You are on page 1of 10

Computers and Chemical Engineering 35 (2011) 28762885

Contents lists available at ScienceDirect

Computers and Chemical Engineering


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

Modelling and dynamic optimization of thermal cracking of propane for


ethylene manufacturing
Mehdi Berreni, Meihong Wang
Process Systems Engineering Group, School of Engineering, Craneld University, MK43 0AL, UK

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

Article history: In tubular reactors inside a cracking furnace, heat transfer, thermal cracking reactions and coke buildup
Received 8 September 2010 take place and closely interact with each other. It is important to understand the process and optimize
Received in revised form 21 April 2011 its operation. A 1-dimensional (1D) pseudo-dynamic model was developed based on rst principle and
Accepted 13 May 2011
implemented in gPROMS . Coke buildup inside the tube wall was also accounted for. The model was
Available online 20 May 2011
validated dynamically. The impact of process gas temperature prole, and constant tube outer wall tem-
perature prole on product yields and coking rate are assessed. Finally, dynamic optimization was applied
Keywords:
to the operation of this tubular reactor. The effects of coking on reduction of production time and the
Mathematical modelling
Dynamic optimization
decoking cost have been considered. The tube outer wall temperature prole and steam to propane ratio
Tubular reactor in the feed were used as optimization variables. Dynamic optimization investigation indicates that it can
Case study improve operating prot by 13.1%.
Ethylene 2011 Elsevier Ltd. All rights reserved.
Thermal cracking

1. Introduction 1.1. Literature review

Ethylene is one of the major building blocks in the petro- This section is to critically assess research progress in different
chemical industry because it is a very reactive intermediate to aspects of thermal cracking for ethylene production.
produce plastics, resins and bers. Thermal cracking of naphtha,
ethane or propane is the most widely accepted technology to pro- 1.1.1. Thermal cracking reaction mechanism
duce ethylene. Thermal cracking of propane in tubular reactor will Thermal cracking of propane proceeds via free-radical reactions.
be the topic of this paper. The main advantage of using propane as The reaction mechanism has been studied and improved several
feedstock is that the costs for the separation before and after the times by different researchers. Zou, Lou, and Liu (1988) proposed a
cracking process are lower compared to thermal cracking of naph- reduced free-radicals reaction scheme for the pyrolysis of propane
tha, moreover the total yield for ethylene and propylene is higher which includes ten reactions. Three main reaction types namely
compared to thermal cracking of ethane (Sundaram, Shreehan, and initiation, propagation and termination were included in this free-
Olszewski, 2001). radical reaction mechanism. Due to numerical reasons, molecular
The reaction from propane to ethylene is endothermic. There- reaction schemes have been widely used in simulations to approx-
fore it needs lots of external energy. The furnace is made up of imate the behaviour of the reactions. Based on experiments from
two main sections. The mixture of steam and propane is preheated Van Damme, Narayanan, and Froment (1975) and Sundaram and
in the convection section. It is then introduced into the radiation Froment (1977) developed a complete molecular scheme including
section which consists of long tubular reactors for the cracking reac- 9 reactions and 9 species. The same authors also investigated the
tions to take place. Floor and/or side burners apply a high heat ux kinetics of coking rate (Sundaram and Froment, 1979). They con-
to the tube outer wall so that a suitable process gas temperature sidered propylene as the only precursor and two reactions were
prole can be achieved. During operation, coke is formed inside the added to their previous reaction scheme to account for the coke
tube wall and the reactor must be put ofine regularly to remove formation.
the coke buildup. In order to reduce the coking rate, steam is mixed
with the propane as diluents (Sundaram et al., 2001). 1.1.2. Steady state modelling and steady state optimization
Shahrokhi and Nejati (2002) developed a 1D steady state model
of thermal cracking of propane. Their aim was to determine the
Corresponding author. Tel.: +44 1234 754655; fax: +44 1234 754685. optimal process gas temperature prole based on the molecular
E-mail addresses: meihong.wang@craneld.ac.uk, reaction scheme of Sundaram and Froment (1979). They consid-
wang 2003 uk@yahoo.co.uk (M. Wang). ered both ethylene and propylene production using an objective

0098-1354/$ see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compchemeng.2011.05.010
M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885 2877

Ethylene furnace was investigated with 3D model using com-


Nomenclature putational uid dynamics (CFD) technique by Detemmerman and
Froment (1998) and Lan, Gao, Xu, and Zhang (2007) respectively.
Bend angle [ ] The authors investigated the inuence of the position of the gas
Hr,j (z) Reaction heat [kJ/kmol] burners as well as the ue gas patterns in the rebox. The studies
(z) Bend factor [-] were carried out for process design purposes.
 i,j Stoichiometry coefcient [-]
(z) Coke thickness [m]
c Coke density [kg m3 ] 1.1.3. Dynamic modelling and dynamic optimization
c Conduction heat transfer coefcient of coke Shahrokhi and Nejati (2002) also developed 1D dynamic model
t (z) Conduction heat transfer coefcient of tube wall for the PFR. By adding the heat transfer from furnace wall to the
[kJ/(s m K)] PFR in the dynamic model, they evaluated the performance of a PID
Ac Pre-exponential factor for coking reaction controller for furnace wall temperature control in thermal cracking
[kg m3 /(kmol m2 s)] of propane. Gao, Wang, Pantelides, Li, and Yeung (2009) developed
Aj Pre-exponential factor (kmol1 m3 s1 for a 2nd a detailed 1D dynamic model in gPROMS for naphtha cracking. The
order reaction or s1 for a 1st order reaction) model includes dynamics in mass balance, energy balance and coke
CC3 H6 (z) Molar concentration of propylene [mol/m3 ] buildup, then steady-state optimization was applied for optimal
Ci (z) Mass concentration [kg m3 ] operation.
Cpmi (z) Specic heat capacity [kJ/(kg K)] Ghashghaee and Karimzadeh (2007) studied dynamic modelling
Di Internal diameter of the pipe [m] of thermal cracking furnaces which is capable of describing and
Dt Outer diameter of the pipe [m] predicting the unsteady-state behaviour of crackers during start-
Ec Activation energy of coking reaction [kJ/kmol] up. The mathematical model includes four sub-models: convection
Eaj Activation energy [kJ/kmol] model, tubular coil, rebox model and tube skin model. Based on
F0 Outlet mass owrate of propane [kg/s] these sub-models, the authors developed a two-dimensional (2D)
Fi Outlet mass owrate of desirable product i [kg/s] zone model to predict fuel consumption, heat transfer rates and
Fi (z) Mass owrate [kg/s] cracking temperatures during start-up.
Fr (z) Friction factor [m1 ] Dynamic optimization is dened as the establishment of the
Fs Intlet mass owrate of steam [kg/s] best procedure for correcting the uctuations in a process (Babu &
NC Number of components [-] Knovel, 2004). Sowers and Reed (2001) studied the dynamic opti-
NR Number of reaction [-] mization of ethylene furnaces cracking propane using the software
nij Reaction order [-] SPYROTM . Most particularly they focused on the optimal conver-
Nsteam Molar owrate of steam [kmol/s] sion prole and owrate as a function of time in order to maximize
P(z) Pressure [Pa] operating prot. The model they used was very simple in com-
Q (z) Heat transfer rate [kJ/s] parison to the one that is developed in this paper. They observed
Qt Total heat transfer rate [kJ/s] 12% improvement in cumulative prot compared with steady-
R Ideal gas constant [kJ/(K m)] state optimization.
R1 (z) Thermal resistivity relative to gas convection
[K/kW] 1.2. Scope of this study
R2 (z) Thermal resistivity relative to coke conductivity
[K/kW] Only the radiant section of the cracking furnace will be consid-
R3 (z) Thermal resistivity relative to tube conductivity ered since this is where the reactions take place. No downstream
[K/kW] separation and quenching are taken into account. This paper pro-
Rb Tube bend radius [m] vides detailed 1D mathematical modelling of the tubular reactors
Re(z) Reynolds number [K] based on rst principles. The molecular reaction scheme for ther-
rc (z) Coking rate [kg/(m2 s)] mal cracking of propane and the coke formation developed by
rj (z) Reaction rate [kmol/(s m3 )] Sundaram and Froment (1979) is used. The model also takes into
T(z) Process gas temperature [K] account the impact of coking on the product yield and on the run
Tc (z) Coke surface temperature [K] length (i.e. the time between two consecutive decoking operations).
Twe (z) Tube outer wall temperature [K] The model can predict accurately the product yields, coke thickness,
Twi Tube inner wall temperature [K] run length, pressure drop and heat transfer rate. The impact of vary-
ing process gas temperature on the furnace operation is assessed.
This is followed by investigating the impact of applying a constant
function that calculates operating prot. Masoumi, Sadrameli, tube outer wall temperature prole. Finally dynamic optimization
Towghi, and Niaei (2006) developed a 1D steady-state model for was carried out with operating prot as objective function. Tube
tubular reactors for thermal cracking of naphtha. A free-radical outer wall temperature prole and steam-to-propane ratio are used
reaction scheme including 90 species and 543 reactions was used. as optimization variables. The overall aim of this paper is to provide
A steady-state optimization study was carried out in an effort to insights for optimal operations in the ethylene industry.
maximise operating prot.
Gao, Wang, Ramshaw, Li, and Yeung (2009) applied steady-state 1.3. Novelties
optimization for naphtha cracking based on steady state model in
Aspen HYSYS. The results were very encouraging since the opti- The novel contributions of this paper include (a) 1D pseudo-
mization improved the operating prot signicantly. Gao, Wang, dynamic modelling of the tubular reactor and model validation
Ramshaw, et al. (2009) assumed the process to be under steady- from different aspects such as product yields, pressure drop, tube
state conditions, but in reality, the coke buildup takes place with outer wall temperature prole and their changes with produc-
time and affects performance. Moreover, they neglected the impact tion time (i.e. dynamic validation); (b) dynamic optimization based
of the coke thickness on heat transfer coefcient. on the model developed. The use of tube outer wall temperature
2878 M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885

prole as one of the optimization variables makes the optimal oper- 2.3. Energy balance
ation study more realistic.
The steady-state energy balance on the tubular reactor with
1.4. Outline of the paper respect to differential length can be written as:

Mathematical modelling is presented in Section 2. Then an 


NC
T (z) Q (z)  (Di 2(z))
2
Fi (z)Cpmi (z) = +
appropriate benchmark is introduced in Section 3, and the model is z L 4
validated dynamically by comparison to industrial data after 100, i=1

300 and 700 h of operation. Two case studies are performed in Sec- 
NR
tion 4. Section 5 focuses on dynamic optimization of the process. rj (z) (Hr,j (z)) (6)
The optimization results have also been discussed in great detail. j=1
Conclusions are drawn in the end.
Similar to an electrical system, the gaseous phase, the coke
2. Mathematic modelling thickness and the tube thickness can be assimilated to thermal
resistances R1 , R2 and R3 respectively as shown in Fig. 1. The overall
2.1. Assumptions heat transfer can be written as:
1
In order to develop the 1D pseudo-dynamic model, the following Q (z) = (Twe (z) T (z)) (7)
R1 (z) + R2 (z) + R3 (z)
assumptions were made:

2.4. Momentum balance


Negligible radial gradient (i.e. variation only along Z-axis)
Inertness of steam in the furnace
The steady-state momentum balance is determined as following
Only vapour or gas inside the PFR (i.e. no liquid)
based on Shahrokhi and Nejati (2002), Masoumi et al. (2006), and
All cracking reactions take place in the tubular reactors only (i.e.
Sundaram and Froment (1979).
cracking reactions start at the entrance of the reactor and stop at

 
the exit) 1 1 1 T (z)
)/z + + Fr(z)
No hydrodynamic or thermal entrance region effects (i.e. no inu- P(z) Mm(z) Mm(z) T (z) z
= (8)
ence from the environment) z 1
P(z)
Only the coke buildup is considered dynamically (i.e. heat or mass Mm(z)P(z) G(z)2 RT (z)

accumulation inside the reactor is ignored) and coke thickness The friction factor is calculated in Eq. (9). For straight section,
keeps constant within each hour the bend coefcient is 0, and for bended section, it is calculated
Ideal gas behaviour is assumed only when component concen- with Eq. (10). Both equations were extracted from Shahrokhi and
trations are calculated (i.e. relevant with Eq. (2) only) Nejati (2002) and Masoumi et al. (2006).

2.2. Material balance 0.092 Re(z)0.2 (z)


Fr(z) = + (9)
Di 2(z) Rb
The amount of coke accumulated is tiny compared to the large  0.35
 Di 2(z)

mass owrate of propane and/or products. The residence time for (z) = 0.7 + 0.051 + 0.19 (10)
90 Rb
the PFR is also very short. Consequently, the change of concentra-
tion with time can be neglected. Steady-state material balance is
written as follows: 2.5. Physical properties

Fi (z) 
NR 2
 (Di 2(z)) Physical properties such as dynamic viscosity, mass density and
= i,j rj (z) Mwi (1)
z 4 enthalpies were calculated with Multiash using PengRobinson
j=1
equation of state. gPROMS and Multiash can communicate
The molar concentration is calculated based on ideal gas law. through external property package interfaced as a Foreign Object.
  With this software, hypothetical component C6 could be cre-
Fi (z)/Mwi
 P(z) 
ated to account for all the organic chemicals with 5 or more
Ci (z) = NC
RT (z)
(2)
carbon atoms. The molecular weight was xed at 76 g mol1
Nsteam + i=1
(Fi (z)/Mwi )
based on conservation of mass principle and the boiling point
The reaction scheme used in this paper (developed by Sundaram was set at 353.25 K which is the boiling point of benzene.
and Froment, 1977) assumes that the reactions are elementary. Indeed, the fraction C6 is mainly benzene in thermal cracking of
Therefore, the reaction rates are determined according to Arrhenius propane.
law.

NC 3. Model validation
rj (z) = Aj eEaj /RT (z) Ci nij (z) (3)
i=1 3.1. Industrial data
Coking rate is calculated based on Arrhenius law as well
In order to validate the model, industrial data from Sundaram
(Sundaram and Froment, 1979).
and Froment (1979) is used. This industrial reactor was used
rc (z) = Ac eEc/RTc (z) .CC3 H6 (z) (4) several times in the literature. Most recently it was used by
As for the coke thickness (Fig. 1), it is determined as following Shahrokhi and Nejati (2002). The technical specications for the
(Sundaram and Froment, 1979). reactor that is simulated are presented in Table 1 and the tem-
perature prole of the base case are presented in Fig. 2. The
d(z) rc (z) reaction scheme used here was developed by Sundaram and
= (5)
dt c Froment (1977), who added later two reactions in Sundaram and
M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885 2879

Fig. 1. 1D pseudo-dynamic model in a plug ow reactor.

Process gas temperare [K]


Froment (1979). Those two reactions involved C6 , a hypotheti-
1150
cal component, which represents the fraction of all the organic
1100
chemicals with 5 or more carbon atoms. In the thermal crack-
ing of propane, it mainly consists of benzene, that is why it 1050
is referred as C6 . The whole reaction scheme is described in 1000
Table 2. 950
900
Table 1
850
Industrial reactor (Sundaram and Froment, 1979). 0 20 40 60 80 100
Paramaters/variable Value Tube length [m]
Length of the coil in the radiant section 95 [m]
Length of the straight portion of the coil 8.85 [m] Fig. 2. Process gas temperature prole of the base case (Sundaram and Froment,
Length of the bends 0.554 [m] 1979).
Radius of the bends 0.178 [m]
Tube internal diameter 0.108 [m]
Wall thickness 0.008 [m]
Total feed per coil 0.7635 [kg/s] 3.2. Simulated results vs industrial data
Steam dilution rate 0.4 [kg steam/kg propane]
Inlet pressure 3 [bar]
Outlet pressure 2 [bar]
All the results presented in this section were obtained by sim-
Inlet temperature 873.15 [K] ulating the pseudo-dynamic model for PFR under the conditions
Outlet temperature 1111.15[K] given in Table 1 for the reactor and the process gas temperature
Cokes density 1600 [kg/m3 ] prole given in Fig. 2. Table 3 summarizes the simulated product
Thermal conductivity of the coke 0.00154 [kcal/(m s K)]
yields at the outlet of the reactor vs the industrial data. It can be

Table 2
Reaction scheme for propane cracking (Sundaram and Froment, 1977) and for coking reaction (Sundaram and Froment, 1979).

Reactions A(s1 or *k mol1 m3 s1 or **kg mol1 m1 s1 ) E (kJ mol1 )

1 C3 H8 C2 H4 + CH4 4.692 1010 211.7


2 C3 H8  C3 H6 + H2 5.888 1010 9.04 x105 214.695.3
3 C3 H8 + C2 H4 C2 H6 + C3 H6 *2.536 1013 247.1
4 2C3 H6 3 C2 H4 1.514 1011 233.5
5 2C2 H6 0.5C6 + 3CH4 1.423 109 190.4
6 C3 H6  C2 H2 + CH4 3.794 1011 2.32 107 248.5123.7
7 C3 H6 + C2 H6 C4 H8 + CH4 *5.553 1014 251.1
8 C2 H6  C2 H4 + H2 4.652 1013 9.97 108 272.8138.0
9 C2 H4 + C2 H2 C4 H6 *1.026 1012 172.6
10 C4 H8 C6 **6.96 107 143.6
11 C3 H6 Coke **5.82 1014 308.0

Table 3
Comparison of main product yields between pilot plant data and simulated data.

gPROMS simulation Industrial data Absolute difference (wt%) Relative difference (%)

Propane conversion wt% 90.19 90.60 0.41 0.45


CH4 wt% 25.65 24.00 1.65 6.43
C2 H4 wt% 35.08 34.50 0.58 1.65
C3 H6 wt% 14.44 14.70 0.26 1.77
C3 H8 wt% 9.81 9.30 0.51 5.20
H2 wt% 1.68 1.20 0.48 28.57
2880 M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885
Propane conversion [%]

91 Simulated data
90
Industrial data
89
88
87
86
85
84
0 100 200 300 400 500 600 700 800
Producon me [h]

Fig. 3. Propane conversion with production time. Fig. 6. Tube outer wall temperature prole at t = 0 h and t = 700 h.

3.2 Simulated data pared to the simulated results in Fig. 6 and are again in good
3 Industrial data agreement.
Pressure [bar]

2.8 The validation is ended with investigating the diameter reduc-


2.6 tion after 100 h, 300 h and 700 h in Fig. 7. The change of coke
2.4 thickness is important since it dramatically affects the heat transfer,
subsequently the process operation.
2.2
In summary, Figs. 3, 57 present results simulated under the
2
0 10 20 30 40 50 60 70 80 90 100 specic design/operating conditions (Table 1) and temperature
tube length [%] prole (Fig. 2). They indicate how coke thickness, pressure drop,
propane conversion and tube outer wall temperature change with
Fig. 4. Pressure prole at production time 0 h (i.e. for clean tube). production time. These form dynamic validation of the pseudo-
dynamic model.
When the pressure drop or the reduction of diameter reaches a
seen that the result for the two most valuable products (ethylene pre-determined value, the PFR has to be shutdown for de-coking.
and propylene) are in good agreement with the industrial data. Pressure drop and the reduction of diameter are closely related.
The decrease of the propane conversion with the production However diameter reduction will be used in this paper as indicator
time calculated by the simulation is very close to the one observed because it is more convenient for simulation purposes. In this paper,
in a real life reactor as can be seen from Fig. 3. As the coke builds up it was assumed that the shutdown of the process is carried out
with production time, inner diameter is reduced and propane has when the reduction of diameter reaches 25% at the outlet. For the
less time to react inside the reactor (i.e. shorter residence time). case simulated in this section, the run length of the reactor is 39.8
This results in a decreasing propane conversion with production days.
time.
The prole of the pressure along the reactor (for clean tube, i.e. 4. Case studies
production time at 0 h) is drawn in Fig. 4. Each step corresponds
to a U-bend where the pressure loss is bigger because of an added 4.1. Impact of process gas temperature prole
term for the component loss as seen in Eq. (10). The tube length
was actually divided into 300 sections to achieve such an accurate The aim of this case study is to look at the impact of the process
prole. gas temperature prole on the process performances. The process
From Fig. 5, the pressure drop across the reactor increases with gas temperature at the outlet (i.e. coil outlet temperature COT)
the production time due to the coke buildup. In order to get the will be used as an input variable and output variables will be prod-
same process gas temperature prole, the tube outer wall temper- uct yields, propane conversion, run length and coking rate. COT
ature is also increased over the production time as shown in Fig. 5. will vary from 973.15 K up to 1143.15 K. This range has been cho-
This is to compensate for the higher thermal resistance caused by sen because this is the range of temperature for which Van Damme
thicker coke layer. et al. (1975) have carried out their experiments, later reconsidered
Sundaram and Froment (1979) also provided the whole tube by Sundaram and Froment (1979) in order to derive the reaction
outer wall temperature proles for clean tube (i.e. production time scheme selected for this simulation. The steam-to-propane ratio
at 0 h) and after operations of 700 h. The industrial data are com- will remain constant at 0.4 kg/kg. The pressure at the inlet of the
reactor is xed at 3 bars and Eqs. (8)(10) were used to calculate
the pressure drop along the reactor.
1250
1.9
Maximum tubeouter wall

Twe max
Pressure drop [bar]

1.7 1200 20
Diameter reducon [%]
temperature [K]

18 Simulated data
1.5 16
1150 Industrial data
14 700 hrs
1.3 P
1100 12
1.1 10
8
Simulated data 1050 300 hrs
0.9 6
Industrial data 4
0.7 100 hrs
1000 2
0 200 400 600 0
0 20 40 60 80 100
Producon me [h]
Reactor length [%]
Fig. 5. Pressure drop and maximum tube skin temperature as a function of produc-
tion time. Fig. 7. Diameter reduction after various production time.
M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885 2881
Process gas temperature [K]

Porpane conversion (%)


100
1150 COT=1143.15 K
COT=1111.15 K 80
1100
COT=1073.15 K
1050 60
COT=1043.15 K
1000 COT=1013.15 K 40
COT=973.15 K
950 20
900
0
850 950 1000 1050 1100 1150
0 20 40 60 80 100
Coil Outlet Temperature [K]
Tube length [%]
Fig. 10. Propane conversion under clean tube condition (i.e. no coking) under vari-
Fig. 8. Process gas temperature proles. ous COTs.

45 1000000 0.025

Coking rate rate [kg/(m.h)]


40 C2H4
Run length
35 0.02
100000 Coking rate

Run length [h]


Yield [wt%]

30 CH4
25 0.015
20 10000
15 0.01
C3H6
10 1000
5 0.005
0
950 1000 1050 1100 1150 100 0
950 1000 1050 1100 1150 1200
Coil Outlet Temperature (K)
Coil Oulet Temperature [K]
Fig. 9. Main product yields under clean tube condition (i.e. no coking) under various
Fig. 11. Coking rate under clean tube condition (i.e. no coking) and run length.
COTs.

Based on the PFR inlet temperature (xed at 873.15 K) and outlet should be avoided because they result in propane conversion lower
temperature (varying between 973.18 K and 1143.15 K), the strat- than 50%.
egy adopted enables to derive process gas temperature proles The effect of COT on coking rate can be seen in Fig. 11. As
with similar shape as the process gas temperature prole provided could be expected, as the COT increases, so does the coking rate
by Sundaram and Froment (1979). at the outlet but it also starts decreasing for the highest tempera-
In order to get a prole that will achieve for example 1043.15 K ture. At 973 K, the coking rate is very small 1.1 104 kg/(m2 h).
at the outlet, the ratio between the original and the new prole Then, it increases faster and faster and reaches its maximum
is calculated i.e. r = (1043.15/1108.15) = 0.94. This ratio is weighted value of 2.3 102 kg/(m2 h) at 1123 K before decreasing to
with a variable factor along the tube length so that the temperature 2.2 102 kg/(m2 h) at 1143 K. This behaviour is due to the fact that
of the base case is multiplied by 1 at the inlet and 0.94 at the outlet. the coking rate depends upon not only the process gas temperature,
Thus, the temperature remains at 873.15 K at the inlet, and it is but also the molar concentration of the coke precursor propylene.
then gradually decreased compared to the base case until the outlet As we have seen from Fig. 9, the yields of propylene and conse-
where the temperature of the base case is multiplied by 0.94. A quently its concentration, starts decreasing gradually at 1073 K. At
slowly changing factor enables us to generate smooth proles as the beginning, the decrease of concentration is compensated by the
can be seen in Fig. 8. The ratio is weighted with Eq (11) and the increase of temperature. But at 1123 K, the increase in temperature
temperature at any position x is then determined with Eq. (12). is not sufcient any more and the coking rate decreases. Fig. 11 also
shows the run length determined dynamically for various COTs. It
Weight factor (x) can be seen that, the faster the coking rate, the shorter the run
Temperature (x) Temperature (l = 0) length. At 1043 K, the run length is around 5526 h (230.3 days), but
= (11) it then goes down to 986 h (31.1 days) at 1143 K. For COT below
Temperature (l = 95) Temperature (l = 0)
1043 K, run length is very long, but the cracking reactions are also
very slow.
T (l = 95)desired In summary, process gas temperature affects the product yields
T (x) = Weight factor (x) (12)
1108.15 signicantly. Low temperatures result in low ethylene yield. High
Fig. 9 shows the propylene, ethylene and methane yields at the temperatures increase the total yield of ethylene and propylene,
outlet of the reactor for clean tube conditions (i.e. operating hours and also favour coke formation. A compromise in determining pro-
at 0 h) under various COTs. It can be observed that all the yields cess gas temperature prole is necessary.
are very low at 973 K, 6.8 wt% for ethylene, 8.12 wt% for propylene
and 4.11 wt% for methane. Ethylene and methane yields increase 4.2. Impact of constant tube outer wall prole
steadily with the increase of COT, respectively up to 40.7 wt% and
30.2 wt% at 1143 K. As for propylene yield, it reaches its maximum Here the tube outer wall temperature prole along the length of
value of 19.5 wt% around 1073 K. Afterwards it decreases down to the reactor is xed, the process gas temperature prole is calculated
6.7 wt% at 1143 K. from the pseudo-dynamic model. The impact of xed tube outer
In Fig. 10, the impact of the COT on the propane conversion wall temperature prole on the process will be studied over the
has been investigated. It is observed that the propane conversion whole run length. The conditions applied are a steam-to-propane
increases quite steadily from 20% up to nearly complete conversion ratio of 0.4 kg/kg and a tube outer wall temperature prole shown
(97.5%) for COTs from 973.15 K and 1143.15 K. It should be empha- in Fig. 12 that generates a COT of 1111.15 K under clean tube con-
sized that based on Fig. 10, we can infer that COTs lower than 1050 K ditions.
2882 M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885

Process gas temperature at


Tube wall temperatue [K]
1180
0.025

Coking rate [kg/(m.s)]


1160 1120 Process gas temperature at
the outlet 0.02

the outlet [K]


1140 1110 Coking rate

1120 1100 0.015


1100 1090
0.01
1080 1080

1060 1070 0.005


0 20 40 60 80 100 0 200 400 600 800 1000 1200 1400 1600
Producon me [h]
Reactor length [%]
Fig. 15. Process gas temperature at the outlet and coking rate with production time.
Fig. 12. Tube outer wall temperature prole.

Producon in 100 hours [kg]


80000
Since the overall heat transfer coefcient will decrease with
70000
higher coke thickness, the process gas temperature prole will
60000
decrease with production time when the tube wall temperature
50000
prole remains constant. Because of the reduced process gas tem-
40000
perature, the run length is 1681 h (70 days). The impact of a constant
30000
tube outer wall temperature prole on the yields of ethylene and
20000 Ethylene
propylene with time is shown in Fig. 13. On one hand, the ethy-
10000 Propylene
lene yields decreases rapidly from 35 wt% at the beginning down
0
to 25.1 wt% after 1681 h. On the other hand, the propylene yield 0 200 400 600 800 1000 1200 1400 1600
increases from 14.34 wt% to 18.7 wt% in the rst 700 h and then it Producon me [h]
keeps increasing slowly up to 19.5 wt% at the end of the run length.
Propane conversion with production time is shown in Fig. 14. The Fig. 16. Amount of ethylene and propylene produced within every 100 h with pro-
propane conversion decreases during the production time. Right duction time.
after decoking operation, the propane conversion is 90%, but it
decreases to 70.5% after 1681 h (i.e. 70 days) of operation. time rapidly from 0.022 kg/(m2 s) down to 0.0077 kg/(m2 s) at the
Fig. 15 shows that the coil outlet temperature (COT) of the pro- end of the run length. This can be explained by the trend of the
cess gas decreases signicantly with the production time. Initially process gas temperature, and the low coking rate is the reason for
the process gas temperature at the outlet was 1111.15 K. But after this unusually high run length.
1000 h of operation, it is only 1080.9 K. Finally the process gas tem- Fig. 16 plots the production at a given time minus the production
perature at the outlet drops to 1072.2 K at the end of the run length. 100 h earlier for different production time. In this way, it is pos-
This means that the reactor is going to operate under conditions far sible to analyse the amount of ethylene and propylene produced
from the one initially set by the operator. Consequently, the prod- every 100 h. At the very beginning of the operation, the furnace
uct yields are dramatically affected as shown in Fig. 13. In Fig. 15, produced 67,400 kg of ethylene in the rst 100 h, but at the end of
it can also be seen that the coking rate at the outlet decreases with the run length, the reactor produces only 49,900 kg in 100 h. On the
contrary, the furnace produces 29,400 kg of propylene in the rst
40 100 h and 38,100 kg at the end of the run length within the same
period of time (100 h). The production of propylene increases, but
35 the production of ethylene decreases. Overall, the amount of valu-
30 able products (ethylene plus propylene) produced during a given
Yields [wt%]

period decreases signicantly with the production time.


25 In conclusion, this case study shows the importance of tuning
the outer wall temperature prole dynamically. Although xing a
20
constant value is much easier from a control point of view, the
15 losses of production are signicant, and may not be acceptable.
10
0 200 400 600 800 1000 1200 1400 1600 5. Process operation optimization
Producon me [h]
Steam-to-propane ratio and process gas temperature prole are
Fig. 13. Product yields. major operating variables and some trade-offs in selecting them are
necessary. However in actual plants, operators adjust the fuel burn
95 rates in side or bottom burners, thus changing the tube outer wall
Propane conversion [%]

temperature prole, which will subsequently affect the process gas


90
temperature prole. Therefore, tube outer wall temperature prole
85 and steam-to-propane ratio are used as optimization variables in
this section.
80

75 5.1. Methodology
70
0 200 400 600 800 1000 1200 1400 1600 Operating prot over a period of 1 year is taken as objective
Producon me [h] function to maximise. In this dynamic optimization, the following
assumptions were made: (a) No downstream product separation
Fig. 14. Propane conversion. costs were considered; (b) Coke thickness is updated every hour;
M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885 2883

Steam-to-propane rao [kg/kg]


Table 4
Optimization parameters. 0.41
0.4
Parameters Physical meaning Values Units
0.39
Base case
tp Yearly production time 8160 h 0.38
td Decoking time per cycle 48 h Opmal case
0.37
C0 Propane price factor 0.596 $/kg 0.36
C1 Ethylene price factor 1.356 $/kg
0.35
C2 Propylene price factor 1.576 $/kg
0.34
CS Steam price factor 0.0129 $/kg
CQ Heat price factor 1.26 105 $/kJ
0.33
0 200 400 600 800 1000 1200
DCC Decoking cost per cycle 66,000 $
c Density 1600 kg/m3 Producon me [h]
max Maximum coke thickness 0.0135 m
Fig. 17. Steam-to-propane ratio with production time.

(c) Only ethylene and propylene are assumed to be sold. Methane


5.2. Results
was not included because of its relative low price.
Constraints were set for the optimization variables. Since
Dynamic optimization is performed in gPROMS with a control
Sundaram and Froment (1979) developed their reaction scheme
interval of 50 h. The results are compared to the base case used
for a certain operating range, the same limits were applied:
for the model validation. Table 5 shows a signicant increase of
13.1% of the yearly prot ($851,840 for the base case and $963,279
973 K < coil outlet temperature for process gas < 1143 K
for the optimal case). The run length is increased to 55.6 days com-
0.2 kg/kg < steam-to-propane ratio < 1.0 kg/kg
pared to 39.8 days for the base case. In Table 5, two values are given
for pressure drop, heat transfer rate and propane conversion, they
Moreover, a constraint was applied for the maximum temper- represent values at the start and end of the run length respectively.
ature that can reach the tube metal. The material considered is The prole of the steam-to-propane ratio is investigated in
HK40 and this limit (for tube outer wall temperature) was set at Fig. 17. The optimized ratio increases from 0.34 kg/kg up to
1338.15 K (Albright, Crynes, & Corcoran, 1983). Table 4 summarizes 0.38 kg/kg during the rst 400 h, and then decreases slowly down
the parameters assumed for the optimization. to 0.35 kg/kg at the end of the cycle. The ratio is however always
Price factors, decoking cost, steam price factor, decoking time lower than in the base case. Fig. 18 shows the optimal tube outer
per cycle and yearly production time were taken from Gao, Wang, wall temperature at the outlet changing with the production time.
Ramshaw, et al. (2009). As for chemicals price factors, raw mate- It can be seen that in the optimal case, the temperature is always
rial would be propane and the two most desirable products are between 25 and 30 K lower than in the base case. The process gas
propylene and ethylene. Spot prices from May 2010 have been con- temperature at the outlet is also given in Fig. 18. In the base case,
sidered (Chemical Weekly, 2010). Finally, coke density was taken the process gas temperature at the outlet is maintained constant.
from Sundaram and Froment (1979). Shutdown is assumed to take While in the optimal case, it increases with production time. The
place when coke thickness reaches 25% tube diameter reduction at overall trend is an increase from 1087 K at the beginning to 1100 K
the outlet. It should be emphasized that the diameter reduction is at the end of the run length. A cycle takes place, due to an increase
therefore much lower for the rest of the tube. of the tube outer wall temperature, roughly every 2 days. Fig. 19
A new objective function was developed in this paper based on shows the tube outer wall temperature prole at the beginning of
Gao, Wang, Ramshaw, et al. (2009). The principle is to calculate the run length and at the end. Figs. 20 and 21 show the ethylene
the prot that is made every hour, and to accumulate those prots yield and propylene yield in both optimal and base cases. In the
over the whole duration of a cycle. A cycle includes the time of
operation of the reactor plus the time of decoking operation. Thus,
the decoking frequency is also taken into account. The objective 1220
Tube outer wall temperature
Temperature at the outlet [K]

function is as follows: 1200


Prot 1180
= Income Cost (13)
t
 1160 Opmal case
tc
Income = Fi Ci (14) 1140 Base case
tc + td
tc
 DCC
 1120 Process gas temperature

Cost = F0 C0 + Fs Cs + Qt CQ + (15) 1100


tc + td tc
1080
More particularly, the run length (i.e. the time between two
0 200 400 600 800 1000 1200
consecutive decoking operations) tc can be determined as follows: Producon me [h]
max c
tc = (16) Fig. 18. Tube outer wall temperature and process gas temperature at the outlet with
rc (z = 95) production time.

Table 5
Optimization results.

Parameter/variable Unit Dynamic optimization Base case

Pressure drop Pa 95,612183,255 95,758181,700


Radiant heat transfer rate kJ/h 1.65 106 1.74 106 1.77 106 1.72 106
Propane conversion % 90.286 82.381.6
Objective function $/year 963,279 851,840
Run length days 55.6 39.8
2884 M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885
Tube outer wall temperature [K]
1240 Base case sented in Table 5. This can explain why the yearly prot for the
1220 Opmal optimal case ($963,279) is much higher than for the base case
1200
case ($851,840).
The steam-to-propane ratio has impact on the amount of desired
1180
products manufactured and the run length. The optimization study
1160
was carried out when the molar ow rate in the feed to PFR was
1140
xed at 87.75 kmol/h. This is to consider the maximum capacity
1120
in the reactor in terms of volumetric ow rate. Under this con-
1100
dition, the increase of steam-to-propane ratio means more steam
1080
is added, then the propane in the feed is reduced. This indicates
1060
0 20 40 60 80 100 lower amount of products will be manufactured. Secondly, increase
Tube length [%] of steam-to-propane ratio can reduce coking rate. This has been
studied in detail in Berreni (2010).
Fig. 19. Tube outer wall temperature prole at the beginning and at the end of the Process gas temperature also has impact on the amount of
run length. desired products manufactured and the run length (Berreni, 2010).
On one hand, higher temperature means faster cracking reactions
36 (i.e. more desired products manufactured at unit time). On the
Ethylene yield [wt%]

35 Base case other hand, this higher temperature speeds up coking, which sub-
34 Opmal case
sequently makes run length shorter.
33 Following the discussions on the impacts of steam-to-propane
32
ratio in the feed and process gas temperature prole on the desired
products manufactured and the run length, the explanations for
31
the optimal steam-to-propane ratio prole (Fig. 17) and process
30
gas temperature prole (Fig. 18) will be easier to understand.
29
At the start (i.e. production time 0 h in Fig. 17), the steam-to-
28 propane ratio for the optimal case is 0.34 compared with 0.40 for
0 200 400 600 800 1000 1200
the base case. The process gas temperature (Fig. 18) is still quite low
Producon me [h] (1087 K). At this temperature, the coking rate is quite slow, using
lower steam-to-propane ratio would get more benets since more
Fig. 20. Ethylene yield with production time.
desired products will be manufactured.
From 0 h to 400 h (Figs. 17 and 18), the process gas temper-
optimal case, ethylene and propylene yields are maintained con- ature increases all the time, the cracking reactions speed up, so
stant respectively around 30.5 wt% and 18 wt%. On the contrary, in does the coking reaction. Generally the reaction rate doubles every
the base case, the propylene yield increases and the ethylene yield 10 K increase of the temperature. Steam-to-propane ratio has been
decreases considerably. increased to compensate. So the increase of the coking rate would
not be detrimental to the increase of yearly prot.
5.3. Discussion At around 400 h (Figs. 17 and 18), the process gas temperature
has reached a point, at which the economic benet brought by
This section is to explain the results presented in Section 5.2. continuous increase of steam-to-propane is not higher than that
Here the product prices and the raw material price are xed brought by keeping or even reducing steam-to-propane ratio. This
(Table 4). In order to maximize the objective function (i.e. Eq. is because lower steam-to-propane ratio means more propane is
(13)) for higher yearly prot, we can either increase the amount fed to the PFR to get more products. That is why the steam-to-
of desired products (i.e. ethylene and propylene) manufactured or propane ratio reaches its maximum.
increase the run length. It should be noted that the results presented in Table 5 and from
Even though the ethylene product yield for the optimal case is Figs. 17 to 21 were obtained assuming xed product prices and raw
always lower than for the base case (Fig. 20), the propylene yield for material price (Table 4). In real life, these prices uctuate all the
the optimal case is always higher than for the base case (Fig. 21). time. Therefore the methodology developed in this paper should
Ethylene and propylene are all desired products and their prices be very helpful for determining most protable operating strategies
are close to each other (Table 4). Adding the two product yields under different market conditions.
together, the total for the optimal case is slightly lower than for
the base case. However, the run length for the optimal case (55.6
days) is much longer than for the base case (39.8 days), as pre- 6. Conclusions

20 In this paper, modelling, simulation and optimization of ther-


mal cracking of propane in a tubular reactor were implemented
Propylene yield [wt%]

19
18 successfully in gPROMS . A 1D pseudo-dynamic model consider-
17 ing the radiant section alone was developed and the model was
16 validated. The simulated results were in good agreement with the
15 industrial data. Case studies were then carried out to investigate
14 Base case the behaviour of the process under different process gas tempera-
13
ture proles and under a xed tube outer wall temperature prole.
Opmal case
Finally dynamic optimization was performed with steam-to-
12
0 200 400 600 800 1000 1200 propane ratio and tube outer wall temperature prole as optimiza-
Producon me [h] tion variables. Operating prot was dened as objective function.
The impact of the coke buildup on the run length and on the prod-
Fig. 21. Propylene yield with production time. ucts yields was taken into account. The cumulative operating prot
M. Berreni, M. Wang / Computers and Chemical Engineering 35 (2011) 28762885 2885

over a period of 1 year was increased by 13.1% compared to the base Gao, G.-Y., Wang, M., Pantelides, C. C., Li, X.-G., & Yeung, H. (2009). Mathemati-
case, proving the efciency of the method developed in this paper. cal modeling and optimal operation of industrial tubular reactor for naphtha
cracking. Computer Aided Chemical Engineering, 27, 501506.
Gao, G.-Y., Wang, M., Ramshaw, C., Li, X.-G., & Yeung, H. (2009). Optimal operation
Acknowledgements of tubular reactors for naphtha cracking by numerical simulation. Asian Pacic
Journal of Chemical Engineering, 4(6), 885892.
Ghashghaee, M., & Karimzadeh, R. (2007). Dynamic modelling and simulation of
The authors would like to thank Dr Praveen Lawrence from PSE steam cracking furnaces. Chemical Engineering and Technology, 30(7), 835843.
Ltd for help to use optimization tool in gPROMS . Meihong Wang Lan, X., Gao, J., Xu, C., & Zhang, H. (2007). Numerical simulation of transfer and reac-
would like to acknowledge the helpful discussions with Professor tion processes in ethylene furnaces. Chemical Engineering Research and Design,
12(85), 15651579.
Costas C. Pantelides from Centre for Process Systems Engineering Masoumi, M. E., Sadrameli, S. M., Towghi, J., & Niaei, A. (2006). Simulation, opti-
(CPSE), Department of Chemical Engineering, Imperial College mization and control of a thermal cracking furnace. Energy, 31(4), 516527.
London. Shahrokhi, M., & Nejati, A. (2002). Optimal temperature control of a propane thermal
cracking reactor. Industrial & Engineering Chemistry Research, 41(25), 65726578.
Sowers, G., & Reed, C. (2001). Dynamic optimization of ethylene furnaces cracking
References propane. In Proceedings of AIChE Ethylene Producers Conference Houston, TX, (pp.
366405).
Albright, L. F., Crynes, B. L., & Corcoran, W. H. (1983). Pyrolysis Theory and Industrial Sundaram, K. M., & Froment, G. F. (1977). Modelling of thermal cracking kinetics-I.
Practice. New York, USA: Academic Press Inc. Thermal cracking of ethane, propane and their mixture. Chemical Engineering
Babu, B. V., & Knovel. (2004). Process plant simulation. New Delhi, India/New York: Science, 32(6), 601608.
Oxford University Press. Sundaram, K. M., & Froment, G. F. (1979). Kinetics of coke deposition in the thermal
Berreni, M. (2010). Modelling, simulation and dynamic optimization of thermal cracking cracking of propane. Chemical Engineering Science, 34(5), 635644.
of propane for ethylene industry. MSc Thesis, Craneld University, UK. Sundaram, K. M., Shreehan, M. M., & Olszewski, E. F. (2001). Ethylene. In J. I.
Chemical Weekly (May 2009May 2010). CW PRICE REPORT, available at Kroschwitz, A. Seidel, & R. E. Kirk, et al. 5th (edtors), Kirk-Othmer Encyclopedia
http://web.ebscohost.com/ehost/pdfviewer/pdfviewer?vid=4&hid=11&sid=123 of Chemical Technology. Hoboken, NJ: Wiley-Interscience.
b8038-b568-4a5c-8f37-d95f56e765c6%40sessionmgr14 (last accessed in July Van Damme, P. S., Narayanan, S., & Froment, G. F. (1975). Thermal cracking of propane
2010). and propane-propylene mixtures: Pilot plant versus industrial data. AIChE Jour-
Detemmerman, T., & Froment, F. (1998). Three dimensional coupled simulation of nal, 21(6), 10651073.
furnaces and reactor tubes for the thermal cracking of hydrocarbons. Oil & Gas Zou, R., Lou, Q., & Liu, Z. (1988). Study of kinetic models for the pyrolysis reaction of
Science and Technology Review. IFP, 2(2), 181194. propane. Journal of Analytical and Applied Pyrolysis, 13(3), 183190.

You might also like