You are on page 1of 8

Fuel Processing Technology 134 (2015) 1825

Contents lists available at ScienceDirect

Fuel Processing Technology


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

CFD modeling of phase change and coke formation in petroleum


rening heaters
Xue-Gang Li a,b, Lu-Hong Zhang a,b,, Rong-Ya Zhang a,b, Yong-Li Sun a,b, Bin Jiang a,b,c,
Ming-Fang Luo c, Xin-Gang Li a,b,c
a
b
c

School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, People's Republic of China
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin University, Tianjin 300072, People's Republic of China
National Engineering Research Centre of Distillation Technology, Tianjin University, Tianjin 300072, People's Republic of China

a r t i c l e

i n f o

Article history:
Received 14 August 2014
Received in revised form 4 March 2015
Accepted 6 March 2015
Available online 20 March 2015
Keywords:
CFD
Phase change
Coke formation
Multiphase ow
Fired furnace

a b s t r a c t
A three dimensional computational uid dynamics (CFD) model is developed to predict the phase change and
reactions inside the tubes of red furnace. The model is constructed within the Eulerian framework in which
the liquid phase is treated as a continuous phase while gas and solid phases are both considered as dispersed
phases, and all phases are treated as interpenetrating continuum having separate transport equations. The thermal cracking kinetic model is validated against experimental data. Calculations are performed using the commercial packages ANSYS CFX 14.0. The temperature distribution, ow eld, liquidgas phase separation, and coke
formation are predicted for different severities of petroleum type and tube wall temperature. The information
predicted by the CFD model can be utilized in the optimal design of industrial red furnaces.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Fired tube furnace is a common type of heating units in the petroleum rening industry, including atmospheric & vacuum furnace in
crude unit, delayed coking furnace, and catalytic reforming furnace.
Oil runs inside the tube pass accompanied with heat transfer, and sometimes phase change and thermal reactions. A chronic problem usually
encountered is the coke formation inside the tubes of red furnaces
caused by a combination of high operating temperature, long residence
time, and inherent instability of the oil [15]. Coke will deposit and
accumulate on the inner wall of the tube. The coke layer causes progressively high tube metal temperatures (TMTs), in turn, promoting coke
formation. As the TMTs get close to the tube metallurgical temperature
limit, the furnace must be shut down to remove coke [6,7]. It has long
been established that the prediction of temperature distribution and
coke formation inside the tubes of red furnaces is necessary for the
operational safety and overall performance. Besides, the understanding
of these heat transfer and thermal reaction phenomena is helpful for the
ne-tuning and optimization of modern furnace design.
Because of progress in computer hardware and software and the
consequent increase of the calculation speed, the computational uid
dynamics (CFD) modeling technique becomes a powerful and effective
tool for understanding the complex transport and reaction phenomena
Corresponding author. Tel./fax: +86 22 27400199.
E-mail address: zhanglvh@tju.edu.cn (L.-H. Zhang).

http://dx.doi.org/10.1016/j.fuproc.2015.03.005
0378-3820/ 2015 Elsevier B.V. All rights reserved.

in chemical engineering processes. It provides a theoretical basis and a


computational technique for predicting the ow inside the tubes of
red furnaces. Since the oil is composed of a complex mixture of hydrocarbons, many of which unknown, it is difcult to predict the real reactions that occur for each real chemical species. A way to circumvent this
problem is by considering that the feed molecules can have its behavior
represented by a small number of representative molecules, called
pseudo-components. Kseoglu and Phillips [811] presented three
kinetic models with a group of pseudo-components for catalytic hydrocracking of Athabasca bitumen. The reaction rate constants were represented by an Arrhenius type expression. Activation energies and the
frequency factors were determined for each reaction model proposed.
Takatsuka et al. [12] proposed a ternary compatibility control diagram
for predicting coke formation based on their own reaction model.
Sandra et al. [13] developed a CFD model to study the coke formation induced by liquid hydrocarbon droplets on a hot surface, and they found
that the impact energy of the droplet and the hot surface temperature
are found to determine the impact behavior. Fontoura et al. [14]
attempted to model the thermal cracking and phase change inside the
tubes of red heaters by using CFD simulations, however only temperature eld and velocity streamlines were discussed without involving
coke formation aspect in their results and conclusions.
In this paper, a three-dimensional gasliquidsolid three-phase CFD
model is developed for predicting the phase change and reactions inside
the tubes of red furnace. The model is constructed within the Eulerian
framework in which the liquid phase is treated as a continuous phase

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

while gas and solid phases are both considered as dispersed phases.
Within the Eulerian multiphase ow model, all phases are considered
as interpenetrating continuum having separate transport equations,
and certain interphase transfer terms used in momentum, heat, and
other interphase transfer models, are modeled using the Particle
Model and the Free Surface Model. The temperature distribution, ow
eld, liquidgas phase separation and coke formation are predicted for
different severities of petroleum type and tube wall temperature. The
objective of this work is extending the studies by Kseoglu and Phillips
and Fontoura et al., and examining the extent to which CFD models can
be used as an investigative tool in industrial practice.
2. Mathematical model development

19

where Di,l represents the kinematic diffusivity of pseudo-component i in


the liquid phase and can be approximated by the kinematic viscosity in
turbulent ow.
Di;l i;l :

10

2.2. Eulerian multiphase ow model


The model considers the liquid phase as continuous phase, gas and
solid phase as dispersed phase, and all of them are treated as interpenetrating continuum having separate transport equations. For each phase
the volume-averaged continuity (mass) and momentum conservation
equations are shown as follows.

2.1. Thermal cracking kinetic model


2.2.1. Continuity equations
The model used in this work is based on the modied model of
Kseoglu and Phillips [811], in which the petroleum was represented
by S.A.R.A. (saturates, aromatics, resin, asphaltenes) and the pseudocomponents of distillate and coke were introduced to account for phase
change and thermal reaction respectively. As is known, asphaltenes are
originally present in crude oil but don't lead to coke formation or precipitates in a stable oil system unless changes occur in temperature, pressure
or mixture composition. In the present study, another pseudo-component
of mesoasphaltenes was introduced and considered to be the only reactant of coke formation, so the contribution of the asphaltenes in feed to
coke formation was deducted. The complete kinetic net is given in Fig. 1.
According to Fig. 1, the source terms for the pseudo-components are
expressed by:
Sdistillate k1 Y saturates;l

Ssaturates k2 Y aromatics;l k1 Y saturates;l

Saromatics k3 Y resins;l k2 Y aromatics;l

Sresins



k5 Y asphaltenes;l Y mesoasphaltenes;l k3 2k4 Y resins;l

r r u S
t

11

where , r, and u represent the density, volume fraction, and velocity


of phase respectively, S represents the mass source term. For gas
phase, Sg represents the mass source term induced by phase change.
Sg Sdistillate k1 Y saturates;l

12

For solid phase, Ss represents the mass source term induced by coke
formation.
Ss Scoke k6 Y mesoasphaltenes;l

13

For liquid phase, Sl represents the mass source term induced by both
phase change and coke formation.
Sl Scoke Sdistillate k6 Y mesoasphaltenes;l k1 Y saturates;l

14

4
2.2.2. Momentum equations

Sasphaltenes k4 Y resins;l k5 Y asphaltenes;l

Smesoasphaltenes k4 Y resins;l k5 k6 Y mesoasphaltenes;l

Scoke k6 Y mesoasphaltenes;l

where Yi,l represents the mass fraction of pseudo-component i in


liquid phase, k16 are the reaction rate constants. k1 and k6 account
for the phase change and coke formation rates respectively. k3 and
k 2 are the reaction rate constants of the sequential reaction,
resins aromatics saturates. The interconversion reactions of
mesoasphaltenes and asphaltenes with resins are considered to
have the same reaction rate constant as k 4 and k 5. All the reaction
rate constants are given by Arrhenius equations as:


1
B
exp A
ki
3600
RT

15

where M represents the interphase momentum transfer acting on


phase due to interaction with phase, and g is the gravitational force.
The Eulerian multiphase ow model only considers the interphase momentum exchange between the continuous phase and the dispersed
phase, in other words, only Mlg and Mls are considered while Mgs is ignored. Note that interfacial forces between the two phases are equal
and opposite, so the net interfacial forces sum to zero.
M M

16

where A is the pre-exponential factor and B is the activation energy.


The values of A and B for k16 are listed in Table 1.
The reactions for phase change (Reaction 1) and coke formation (Reaction 6) involve interphase transfer, so they are introduced as source
terms of continuity equations. Other reactions only occur within the
liquid phase, and so are introduced as source terms of mass fraction
transfer equations.
 



Si;l
r Y  r l l ul Y i;l l Di;l Y i;l
t l l i;l





T
r u  r u u r eff ; u u
t
r p M r g

2.2.3. Closure relationships


In order to solve the continuity and momentum equations for velocities, pressure, and volume fractions, additional equations are indispensable to relate the interphase momentum transfer term M and the
turbulent viscosities to the mean ow variables. The interphase momentum transfer term includes drag force, lift force, wall lubrication
force, virtual mass force, and turbulent dispersion force. For liquid
solid interphase interaction, only drag force is considered as the solid
phase fraction is very limited. The drag force coefcient is calculated
on the basis of the model proposed by Schiller and Naumann [15]
which is suitable for the modeling of sparsely distributed solid particles.

20

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

Fig. 1. Kinetic net of the thermal cracking and phase change of petroleum.

For liquidgas interphase interaction, the drag force coefcient is calculated by the Grace drag model [16] with a volume fraction correction exponent of 4. Lift force and wall lubrication force are calculated using the
model by Tomiyama [17], and the turbulent dispersion force is estimated using the model by Burns et al. [18]. Since the density difference between liquid and gas phases is large, the virtual mass force is ignored in
the simulation.

Fig. 2. Geometry model. (a) Geometry and boundary conditions; (b) mesh generation.

The model assumes that the turbulence viscosity is linked to the


turbulence kinetic energy and dissipation via the relation:
2

2.2.4. Free-surface model


A VOF-like method [19] called the Free Surface Model, which is
used to allow phase separation when incorporated with an inhomogeneous model, provided by CFX for multiphase calculations, is introduced
in the calculation of the interface between the uids. This model is applicable to a multiphase ow situation where the phases are separated
by a distinct interphase.

2.2.5. Turbulence model


The turbulence eld is treated as a homogeneous multiphase ow in
the present study, as generally recommended for Free Surface Model in
the CFX-14.0 User Manual [20]. The homogeneous model implies that a
single turbulence eld is solved using a single turbulence model, and
bulk turbulence equations are solved which are the same as the single
phase equations, except that the mixture density and the mixture viscosity are used. In this work, the standard turbulence model was
employed to model the turbulence eld and the values of and
come directly from the differential transport equations for the turbulence kinetic energy and turbulence dissipation rate:

 

u tur P P b

17


 

u tur C 1 P C 2 C 1 P b 18

t
where C1, C2, , and are constants, Pb and Pb represent the inuence of the buoyancy force, and Pis the turbulence production due to
viscous forces.

tur

!
19

where C is a constant. Then the effective viscosity is the sum of the molecular viscosity with the turbulent viscosity and can be expressed by:
eff tur :

20

In order to model the particle induced turbulence in the liquid phase


due to the presence of wakes behind the particles, the Sato Enhanced
Eddy Viscosity model [21] is used for turbulence enhancement in gas
liquid uid pair.
2.3. Heat transfer model
Since only incompressible and low speed compressible ows are involved, thermal energy model is employed to model heat transfer in
which the transport of enthalpy through the uid domain is calculated
while the effects of mean ow kinetic energy are not included. The phasic governing equation for static enthalpy can be written as:

r H  r u H  r T r
t

XN 

p
: u SE Q
Hs Hs
1

21

where H, T, and denote the static enthalpy, the temperature, and


the thermal conductivity of phase , SE and Q describes external
heat sources and interphase heat transfer to phase across interfaces
+
with other phases, and the term (+
Hs Hs) represents heat
transfer induced by interphase mass transfer.
Table 2
Mass fraction of pseudo-components at inlet.

Table 1
Characteristic parameters for the reactions.
Reaction

B (kJmol1)

1
2
3
4
5
6

29.9
23.43
13.57
17.07
19.34
27.54

175
136
85
96
103
168

Pseudo-component

AR, Yi

VR, Yi

Distillate
Saturates
Aromatics
Resins
Asphaltenes
Mesoasphaltenes
Coke

0
0.57
0.22
0.16
0.05
0
0

0
0.18
0.34
0.32
0.16
0
0

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

21

Fig. 3. Comparison between experimental and predicted coke formation rates at different temperatures. (a) 703 K; (b) 718 K; (c) 733 K; (d) 748 K.

In the present study, constant tube wall temperature is considered


and the external heat is added as the source term of governing equation.


SE hc A T w T nw;

22

Two-resistance model is employed for interphase heat transfer calculations between the continuous phase and the dispersed phases. A
zero resistance condition is specied on the dispersed phase sides.
This is equivalent to an innite uid specic heat transfer coefcient

where Tw is the wall temperature, and hc, A, and Tnw, denote the heat
transfer coefcient, wall contact area, and near-wall uid temperature
of phase . hc involves the use of turbulent wall functions.


hc

cp u
T

23

where cp is the heat capacity, and T+ and u are the dimensionless


temperature and the velocity scale relevant to wall function respectively.

Fig. 4. Comparison of pressure drop by CFD simulation and correlation.

Fig. 5. Gas phase holdup. (a) Gas phase holdup prole of AR, Tw = 873 K; (b) gas phase
holdup at different severities of petroleum type and tube wall temperature.

22

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

3.2. Boundary conditions


3.2.1. Inlet
A uniform liquid inlet velocity prole of 2 m/s is imposed at the entrance. In order to evaluate the effect of oil composition on phase change
and coke formation, comparative simulation cases are conducted with
vacuum residuum (VR) and atmospheric residue (AR) of Arabian light
as listed in Table 2. The oil temperature is 673 K at the beginning of
the tube and there is only liquid phase present.
3.2.2. Outlet
The outlet boundary is specied as average static pressure of atmospheric pressure with pressure prole blend of 0.5 over whole outlet.
This is the most commonly used option. The average constraint is applied by comparing the area weighted pressure average over the entire
outlet to the specied value. The pressure prole at the outlet is shifted
by this difference such that the new area weighted pressure average will
be equal to the specied value. The ow direction is an implicit result of
the computation.
3.2.3. Wall
Constant temperature boundary condition is imposed at the tube
wall. In order to investigate the effect of tube wall temperature on
phase change and coke formation, comparative simulation cases with
tube wall temperature of 773 K and 873 K are conducted. A no-slip
wall boundary condition is specied for the liquid phase and a freeslip wall boundary condition for the vapor and solid phases.
3.3. Solution algorithms

Fig. 6. Liquid temperature. (a) Liquid temperature prole of VR, Tw = 873 K; (b) liquid
temperature distribution at different severities of petroleum type and tube wall
temperature.

and its effect is to force the interfacial temperature to be the same as the
phase temperature. RanzMarshall Correlation [22] is used to calculate
the dimensionless Nusselt number of the liquid phase side.

0:5

Nu 2 0:6Re

0:3

Pr

0 Re b 200 0 Pr b 250

24

This is based on boundary layer theory for steady ow past a spherical particle and is the most well tested correlation.

Good initial guesses of the ow variables are important not only to


avoid a signicantly longer computational time but also in some cases
to avoid divergence. At the start of a simulation, the tube shown in
Fig. 1 is lled with a uniform liquidgassolid dispersion, which has
the same temperature and composition as the feed, and each phase
has a Cartesian velocity component U of 2 m/s and component V/W of
0 m/s.
The simulations are carried out using the commercial packages
ANSYS CFX 14.0 and run on a HP Z800 workstation with two Intel
Xeon X5687 3600 MHz processors used in parallel and 96 GB RAM. A
high resolution scheme is used to discretize the convection terms in
order to reduce the numerical diffusion errors. To obtain converged results, the target value of the normalized residual for each variable is set
to 105, as generally recommended in the CFX-14.0 User Manual [20].
Functions of averaged liquid phase temperature and vapor phase
volume fraction at outlet are wrote using CEL [20] and taken as monitor
expressions in the output control. During the simulation the two

3. Geometry model and solution algorithms


3.1. Geometry model and mesh generation
As shown in Fig. 2(a), a horizontal cylindrical geometry of 18 m in
length and 0.152 m in diameter is constructed to represent a segment
of red furnace tube. The computation domain is discretized using a
structured mesh by means of ICEM CFD 14.0. The grids near the tube
wall are rened in order to consolidate the calculation in the boundary
layer region as shown in Fig. 2(b). The whole computation domain is
represented using the mesh consisting of 1.80 106 cells. Mesh independence studies are performed to establish the effect of mesh size on
the calculated results. Calculations are performed with three mesh
sizes consisting of 7.33 105, 1.26 106, 1.80 106, and 2.41 106
cells to examine the effect on the solutions. The temperatures obtained
from calculations using 2.41 106 cells are nearly identical to those obtained using 1.80 106 cells. This mesh is employed in all further
simulations.

Fig. 7. 3D streamline prole of liquid phase velocity of AR, Tw = 873 K.

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

23

Fig. 8. Gas phase velocity proles of AR, Tw = 873 K. (a) Line locations; (b) gas supercial velocity X prole; (c) surface gas velocity streamline in XZ-plane; (d) gas supercial velocity Z
prole.

expressions are monitored and a quasi-steady state is assumed to prevail if the values of them remain constant or show no appreciable
change for a period long enough to determine the time-averaged values
of the various parameters.
4. Results and discussion
The thermal cracking kinetic model is validated against experimental data to examine its reliability. Sawarkar et al. [23] carried out batch
reaction experiments pertaining to the coking of Arabian mix vacuum
residue (AMVR) at different severities of temperature and reaction
time. The S.A.R.A. mass fractions of AMVR they used are 15.00%,
66.28%, 7.60%, and 11.12% respectively. In their experiments, the
reactor was heated by molten tin bath and rocked manually for

Fig. 9. Averaged mass fraction distribution of coke in liquid phase at different severities of
petroleum type and tube wall temperature.

temperature uniformity. Transient simulations are performed under


adiabatic condition with xed liquid temperature and the generated
vapor phase exits from the computational domain at a degassing outlet
boundary. Backup results are output with reaction time interval of 300 s
and the calculated total reaction time is 1200 s. Both the calculated and
experimental coke formation rates are shown in Fig. 3. It can be observed that the predicted coke formation rates show strong disagreement with experimental values in Fig. 3(a) and (b). The induction
period [2426] for the thermolysis of asphaltenes at lower temperature
is believed to cause this divergence. As the temperature increases, the
induction period tends to disappear, and the predicted coke formation
rates show good agreement with experimental values as shown in
Fig. 3(c) and (d). In the present study, the coke mainly generates near
the hot tube wall (773 K/873 K), thus the thermal cracking kinetic
model sufces in the simulations. It is worth noting that if the temperature is far beyond the values in the present study, new reaction constants from experiments can be inputted to cooperate with the
present model framework. To be more specic, if the temperature distribution varies greatly, a segment algorithm can be employed for
more precise calculation.
The ow model was validated by comparing the predicted pressure
drop with the values estimated by the Fanning equation with Blasius
friction factor correlation [27] as shown in Fig. 4. It can be seen that
the results exhibit a very good agreement. The relative error was
below 5% for all the gas phase mass fractions analyzed.
As shown in Fig. 5(a), accumulation of gas phase in the upper domain as a consequence of gravity is evident, indicating that the implemented free-surface model succeeds in modeling the phase separation
and capturing the gasliquid interphase. As shown in Fig. 5(b), one
can see that the gas phase holdup of atmospheric residue of Arabian
light with the tube wall temperature of 873 K is far above other severities, with a gas phase holdup of 10.69 near the outlet. The combined effect of high saturates content and high tube wall temperature is
expected to cause the difference. When the tube wall temperature increases from 773 K to 873 K, vacuum residue of Arabian light shows
an indistinctive increase in gas phase holdup as a result of its low content of saturates.

24

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

As demonstrated in Fig. 6(a), the oil temperature near the tube wall
in each cross section is higher than the central zone as a result of wall
heat transfer, and a gradual increase of bulk oil temperature in the
ow direction is evident. It can be observed from Fig. 6(b) that the
tube wall temperature dominates the liquid phase temperature, as the
liquid temperature of the severities with tube wall temperature of
873 K is far above that of 773 K. While with the same tube wall temperature, AR and VR show very close liquid temperature distributions. It is
observed that with the same tube wall temperature, the liquid temperature of AR is slightly higher than that of VR at the latter part of the tube.
That is because with the same tube wall temperature, the gas phase
holdup of the former is higher than that of the latter, and the convective
heat transfer is enhanced and more heat is transferred from the tube
wall to the liquid phase.
As shown in Fig. 7, the ow eld is uniform at the beginning of the
tube, and straight streamlines are apparent. As the gas phase appears
and accumulates in the tube, the existence of rotational ow near the
tube wall is apparent, while the streamlines near the centerline of the
tube waves in radial direction but no extensive rotation appears. In addition, liquid velocity near the centerline of the tube is far above that
near the tube wall as a result of boundary layer distribution. It can be observed that liquid velocity near the centerline exceeds the inlet velocity
of 2 m/s. Pressure increase induced by the gas phase is expected to
speed up the liquid ow.
Fig. 8 shows the gas phase velocity proles, in which (b) and
(d) show the gas supercial velocity X and Z respectively at four vertical
lines as marked in (a). Fig. 8(c) shows the surface gas velocity streamline in XZ-plane. Velocities X and Z represent the gas velocity at the vertical direction and uid ow direction respectively. The magnitude
saltation locations indicate the interphase between the continuous liquid phase and the continuous gas phase. From line 1 to line 4, the saltation moves in minus X-direction gradually indicating that the gas phase
is accumulating in the upper domain and occupying a larger room in Xdirection. It can be observed that both the supercial velocities X and Z
of the gas phase dispersed in liquid phase are fairly small in magnitude
compared with those of the continuous gas phase, and this is also indicated in the streamline prole shown in Fig. 8(c). Supercial velocity X
in minus X-direction is observed, indicating the gas rebound from the
upper tube wall. It also can be observed from Fig. 8(d) that the magnitude of gas supercial velocity Z increases gradually from line 1 to line
4 as a result of gas accumulation and consequent higher gas phase pressure head.
As shown in Fig. 9, the magnitude of the coke mass fractions in the
tube is small compared with that in the experimental data of Sawarkar
et al. [23], that is because the oil residence time in the tube is very short
(about 9 s, tube length divided by inlet velocity). The coke amount accumulates as the oil ows forward. However, due to the turbulent ow behavior, the distribution of coke in the ow direction does not increase
steadily. As a result, an increasing trend with random uctuation of
the coke mass fraction in the ow direction is observed in Fig. 9. It is
worth noting that since the points in each curve were calculated by averaging over the cross sections with an interval of 1 m, the curves in
Fig. 9 did not exhibit the complete variation trend of the coke mass fraction in the ow direction. However, it still indicates that with the same
tube wall temperature, the coke formation rate of VR is higher than that
of AR. At the beginning of the tube, the oil temperature for all the severities is low, so the coke formation rates are quite close. As the oil temperature increases, the coke formation rates begin to diverge, that is,
the coke formation rates of the severities with higher tube wall temperature and higher contents of asphaltenes and resins increase more
quickly.

cracking kinetic model is validated against experimental data and is


proved to be efcient for the coke formation prediction in the present
study. The phase change and phase separation as well as coke formation
are predicted for different severities of petroleum type and tube wall
temperature. Asymmetric behavior of the two-phase ow eld is observed due to gravity. Phase change is found to contribute to heat transfer between uid and hot tube wall. Heavy oil encounters more rapid
coke formation rate and rises with increasing tube wall temperature.
Additional insights into the gas and liquid ow behavior, gas holdup
and liquid temperature eld are gained with the aid of charts, streamlines and snapshots of distribution proles. The information predicted
by the CFD model can be utilized in the optimal design of industrial
red furnaces.

5. Conclusions

Acknowledgment

In this work we have attempted to predict the phase change and reactions inside the tubes of red furnace by means of CFD. The thermal

We are grateful for the nancial support from the National Science
Foundation of China (No. 21336007).

Nomenclature
A, B
characteristic parameters for the reactions
A
wall contact area of phase
cp
heat capacity (Jkg1K1)
C1, C2, C constants of k turbulence model
Di,l
kinematic diffusivity of pseudo-component i in liquid phase
(kgm1s1)
g
gravitational force (9.8 ms2)
h
heat transfer coefcient
H
static enthalpy (J)
k16
reaction rate constants (s1)
M
interphase momentum exchange (Nm3)
Nu
Nusselt number
P
pressure (Nm2)
Pk
turbulence production due to viscous forces (Pas1)
Pkb, Pb turbulence production due to buoyancy (Pas1)
Pr
Prandtl number
Q
interphase heat transfer (Js1)
r
volume fraction of phase
R
universal gas constant (Jkg1K1)
Re
Reynolds number
Si,
source term of pseudo-component i in phase
t
time (s)
T
temperature (K)
u
velocity vector (ms1)
Yi,l
mass fraction of pseudo-component i in liquid phase

Greek letters

density of phase (kgm3)

kinematic viscosity (Ps1)

turbulent dissipation rate (m2s3)


,
constants of turbulence model

thermal conductivity (Wm1K1)

turbulent kinetic energy (m2s2)

interphase mass ow rate (kgs1)

Subscripts
eff
effective
g
referring to gas phase
i
referring to pseudo-component i or reaction i
l
referring to liquid phase
tur
turbulent
,
referring to phase and phase respectively

X.-G. Li et al. / Fuel Processing Technology 134 (2015) 1825

References
[1] J.X. Zhou, H. Xu, X.J. Luan, X. Ling, Inuence of the SiO2/S coating and sulfur/
phosphorus-containing coking inhibitor on coke formation during thermal cracking
of light naphtha, Fuel Process. Technol. 104 (2012) 198203.
[2] X.C. Tan, C.C. Zhu, Q.K. Liu, T.Y. Ma, P.Q. Yuan, Z.M. Cheng, W.K. Yuan, Co-pyrolysis of
heavy oil and low density polyethylene in the presence of supercritical water: the
suppression of coke formation, Fuel Process. Technol. 118 (2014) 4954.
[3] C.C. Li, G.H. Hu, W.M. Zhong, W.L. He, W.L. Du, F. Qian, Coke deposition inuence
based on a run length simulation of a 1,2-dichloroethane cracker, Ind. Eng. Chem.
Res. 52 (2013) 1750117516.
[4] A.J. Guo, Z.Q. Wang, H.J. Zhang, X.J. Zhang, Z.X. Wang, Hydrogen transfer and coking
propensity of petroleum residues under thermal processing, Energy Fuel 24 (2010)
30933100.
[5] T. Gentzis, P. Rahimi, R. Malhotra, A.S. Hirschon, The effect of carbon additives on the
mesophase induction period of Athabasca bitumen, Fuel Process. Technol. 69 (2001)
191203.
[6] G. Martin, T. Barletta, Vacuum unit red heater coking-avoid unscheduled shutdowns, Pet. Technol. Q. (2001) 123127.
[7] T. Barletta, Why vacuum unit red heaters coke, Pet. Technol. Q. (2002) 123128.
[8] R.. Kseolu, C.R. Phillips, Kinetics and product yield distributions in the
CoOMoO3Al2O3 catalysed hydrocracking of Athabasca bitumen, Fuel 67 (1988)
14111416.
[9] R.. Kseoglu, C.R. Phillips, Hydrocracking of Athabasca bitumen: kinetics of formation of gases, Fuel 67 (1988) 552556.
[10] R.. Kseoglu, C.R. Phillips, Effect of reaction variables on the catalytic hydrocracking
of Athabasca bitumen, Fuel 67 (1988) 12011204.
[11] R.. Kseolu, C.R. Phillips, Kinetic models for the non-catalytic hydrocracking of
Athabasca bitumen, Fuel 67 (1988) 906915.
[12] T. Takatsuka, Y. Wada, S. Hirohama, Y. Fukui, A prediction model for dry sludge
formation in residue hydroconversion, J. Chem. Eng. Jpn. 22 (1989) 298303.
[13] S.C.K. De Schepper, G.J. Heynderickx, G.B. Marin, Modeling the coke formation in the
convection section tubes of a steam cracker, Ind. Eng. Chem. Res. 49 (2010)
57525764.

25

[14] D. Fontoura, E. Matos, J. Nunhez, A three-dimensional two-phase ow model with


phase change inside a tube of petrochemical pre-heaters, Fuel 110 (2013) 196203.
[15] L. Schiller, A. Naumann, A drag coefcient correlation, Vdi Ztg. 77 (1935) 318320.
[16] R. Clift, J.R. Grace, M.E. Weber, Bubbles, Drops, and Particles, Courier Dover Publications, 2005.
[17] A. Tomiyama, Struggle with computational bubble dynamics, Multiph. Sci. Technol.
10 (1998) 369405.
[18] A.D. Burns, T. Frank, I. Hamill, J.-M. Shi, The Favre averaged drag model for turbulent
dispersion in Eulerian multi-phase ows, 5th International Conference on Multiphase Flow, Yokohama, Japan, 2004.
[19] C.W. Hirt, B.D. Nichols, Volume of uid (VOF) method for the dynamics of free
boundaries, J. Comput. Phys. 39 (1981) 201225.
[20] ANSYS CFX 14.0 User Manual, in: ANSYS Inc.
[21] Y. Sato, K. Sekoguchi, Liquid velocity distribution in two-phase bubble ow, Int. J.
Multiphase Flow 2 (1975) 7995.
[22] W. Ranz, W. Marshall, Evaporation from drops, Chem. Eng. Prog. 48 (1952)
141146.
[23] A.N. Sawarkar, A.B. Pandit, J.B. Joshi, Studies in coking of Arabian mix vacuum residue, Chem. Eng. Res. Des. 85 (2007) 481491.
[24] A.J. Guo, X.J. Zhang, Z.X. Wang, Simulated delayed coking characteristics of petroleum residues and fractions by thermogravimetry, Fuel Process. Technol. 89
(2008) 643650.
[25] K. Chen, H. Liu, A.J. Guo, D.N. Ge, Z.X. Wang, Study of the thermal performance and
interaction of petroleum residue fractions during the coking process, Energy Fuel 26
(2012) 63436351.
[26] A. Ambalae, N. Mahinpey, N. Freitag, Thermogravimetric studies on pyrolysis and
combustion behavior of a heavy oil and its asphaltenes, Energy Fuel 20 (2006)
560565.
[27] P. Bhramara, V. Rao, K. Sharma, T. Reddy, CFD analysis of two phase ow in a horizontal pipeprediction of pressure drop, Momentum 10 (2009) 4.

You might also like