gas
and liquid phases causes the reaction heat in the liquid phase
to only exchange with the solid phase. Thus, due to lower
heat loss in the liquid phase, the phase temperature increases
along the bed (Fig. 15B) which leads to the reaction conversion shows higher values than expected. In Fig. 16, contour of
temperature is shown with and without effect of gasliquid
interphase heat transfer during the CFD simulation at sim-P3.
As it can be inferred from the gure, neglecting the gasliquid
thermal interaction affects signicantly on the local temperature distribution, so that the location and values of hot spots
change in the bed. Subsequently, the entire phenomena such
as interphase mass transfer, phases properties and local reactants conversion are affected by difference at temperature
distribution. Therefore, it is safe to say that utilizing gasliquid
interphase heat transfer signicantly improves the accuracy of
TBR performance prediction.

5.

Conclusion

CFD simulations of hydrotreating TBR were carried out


by means of EulerianEulerian multiphase approach. The
main objective of this work was the CFD simulation of
non-isothermal reactor by implementation of appropriate
interphase heat transfer coefcient. The modied solid-uids

heat transfer coefcient and a new gasliquid interphase heat


transfer model were developed to accurately predict the reactions conversions and local properties through the bed. Based
on the CFD simulations, the following results can be inferred:
The reaction kinetics constants were optimized based on
developed CFD model. The optimization was done to nd
the best values that present intrinsic chemical behavior of
reactions on the basis of appropriate mathematical model
and real conditions of studied TBR. Results showed good
compromise between numerical and experimental results.
It was shown that the radial temperature of the bed affected
by the bed porosity and the highest temperature occurred
at positions near the bed center. Results showed that the
maximum heat loss occurs at the vicinity of the wall; in
this zone, high velocities of the phases and constant wall
temperature cause the maximum heat loss in the bed.
At the adiabatic and constant wall temperature, a large
difference was observed between the bed temperatures. It
was shown that there is a difference about 16 K between
maximum bed temperatures at adiabatic and constant wall
temperature conditions. In addition, about 16% improvement was observed in HDS reaction conversion at adiabatic
conditions.
Finally, it was shown that neglecting the gasliquid interphase heat transfer affects on the prediction of reactions
conversion and local parameters such as bed temperature.
Therefore, using appropriate model to describe gasliquid
interphase heat transfer seems inevitable.

Acknowledgment
The authors would like to thank Pars Oil and Gas Company
(POGC - Islamic Republic of Iran) for the nancial support of
this work (Grant No. 91-203/TP).

562

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

References
Ahmed, T.H., 2007. Equations of State and PVT Analysis:
Applications for Improved Reservoir Modeling. Gulf Pub.,
Houston, TX.
M.P., 1995. Catalyst wetting
Al-Dahhan, M.H., Dudukovic,
efciency in trickle-bed reactors at high pressure. Chem. Eng.
Sci. 50, 23772389.
Alvarez, A., Ancheyta, J., 2008. Simulation and analysis of
different quenching alternatives for an industrial vacuum
gasoil hydrotreater. Chem. Eng. Sci. 63, 662673.

Alvarez, A., Ancheyta, J., Munoz,


J.A.D., 2007. Comparison of
quench systems in commercial xed-bed hydroprocessing
reactors. Energy Fuels 21, 11331144.
Atta, A., Roy, S., Nigam, K.D.P., 2007a. Investigation of liquid
maldistribution in trickle-bed reactors using porous media
concept in CFD. Chem. Eng. Sci. 62, 70337044.
Atta, A., Roy, S., Nigam, K.D.P., 2007b. Prediction of pressure drop
and liquid holdup in trickle bed reactor using relative
permeability concept in CFD. Chem. Eng. Sci. 62, 58705879.
Atta, A., Roy, S., Nigam, K.D.P., 2010. A two-phase Eulerian
approach using relative permeability concept for modeling of
hydrodynamics in trickle-bed reactors at elevated pressure.
Chem. Eng. Res. Des. 88, 369378.
Attou, A., Ferschneider, G., 1999. A two-uid model for ow
regime transition in gasliquid trickle-bed reactors. Chem.
Eng. Sci. 54, 50315037.
Attou, A., Ferschneider, G., 2000. A two-uid hydrodynamic
model for the transition between trickle and pulse ow in a
cocurrent gasliquid packed-bed reactor. Chem. Eng. Sci. 55,
491511.
Aydin, B., Larachi, F., 2005. Trickle bed hydrodynamics and ow
regime transition at elevated temperature for a Newtonian
and a non-Newtonian liquid. Chem. Eng. Sci. 60, 66876701.
Aydin, B., Larachi, F., 2008. Trickle bed hydrodynamics for
(non-)Newtonian foaming liquids in non-ambient conditions.
Chem. Eng. J. 143, 236243.
Babu, B.V., Shah, K.J., Govardhana Rao, V., 2007. Lateral mixing in
trickle bed reactors. Chem. Eng. Sci. 62, 70537059.
Baussaron, L., Julcour-Lebigue, C., Wilhelm, A.-M., Delmas, H.,
Boyer, C., 2007a. Wetting topology in trickle bed reactors.
AIChE J. 53, 18501860.
Baussaron, L., Julcour-Lebigue, C., Wilhelm, A.M., Boyer, C.,
Delmas, H., 2007b. Partial wetting in trickle bed reactors:
measurement techniques and global wetting efciency. Ind.
Eng. Chem. Res. 46, 83978405.
Bazmi, M., Hashemabadi, S.H., Bayat, M., 2011. CFD simulation
and experimental study for two-phase ow through the
trickle bed reactors, sock and dense loaded by trilobe
catalysts. Int. Commun. Heat Mass Transf. 38, 391397.
Bazmi, M., Hashemabadi, S.H., Bayat, M., 2012. CFD simulation
and experimental study of liquid ow mal-distribution
through the randomly trickle bed reactors. Int. Commun. Heat
Mass Transf. 39, 736743.
Boelhouwer, J.G., Piepers, H.W., Drinkenburg, A.A.H., 2001.
Particle-liquid heat transfer in trickle-bed reactors. Chem.
Eng. Sci. 56, 11811187.
Borremans, D., Rode, S., Wild, G., 2004. Liquid ow distribution
and particleuid heat transfer in trickle-bed reactors: the
inuence of periodic operation. Chem. Eng. Process. 43,
14031410.
Boyer, C., Volpi, C., Ferschneider, G., 2007. Hydrodynamics of
trickle bed reactors at high pressure: two-phase ow model
for pressure drop and liquid holdup, formulation and
experimental validation. Chem. Eng. Sci. 62, 70267032.
Brkljac, B., Bludowsky, T., Dietrich, W., Grunewald, M., Agar, D.W.,
2007. Modelling of unsteady-state hydrodynamics in
periodically operated trickle-bed reactors: inuence of the
liquid-phase physical properties. Chem. Eng. Sci. 62,
70117019.
Carberry, J.J., 2001. Chemical and Catalytic Reaction Engineering.
Dover Publications, Mineola, NY.

Carberry, J.J., White, D., 1969. On the role of transport phenomena


in catalytic reactor behavior. Ind. Eng. Chem. 61, 2735.
Cheng, Z.-M., Kong, X.-M., Zhu, J., Wang, Z.-Y., Jin, J., Huang, Z.-B.,
2012. Hydrodynamic modeling on the external liquidsolid
wetting efciency in a trickling ow reactor. AIChE J. 59,
283294.
Chowdhury, R., Pedernera, E., Reimert, R., 2002. Trickle-bed
reactor model for desulfurization and dearomatization of
diesel. AIChE J. 48, 126135.
Doraiswamy, L.K., Tajbl, D.G., 1974. Laboratory catalytic reactors.
Cat. Rev. Sci. Eng. 10, 177219.
Flvio Pinto Moreira, M., do Carmo Ferreira, M., Freire, J.T., 2006.
Evaluation of pseudohomogeneous models for heat transfer
in packed beds with gas ow and gasliquid cocurrent
downow and upow. Chem. Eng. Sci. 61, 20562068.
Goto, S., Smith, J.M., 1975. Trickle-bed reactor performance. Part I.
Holdup and mass transfer effects. AIChE J. 21, 706713.
Guardo, A., Coussirat, M., Recasens, F., Larrayoz, M.A., Escaler, X.,
2006. CFD study on particle-to-uid heat transfer in xed bed
reactors: convective heat transfer at low and high pressure.
Chem. Eng. Sci. 61, 43414353.
Guardo, A., Coussirat, M., Recasens, F., Larrayoz, M.A., Escaler, X.,
2007. CFD studies on particle-to-uid mass and heat transfer
in packed beds: free convection effects in supercritical uids.
Chem. Eng. Sci. 62, 55035511.
Gunjal, P.R., Ranade, V.V., 2007. Modeling of laboratory and
commercial scale hydro-processing reactors using CFD. Chem.
Eng. Sci. 62, 55125526.
Gunjal, P.R., Ranade, V.V., Chaudhari, R.V., 2003. Liquid
distribution and RTD in trickle bed reactors: experiments and
CFD simulations. Can. J. Chem. Eng. 81, 821830.
Habtu, N., Font, J., Fortuny, A., Bengoa, C., Fabregat, A., Haure, P.,
Ayude, A., Stber, F., 2011. Heat transfer in trickle bed column
with constant and modulated feed temperature: experiments
and modeling. Chem. Eng. Sci. 66, 33583368.
Heidari, A., Hashemabadi, S.H., 2013. Numerical evaluation of the
gasliquid interfacial heat transfer in the trickle ow regime
of packed beds at the micro and meso-scale. Chem. Eng. Sci.
104, 674689.
Heidari, A., Hashemabadi, S.H., 2014. CFD simulation of
isothermal diesel oil hydrodesulfurization and
hydrodearomatization in trickle bed reactor. J. Taiwan Inst.
Chem. Eng. 45, 13891402.
Hickman, D.A., Holbrook, M.T., Mistretta, S., Rozeveld, S.J., 2013.
Successful scale-up of an industrial trickle bed hydrogenation
using laboratory reactor data. Ind. Eng. Chem. Res. 52,
1528715292.
Iliuta, I., Grandjean, B.P.A., Larachi, F., 2002. Hydrodynamics of
trickle-ow reactors: updated slip functions for the slit
models. Chem. Eng. Res. Des. 80, 195200.
Iliuta, I., Larachi, F., 2002. Hydrodynamics of power-law uids in
trickle-ow reactors: mechanistic model, experimental
verication and simulations. Chem. Eng. Sci. 57, 19311942.
Iliuta, I., Larachi, F., 2005. Modelling the hydrodynamics of
gasliquid packed beds via slit models: a review. Int. J. Chem.
React. Eng. 3, 125.
Iliuta, I., Larachi, F., Al-Dahhan, M.H., 2000a. Double-slit model
for partially wetted trickle ow hydrodynamics. AIChE J. 46,
597609.
Iliuta, I., Larachi, F., Al-Dahhan, M.H., 2000b. Multiple-zone model
for partially wetted trickle ow hydrodynamics. Chem. Eng.
Res. Des. 78, 982990.
Jarullah, A.T., Mujtaba, I.M., Wood, A.S., 2011. Kinetic parameter
estimation and simulation of trickle-bed reactor for
hydrodesulfurization of crude oil. Chem. Eng. Sci. 66, 859871.
Jarullah, A.T., Mujtaba, I.M., Wood, A.S.,2012a. Economic analysis
of an industrial rening unit involving hydrotreatment of
whole crude oil in trickle bed reactor using gproms. In: The
22nd European Symposium on Computer Aided Process
Engineering. Elsevier, London, pp. 652656.
Jarullah, A.T., Mujtaba, I.M., Wood, A.S., 2012b. Whole crude oil
hydrotreating from small-scale laboratory pilot plant to

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

large-scale trickle-bed reactor: analysis of operational issues


through modeling. Energy Fuels 26, 629641.
Kallinikos, L.E., Bellos, G.D., Papayannakos, N.G., 2008. Study of
the catalyst deactivation in an industrial gasoil HDS reactor
using a mini-scale laboratory reactor. Fuel 87,
24442449.
Kam, E.K.T., Al-Shamali, M., Juraidan, M., Qabazard, H., 2005. A
hydroprocessing multicatalyst deactivation and reactor
performance modelpilot-plant life test applications. Energy
Fuels 19, 753764.
Kesler, M.G., Lee, B.I., 1976. Improve Prediction of Enthalpy of
Fractions, Hydrocarbon Processing. Gulf Pub., Houston, pp.
153158.
Korsten, H., Hoffmann, U., 1996. Three-phase reactor model for
hydrotreating in pilot trickle-bed reactors. AIChE J. 42,
13501360.
Kundu, A., Nigam, K.D.P., Verma, R.P., 2003. Catalyst wetting
characteristics in trickle-bed reactors. AIChE J. 49, 22532263.
Lamine, A.S., Gerth, L., Le Gall, H., Wild, G., 1996. Heat transfer in
a packed bed reactor with cocurrent downow of a gas and a
liquid. Chem. Eng. Sci. 51, 38133827.
Lappalainen, K., Alopaeus, V., Manninen, M., Aittamaa, J., 2008.
Improved hydrodynamic model for wetting efciency,
pressure drop, and liquid holdup in trickle-bed reactors. Ind.
Eng. Chem. Res. 47, 84368444.
Lappalainen, K., Manninen, M., Alopaeus, V., 2009. CFD modeling
of radial spreading of ow in trickle-bed reactors due to
mechanical and capillary dispersion. Chem. Eng. Sci. 64,
207218.
Larachi, F., Belfares, L., Iliuta, I., Grandjean, B.P.A., 2003. Heat and
mass transfer in cocurrent gasliquid packed beds. Analysis,
recommendations, and new correlations. Ind. Eng. Chem. Res.
42, 222242.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2007. Trickle-bed CFD studies
in the catalytic wet oxidation of phenolic acids. Chem. Eng.
Sci. 62, 70457052.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2008. Three-dimensional
numerical simulation of pressure drop and liquid holdup for
high-pressure trickle-bed reactor. Chem. Eng. J. 145,
112120.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2009a. CFD modelling of
multiphase ow distribution in trickle beds. Chem. Eng. J. 147,
342355.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2009b. Numerical simulation
of trickle-bed reactor hydrodynamics with RANS-based
models using a volume of uid technique. Ind. Eng. Chem.
Res. 48, 17401748.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2010a. Assessment of CFD
EulerEuler method for trickle-bed reactor modelling in the
catalytic wet oxidation of phenolic wastewaters. Chem. Eng. J.
160, 293301.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2010b. Evaluation of
multiphase CFD models in gasliquid packed-bed reactors for
water pollution abatement. Chem. Eng. Sci. 65, 291297.
Lopes, R.J.G., Quinta-Ferreira, R.M., 2010c. Numerical studies of
catalyst wetting and total organic carbon reaction on
environmentally based trickle-bed reactors. Ind. Eng. Chem.
Res. 49, 1073010743.
Lopes, R.J.G., Silva, A.M.T., Quinta-Ferreira, R.M., 2007. Kinetic
modeling and trickle-bed CFD studies in the catalytic wet
oxidation of vanillic acid. Ind. Eng. Chem. Res. 46, 83808387.
Mapiour, M., Sundaramurthy, V., Dalai, A.K., Adjaye, J., 2010.
Effects of the operating variables on hydrotreating of heavy
gas oil: experimental, modeling, and kinetic studies. Fuel 89,
25362543.
Mariani, N.J., Mazza, G.D., Martnez, O.M., Cukierman, A.L.,
Barreto, G.F., 2003. On the inuence of liquid distribution on
heat transfer parameters in trickle bed systems. Can. J. Chem.
Eng. 81, 814820.
Martnez, M., Pallares, J., Lpez, J., Lpez, A., Albertos, F., Garca,
M.A., Cuesta, I., Grau, F.X., 2012. Numerical simulation of the
liquid distribution in a trickle-bed reactor. Chem. Eng. Sci. 76,
4957.

563

Mederos, F.S., Ancheyta, J., 2007. Mathematical modeling and


simulation of hydrotreating reactors: cocurrent versus
countercurrent operations. Appl. Catal. A 332, 821.
Mederos, F.S., Ancheyta, J., Chen, J., 2009a. Review on criteria to
ensure ideal behaviors in trickle-bed reactors. Appl. Catal. A
355, 119.
Mederos, F.S., Ancheyta, J., Elizalde, I., 2012. Dynamic modeling
and simulation of hydrotreating of gas oil obtained from
heavy crude oil. Appl. Catal. A 425-426, 1327.
Mederos, F.S., Elizalde, I., Ancheyta, J., 2009b. Steady-state and
dynamic reactor models for hydrotreatment of oil fractions: a
review. Cat. Rev. Sci. Eng. 51, 485607.
Mederos, F.S., Rodrguez, M.A., Ancheyta, J., Arce, E., 2006.
Dynamic modeling and simulation of catalytic hydrotreating
reactors. Energy Fuels 20, 936945.
Mills, P.L., Dudukovic, M.P., 1981. Evaluation of liquidsolid
contacting in trickle-bed reactors by tracer methods. AIChE J.
27, 893904.
Mousazadeh, F., van den Akker, H.E.A., Mudde, R.F., 2012. Eulerian
simulation of heat transfer in a trickle bed reactor with
constant wall temperature. Chem. Eng. J. 207208,
675682.

Munoz,
J.A.D., Alvarez, A., Ancheyta, J., Rodrguez, M.A.,
Marroqun, G., 2005. Process heat integration of a heavy crude
hydrotreatment plant. Catal. Today 109, 214218.
Murali, C., Voolapalli, R.K., Ravichander, N., Gokak, D.T.,
Choudary, N.V., 2007. Trickle bed reactor model to simulate
the performance of commercial diesel hydrotreating unit.
Fuel 86, 11761184.
Pacheco, M.E., Martins Salim, V.M., Pinto, J.C., 2011. Accelerated
deactivation of hydrotreating catalysts by coke deposition.
Ind. Eng. Chem. Res. 50, 59755981.
Patankar, S.V., 1980. Numerical Heat Transfer and Fluid Flow.
Hemisphere, London.
Perry, R.H., Green, D.W., 2008. Perrys Chemical Engineers
Handbook, 8th ed. McGraw-Hill, New York.
Pironti, F., Mizrahi, D., Acosta, A., Gonzalez-Mendizabal, D., 1999.
Liquidsolid wetting factor in trickle-bed reactors: its
determination by a physical method. Chem. Eng. Sci. 54,
37933800.
Propp, R.M., Colella, P., Crutcheld, W.Y., Day, M.S., 2000. A
numerical model for trickle bed reactors. J. Comput. Phys. 165,
311333.
Ranade, V.V., Chaudhari, R.V., Gunjal, P.R., 2011. Trickle Bed
Reactors: Reactor Engineering & Applications. Elsevier.
Reid, R.C., Prausnitz, J.M., Poling, B.E., 1987. The Properties of
Gases and Liquids, 4th ed. McGraw-Hill, New York;
London.
Riazi, M.R., 2005. Characterization and Properties of Petroleum
Fractions. ASTM International, W. Conshohocken, PA.
Riazi, M.R., Faghri, A., 1985. Thermal conductivity of liquid and
vapor hydrocarbon systems: pentanes and heavier at low
pressures. Ind. Eng. Chem. Process Des. Dev. 24, 398401.
Ring, Z.E., Missen, R.W., 1991. Trickle-bed reactors: tracer study of
liquid holdup and wetting efciency at high temperature and
pressure. Can. J. Chem. Eng. 69, 10161020.
Salimi, M., Hashemabadi, S.H., Noroozi, S., Heidari, A., Bazmi, M.,
2013. Numerical and experimental study of catalyst loading
and body effects on a gasliquid trickle-ow bed. Chem. Eng.
Technol. 36, 4352.
Sattereld, C.N., 1975. Trickle-bed reactors. AIChE J. 21, 209228.
Sie, S.T., Krishna, R., 1998. Process development and scale up III.
Scale-up and scale-down of trickle bed processes. Rev. Chem.
Eng. 14, 203252.

A.M., Meyn, V.W., Severin, D.K.,


Skala, D.U., Saban,
M.D., Orlovic,
M.V., 1991. Hydrotreating of
Rahimian, I.G.H., Marjanovic,
used oil: prediction of industrial trickle-bed operation from
pilot-plant data. Ind. Eng. Chem. Res. 30, 20592065.
Stanislaus, A., Cooper, B.H., 1994. Aromatic hydrogenation
catalysis: a review. Cat. Rev. Sci. Eng. 36, 75123.
Strasser, W., 2010. CFD study of an evaporative trickle bed reactor.
Mal-distribution and thermal runaway induced by feed
disturbances. Chem. Eng. J. 161, 257268.

564

chemical engineering research and design 9 4 ( 2 0 1 5 ) 549564

Tsonopoulos, C., Heidman, J.L., Hwang, S.-C., 1986.


Thermodynamic and Transport Properties of Coal Liquids.
Wiley, New York.
van der Merwe, W., Nicol, W., de Beer, F., 2007. Three-dimensional
analysis of trickle ow hydrodynamics: computed
tomography image acquisition and processing. Chem. Eng.
Sci. 62, 72337244.
van Houwelingen, A.J., Sandrock, C., Nicol, W., 2006. Particle
wetting distribution in trickle-bed reactors. AIChE J. 52,
35323542.

Wang, Y., Chen, J., Larachi, F., 2013. Modelling and simulation of
trickle-bed reactors using computational uid dynamics:
a state-of-the-art review. Can. J. Chem. Eng. 91,
136180.
Zeiser, T., Lammers, P., Klemm, E., Li, Y.W., Bernsdorf, J., Brenner,
G., 2001. CFD-calculation of ow, dispersion and reaction in a
catalyst lled tube by the lattice Boltzmann method. Chem.
Eng. Sci. 56, 16971704.

You might also like