You are on page 1of 13

See

discussions, stats, and author profiles for this publication at:


https://www.researchgate.net/publication/279897901

Simulation and optimization of a naphtha


thermal cracking pilot plant

Article in IRANIAN JOURNAL OF CHEMISTRY & CHEMICAL ENGINEERING-INTERNATIONAL ENGLISH


EDITION January 2003

CITATIONS READS

7 80

4 authors, including:

S.M. Sadrameli Jafar Towfighi


Tarbiat Modares University Tarbiat Modares University
80 PUBLICATIONS 914 CITATIONS 139 PUBLICATIONS 1,148 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Production of valuable chemicals View project

Simulation and numerical investigation of Propane thermal cracking Furnace and Reactor
View project

All content following this page was uploaded by S.M. Sadrameli on 22 November 2015.

The user has requested enhancement of the downloaded file.


Energy 31 (2006) 516527
www.elsevier.com/locate/energy

Simulation, optimization and control of a thermal cracking furnace


M.E. Masoumia, S.M. Sadramelia,*, J. Towfighia, A. Niaeib
a
Chemical Engineering Department, Tarbiat Modarres University, P.O. Box 14115-143, Tehran, Iran
b
Department of Applied Chemistry, University of Tabriz, 51666-14766 Tabriz, Iran
Received 2 November 2003

Abstract
The ethylene production process is one of the most important aspect of a petrochemical plant and the cracking
furnace is the heart of the process. Since, ethylene is one of the raw materials in the chemical industry and the
market situation of not only the feed and the product, but also the utility is rapidly changing, the optimal operation
and control of the plant is important. A mathematical model, which describes the static and dynamic operations of
a pilot plant furnace, was developed. The static simulation was used to predict the steady-state profiles of
temperature, pressure and products yield. The dynamic simulation of the process was used to predict the transient
behavior of thermal cracking reactor. Using a dynamic programming technique, an optimal temperature profile
was developed along the reactor. Performances of temperature control loop were tested for different controller
parameters and disturbances. The results of the simulation were tested experimentally in a computer control pilot
plant.
q 2005 Elsevier Ltd. All rights reserved.

1. Introduction

The bulk of the worldwide annual commercial production of ethylene is based on thermal cracking of
petroleum hydrocarbons with steam; the process is commonly called pyrolysis or steam cracking.
Hydrocarbon feed mixed with process steam is introduced into the tubular reactors (cracking coils) with
short residence time and at a high temperature. Steam is used to increase the olefin selectivity and to
reduce the coke formation by decreasing the hydrocarbon partial pressure.

* Corresponding author. Visiting Faculty: Chemical Engineering, University of Florida, Chemical Engineering Building,
Gainesville, FL 32611-6005, USA. Tel.: C1 352 392 0862; fax: C1 352 392 9513.
E-mail address: sadrameli@che.ufl.edu (S.M. Sadrameli).
0360-5442/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2005.04.005
M.E. Masoumi et al. / Energy 31 (2006) 516527 517

Nomenclature
Ci concentration of coke precursors, mole/m3
Cp heat capacity, J/mole K
dt tube diameter, m
F molar flow rate, mole/h
Fr friction factor
G total mass flux of the process gas, kg/m2s
DH heat of reaction, J/mole
Kc proportional constant
Mm average molecular weight, kg/mole
Pt total pressure, kPa
PI performance index
PI proportional integral control
PID proportional integral derivative control
Q heat flux, W/m2
Rb radius of the tube bend, m
Re reynolds number
r tube radius, m
rc coking reaction rate, kg/m3s
rri reaction rate in pyrolysis process, mole/m3s
tc coke thickness, m
t time, h
Td derivative control constant
TI integral Control constant
sij stoichiometry factor
T temperature, K
z axial reactor coordinate, m
Greek letters
a coking factor
L angle of bend 0
rc coke density, kg/m3
h unit conversion factor
Abbreviations
COT coil outlet temperature
XOT cross over temperature
S/HC steam to hydrocarbon ratio
IAE integral of absolute value of error

The paraffinic feedstock is thermally cracked into mainly olefins, aromatics, methane and hydrogen.
The homogeneous cracking reactions are endothermic and energy input is required in order to reach the
gas temperature as high as 800900 8C at the coil outlet. The required energy is supplied by a thermal
cracking furnace. In fact, hydrocarbon feed is mixed with steam in the convection section of the furnace
518 M.E. Masoumi et al. / Energy 31 (2006) 516527

and the temperature of the mixture is raised to the cracking temperature. The mixture is then fed into the
radiant section of the furnace, where the temperature of the gas mixture increases rapidly to the desired
cracking temperature. In the radiant section, the hydrocarbon is cracked to a combination of olefins,
aromatics, pyrolysis fuel oil and other heavier hydrocarbons. Upon leaving the radiant section of the
furnace the cracked gas is cooled rapidly to stop the undesired reactions [1].
Coil Outlet Temperature (COT) is an important parameter affecting yields of ethylene production
thus, should be controlled. Since, furnaces are the first step of the production process, the entire process
is affected by disturbances that occur due to the furnace operation. The heat input to the furnace reactor is
manipulated and controlled by measuring the cracked gas temperature in the coil outlet (COT). This loop
is one of the most important loops in the control of thermal cracking furnaces.
A pilot plant is designed and set up to investigate the effects of different parameters on the products
yields. A computer program simulates the parameters of the furnace before running the pilot plant. In
this work, the effect of furnace parameters on the reactor yields is studied.

2. Pilot plant

The schematic diagram of the pilot plant is shown in Fig. 1. The reactor feed contains at least two
streams: the hydrocarbon and the dilution steam. Liquid hydrocarbons and water, used to produce
dilution steam, are fed by means of two dosing pumps. The feed flow rates and S/HC (steam to
hydrocarbon) ratio can be altered between 515 g/min and 0.30.8, respectively.
The furnace preheaters consist of two electrical coils for heating water and hydrocarbon feeds. The
reaction section is divided into eight zones, which can be heated independently to achieve the desired
temperature profile. The reactor is a tube of 1 m lengths, and 0.01 inside diameter made of Inconel (alloy
600 HS 2). Temperature of different parts is measured by eight thermocouples as shown in Fig. 1.
After cooling, the reactor effluent to the appropriate temperature in a double pipe heat exchanger,
liquids and tars are separated from the cracked gas. A fraction of the product gas is then withdrawn for
the on-line analysis via a gas chromatograph, while the rest is sent directly to the flare. The system is
connected to a personal computer through the interface cards for monitoring and control purposes.

3. Static simulation

The heart of an ethylene plant is the cracking furnace. For finding an optimal operating strategy, it is
necessary to investigate the influences of the operating parameters, which can be satisfactorily calculated
through the rigorous modeling. Rao et al. [2] simulated the reactor and the radiant box simultaneously.
Several packages were developed by other researchers [36].
In almost all of the ethylene plants in Europe, naphtha is the most widely used feed material for the
thermal cracking furnaces. Naphtha is a mixture of complex hydrocarbon materials, which ranges mostly
from C5 to C10 paraffins. In the reactor, numerous cracking reactions occur to produce ethylene and
propylene. The reaction mechanism of thermal cracking of hydrocarbons is free-radical chain reaction
[7]. In this work, a free-radical reaction set with the kinetic parameters for 90 species and 543 reactions
has been used.
M.E. Masoumi et al. / Energy 31 (2006) 516527 519

INPUT
SIGNALS

TT TT
TT
SAFETY INTERFAC
E CARD
WATER
VALVE

LIQUID OUTPUT
WATER SIGNALS
FEED PRE TER TT
HEA
BALANCE BALANCE PUMP
PUMP
TT
WT WT
TT

F UR NA CE
1
1 TT

2
2 TT
FEED
3 TT

4
4 TT

PRE 5 TT
HEA TER

ZO NE
COMPRESSOR TT TT
6
6 TT
AIR FILTER TT 7 TT
AIR FOR DECOKING

8
8 TT

TT

TT

VENT
FLARE

EXCHANGER
Gas
GAS
COUNTOR
Flowmeter

G.C. EXCHANGER

H.C. WATER
GASEOUS FEED CONDENSATE CONDENSATE

Fig. 1. Schematic diagram of the thermal cracking pilot plant.

3.1. Reactor model

The geometry of the model configuration for the reactor tube is shown in Fig. 2. The following
assumptions have been considered for the mathematical model:

1. One dimensional flow


2. Plug flow and laminar regime
3. Radial concentration gradients and axial dispersion are negligible
4. Ideal gas behavior
5. Inertness of the steam diluents in feed
6. No hydrodynamic or thermal entrance region effects
7. Quasi steady-state in Coke deposition model.

In this form, the coking rate model is pseudo steady-state with respect to time. In other words, coking
rate is assumed to be constant over a time step and the effect of coke formation through coking equation
520 M.E. Masoumi et al. / Energy 31 (2006) 516527

Fig. 2. Differential element of a cracking coil.

is updated explicitly at the end of each time step. This pseudo steady-state assumption would be indeed
valid as long as the coke formation rate does not change appreciably over a sufficiently small time step.
The set of continuity equations for the various process gas species is solved simultaneously with the
energy, momentum and the coking rate equations required [11]. These equations are as follows:
Mass balance:
!
dFj X pdt2
Z sij rri (1)
dz i
4

Energy balance:
X dT pd2 X
Fj cpj hQzpdt C t rri KDHi (2)
j
dz 4 i

Momentum balance:
     
1 Pt dpt d 1 1 1 dT
K Z C C Fr (3)
Mm Pt hG2 RT dz dz Mm Mm T dz
With the friction factor:
ReK0:2 z
Fr Z 0:092 C (4)
dt pRb
And for the tube bends as:
  
L dt
z Z 0:7 C 0:35 0 0:051 C 0:19 (5)
90 Rb
where Rb and L are tube bend radius and bend angle, respectively. Since, the coking is slow, quasi
steady-state conditions may be assumed, so that, we can write the rate of coke formation:
vC ar
Z dt K 2tc c (6)
vt rc
M.E. Masoumi et al. / Energy 31 (2006) 516527 521

Table 1
Specification of naphtha feed (wt%)
Carbon no. n-Paraffins Iso-paraffins Naphthenes Aromatics
4 0.22 2.67
5 25.22 17.94 4.19
6 14.88 23.41 2.82 2.0
7 1.67 3.27 0.97
8 0.57 0.2
Total 41.99 47.83 7.01 3.17

Using the mathematical model, the amount of coke deposited on the internal wall of the reactor tubes
has been calculated with a limiting value for tube skin temperature (1100 8C). The governing mass,
energy, and momentum balance equations for the cracking coil constitute the two-point boundary value
problem, which is highly stiff. The implicit Euler method [8] is used for solving the equations. The rate
of coke formation has been taken into account [911]. The tuning parameters, such as overall heat
transfer coefficient and coking rate factor can be adjusted to make the model prediction close to the
actual data [12]. The developed software receives the feed specifications and provides products yield and
gas temperature profile. Details of the static simulation were published recently by Masoumi et al. [13].

3.2. Simulation results

Specifications of naphtha feed and operating conditions are given in Tables 1 and 2, respectively.
After running, the program under these conditions, furnace temperature, gas temperature and product
yield profiles are obtained along the reactor. Different temperature profiles were applied to the reactor
and the simulation results at the reactor outlet are presented in Table 3. Fig. 2 shows that increasing COT
will increase the ethylene yield while decreasing the propylene yield. The gas temperature profiles
resulted from the simulation are illustrated in Fig. 3.

3.3. Optimal reactor temperature

During a production period, coke is formed and deposited on the inner tube surface of the cracking
coil. The coke formation reduces the olefin selectivity mainly due to the pressure drop increase in the
cracking coil. The coke formation also leads to excitation of one or more of the furnace operational
constraints at the end of production period. Coke is burned-off with steam and air during a 14 days

Table 2
Operating variables
Inlet temperature (8C) 600
Inlet pressure (atm. abs.) 3
Feed flow rate (g/min) 10
Inlet pressure (atm. abs.) 3
COT (8C) 83070
S/HC ratio 0.7
522 M.E. Masoumi et al. / Energy 31 (2006) 516527

Table 3
Static simulation/experimental results
COT (8C) C2H4 yield (wt%) C3H6 yield (wt%) Furnace temp. (8C)
800 /19.76 /18.32 870
810 /21.15 /18.74 /880
820 /21.42 /18.73 /890
830 26.79/22.00 11.80/18.63 913.0/900
840 28.81/22.61 12.07/18.77 924.5/910
850 30.66/23.42 12.17/18.63 935.5/920
860 32.15/24.02 12.10/18.59 945.5/930
870 33.43/24.83 11.86/18.52 954.0/940
880 34.40/23.97 11.51/18.35 962.0/950
890 35.18/26.43 11.02/17.55 970.0/960
900 35.68/27.72 10.48/16.99 977.0/970

decoking period. The decoking operation has to be done every 3080 days depending on furnace design
and operating philosophy [14,15].
While the reactor is in production, the thickness of the carbon deposit increases monotonically. At the
same time, the instantaneous net earning from the operation decreases. For chemical and petrochemical
plants, it seems that maximizing the operating profit would always give a satisfactory solution. The
objective function must combine the favorable effect of higher valuable product yield with the negative
effect of the coking rate on the profit.
Increasing the reactor gas temperature increases the ethylene yield and consequently the income. On
the other hand, increasing temperature will also increase the rate of coke deposition as shown in Fig. 4
which results in higher cost. By maximizing the defined objective function using dynamic programming
technique, the optimal temperature profile along the reactor is obtained. The objective function proposed
by Towfighi [16] is used as follows
PI Z Income K cost
Income Z Ethylene mass flow rate  ethylene price  yearly operating time;
Cost Z L  ethylene mass flow rate  ethylene price  yearly decoking period;

Yield (wt%)
40
C2H4, Ex p.
C2H4, Sim.
C2H6, Ex p.
C2H6, Sim.
30

20

10
800 820 840 860 880 900
Coil outlet Temperature (C)

Fig. 3. Effects of COT on the product yields.


M.E. Masoumi et al. / Energy 31 (2006) 516527 523

Gas Temperature (C)


900
900C

800 830C

700

600
0 0.2 0.4 0.6 0.8 1
Reactor Length (m)

Fig. 4. Gas temperature profiles along the reactor tube at different COTs.

The labor and utility costs necessary for decoking, are expressed as a fraction, L, of term representing
the benefit lost due to the interruptions. The iterative dynamic programming procedure using systematic
region contraction is used for the optimization. The optimization criteria are the maximization of benefit,
since increasing COT, increase the coke and ethylene production. Increasing of ethylene production,
increases the income, but subsequently increases coke production, that results increase of decoking cost.
The details of the optimization procedure are discussed in [13]. The optimal gas temperature profile
obtained through simulation is shown in Fig. 5. As can be seen, the optimal COT is 877.34 8C. At this
condition the ethylene yield in the reactor effluent, is 33.737%.

4. Dynamic simulation

In order to investigate the performance of a control scheme, it is necessary to have a dynamic model
of the process. By applying conservation laws, such a model can be obtained. For simulating the system,
the unsteady mass and energy balances must be solved simultaneously with pressure drop equation. It is
difficult to measure the gas temperature inside the reactor due to the coke deposition. If the desired
furnace wall temperature corresponding to desired reactor temperature is known, it would be much
easier to measure and control the furnace wall temperature than the reactor temperature. Therefore, if

Coke deposition (gr/min)


0.016

0.012

0.008

0.004

0
830 840 850 860 870 880 890 900
Coil outlet temperature (C)

Fig. 5. Effect of COT on the coke deposition.


524 M.E. Masoumi et al. / Energy 31 (2006) 516527

the gas temperature in the reactor is known then furnace wall temperature can be obtained by back
calculation. Using the optimal temperature profile obtained in Section 3, furnace wall temperature is
calculated. Furnace is divided into eight thermal zones. The desired optimal furnace wall temperature
can be approximated by eight step functions. Electrical heaters are used to provide necessary heat. The
transfer function between electrical voltage of element and related temperature of each zone is assumed
a first order plus lag dynamic and its parameters are obtained experimentally as: KZ100, TZ5,
TdZ2 min. Detail of the dynamic simulation are presented by Shahroki and Nedjati [17].

4.1. Furnace wall temperature control

As mentioned, the furnace wall temperature can be obtained from the optimal reactor gas temperature
using by-back calculation. Discretization of this profile provides the desired temperature of each zone.
For temperature control of each zone, a PI controller can be used. The controller parameters are
calculated based on open loop response and minimizing the integral of absolute value of the error (IAE).
In this research controller parameters, Kc and TI are set to 0.1 and 0.004, respectively.

4.2. Reactor temperature control

It is almost impossible to measure the gas temperature inside the reactor due to the coke deposition. In
this case, the gas temperature measurement is available at the end of each zone. The set point of each
controller is the temperature at the outlet of corresponding zone. To evaluate the performance of control
system, the reactor temperature is changed and the controllers activate one at a time. In this case the
temperature overshoot is high. To get a better transient response, it is proposed to activate the first zone
controller and freeze the output of the other controllers. When the temperature error in the first zone
becomes less than a predetermined value, the second controller is activated and so on. In this case the
temperature overshoot is considerably smaller than the previous case. This control strategy cannot be
applied to the pilot plant due to the impossibility of the temperature measurement in the end of each
zone. Therefore, only simulation studies have been done and the results are published by Shahrokhi and
Nedjati [17].

4.3. COT control

One of the main objectives of this work is COT control. Therefore, furnace wall temperature will be
constant in all zones and furnace acts as a single zone furnace. The transfer function between electrical
voltage of element and COT is assumed a second order plus lag dynamic and its parameters are obtained
experimentally as: KZ100, TZ5, TdZ2 min. A PID controller, which was tuned by open loop
response, is used for system controlling. Controller parameters are 0.1 for proportional gain, 0.004 min
for integral time and 0.0001 for derivative time.

4.4. Ethylene yield control

Control of ethylene yield can be achieved by measuring of ethylene yield directly in coil outlet or by
measuring of COT and then evaluation of ethylene yield from COT (inferential control). If there is a
relation between COT and ethylene yield, or system is observable, then inferential control is practical.
M.E. Masoumi et al. / Energy 31 (2006) 516527 525

Gas temperature (C)


900

800

700

600
0 0.2 0.4 0.6 0.8 1
Reactor Length (m)

Fig. 6. Optimal gas temperature profile along the reactor.

In this work, the GC run time for cracked gas analysis is about 35 min that makes the on-line ethylene
yield control impossible. Therefore, this control strategy is eliminated from this work.

5. Experimental studies

Experiments are performed with typical naphtha as feed, which was supplied from Arak
Petrochemical Company (APC). The feed flow rates and S/HC ratio are 10 g/min and 0.7, respectively.
Preheaters and the furnace temperatures are set to desired values by using PID digital controllers. To set
the XOT to 600 8C, water and hydrocarbon preheater set points are set at 750 8C. By changing the
furnace set point, different COT values will be achieved. This enables different temperature profiles to be
applied to the reactor.
While running the pilot under these conditions, the effluent cracked gas is sent to the GC for the
analysis. The experimental results of this investigation are shown in Table 3 and Figs. 38. As can be
seen from Fig. 2, increasing COT, increases the ethylene and at the same time decreases the propylene
yield. XOT, COT, furnace temperature, and furnace set point are given in Figs. 7 and 8. As can be seen

Temperature (C)
1000
setpoint
furnace temp.
950

900

850

800
0 1 2 3 4 5 6 7 8 9
Time (hr)

Fig. 7. Temperature control of zone 8.


526 M.E. Masoumi et al. / Energy 31 (2006) 516527

Temperature (C)
1000
XOT
COT
900

800

700

600

500
0 1 2 3 4 5 6 7 8 9
Time (hr)

Fig. 8. Reactor inlet (XOT) and COT temperature diagrams.

from the figures the performance of control system in keeping constant XOT and tracking the set point
by changing COT and furnace wall temperature is fairly well. On the other hand, results of Fig. 3 show
that increasing COT, increases ethylene yield while decreases propylene yield which is confirmed by
the literature data. The XOT is approximately constant about 600 8C. Furnace temperature tracks furnace
set points with good competition.

6. Conclusions

A computer program for simulating the behavior of a thermal cracking pilot is developed based
on a kinetic model. This kinetic model consists of 543 radical reactions. The developed software is
used to simulate the static behavior of the process and find the effects of different parameters on
product yields. Simulation results indicate that increasing COT increases the ethylene yield and the
rate of coke deposition. A thermal cracking pilot plant including eight zoned electrical furnace was
designed and set up to investigate the influence of the furnace parameters on the product yields.
Finally the new developed optimal temperature profile for naphtha cracking was obtained in the
range of COT between 800 and 900 8C which have not been reported before in the literature.
Simulation results were tested experimentally by the pilot. Comparison of the simulation and
experimental results show similar trends. The discrepancies between the simulation and experimental
results are due to the nonlinear and unknown furnace dynamic model, and some measurement
experimental errors.

Acknowledgements

The authors are grateful to Dr M. Pishvaei and Mr A. Nedjati for the preparation of the control
software. The financial support by APC is also gratefully acknowledged.
M.E. Masoumi et al. / Energy 31 (2006) 516527 527

References

[1] Cugini J. Computer control of ethylene plant cracking furnaces. In: Grossman H, editor. Symposium on characterization of
thermodynamics and transport properties of polymer system, New Orleans, LA, USA.
[2] Rao RMV, Plehiers MP, Froment GF. The coupled simulation of heat transfer and reaction in a pyrolysis furnace. Chem
Eng Sci 1998;42(6):12239.
[3] Dente M, Ranzi E, Goossens GA. Detail prediction of olefin yields from hydrocarbon pyrolysis through a fundamental
simulation model (SRYRO). Comput Chem Eng 1979;3:6172.
[4] Towfighi J, Nazari H, Karimzadeh R. Development of mechanistic model for pyrolysis of naphtha.. APCCHE/CHEM-
ECA 93, Melbourne, Australia. vol. 3 1993 p. 33742.
[5] Joo E, Lee K, Lee M, Park S. CRACKER-a PC-based simulator for industrial cracking furnaces. Comput Chem Eng 2000;
24:15238.
[6] Goethem MWM, Kleinendorst FI, Leeuwen CV, Velzen NV. Equation-Based SPYRO model and solver for the simulation
of the steam cracking process. Comput Chem Eng 2001;24:90511.
[7] Joo E, Park S. Pyrolysis reaction mechanism for industrial naphtha thermal cracking furnaces. Ind Eng Chem Res 2002;40:
240915.
[8] Thomas P. Simulation of industrial process. London: Butterworth; 1999.
[9] Kopinke FD, Zimmermann G, Nowak S. On the mechanism of coke formation in steam cracking, conclusion from results
obtained by trace experiments. Carbon 1998;56(2):11724.
[10] Kopinke FD, Zimmermann G, Reyners G, Froment GF. Relative rate of coke formation from hydrocarbon in steam
cracking of naphtha, 2. parafins, naphthenes, mono, di and cyclo olefins and acetylenes. Ind Eng Chem Res 1993;32:
5661.
[11] Kopinke FD, Zimmermann G, Reyners G, Froment GF. Relative rate of coke formation from hydrocarbon in steam
cracking of naphtha 3. Aromatic hydrocarbons. Ind Eng Chem Res 1993;32:26205.
[12] Towfighi J, Niaei A, Hoseini S. Development of kinetic model for coke formation in thermal cracking of naphtha. Iran
J Sci Technol, Trans B 2001;23(B2):27584.
[13] Masoumi ME, Shahrokhi M, Sadrameli M. Simulation and optimization of an ethylene cracking pilot plant. Iran J Chem
Chem Eng (IJCCE) 2003;23(1):34555.
[14] Newberger MR, Kadl RH. Optimal operation of a tubular chemical reactor. AIChE 1971;17(6):13817.
[15] Ranzi E, Dente M, Pierucci S, Barendregt S. Optimization of pyrolysis furnace operation including run-time constraint.
Ingegnere Chimico Italiano 1981;17(11-12):7681.
[16] Towfighi J, Karimaei M, Karimzadeh R. Application of dynamic programming method in optimal control of light
hydrocarbons pyrolysis reactors. Sci Iran 1998;5:18.
[17] Shahrokhi M, Nedjati A. Optimal temperature control of propane thermal cracking reactor. Ind Eng Chem Res 2002;
41(25):65728.

View publication stats

You might also like