You are on page 1of 11

Chemical Engineering Science 123 (2015) 7080

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Numerical simulation of the ue gas side of rening vacuum furnace


using CFD
Xuegang Li a,b, Luhong Zhang a,b,n, Yongli Sun a,b, Bin Jiang a,b,c, Xingang Li a,b,c, Jun Wang a,b
a

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
c
National Engineering Research Centre of Distillation Technology, Tianjin University, Tianjin 300072, People's Republic of China
b

H I G H L I G H T S






A 3-D CFD model was developed for a rening vacuum furnace.


The ue gas ow, temperature and NO mass concentration elds were predicted.
A comparative simulation case with larger excess air coefcient was carried out.
The calculated results were compared with industrial values.

art ic l e i nf o

a b s t r a c t

Article history:
Received 30 April 2014
Received in revised form
15 October 2014
Accepted 1 November 2014
Available online 10 November 2014

In this work, we aimed to predict the ue gas side performance of a rening vacuum furnace with oor
gas burners. The computational uid dynamics (CFD) approach was employed to simulate the ow,
combustion, heat transfer and NO emission. Detailed insights into the ue gas velocity, temperature eld
and NO mass concentration distribution were obtained with the aid of velocity vectors and contour
snapshots. The standard k- model was applied to turbulence simulation. The non-premixed turbulent
ames and NO emission were predicted using the Laminar Flamelet model. The discrete transfer model
(DTM) was applied to the radiative heat transfer simulation. Comparative simulation cases were carried
out to investigate the effect of excess air amount on the ue gas temperature distribution and NO
emission. Calculations were performed using the commercial packages ANSYS CFX 14.0.
& 2014 Elsevier Ltd. All rights reserved.

Keywords:
CFD
Rening vacuum furnace
Radiation simulation
Combustion simulation

1. Introduction
Rening vacuum furnace is one of the key facilities in atmospheric
and vacuum distillation units. It is used to raise the atmospheric
residuum temperature high enough to meet the vacuum gas oil
(VGO) yield target, and supply sufcient wash oil ow to prevent the
wash section packing from coking. The combined effect of high
operating temperature and heavy feedstock tends to increase difculties in design, especially in deep-cut unit in which the atmospheric
residue is heated to above 420 1C. Furnaces with design defects could
carry the risk of coking and result in relatively shorter operating life
cycles. It has been long established that the prediction of ue gas side
behavior is necessary for predicting the thermal efciency, coking
n
Corresponding author at: School of Chemical Engineering and Technology,
Tianjin University, Tianjin 300072, People's Republic of China.
Tel./fax: +86 22 27400199.
E-mail address: zhanglvh@tju.edu.cn (L. Zhang).

http://dx.doi.org/10.1016/j.ces.2014.11.001
0009-2509/& 2014 Elsevier Ltd. All rights reserved.

tendency and overall performance of the furnace. Besides, the understanding of the ue gas ow pattern and temperature eld distribution in the furnace is helpful for the ne tuning and optimization of
modern furnace design.
Due to progresses in computer hardware and software and
consequent increase of the calculation speed, the computational
uid dynamics (CFD) modeling technique becomes a powerful and
effective tool for understanding the complex chemical engineering
processes. It provides a theoretical basis and a computational
technique for predicting the ue gas velocity, temperature eld,
species concentrations as well as the process side conditions in a
tube furnace. The simulation of a tube furnace can be divided into
two parts. One is the ue gas side simulation including ue gas
ow, fuel combustion and heat transfer. The other is the process
side simulation including heater transfer, two-phase ow, phase
change and thermal cracking reaction. There have been many
attempts to model petrochemical furnaces by the use of CFD.
Heynderickx et al. (2001) used a 3-D CFD model, which was

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

71

Fig. 1. Conguration of the radiant section of the furnace. (a) 3D view, (b) Top view.

combined with a reactor model for the cracking tubes and a


radiation model for the radiant heat transfer to calculate the ue
gas ow pattern and the corresponding temperature eld for a
pyrolysis furnace. Though typical ue-gas circulation proles in
the furnace were calculated, no large vertical recirculation proles
of the ue gas in the furnace were obtained. Stefanidis et al. (2006)
developed a 3-D mathematical model allowing the detailed
investigation of the ame structure, and concluded that more
sophisticated turbulencechemistry interaction models like the
EDC model and more detailed reaction kinetics should be used for
combustion modeling. Habibi et al. (2007a) performed NOx emission calculations as a post-processing step with the ow eld,
temperature and species concentrations xed. Their results were
in line with what is reported by Stefanidis et al. (2006) and Habibi
et al. (2007b). Sanders et al. (1997) explored the potential of using
the laminar amelet model for NO emission prediction from nonpremixed turbulent ames, and the calculations were compared
with the experimental data of Driscoll et al. (1992). Their studies
revealed that the amelet approach was appropriate for high
Damkhler number turbulent ames and the prediction of NO
emission improved when Damkhler number increases. There are
also many other researchers (De Schepper et al., 2009a, 2010;
Detemmerman and Froment, 1998; Guihua et al., 2011;
Heynderickx and Froment, 1998; Hu et al., 2011; Lan et al., 2007;
Masoumi et al., 2006; Oprins and Heynderickx, 2003; Souza et al.,
2006; Stefanidis et al., 2008) attempted to model the cracking
furnaces by the use of CFD method, however, no similar reports on
rening vacuum furnace were found in the literature.
In this article, a three-dimensional CFD model was developed for
the ue gas side simulation of a rening vacuum furnace which was
newly installed in a renery plant in China. The rening vacuum
furnace usually has a much larger scale than cracking furnace and
consists of convection section and radiant section. The radiant section
takes up more than 80% of the total heat load and involves even more
complex heating conditions than the convection section. Thus, only the
radiant section simulation was focused on in the present work. The
conguration of the radiant section is illustrated in Fig. 1. The oil makes
six passes through the furnace and four ared tubes with the tube
diameter increasing from 152 mm to 325 mm are used in each pass
(see Fig. S1 in Supporting Information). Adjacent tubing passes have
inverse oil ow directions. At the furnace oor, 18 gas burners are
divided into three rows equally, each of which is surrounded by two
tubing passes. Simulations were carried out to predict the ow,
combustion, heat transfer and NO emission. The objective of this work
is examining the extent to which CFD models can be employed as an
investigative and design tool in industrial practice.

Table 1
Meanings of , , and S in governing equations.
Equation

Continuity equation

Momentum equation

ui

Enthalpy equation
Species equation
k-equation

h
Zi
k

ht
D
kt

-equation
Here: t

C k2
,

i
P k t u
xj




P
 x
x i i xi j g i
i

Sh
0
Pk -
(/k)(C1Pk-C2)

uj
ui
xj xi

Table 2
Arrhenius coefcients for kt and kp.
Reaction rate constant

A (1/s)

Ta (K)

kt
kp

1.8  1011
6.4  106

38370
36510

Fig. 2. Model geometry and boundary conditions.

2. CFD model development


2.1. Flow model
The model used to calculate the ue gas ow in the furnace was
based on the Reynolds-Averaged Navier-Stokes (RANS) equations,
containing continuity, momentum, energy and species equations.

72

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

The standard k- model (Jones and Launder, 1972) with scalable


wall function (ANSYS, 2011; Launder and Spalding, 1974) was
adopted in order to account for turbulence. The general form of
these equations can be written as:


uj

1
S
t
xj
xj
xj
The means of , , and S for each equation are summarized
in Table 1.
2.2. Combustion model
In this work, the furnace was red with non-premixed burners
at oor and methane was used as fuel. Many researchers (Guihua
et al., 2011; Habibi et al., 2007b; Hu et al., 2011; Magnussen and
Hjertager, 1977; Yang et al., 2012) performed similar combustion
calculations using the combined EDM/FRC model (Eddy Dissipation model/Finite Rate Chemistry). A limitation of this model is
that the radical or intermediate species, such as CO and NO, cannot
be calculated with adequate accuracy. Besides, the O concentration is estimated using the O2 concentration and temperature
instead of calculating, and this may lead to inaccuracy for the
prediction of NO emission.
Due to the fact that the fuel and oxidizer (air) undergo fast
combustion which means the combustion rate is dominated by the
rate of mixing of the materials, the Laminar Flamelet model
(Peters, 2000; Sanders et al., 1997) is adopted for its good
performance in modeling turbulent ames at a high Damkhler
number. Flamelet equations are derived from energy and species
transport equations by applying a coordinate transformation and
assuming one-dimensional behavior of the combustion phenomena in the direction normal to the ame front. In these equations,
temperature and species mass fractions are functions of a conserved scalar known as mixture fraction (Z). Non equilibrium
effect, which means the inuence of the outer ow eld on the
inner reaction zone, is described by the scalar dissipation rate ()
at stoichiometric mixture:
2DZ2

be written as:
T 4 s
!!!
!!

I r ; s s  s I r ; s n2

4
0

0
! !
! !0
I r ; s s U s d
0

3
In the context of typical combustion systems, the dominant
emitters of radiation are carbon dioxide and water vapor (although
hydrocarbons, CO and SO2 also make a minor contribution) in ue gas.
The weighted sum of gray gases model (WSGGM) (Heynderickx and
Nozawa, 2005; Liua et al., 1998; Stefanidis et al., 2007) was used for
the calculation of the absorption coefcient of ue gas. In the WSGGM,
the emissivity of the real gas is expressed as the weighted sum of the
emissivities of a number of gray gases and a transparent gas, and it is
presented as:
I

;i T1  e  ki sP

i0

where ;i is a weighting factor and the bracketed quantity is the ith


ctitious gray gas emissivity.
Because emissivity must be proportional to absorptivity by Kirchoffs' law, it follows that the emissivity of the real gas must approach
unity as sP approaches to innite. This imposes a constraint on the
gray gas weights or amplitudes:
I

;i 1

i0

2.4. NO formation model


The formation of NO is a complex process involving several
different mechanisms including thermal NO, prompt NO, fuel
nitrogen, reburn (destruction of NO) and so on. The thermal NO
mechanism is a predominant source of NOx in gas ames at
temperatures above 1800 K. This kind of NO is formed from the
combination of free radical O and N species, which are abundant
at high temperature. The two-step mechanism, referred to as the
Zeldovich mechanism, is thought to dominate the process:
O N2-NO N

(6)

The Flamelet model requires the availability of a amelet library


suited for the required fuel/oxidizer combination over the pressure
and temperature ranges of interest. Only a transport equation for
the mixture fraction needs to be solved during the CFD simulation,
while species and temperature can be evaluated from the amelet
libraries. In the present work, the amelet library was generated
using CFX-RIF. In addition, this model can simulate the products of
incomplete combustion; for this reason, it generally provides a
more accurate solution than the Eddy Dissipation model (ANSYS,
2011). For further information on the Laminar Flamelet model, the
papers of Peters (1984; 1988; 2000), Chou et al. (1998), Pitsch et al.
(2003), Pitsch and Peters (1998) should be consulted.
2.3. Radiation model
In the radiant section of petroleum furnace, radiative heat
transfer dominates the heat transfer process. The proper choice
of the thermal radiation model will affect not only the accuracy of
the solution, but also the computational time it requires. In this
work, the discrete transfer model (DTM) was employed. This
model is based on tracing the domain by multiple rays leaving
from the bounding surfaces and depends upon the discretization
of the equation of transfer along rays. The path along a ray is
discretized by using the sections formed from breaking the path at
element boundaries. The physical quantities in each element are
assumed to be uniform. The radiative transfer equation of DTM can

Fig. 3. Illustration of the computational mesh.

Table 3
The effect of the grid size on the outlet average temperature of ue gas.
No. of cells
Grid 1
Grid 2
Grid 3

5.74  10
7.06  106
8.19  106

Average temperature of outlet ue gas (K)


864.15
977.57
980.32

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

N O2-NOO

(7)

In sub or near stoichiometric conditions, a third reaction may


be important:
OH N-NO H

(8)

At temperatures lower than 1800 K, hydrocarbon ames tend


to have an NO concentration that is too high to be explained by the
Zeldovich mechanisms. Hydrocarbon radicals can react with
molecular nitrogen to form HCN, which may be oxidized to NO
under lean ame condition.
CH N2-HCN N

(9)

HCN O2-NO

(10)

In this work, both thermal NO and prompt NO were modeled


by means of generic reactions with Arrhenius Temperature PDF
reaction rates. The thermal NO formation source was described by
a single reaction rate proposed by Warnatz et al. (2006) as:
SNO;t 2M NO kt ON 2 

11

It is worth noting that Eq. (11) is derived from the reaction rate
of Reaction (6) multiplying by 2. This is based on the fact that
Reaction (6) tends to be the rate limiting step, producing both an
NO and N species, while Reaction (7) and (8) are assumed to be
fast, so one step of Reaction (6) will produce two NO molecules.
The prompt NO formation source was described by a single
reaction rate proposed by De Soete (1975) as:


M ave 1:5
SNO;p M NO kp O2 0:5 N2 CH 4 
12

kt and kp in Eqs. (11) and (12) were the reaction rate constants and
they could be related to temperature by Arrhenius equation as:
k Ae  T a =T

13

where A was the pre-exponential factor, Ta was the activation


temperature. Warnatz et al. (2006) and De Soete (1975) proposed
the values of Arrhenius coefcients for kt and kp respectively as
listed in Table 2.

73

Unlike in EDM, the O concentration was calculated from the


amelet library, where its concentration information was directly
available in Laminar Flamelet model. The Flamelet conguration
utilized chemistry post-processing in solving the concentration of
NO. It is worth noting that there is no convergence procedure
needed as the calculated eld will hardly be affected by the small
amount of NO.
2.5. Geometry and boundary conditions
To save computational time and machine memory, only half of
the geometry was modeled due to the structural similarity. The
model geometry and boundaries are illustrated in Fig. 2.
2.5.1. Air inlet
There were two groups of air inlets, named the primary air inlet
and the secondary air inlet respectively, and uniform velocity
proles were imposed separately for them.
U air;pri

Q air;pri
Aair;pri

14

U air;sec

Q air;sec
Aair;sec

15

2.5.2. Fuel gas inlet


To obtain a fully developed ow of fuel gas before entering the
nozzle holes, the fuel gas inlets were extended beyond the primary
air inlets as shown in Fig. 2. The nozzles and the inlet pipes above
the primary air inlets were taken as inner walls. The same and
Table 5
Comparison of simulation results and industrial data.
Items

Industrial data

Simulation results

Temperature of ue gas at outlet (K)


Excess air coefcient, q

973.15
1.20

977.57
1.22a

It was obtained based on the calculated average O2 concentration in the ue gas.

Table 4
Furnace dimension and operating conditions.
Items
Furnace segment
Height (y-direction) (m)
Length (x-direction) (m)
Width (z-direction) (m)
Number of burners
coils
Number of passes
Number of tubes per pass
Tube length (m)
Tube diameter (mm)  number per pass
Tube pitch (mm)
Firing conditions
Fuel composition
Fuel gas ow rate (kg/s)
Feed temperature of fuel (K)
Feed temperature of air (K)
Air composition (wt%)
O2
N2
Primary air ow rate (kg/s)
Secondary air ow rate (kg/s)
Excess air coefcient, q
Total heat input (kW)
Material properties
Emissivity of chamber wall
Emissivity of tube skin

Data

13.14
7.888
4.224
12 (6 of half)
6
24
1.2
152  17/168  2/219  2/273  2/325  1
304/336/438/546/650
Methane
0.354
300 K
370 K
23.2
76.8
4.38
2.92
1.20
19768
0.8
0.9

74

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

uniform velocity prole was imposed at the fuel gas inlets.


Q f uel
U f uel
Af uel

16

where Qfuel was the total fuel gas ow rate and Afuel was the total
cross area of all inlets.
2.5.3. Flue gas outlet
The ue gas outlet boundary was specied as average static
pressure of  30 Pa (gauge pressure) with pressure prole blend of
0.5 over whole outlet. The average constraint was applied by
comparing the area weighted pressure average over the entire
outlet to the specied value. The pressure prole at the outlet was
shifted by this difference such that the new area weighted
pressure average would be equal to the specied value. The ow
direction was an implicit result of the computation.
2.5.4. Wall and symmetry boundary conditions
In this work, no coupled furnace-reactor simulations (De
Schepper et al., 2009b; Guihua et al., 2011; Heynderickx et al.,
2001; Hu et al., 2011; Oprins and Heynderickx, 2003; Yang et al.,
2012) were performed, since only the ue gas side of the furnace
was simulated. Therefore, a xed tube skin temperature prole
(see Table S1 in Supporting Information), taken from industrial
operating condition, was applied and considered to be part of the
boundary conditions of the simulations. The inner wall of the
chamber was assumed to be adiabatic. A no-slip and smooth wall
boundary condition was imposed at all the walls and the
emissivity was specied the same as 0.8. The symmetry plane
of the geometry was specied as symmetry boundary as shown
in Fig. 2.

carried out to obtain converged results with the target value of the
normalized residual for each variable to be 10-5, as generally
recommended in the ANSYS CFX 14.0 User Manual (2011). The
average temperature of the ue gas at the outlet was taken as a
monitor parameter in the output control. During the simulation this
parameter was monitored and quasi-steady state was assumed to
prevail if the value remained constant for a period long enough to
determine the time-averaged values of the various parameters.

4. Results and discussion


Table 5 shows a comparison between simulation results and
industrial data under the operating conditions shown in Table 4.
One can see that the calculated outlet average temperature of ue
gas and excess air coefcient are in good agreement with the
industrial data.
4.1. Flue gas velocity proles
Fig. 4 shows the 3D ue gas velocity vector snapshots at
different heights from the furnace bottom. Fig. 4(c) is close to
the burners. It can be seen that the velocity vectors mainly have a
vertical component which has an overwhelming magnitude as a
consequence of the high vertical velocity of fuel gas and air
through the burners. Surrounding the vertical velocity vectors
are those having a small vertical and a limited horizontal components. This behavior results in vertical ames and prevents ame
lick on the tubing. With increasing of the furnace height, both the

2.6. Mesh generation


The computation domain was discretized using an unstructured mesh by means of ICEM CFD 14.0. The grids near the tube
skins and burners were rened in order to capture the geometry
details as shown in Fig. 3. In order to ensure the solution
independency from the grid size, the geometry was meshed using
three different grid sizes and the predicted outlet average temperature of ue gas was compared as shown in Table 3. It can be
observed that, due to the ner grids in Grid 2 setup, the calculated
temperature is approximately 13% higher than the Grid 1 setup.
However, the values of calculated temperature using the Grids
2 and 3 setups are nearly identical to each other. Therefore, the
Grid 2 setup was chosen due to the lower required computation
time. In this mesh conguration, the domain was divided into
7.06  106 cells.

3. Operating conditions and simulation procedure


The basic geometries of the furnace and coils, as well as the
operating conditions are given in Table 4.
The simulations were carried out using the commercial packages
ANSYS CFX14.0 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 (ANSYS, 2011) was used to discretize the
convection terms. This discretization scheme involves a blend factor
which varies throughout the domain based on the local solution eld.
In ow regions with low variable gradients, the blend factor will
be close to 1 (tends to second order scheme) for accuracy. In areas
where the gradients change sharply, the blend factor will be close to
0 (tends to rst-order upwind scheme) to prevent overshoots and
undershoots and maintain robustness. Steady-state simulations were

Fig. 4. 3D ue gas velocity vectors in horizontal cross sections at different furnace


heights. (a) Y=6.57m, (b) Y=1.6m, (c) Y=0.3m.

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

direction and magnitude of the ue gas velocity tend to become


chaotic and uniform due to the turbulent ow behavior. At the
height of 1.6 m, as shown in Fig. 4(b), the zonal distribution of the
vertical vectors corresponding to the burner conguration is
evident. In the mid-height cross section, the chaotic manner of
the ue gas velocity vectors is observed as shown in Fig. 4(a), and a
more uniform distribution exists in the upper region (not shown).
In order to have a clearer view of the ue gas ow eld, 2D ue
gas velocity vectors in XY- and YZ-plane are shown in Figs. 5 and 6
respectively. Fig. 5(a) is situated at the symmetry plane of the
computational domain, encountering a row of burners in the
sweeping direction. Two approximately symmetrical recirculation
zones are observed because of the high inlet velocities. Similar
recirculation zones are also observed in Fig. 5(d), which is situated
across another row of burners near the wall, while symmetrical
prole does not appear. This is because that the burners here are
installed in two lines as shown in Fig. 1(b), and the irregular
arrangement is believed to cause the asymmetrical ow pattern. In
the region between the two rows of burners, as shown in Fig. 5(b)
and (c), the ue gas velocity vectors present chaotic behavior and
more small recirculation zones are observed due to the turbulent
ow as well as the disturbance caused by the tubing. Fig. 6(a)
and (b) present the ue gas velocity vector snapshots in YZ-plane,
which are situated in the middle and 1.5 m from the middle
respectively, and the latter encounters two burners in the sweeping
direction. It can be observed that the two burners present different

75

velocity vector proles. The ame situated at the symmetry plane


presents perpendicular prole as a result of domain symmetry. For
the ame near the furnace wall, the ame core also presents
perpendicular prole as a result of the high vertical velocity of feed
gas, while the ame tail tends to be unstable and presents wobbly
prole due to the turbulent ow and asymmetry location. It is
observed that the highest ue gas velocity in the ame locates at
about 1 m above the furnace oor. The ue gas expansion due to
temperature rise is expected to cause the acceleration of the ue gas.
It can also be seen that the inuence of the ames is felt over the
entire furnace height because the ue gas velocities above the
burners remain at a higher level than those in surrounding region.
In the vertical cross sections away from the burners, as shown in
Fig. 6(c) and (d), uniform velocity proles of ue gas can be
observed.
Fig. 7 shows Y-velocity distribution in Z-axis direction above
two burners (X4.444 m) at different heights from the bottom,
which are 0.1, 1, 3, 6, 9 and 12 m. The jet zone near the furnace
bottom is apparent. At the height of 0.1 m from the bottom
burners, the Y-velocity is about 25 m/s. With increasing furnace
height, Y-velocity decreases signicantly and the jet zone tends to
disappear. It also shows that recirculation occurs near the tubing
and the furnace wall as the Y-velocity evolves into minus
Y-direction. At the height of 12 m nearing the furnace top and
side walls, the recirculation velocity reaches to 3.3 m/s as a
consequence of wall reection.

Fig. 5. Flue gas velocity vectors in XY-plane. (a) Z=0m, (b) Z=1.346m, (c) Z=2.1515m, (d) Z=2.792m.

76

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

4.2. Snapshots of ue gas temperature


The ue gas temperature distribution in the chamber is inuenced
by the velocity eld. Fig. 8 shows the ue gas temperature snapshots
of various vertical cross sections. One can see that these proles are in
close accordance with those shown in Figs. 5 and 6. The ue gas
temperature above or near the burners is obviously higher than
surrounding regions. Due to the turbulent ow as well as the
disturbance caused by the tube rows, the ue gas velocity eld near

the tubing is more uniform. As a consequence, the ue gas here also


presents a uniform temperature eld. This situation contributes to the
uniform heating of tubes and avoiding coking problem.
In order to quantify the inhomogeneity of the ue gas temperature distribution in the symmetry plane, temperature prole
along the centerline of each burner is presented in Fig. 9. Due to
the approximately symmetric locations of Burner 1 versus Burner
6, they present similar proles, but not completely the same. Be
more specic, their proles diverge in the ame tails and the

Fig. 6. Flue gas velocity vectors in YZ-plane. (a) X=3.944m, (b) X=5.444m, (c) X=7.244m, (d) X=7.744m.

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

upper domain of the chamber, while below the ame height


(about 1 m) are nearly identical to each other. In addition to the
location effect, the divergence above the ame height is mainly
induced by the turbulent ow behavior. Although at the ame core
area, the ow eld is dominated by the same inlet velocity of feed
gas, so the proles are nearly the same. Similar explanation also
applies to Burner 2 versus Burner 4 and Burner 3 versus Burner 5.
4.3. NO concentration prole
Fig. 10 shows the NO mass concentration contours in vertical
cross sections. When compared with Fig. 8(e) and (f), one can see
that the NO concentration prole agrees well with the ue gas
temperature prole, indicating that the highest NO concentration
presents in the region of the highest ue gas temperature. This
behavior is further revealed in Fig. 11, which shows the NO mass
concentration snapshots at different heights from the bottom. One
can see that most NO is formatted in the domain above the

Fig. 7. Y-velocity distribution at different heights in plane X 4.444 m

77

burners, which is also in close accordance with the changing trend


of the ue gas velocity vector shown in Fig. 4.
For further insights into the NO distribution, NO mass concentration prole along the centerline of each burner in the symmetry
plane is presented in Fig. 12. One can see that the NO mass
concentration reaches its peak value at the height of about 1 m
from the furnace bottom, where the ue gas temperature also
reaches its peak value as shown in Fig. 9. With increasing height
above 1 m from the furnace bottom, the NO mass concentration
tends to decrease. Although NO is generated constantly as the ue
gas owing up from the bottom and the NO amount is accumulated, the NO mass concentration is diluted due to the turbulent
ow and recirculation of the ue gas in the upper domain.
4.4. Effect of excess air amount
In order to investigate the effect of excess air amount on the
ue gas temperature eld, a comparative simulation case with

Fig. 9. Flue gas temperature proles along centerlines of the burners in the
symmetry plane. Burners are numbered in the minus X-direction.

Fig. 8. Flue gas temperature snapshots of vertical cross sections. (a)  (d) XY-plane; (e)  (h) YZ-plane. (a) Z=m, (b) Z=1.346m, (c) Z=2.1515m, (d) Z=2.792, (e) X=3.944m, (f)
X=5.444m, (g) X=7.244m, (h) X=7.744m.

78

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

excess air coefcient q 1.5 was conducted. Fig. 13 shows the ue


gas temperature proles of the comparative cases at the symmetry
plane. It can be observed that the ue gas temperature above the
burners with q 1.5 in Fig. 13(b) is lower than that with q 1.2 in
Fig. 13(a). This is because the extra air introduced takes out much
heat from the outlet and results in a lower average temperature of
the ue gas in the chamber. In addition, the extra air feed at low
temperature (300 K) also absorbs heat during the temperature
increase. In order to quantify the differences between the two
cases, the average ue gas temperature distributions along the
furnace height were plotted in Fig. 14. One can see that the peak
value of the average ue gas temperature appears at the height of
about 1.8 m from the furnace bottom, higher than the location of
the peak values along the burner centerlines shown in Fig. 9.
Fig. 14 shows that with increasing of the height from the furnace
bottom, the ue gas temperature rst increases rapidly below the
ame height, and then decreases gradually. When q 1.5, the ue
gas temperature in the ame region decreases signicantly, while
the ue gas temperature at the mid-height is more uniform
compared to q 1.2. This is because the turbulence intensity of
the ue gas increases as more excess air is introduced and the
mixing process is enhanced, leading to more uniform velocity and
temperature elds which are benecial to the uniformity of the
heating process. Although the outlet ue gas temperature of q 1.5
is lower as shown in Fig. 14, the total ue gas ux increases and
the waste heat it takes out from the outlet also increases, reducing
the thermal efciency of furnace.
Fig. 15 shows the average NO mass concentration distribution
along the furnace height. The cap prole is clearly evident.
Within the furnace height between 1 m to 6 m from the bottom,
the NO mass concentration of q 1.2 is higher than that of q 1.5
due to the higher ame temperature of the former, while they
tend to be identical as getting closer to the top of the furnace.

ue gas velocity, temperature eld and NO mass concentration


distribution were obtained with the aid of velocity vectors and
contour snapshots. It is found that the feed gas ow through
burners greatly affects the ow eld inside the radiant section,
thus affects the temperature eld and NO mass concentration
distribution, and this inuence tends to weaken with the increase
of furnace height due to the turbulent ow behavior. Comparative
simulation cases show that the excess air will decrease the average
temperature in the radiant section and reduce the thermal

5. Conclusions
In this article, we have attempted to predict the ue gas side
behavior of the radiant section of a rening vacuum furnace using
CFD. A steady-state 3D model was developed for modeling of
combustion and radiative heat transfer. Detailed insights into the

Fig. 11. NO mass concentration prole snapshots of horizontal cross sections


at different furnace heights. (a) Y=5m, (b) Y=3m, (c) Y=1m.

Fig. 10. NO mass concentration contours in vertical cross sections. (a) X=3.944, (b) X=4.444.

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

efciency of the furnace. The information predicted by the CFD


model will provide a theoretical basis for the optimization of the
geometrical structure and operating parameters of the rening
vacuum furnace.

P
Pk

79

pressure and sum of partial pressures of all absorbing


gases (Pa)
turbulence production due to viscous forces (J  m-3  s-1)

Nomenclature
A
pre-exponential factor
total area of primary air inlet (m2)
Aair,pri
Aair,sec
total area of secondary air inlet (m2)
Afuel
total area of fuel gas inlet (m2)
C1, C2, C constants of k- turbulent model
D
diffusion coefcient of mixture fraction equation
h
enthalpy (J  kg-1)
I
radiation intensity (J  m-2  s-1)
k
turbulent kinetic energy (m2  s-2)
ki
absorption coefcient of the ith gray gas (m-1)
kp
reaction rate constant for prompt NO formation (s-1)
kt
absorption coefcient of the ith gray gas (m3  kmol-1  s-1)
MNO
molar mass of NO (kg  kmol-1)
Mave
mean molar mass of the mixture (kg  kmol-1)
n
refractive index

Fig. 12. NO mass concentration distribution along the centerline of each burner in
the symmetry plane.

Fig. 14. Average ue gas temperature distribution along the furnace height.

Fig. 15. Average NO mass concentration distribution along the furnace height.

Fig. 13. Comparison of ue gas temperature snapshots of different excess air coefcients. (a) q=1.2, (b) q=1.5.

X. Li et al. / Chemical Engineering Science 123 (2015) 7080

80

q
Qair,pri
Qair,sec
Qfuel
!
r
!
s
s
SNO,t
SNO,p
S
t
T
Ta
Uair,pri
Uair,sec
Ufuel
uj
xj
Z

excess air coefcient


total primary air ow rate (m3  s-1)
total secondary air ow rate (m3  s-1)
total fuel gas ow rate (m3  s-1)
position vector
direction vector
path length
thermal NO formation source (kg  m-3  s-1)
prompt NO formation source (kg  m-3  s-1)
source term of
time (s)
temperature (K)
activation temperature (K)
inlet velocity prole of primary air (m  s-1)
inlet velocity prole of secondary air (m  s-1)
inlet velocity prole of fuel gas (m  s-1)
velocity component in the j-direction (m  s-1)
coordinate direction in the j-direction (m)
mixture fraction

Greek Letters

'

absorption coefcient (m-1)


gas density (kg  m-3)
dynamic viscosity (Pa  s)
turbulent viscosity (Pa  s)
Stefan Boltzmann constant (5.672  10-8 W  m-2  K-4)
scattering coefcient (m-1)
phase function
solid angle
turbulent dissipation rate (m2  s-3)
gas emissivity
diffusion coefcient of
scalar dissipation rate

Subscripts
air,pri
air,sec
fuel

primary air
secondary air
fuel

Acknowledgment
We are grateful for the nancial support from the National
Science Foundation of China (No. 21336007).
Appendix A. Supporting information
Supplementary data associated with this article can be found in
the online version at http://dx.doi.org/10.1016/j.ces.2014.11.001.
References
Chou, C.-P., Chen, J.-Y., Yam, C.G., Marx, K.D., 1998. Numerical modeling of NO formation
in laminar bunsen ames-a amelet approach. Combust. Flame 114, 420435.

De Schepper, S.C., Heynderickx, G.J., Marin, G.B., 2009a. Modeling the evaporation of a
hydrocarbon feedstock in the convection section of a steam cracker. Comput. Chem.
Eng. 33, 122132.
De Schepper, S.C., Heynderickx, G.J., Marin, G.B., 2009b. Coupled simulation of the
ue gas and process gas side of a steam cracker convection section. Aiche J. 55,
27732787.
De Schepper, S.C., Heynderickx, G.J., Marin, G.B., 2010. Modeling the coke formation in
the convection section tubes of a steam cracker. Ind. Eng. Chem. Res. 49, 57525764.
Detemmerman, T., Froment, G., 1998. Three dimensional coupled simulation of
furnaces and reactor tubes for the thermal cracking of hydrocarbons. Oil Gas
Sci. Technol. 53, 181194.
Driscoll, J.F., Chen, R.-H., Yoon, Y., 1992. Nitric oxide levels of turbulent jet diffusion
ames: effects of residence time and Damkohler number. Combust. Flame 88,
3749.
Guihua, H., Honggang, W., Feng, Q., 2011. Numerical simulation on ow, combustion
and heat transfer of ethylene cracking furnaces. Chem. Eng. Sci. 66, 16001611.
Habibi, A., Merci, B., Heynderickx, G., 2007a. Multiscale modeling of turbulent
combustion and NOx emission in steam crackers. Aiche J. 53, 23842398.
Habibi, A., Merci, B., Heynderickx, G.J., 2007b. Impact of radiation models in CFD
simulations of steam cracking furnaces. Comput. Chem. Eng. 31, 13891406.
Heynderickx, G.J., Froment, G.F., 1998. Simulation and comparison of the run length
of an ethane cracking furnace with reactor tubes of circular and elliptical cross
sections. Ind. Eng. Chem. Res. 37, 914922.
Heynderickx, G.J., Nozawa, M., 2005. Banded gas and nongray surface radiation
models for high-emissivity coatings. Aiche J. 51, 27212736.
Heynderickx, G.J., Oprins, A.J., Marin, G.B., Dick, E., 2001. Three-dimensional ow
patterns in cracking furnaces with longame burners. Aiche J. 47, 388400.
Hu, G., Wang, H., Qian, F., Zhang, Y., Li, J., Van Geem, K.M., Marin, G.B., 2011.
Comprehensive CFD simulation of product yields and coking rates for a oorand wall-red naphtha cracking furnace. Ind. Eng. Chem. Res. 50, 1367213685.
Lan, X., Gao, J., Xu, C., Zhang, H., 2007. Numerical simulation of transfer and reaction
processes in ethylene furnaces. Chem. Eng. Res. Des. 85, 15651579.
Launder, B.E., Spalding, D., 1974. The numerical computation of turbulent ows.
Comput. Method. Appl. M. 3, 269289.
Liua, F., Becker, H., Bindar, Y., 1998. A comparative study of radiative heat transfer
modelling in gas-red furnaces using the simple grey gas and the weightedsum-of-grey-gases models. Int. J. Heat Mass. Tran. 41, 33573371.
Magnussen, B.F., Hjertager, B.H., 1977. On mathematical modeling of turbulent
combustion with special emphasis on soot formation and combustion. Symposium (International) on Combustion, Elsevier, 719729.
Masoumi, M., Sadrameli, S., Towghi, J., Niaei, A., 2006. Simulation, optimization
and control of a thermal cracking furnace. Energy 31, 516527.
Oprins, A., Heynderickx, G., 2003. Calculation of three-dimensional ow and
pressure elds in cracking furnaces. Chem. Eng. Sci. 58, 48834893.
Peters, N., 1984. Laminar diffusion amelet models in non-premixed turbulent
combustion. Prog. Energ. Combust 10, 319339.
Peters, N., 1988. Laminar amelet concepts in turbulent combustion. Symposium
(International) on Combustion, Elsevier, 12311250.
Peters, N., 2000. Turbulent Combustion. Cambridge University Press.
Pitsch, H., Cha, C.M., Fedotov, S., 2003. Flamelet modelling of non-premixed
turbulent combustion with local extinction and re-ignition. Combust. Theor.
Model 7, 317332.
Pitsch, H., Peters, N., 1998. A consistent amelet formulation for non-premixed
combustion considering differential diffusion effects. Combust. Flame 114, 2640.
Sanders, J., Chen, J.-Y., Gkalp, I., 1997. Flamelet-based modeling of NO formation in
turbulent hydrogen jet diffusion ames. Combust. Flame 111, 115.
Souza, B., Matos, E., Guirardello, R., Nunhez, J., 2006. Predicting coke formation due
to thermal cracking inside tubes of petrochemical red heaters using a fast CFD
formulation. J. Petrol. Sci. Eng. 51, 138148.
Stefanidis, G., Merci, B., Heynderickx, G., Marin, G., 2007. Gray/nongray gas
radiation modeling in steam cracker CFD calculations. Aiche J. 53, 16581669.
Stefanidis, G., Van Geem, K., Heynderickx, G., Marin, G., 2008. Evaluation of highemissivity coatings in steam cracking furnaces using a non-grey gas radiation
model. Chem. Eng. J. 137, 411421.
Stefanidis, G.D., Merci, B., Heynderickx, G.J., Marin, G.B., 2006. CFD simulations of
steam cracking furnaces using detailed combustion mechanisms. Comput.
Chem. Eng. 30, 635649.
Warnatz, J., Maas, U., Dibble, R.W., 2006. Combustion: Physical and Chemical Fundamentals, Modeling and Simulation, Experiments, Pollutant Formation. Springer.
Yang, J., Tai, N., Wang, L., Xiao, J., Yang, C., 2012. Numerical simulation of the ue gas
and process side of coking furnaces. Ind. Eng. Chem. Res. 51, 1544015447.
ANSYS CFX 14.0 User Manual, 2011. Canonsburg, PA, USA, ANSYS Inc.,.

You might also like