You are on page 1of 13

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/223648792

Modeling and simulation of petroleum coke


calcination in rotary kilns

ARTICLE in FUEL SEPTEMBER 2001


Impact Factor: 3.52 DOI: 10.1016/S0016-2361(01)00032-1

CITATIONS READS

21 437

3 AUTHORS:

Marcio Ardes Martins Leandro Soares Oliveira


Universidade Federal de Viosa (UFV) Federal University of Minas Gerais
23 PUBLICATIONS 68 CITATIONS 92 PUBLICATIONS 889 CITATIONS

SEE PROFILE SEE PROFILE

Adriana S Franca
Federal University of Minas Gerais
99 PUBLICATIONS 1,228 CITATIONS

SEE PROFILE

Available from: Leandro Soares Oliveira


Retrieved on: 28 September 2015
Fuel 80 (2001) 16111622
www.fuelrst.com

Modeling and simulation of petroleum coke calcination in rotary kilns


Marcio A. Martins, Leandro S. Oliveira*, Adriana S. Franca
Department of Chemical Engineering, Universidade Federal de Minas Gerais, R. Esprito Santo, 35-6 andar, 30160-030 Belo Horizonte, MG, Brazil
Received 7 July 2000; revised 20 December 2000; accepted 16 January 2001

Abstract
A one-dimensional mathematical model was developed for the simulation of petroleum coke calcination in rotary kilns. The model is
comprised of 14 ordinary differential equations derived from mass and energy conservation principles. The system of equations is solved by a
fourth-order RungeKutta method. The model predicts, in the axial direction, temperature proles for the bed of particles, the gas phase and
the kiln internal wall. It also predicts the composition proles for the gas and solid phases. The modeling of the axial ow of the bed includes
the rheological characteristics of the particulate system. The results from the simulation of the calcination systems presented better agreement
with measured industrial data than other simulation results available in the literature. q 2001 Published by Elsevier Science Ltd.
Keywords: Heat and mass transfer; Numerical analysis; Coke combustion

1. Introduction continuous process, stability in achieving a product with


uniform quality is assured. A schematic diagram of a typical
The majority of industrial calcination processes takes kiln is presented in Fig. 1.
place in rotary kilns. The primary functions of a rotary Mathematical models have been developed in the past
kiln are to supply enough heat to raise the solids to reaction decades to describe petroleum coke calcination processes
temperatures and to promote an efcient mixing of these in rotary kilns [15]. The models were usually built upon
solids in order to ensure uniformity of heat transfer. A typi- the basic principles of conservation of heat and mass, and
cal rotary kiln consists of a long and slightly tilted cylinder, took into account chemical reactions. Perron et al. [4,5]
rotating about its axis. The cylinder slope facilitates the developed a one-dimensional model to predict the axial
downward axial motion of the solids bed while hot gas temperature and composition proles for both the gas and
ows upwards. solid phases in a rotary kiln. The simulation results
Carbon electrodes used in the aluminum industry are presented reasonable agreement with measured industrial
made from green petroleum coke, which should be calcined data. The objective of this work is to propose a model in
to remove moisture and volatile matter, and to acquire an order to improve the numerical results. The sub-models,
appropriate crystalline structure. This process is carried out introduced into the conservation equations (main model),
in a calcining rotary kiln. A typical kiln for petroleum coke account for variations in the model parameters and thermo-
is about 60 m long with a diameter of 2.5 m. The kiln is physical properties as a function of the state variables.
tilted by an angle of about 258 and rotates at speeds of 2
4 rpm. Air (primary and secondary) for combustion is
injected through the burner at the solids discharge end of
2. Mathematical modeling
the kiln. A tertiary air is injected into the calcining zone to
aid burning of the volatiles, which evolve from the bed in Petroleum coke calcination in rotary kilns will be
that section. The calcination of green petroleum coke in described by means of a model based on mass and energy
rotary kilns presents low operational costs due to the fact conservation. Heat transfer by conduction, convection and
that the process is energy self-sufcient (i.e. does not require radiation is taken into account. Mass transfer occurring
fuel), and that it presents higher processing capacity between the solid and gas phases is mainly due to natural
compared to the conventional processes. Also, since it is a gas combustion reactions, release of coke volatile compo-
nents (hydrogen, methane and tar), coke combustion and
* Corresponding author. Tel./fax: 155-31-3238-1789. release of coke particles into the gas phase. Fig. 2 presents
E-mail address: leandro@deq.ufmg.br (L.S. Oliveira). the control volumes used to develop the model, representing
0016-2361/01/$ - see front matter q 2001 Published by Elsevier Science Ltd.
PII: S 0016-236 1(01)00032-1
1612 M.A. Martins et al. / Fuel 80 (2001) 16111622

Nomenclature ReD Reynolds number


Rs combustion rate of carbon per unit supercial
A area per unit axial length (m)
Ag cross-sectional area of gas ow in the kiln (m 2) area (kg m 22 s 21)
Sc Schmidt number
Ags chord length (m)
Sh Sherwood number
As cross-sectional area of solids ow in the kiln (m 2)
T temperature (K)
A1 constant for carbon dioxide desorption kinetics
u velocity (m s 21)
equation (kg m 22 s 21 Pa 21)
U gas phase velocity (m s 21)
A2 constant for carbon dioxide desorption kinetics
X mass fraction in coke bed (kg kg 21)
equation (kg m 22 s 21)
Y mass fraction in gas phase (kg kg 21)
c1 characteristic constant for Eq. (3)
Cp heat capacity (kJ kg 21 K 21) Greek symbols
Cn characteristic constant for Eq. (36) (m 3 kg 21)
a absorptivity
a thermal diffusivity (m 2 s 21)
d particle diameter (m)
D external diameter of kiln (m)
1 emissivity
D diffusion coefcient of oxygen in nitrogen (m2 s 21)
1 bed porosity
De hydraulic diameter (m)
f dynamic angle of repose (rad)
e thickness (m)
G angle of solids ll (rad)
ea effective thickness for mixture of oxygen and
h fraction of solids ll
bed particles (m) w angle by which the solids advance axially (rad)
E activation energy (J mol 21)
l relaxation parameter
m viscosity (kg m 21 s 21)
E release energy (J mol 21)
n kinematic viscosity (m 2 s 21)
g acceleration due to gravity (9.8 m s 22)
G mass owrate (kg s 21)
n volatile component
Go heat transfer dimensionless parameter
n kiln slope angle (rad)
r density (kg m 3)
GrD Grashoff number
s StefanBoltzmann constant (5.67 10 28Jm22-
h heat transfer coefcient (kJ m 22 s 21 K 21)
h bed height (m) K 24 s 21)
Hj sensible heat for species j (kJ kg 21)
t residence time (s)
DHk heat of reaction for species k (kJ kg 21)
V form factor for the kiln internal wall and the
solids bed
k thermal conductivity (kJ m 21 K 21 s 21)
v kiln rotational speed (rad s 21)
k0 characteristic constant for Eqs. (38) and (39) (s 21)
Subscripts
k0 velocity constant for Eq. (42) (kg m 22 s 21 Pa 21)
0 initial value
k1 velocity constant for absorption (kg m 22 s 21 Pa 21)
a carbon steel
k2 velocity constant for dessorption (kg m 22 s 21)
K characteristic constant for Eqs. (4) and (32) ap apparent
b coke bed
mi mass owrate of species i per unit length of kiln
c carbon, coke
(kg m 21 s 21)
ce0 convective, external wallambient
Mi molecular weight of species i (kg kmol 21)
cgs convective, gassolid
Nu Nusselt number
cgw convective, gaswall
Num average Nusselt number
cws conductive, wallsolid
Po partial pressure of oxygen (Pa)
d particle in gas phase
Pr Prandtl number
Q heat owrate per unit length of kiln (kJ m 21 s 21) e water, external
g gas phase
Qr;k heat consumption/production rate of species k
i ! e heat conduction in steel wall
(kJ m 21 s 21)
o oxygen
rp particle radius (m)
p particle in solid phase
R universal gas constant (8.314 J mol 21 s 21)
r refractory
Rh hydrogen production rate in the bed per unit
re0 radiative, external wallambient
axial length (kg m 21 s 21)
rgs radiative, gassolid
Ri rate of production or consumption of species i
per reaction (kg m 21 s 21) rgw radiative, gaswall
rws radiative, wallsolid
Ri internal radius of kiln (m)
s solid phase
Rk reaction rate of species k per unit axial length
w internal wall
(kg m 21 s 21)
w ! i heat conduction in refractory layer
Ra Rayleigh number
M.A. Martins et al. / Fuel 80 (2001) 16111622 1613

Fig. 1. Schematic diagram of a typical petroleum coke rotary kiln.

the kiln cross-section with respect to the axial (x) direc- uidization due to ow of volatile components released
tion. The differential control volumes are dened as from the bed and properties dependence on temperature.
dVg Ag dx for the gas phase and dVs As dx for the There is an extensive amount of work on solids ow in
solid phase. Ag and As correspond to the cross-sectional rotary kilns [615]. The solid transport model presented
areas normal to the gas and solid ows, respectively, here is based on the studies developed by Perron and Bui
and dx is the differential length in the axial direction. [15]. Fig. 3 shows a schematic view of the bed and variables
Each dependent variable is represented by an average used for model development.
value in the kiln cross-section, for each control volume. It was observed from measured data in the coke
Assuming steady-state conditions, the mass conservation industry [16] that the axial velocity prole and the
equations can be written as mass of volatiles prole presented the same type of behavior
d d and were therefore correlated. Based on the rheological
u A r mi 1 Ri ; u A r mi 1 Ri 1 behavior of the coke, Perron and Bui [15] assumed that
dx s s i dx g g i
the dynamic angle of repose (f ) varies inversely with the
The energy equations were developed assuming negligible hydrogen production rate in the bed (Rh). Perron et al. [4]
viscous dissipation and that the variations in kinetic and presented the equation:
potential energy are not signicant. Previous studies [4]
have established that such assumptions are valid. The result- 1
ing equations are f 3
1=f0 1 c1 Rh
dTs Xm Xn Xo
Gs Cps Qi 1 Hj mj 1 DHk Rk ; This rate depends on temperature. Perron et al. [4] also
dx i1 j1 k1 presented an equation for evaluating bed velocity in the x
2 direction. However, this equation can be applied only if the
dTg Xm Xn Xo
Gg Cpg Qi 1 Hj mj 1 DHk Rk dynamic angle of repose (f ) is larger than the kiln's slope
dx i1 j1 k1 (n ). At high temperatures (,1600 K), an increase in the
hydrogen release rate leads to a reduction in the dynamic
The sub-models related to the terms in Eqs. (1) and (2) are
angle of repose, attaining values smaller than the kiln's
described in the following sections.
slope. Thus, that equation is not suitable for these conditions
2.1. Solid transport equations and, in the present work, it will be replaced by the equation
presented by Perron and Bui [15]:
Modeling solids ow in a rotary kiln is quite complex,
since physical, chemical and rheological properties vary nR2i
us K 
p 4
along the bed. The major factors affecting bed motion in
h2 cos f 1 2h 2hRi 2 h2 sin f
the axial direction are bed height variation, particles

Fig. 2. Control volumes denition for the rotary kiln model.


1614 M.A. Martins et al. / Fuel 80 (2001) 16111622

Fig. 3. Schematic view of the variables used in the solids transport and mass transfer models in rolling bed ow regime.

2.2. Heat transfer equations internal wall is evaluated according to the relationship given
by Ketlakh and Tsibin [19]:
The equations that describe heat transfer by conduction,
convection and radiation, as well as the equations used for Qrws sA rws 1w 1s 1 2 1g VT w4 2 T 4s 11
temperature evaluation at the kiln internal wall, are generic
Heat transfer by conduction between the coke bed and the
and applicable to coke calcination. Heat uxes are cal-
kiln internal wall is given by
culated using empirical correlations, specic for rotary
kilns. Heat transfer by radiation between the gas phase Qcws hcws Acws Tw 2 Ts 12
and the bed, and between the gas phase and the kiln internal
The heat transfer coefcient (hcws) is evaluated using the
wall, can be evaluated by the equations developed by Hottel
empirical correlation developed by Tscheng and Watkinson
and Sarom [17]. These equations, valid for 1 $ 0.8 can be
[20]:
written as
! !0:3
1g Tg4 2 ag Tk4 kb vR2i G
Qrgk sA rgk 1k 1 1 ; k s; w 5 hcws 11:6 13
2 Acws ab

Heat uxes by convection between the gas phase and the Chemical reactions occurring during petroleum coke
bed, and between the gas phase and the kiln internal wall, calcination involve heat release or consumption. The heat
are evaluated as consumption/production rate can be evaluated as

Qcgk hcgk Acgk Tg 2 Tk ; k s; w 6 Qr;k ^Rk DHk 14

Tscheng [18] studied the effects of gas ow in the axial The temperature at the kiln internal wall is evaluated by
direction, of kiln rotation and of kiln degree of ll on the solving the equations that describe heat transfer through the
heat transfer in a pilot scale rotary kiln. Based on the experi- wall. Heat uxes due to convection, conduction and radia-
mental data obtained for sand and limestone, the heat tion penetrate the internal wall comprised of refractory
transfer coefcients (hcgs and hcgw) were calculated as bricks. The heat is transferred in the radial direction,
through the wall refractory and steel layers. The heat is
kg 0:535 0:104 20:341 then released at the kiln external wall. A schematic repre-
hcgs 0:46 Re Rev h 7
De D sentation of the heat uxes is presented in Fig. 4. The heat
ux by conduction in the angular direction (regenerative
kg 0:575 20:292 ux) was not considered in the model. The experimental
hcgw 1:54 Re Rev 8
De D data obtained by Barr et al. [21] in a pilot scale kiln showed
that the internal kiln wall temperature varied up to 20% in
The hydraulic diameters De, ReD and Rev are evaluated
as follows: the angular direction, for a refractory material with thermal
conductivity of 5 J m 21 K 21 s 21. The refractory materials
Ag that comprise walls of typical kilns used for petroleum
De 4 9
Acgs 1 Acws coke calcination have thermal conductivity of approxi-
mately 0.7 J m 21 K 21 s 21 [4]. According to the results
rg ug D e rg vD2e presented by Barr et al. [21], for this conductivity value,
ReD ; Rev 10 the internal kiln wall temperature varies up to 3.5% in the
mg mg
angular direction. Therefore, the heat ux by conduction in
Heat ux by radiation between the coke bed and the kiln the angular direction is not signicant and can be neglected.
M.A. Martins et al. / Fuel 80 (2001) 16111622 1615

coefcient is evaluated as [23]


D
Num hce0
k0
1=2
" !a #b
0:62ReD Pr 1=3 ReD
0:3 1 11 ; 22
1 1 0:4=Pr2=3 1=4 28 200

5 4
a ; b : 102 , ReD , 2 104 ;
8 5

1
a ; b 1 : 4 104 , ReD , 107
2
where

rDu1 gD3 ur 2 re u
ReD ; GrD ;
Fig. 4. Schematic representation of heat uxes through the bed and kiln m n2 r
wall.
23
mCp
Pr
k0
An energy balance in radial direction leads to the follow- In the case of natural convection Re2D , GrD ; the heat
ing equations: transfer coefcient is evaluated as [23]

Qrws 1 Qcws 1 Qrgw 1 Qcgw Qw!i 15 D 0:386Ra1=6


Num hce0 0:6 1 ;
k0 1 1 0:559=Pr9=16 8=27 24
24 12
10 , Ra , 10
Qw!i Qi!e 16
where
Ra GrD Pr 25
Qi!e Qce0 1 Qre0 17
In order to evaluate the kiln internal wall temperature
Heat uxes due to conduction are evaluated as [22] (Tw), the temperature at the steel/refractory interface (Ti)
and the kiln external wall temperature (Te), Eqs. (5), (6),
2p (11), (12) and (18) are applied to Eq. (15), Eqs. (18) and
Qw!i T 2 Ti 18 (19) are applied to Eq. (16), and Eqs. (19)(21) are applied
kr lnRi 1 er =Ri w
to Eq. (17), generating a system of non-linear equations.
This system of equations is solved by the Picard iteration
scheme coupled to a relaxation formula [24], with T 4w and
2p T 4e approximated by
Qi!e   T 2 Te 19
ka ln Ri 1 er 1 ea =Ri 1 er i
T 4w Tw;n Tw;n21
3
26
Steel thermal conductivity is given by the relationship
47.3 1 1.44 10 22Tw J m 21 K 21 s 21 [15]. The heat losses T 4e Te;n Te;n21
3
27
due to radiation and convection at the kiln external wall
are given by [23] where the subscripts n and n 2 1 correspond to the present
and previous iterations, respectively. This results in a line-
arization procedure, since temperature values at the
Qre0 sA e 1e T 4e 2 T 40 20 previous iterations are known. Convergence is improved
by use of the relaxation equations:
Tw;n lTw;n 1 1 2 lTw;n21 28
Qce0 hce0 Ae Te 2 T0 21
Te;n lTe;n 1 1 2 lTe;n21 29
Heat transfer by convection between the kiln external
wall and the surroundings can be forced or natural. In the with optimal values of l 0:35 and 0.8 for the internal (Tw)
case of forced convection Re2D . GrD ; the heat transfer and external (Te) kiln wall temperatures, respectively.
1616 M.A. Martins et al. / Fuel 80 (2001) 16111622

Table 1
Empirical constants for evaluation of volatile components release and combustion rate [26]

Volatile component k0 (s 21) E (J mol 21) (release) n Cn (m 3 kg 21) a b E (J mol 21) (combustion)

Hydrogen 9.167 10 1 43,681 1.5 20 1 1 75,240


Methane 1.496 10 6 175,017 2.0 30 0.7 0.8 196,400
Tar 1.09 10 21 125,567 1.5 10 1 0.5 131,670

2.3. Mass transfer equations This equation presented good agreement with experimental
data.
Mass transfer during petroleum coke calcination in rotary
kilns is due to transport of water, volatile components and 2.3.3. Kinetics of combustion of the volatile components
coke particles from the bed to the gas phase, to combustion Combustion of volatile components occurs as those
of coke particles in the bed and in the gas phase, and to substances are released from the bed to the gas phase. The
combustion of volatiles in the gas phase. major chemical reactions can be described by the following
equations [26]:
2.3.1. Kinetics of release of water and volatile components 2H2 1 O2 ! 2H2 O 1 DHh 33
from the coke bed
According to the experimental data presented by Lyons et CH4 1 2O2 ! 2H2 O 1 CO2 1 DHm 34
al. [25], water release from the bed to the gas phase can be
described as a rst-order reaction. The resulting equation is C18 H12 1 21O2 ! 6H2 O 1 18CO2 1 DHg 35
 
XG According to Dernedde et al. [26], the combustion rate
Re k0 c c Xe e2E=RTs 30
ub can be described by the empirical equation
where k0 is a characteristic constant (2.549 10 7 s 21) and Rcn Cn Ag tr r2g Yna Yob e2En =RTg 36
E the release energy (41.942 10 3 J mol 21). Volatile
The empirical constants for Eq. (36) are presented in
components are transferred from the solid to the gas phase
Table 1.
as the temperature increases inside the kiln. Studies of
The residence time of each volatile component in the gas
kinetics of volatile decomposition developed by Dernedde
phase (t r) is evaluated as
et al. [26] concluded that the major volatile components
released during coke calcination were hydrogen, methane dx
tr 37
and tar. An empirical correlation was developed for the ug
decomposition of theses volatiles:
    where dx corresponds to the length of the gas phase control
XG Xn n 2E=RTs volume in the axial direction.
Rn k0 c c Xn0 e 31
ub Xn 0
2.3.4. Kinetics of combustion of the coke bed
where k0, n and E were evaluated from thermogravimetric Combustion of petroleum coke particles occurs in the
analysis data and are presented in Table 1. presence of oxygen and can be regarded as a heterogeneous
reaction. Essenhigh et al. [28] developed a model for the
2.3.2. Kinetics of release of coke particles to the gas phase combustion of coke particles based on three different stages.
The turbulent ow of gas inside the kiln and the kiln
rotation cause the release of coke particles from the bed to
the gas phase. A schematic view of the solid particles
release mechanisms is presented in Fig. 5.
Li [2] veried that there are two distinct mechanisms of
release of bed particles into the gas phase. Some particles
are thrown outside the bed due to the wall roughness, and
the smaller particles are carried by the gas. Coke particles
can also be dragged to the gas phase by the vortices
produced by the gas ow in the angular direction, opposite
to the direction of ow of the particles in the bed surface.
Based on these two mechanisms, Li [27] proposed an
empirical correlation for the particles release rate:

Rp 2K2vGc Ri 1=2 Tg3=4 ug rg R2i 1 2 h1=2 32 Fig. 5. Release mechanisms of coke particles to the gas phase.
M.A. Martins et al. / Fuel 80 (2001) 16111622 1617

In the rst stage, the oxygen diffuses through a boundary functions, dened as
layer in the gas phase (combustion and inert gases). The
second stage is characterized by the absorption of oxygen k1 A1 e2E1 =RT ; k2 A2 e2E2 =RT 46
by the particles, with activation energy E1. The last stage 5
Essenhigh et al. [28] evaluated A2 as 1.05 10 , based on
consists of the reaction itself followed by desorption of the experimental data. A1 can be evaluated using the equation:
particles combustion products, with activation energy E2.  
These stages are related to the rate of combustion (Rsc) by MC rg T0 1=2
A1 47
the equation 2Mg2 T
1 1 1 1 A sensitivity analysis was performed using the models
1 1 38
Rsc k0 P o k 1 Po k2 presented by Essenhigh et al. [28] and Lanauze and Jung
[29] for evaluation of the diffusion rate constant. The results
This model was developed based on the following assump-
showed that there is a signicant difference between the
tions: (i) bed particles are spherical with average bed
models for low oxygen concentrations (Yo , 0.05 kg kg 21),
temperature (Ts); (ii) foreign material is inert and does not
low gas velocities (ug , 1 m s 21), and high gas temperatures
affect the rate of combustion; (iii) particle pores are not
(T . 1900 K). The combustion rate dependency on gas
explicitly considered in the coke particles combustion
velocity becomes more signicant at low oxygen con-
model; (iv) combustion of the coke particles occurs solely
centrations (Yo , 0.05 kg kg 21), since this rate is strongly
in the bed layer in contact with oxygen.
affected by the diffusion process. If the gas velocity is zero,
The diffusion rate constant depends on particle size, and
the differences between the models become quite signi-
on the surroundings temperature and velocity:
!   cant. This is due to the differences in evaluation of the
r0 M C D T a21 Nusselt and Sherwood numbers (Eqs. (40) and (43)) when
k0 Nu 39 the Reynolds number is equal to zero. Since the model by
M o rp T0
Lanauze and Jung [29] takes into account the bed porosity, it
The Nusselt number is evaluated as will be used in the present study.
Nu 2 1 0:69Rem Pr n 40 The bed particles combustion rate is a function of the gas
velocity over the particles. Due to the fact that there is no
with solid boundary to force the gas ow through the bed, the
4rg Urp mCpg interstitial velocity within the bed is much lower than the
Re ; Pr 41 gas velocity in the free stream. This interstitial velocity can
mg kg
be neglected in the evaluation of the bed combustion rate
It is noteworthy to mention that the model developed by and of its thermal decomposition. Fig. 3 presents the velo-
Essenhigh et al. [28] was based on the combustion of a city vectors considered over the bed of particles. The vectors
single particle. Therefore, Eqs. (39) and (40) do not account of modulus ug, ub and ur represent the axial velocity of the
for bed porosity. gas phase, the axial velocity of the bed and the particles
Lanauze and Jung [29] evaluated the diffusion rate rolling velocity at the bed surface, respectively. The effec-
constant using experimental data obtained for coke com- tive gas velocity over the coke bed can be calculated by the
bustion. The resulting equation is presented as follows: equation
q
3 103 M0 D ue ug 1 ub 2 1 u2r 48
k0 Sh 42
16RTrp
which in turn can be utilized in the evaluation of the
with the Sherwood number evaluated as Sherwood number in Eq. (43). Once the specic combustion
Sh 21 1 0:69Rem n rate (Rsc) is dened, it is necessary to dene the effective
p Sc 43
combustion area. The equation proposed by Perron et al. [4]
The Reynolds (Rep) and Schmidt (Sc) numbers for the coke is not suitable for the denition of this area since it assumes
bed are dened as a rectangular cross-section and it depends on experimental
4rg Urp mg data related to the active layer of mixture. Therefore, in the
Rep ; Sc 44 present work, a new model for the bed combustion rate is
1mg rg D
proposed.
where 1 is the bed porosity, dened as According to Kohav et al. [30], the petroleum coke bed
rap can be described as a rolling bed, based on the kiln rotation
112 45 speed and degree of ll. A schematic view of a rolling bed is
rp
presented in Fig. 3, v is the kiln rotational speed, D the
where r ap and r p correspond to the apparent and particle thickness of the rolling bed (m), h the bed height in the radial
densities, respectively. The velocity constant for absorption direction (m), uv the average particle velocity in the
and desorption (k1 and k2) are Arrhenius-type temperature stationary layer (m s 21) and ur the average particle velocity
1618 M.A. Martins et al. / Fuel 80 (2001) 16111622

G2
t2 55
v
Once the residence times (t 1 and t 2), the cross-section areas
(A1 and A2) and the velocity of the rolling layer (ur) are
known, the bed combustion rate can be evaluated by
Eq. (49). Notice that in Eq. (49), Rsc1 and Rsc2 are non-
zero since their values are different in each bed layer.
They present different absorption and desorption rates in
each layer (see Eq. (38)). Also, Eq. (49) is generic enough
to describe situations where calcination reactions will occur
within the stationary layer (e.g. limestone).
Fig. 6. Schematic representation of the mass transfer in rolling bed ow
regime.
2.3.5. Combustion kinetics of the coke particles in the gas
phase
in the rolling region (m s 21). In the rolling layer, the rolling
The combustion kinetics of the petroleum coke particles
particles are mixed to the gas phase, with effective velocity
evolved from the bed to the gas phase is described by
evaluated according to Eq. (48). This effective velocity can
Eq. (38) for which the following assumptions were made:
be neglected in the stationary layer. The mass transfer can
(i) the coke particles in the gas phase are spherical, with
be represented by two CSTR reactors connected in series, as
diameter dd, and are at the gas temperature Tg; (ii) the
shown in Fig. 6. t 1 and t 2 correspond to the reactors
relative velocity between the particles in the gas phase
residence times, and V1 and V2 are the reactor volumes per
and the gas stream is zero; (iii) the effective surface area
unit length, corresponding to the cross-section areas of the
of the coke particles in the gas phase is assumed to be equal
rolling and stationary layer, respectively.
to the total surface area of the particles in the gas stream.
The combustion rate can be evaluated as
  Hence, the effective combustion area of the particles in the
61 2 1 Rsc1 A1 t1 1 Rsc2 A2 t2 gas phase can be dened as
Rbc 49
dp t1 1 t2 rg G n
Aegd 6Yd Ag 56
In order to obtain the residence times (t 1 and t 2) and the rd G g dd
cross-section areas (A1 and A2), the velocity of the rolling
region must be evaluated. Kohav et al. [30] presented the 2.4. Model unique features
equation
s The unique features of the model presented in this study
8 Ri g sinG 1 =2 sin f are related to the sub-models and are discussed as follows.
ur 50
5 cos w Comparisons will be made to the model by Perron et al. [4],
where G 1, Ri, f and w are dened in Fig. 3. The cross- since it was the most complete model available in the
section areas (A1 and A2) and the thickness of the rolling literature.
layer (D ) are given by
" !! !! Thermophysical properties of the gas phase: Perron et al.
2 21 D G1 D G1 [4] used constant average values for the thermophysical
A 1 Ri cos 1 cos 2 1 cos properties of each component in the gas phase, whereas
Ri 2 Ri 2
in this study, thermophysical properties of each com-
v
!!2 #
u ponent were described as functions of the temperature
u D G1
t
 12 1 cos (51) and pressure of the gas phase.
Ri 2 Heat transfer in the gas phase: in this study, the radiative
properties (emissivity and absorptivity) of the particles in
A 2 v ur D 52 the gas phase were given by a generic theoretical model
[31], whereas the model by Perron et al. [4] was
empirical and specic for coke particles.
A1 1
R2i G 1 2 sin G 1 2 A2 53
2 Solid transport: the model used in this study takes into
Eqs. (51)(53) comprise a system of equations for which account the bed height and presented better agreement
the unknowns are A1, A2 and D . The residence times with experimental data [13] in comparison to the model
are evaluated using the relationships developed by Kohav used by Perron et al. [4].
et al. [30]: Combustion kinetics of the coke bed: the model used in
s this study presents a more consistent theoretical basis
Ri sinG 1 =2 than the model used by Perron et al. [4]. It takes into
t1 2 54
g sin f cos w account the particle residence time within the stationary
M.A. Martins et al. / Fuel 80 (2001) 16111622 1619

and rolling layers of the bed in a rolling bed type of ow of ow requires a good estimation of the initial value for the
regime. Also, the model for the specic combustion rate dependent variables in the equations describing the ow in
takes into account the porosity of the bed, whereas the the opposite direction to the axial direction. It was also
model by Perron et al. [4] considers the kinetics of a shown that this kind of solution methodology will present
single particle only. convergence problems in cases where there are steep
Combustion kinetics of the volatile components: the gradients in the source terms, in the axial direction, as is
model used by Perron et al. [4] does not take into account the case for the calcination of petroleum coke in rotary kilns.
the volatiles residence time in the gas phase. Even though In the iterative solution methodology, the differential
combustion of the volatile components occurs fast, the equations are solved following the physical direction of
gas phase velocity is high enough (1825 m s 21) so that the ow in each phase. The application of an iterative solu-
the residence and combustion times are of the same order tion methodology to rotary kiln problems consists of rst
of magnitude. Hence, the model used in this study solving the equations for the solids bed in the axial direc-
incorporates the residence time. tion, starting from the solids feed end of the kiln. Since the
equations for both phases are coupled, the initial values for
the composition, temperature and velocity of the gas phase
are guessed. When the procedure reaches the solids
3. Solution methodology discharge end of the kiln, the solution of the equations for
the gas phase is then started in the opposite direction, until
Within rotary kilns, the axial gas ow is in the opposite the solids feed end of the kiln is reached again. The solution
direction to the axial ow of solids (countercurrent). In methodology is then re-started in the direction of the solids
previous studies [15,3234], several researchers presented ow and the values used for composition, temperature and
solution methodologies for their proposed models in which velocity of the gas phase are the ones calculated in the
the ordinary differential equations were solved simulta- previous iteration. This procedure is repeated until con-
neously in the axial direction. An alternative methodology vergence is reached for the composition, temperature and
is to solve the equations in an iterative manner, solving the velocity proles of both phases. In this work, the iterative
equations for one phase, in the direction of the solids ow, solution methodology was employed, since it does not
and then solving for the other phase in the opposite require a good estimate for the initial composition, tempera-
direction. Different methodologies for the solution of ture and velocity values of the gas phase at the solids feed
countercurrent ow problems were discussed by Franks end of the kiln.
[35] in a comparative manner. It was demonstrated that The information ow diagram for the petroleum coke
the simultaneous solution of the equations for both directions calcination model is depicted in Fig. 7. The need for an

Fig. 7. Information ow diagram for the petroleum coke calcination model.


1620 M.A. Martins et al. / Fuel 80 (2001) 16111622

iterative solution procedure is readily seen from the inter- Table 4


action of the sub-models. It is noteworthy to call the Coke composition at the solids load end of the kiln [5]
attention for an inner iterative loop where the internal Component % p/p
wall temperature is evaluated. The numerical methods
tested for the solution of the ODE's comprising the model Carbon 80.00
were a fourth-order RungeKutta method, the fourth-order Water 9.250
Hydrogen 7.104
AdamsBashforth/AdamsMoulton predictioncorrection Methane 3.275
method, and the Hamming predictioncorrection method Tar 0.321
[36]. All these methods presented similar performance and
the fourth-order RungeKutta method was chosen for the
solution in the simulation runs due to its simplicity temperature proles was attained for dx smaller than
compared to the others. 0.75 m.
The schematic diagram of the petroleum coke calcination
rotary kiln employed in the simulation runs is presented in
Fig. 1, and its design specications are presented in Table 2.
4. Implementation
The operational parameters of the calcination kiln and the
composition of the green coke are presented in Tables 3 and
The model for the petroleum coke calcination kiln is
comprised of 14 ordinary differential equations, correspond- 4, respectively.
The predicted temperature prole for the solids bed is
ing to the mass and energy conservation equations (1) and
presented in Fig. 8, where it is compared to measured values
(2). The dependent variables are hydrogen, methane, tar,
by Bui et al. [37]. The predicted prole shows a small
water and carbon mass fractions in the solid bed, hydrogen,
depression at about 5 m from the solids feed end of the
methane, tar, water, particles, oxygen and carbon dioxide
kiln, which is attributed to the beginning of an endothermic
mass fractions in the gas phase, and solid and gas tempera-
evolution of water from the bed. As the solids bed tempera-
tures. The system of equations was solved iteratively by a
ture increases, there is an intense evolution of volatile
fourth-order RungeKutta method. A sensitivity analysis
was performed in order to evaluate the optimal value for components from the bed to the gas phase, starting at
about 10 m and ending at 40 m, after which there is only
the increments in the axial direction (dx). Convergence of
combustion of the coke bed. Beyond 40 m, there is a
Table 2
decrease in the temperature of the bed due to the heat
Rotary kiln design specications [5] transfer between the bed and the gas phase, and between
the bed and the internal wall. The tertiary air entering the
Kiln Value Units kiln at 3641 m does not cause signicant effects on the
Length 60 m temperature prole. This is attributed to the difference in
Internal radius 1.06 m order of magnitude of the heat capacities of the gas phase
External radius 1.23 m and the solids bed, the former being much smaller. The
Kiln slope 2.4 8 tertiary air affects only the gas temperature prole, since it
First tertiary air entrance position a 36.44 m
is injected in the gas stream. The effect of the introduction of
Second tertiary air entrance position a 37.66 m
Third tertiary air entrance position a 38.88 m the tertiary air is mildly felt by the bed up to the feed end of
Fourth tertiary air entrance position a 40.10 m the kiln due mostly to its oxygen content rather than due to
Refractory thickness 120 mm
Kiln wall thickness (carbon steel) 50 mm
a
0 m is dened as the solids load end of the kiln [5].

Table 3
Rotary kiln operational parameters [5]

Parameter Value Units

Petroleum coke owrate 5.555 kg s 21


Primary air owrate 0 kg s 21
Secondary air owrate 5.255 kg s 21
Tertiary air owrate 2.6 kg s 21
System pressure 1 atm
Rotation speed 2.58 rpm
Petroleum coke initial 25 8C
temperature
Secondary air initial temperature 25 8C
Tertiary air initial temperature 25 8C
Fig. 8. Temperature proles for the solids bed, gas phase and internal wall.
M.A. Martins et al. / Fuel 80 (2001) 16111622 1621

temperature difference. The differences between the The temperatures of the wall are signicantly affected by the
predicted temperature prole and that obtained by temperature proles for the solid and the gas phase. The
Perron et al. [5] are justied by the difference in the wall temperature reaches a maximum at about 20 m from
models used for description of the solids bed combus- the solids feed end of the kiln and its value (1796 K) is
tion. Bui et al. [37] have measured the bed temperature within the range of maximum wall temperatures
at nine points in a rotary kiln at ALCAN, in Canada reported in the literature: 16502000 K [5]. The wall
(Fig. 1). The kiln studied in this work presents the same temperature prole is strongly dependent on the gas
design specications as the one at ALCAN, and the and bed temperature proles. Notice that the tempera-
simulation results of this study presented better agree- ture proles for the bed and internal wall cross each
ment with the values measured by Bui et al. [37] than other at about 27 m. In the kiln positions ranging
the results obtained by Perron et al. [5]. from 0 to ,20 m, heat is transferred from the wall to
The predicted temperature prole for the gas phase is also the bed, whereas, at positions greater than 20 m, heat is
presented in Fig. 8. The temperature progressively increases lost from the bed into the wall. This crossing is
from the end where the gas enters the kiln to the end it exits expected, since the temperature differences between
the kiln. Several are the factors contributing to this increase the bed and the wall at the opposite ends of the kiln are
in temperature: heat generated from the combustion of the necessarily in opposite directions. Recall that, between 0
coke particles suspended in the gas phase; heat generated and 20 m, the hot gas stream is the major contributor to
from the combustion of the volatiles evolved from the bed; the heating of the bed, being convection (from gas to bed)
heat transfer from the gas evolved from the combustion of a more efcient heating mechanism than conduction (from
the bed (,1450 K). The decrease in gas temperature wall to bed).
between 35 and 40 m is due to the entrance of the tertiary The simulation results obtained for the mass conservation
air at these points. The tertiary air is at a much lower equations presented similar behavior to those obtained by
temperature than the gas phase at that point (see Table 3). Perron et al. [5] for the mass fraction proles of all the
In the range between 20 and 30 m, there is an increase in the components in the gas and solid phases, except for the parti-
gas temperature due to the combustion of the volatiles cles concentration in the gas phase. The particle mass frac-
evolved from the bed. The gas temperature reaches its maxi- tion prole obtained in this study presented a smoother
mum (1995 K) at the lower end in this range. The maximum behavior than that from Perron et al. [5].
temperature predicted in this study is within the range Simulation results for the composition of the kiln exhaust
reported in the literature: 16502250 K [5]. However, its gas were compared to the values predicted by Perron et al.
value is lower than the one predicted in the simulations by [5] and to data measured by Bui et al. [37], as presented in
Perron et al. [5]. This can be attributed to the differences in Table 5. Except for the hydrogen mass fraction, the values
the models used for the combustion of the volatiles predicted in this study presented better agreement with
evolved from the bed, and for the evaluation of the measured data in comparison to the values predicted by
radiative uxes, where the absorptivity and emissivity Perron et al. [4]. According to Perron et al. [5], the differ-
were calculated in this study taking into account all ence between predicted and measured hydrogen mass frac-
bands for water, carbon dioxide and methane [31]. tions can be attributed to hydrogen absorption by tar
The gas temperature decreases from the point where it particles, which are separated from the exhaust gas by ltra-
reaches a maximum to the point where the gas exits the kiln. tion before the analysis by gas chromatography. The solid
This is due to the heat transferred by convection from the particles ow rate in the gas phase was evaluated by divid-
gas to the relatively cold bed and wall at that portion of the ing the ltered mass of particles by the sampling time [37].
kiln. The obtained value was 19.2 kg s 21. The present study
Fig. 8 presents the temperature prole for the kiln internal predicted a value of 18.75 kg s 21 (2.35% difference)
wall. The predicted prole did not present signicant differ- compared to 22.6 kg s 21 (7.73% difference) predicted by
ences compared to the simulations results by Perron et al. [5]. Perron et al. [5].

Table 5
Composition of the kiln exhaust gas

Mass fraction Percent difference (%)

Component Bui et al. [37] (measured) Perron et al. [5] (predicted) This study (predicted) Perron et al. [5] This study

O2
CO 0.005
CO2 0.116 0.094 0.117 19.18 20.65
H2 0.011 0.0275 0.0280 2150.00 2154.39
CH4 0.015 0.0125 0.0166 16.67 210.42
H2 O 0.184 0.208 0.179 213.29 2.76
1622 M.A. Martins et al. / Fuel 80 (2001) 16111622

5. Conclusions [12] Heiligesntaedt W. Thermique appliquee aux fours industriels. Paris:


Dunod, 1971.
[13] Sullivan JD, Maier CG, Ralston OC. Technical Paper, 384. US
A steady-state one-dimensional model that describes
Bureau of Mines, 1972.
petroleum coke calcination in a rotary kiln was presented. [14] Perry RH, Chilton CH. Manual de engenharia qumica. 5th ed. Rio de
The model provides a detailed description of the kinetics Janeiro: Editora Guanabara Dois SA, 1980.
and ow characteristics in the mass transfer processes [15] Perron J, Bui RT. Can J Chem Engng 1990;68:61.
between the solid bed and the gas phase. The effects of [16] Churchill TR, Thompson CJ. Determination of rate of transport of
temperature and hydrogen rate release on the solid bed materials in a rotary kiln by radioactive tracers at Alcan, 1968.
Rapport Interne No. CPSR-176. Ottawa: Atomic Energy of Canada
rheology were taken into account. A combustion model
Limited.
applicable to any reactive rotary bed system was introduced. [17] Hottel HC, Sarom AF. Radiative transfer. 1st ed. New York:
Conductive, convective and radiative heat transfer between McGraw-Hill, 1967.
the solid bed and the gas phase, between the solid bed and [18] Tscheng SH. PhD Thesis. Vancouver: University of British Columbia,
the kiln internal wall, and between the gas phase and the kiln 1978.
[19] Ketlakh GA, Tsibin IP. Ogneupory 1978;1:17.
wall were described by means of semi-empirical models.
[20] Tscheng SH, Watkinson AP. Can J Chem Engng 1979;57:433.
The simulation results of this study for the temperature [21] Barr PV, Brimacombe JK, Watkinson AP. Metall Trans B. 20B
prole of the solids bed and for the composition of the gas 1989:391.
phase presented better agreement with measured industrial [22] Pitts DR, Sissom LE. 1000 Solved problems in heat transfer
data than other simulation results available in the literature. Schaum's solved problems series. 1st ed. New York: McGraw-Hill,
1991.
[23] O zisik MN. Heat transfer. 2nd ed. Montreal: McGraw-Hill, 1985.
Acknowledgements [24] Reddy JN, Gartling DK. The nite element method in heat transfer
and uid dynamics. 1st ed. Boca Raton, FL: CRC Press, 1994.
[25] Lyons JW, Min HS, Parisot PF, Paul JF. Ind Engng Chem Process Des
The authors acknowledge nancial support from the Dev 1962;2:29.
Brazilian government agency CAPES. [26] Dernedde E, Charette A, Bourgeois T, Castonguay RT. Light Met
Proc Tech Sess AIME 105th Annual Meeting, 1986. p. 589.
[27] Li KW. AIChE J 1974;20:1031.
References [28] Essenhigh RH, Forberg R, Howard JB. Ind Engng Chem 1965;57:32.
[29] Lanauze RD, Jung K. Can J Chem Engng 1983;61:262.
[1] Li KW, Friday JR. Carbon 1974;12:225. [30] Kohav T, Richardson JT, Luss D. AIChE J 1995;41:2465.
[2] Li KW. AIChE J 1974;20:1017. [31] Edwards DK, Balakrisnan A, Int J. Heat Mass Transfer 1973;16:25.
[3] Brooks DG. AIME 118th Annual Meeting. AIME, 1989. p. 461. [32] Thurlby JA. Metall Trans B. 19B 1988:103.
[4] Perron J, Bui RT, Nguyen HT. Can J Chem Engng 1992;70:1108. [33] Silcox GD, Pershing DW. J Air Waste Mgmt 1990;40:337.
[5] Perron J, Bui RT, Nguyen HT. Can J Chem Engng 1992;70:1120. [34] Davis RA. Can J Chem Engng 1996;74:1004.
[6] Bayard RA. Chem Met Engng 1945;52:100. [35] Franks RGE. Modeling and simulation in chemical engineering.
[7] Friedman E, Marshall A. Chem Engng Prog 1949;45:482. 1st ed. New York: Wiley, 1972.
[8] Saeman WC. Chem Engng Prog 1951;10:508. [36] Mathews JH. Numerical methods for mathematics, science and
[9] Kramers H, Croockewit P. Chem Engng Sci 1952;1:259. engineering. 2nd ed. London: Prentice-Hall, 1992.
[10] Ronco JJ. Ind Quim 1965;20:605. [37] Bui RT, Perron J, Read M. Carbon 1993;31:1139.
[11] Zablotny WW. Int Chem Engng 1965;5:360.

You might also like