Вы находитесь на странице: 1из 15

I NTERNATIONAL J OURNAL OF C HEMICAL

R EACTOR E NGINEERING
Volume 2 2004 Article A15

A Laboratory Steam-Cracking Reactor to


Characterize Raw Materials
Attila Pinter∗ Antal Tungler† Lajos Nagy‡
László Vida∗∗ Imre Kovács†† Janos Kerezsi‡‡


BUTE, pinter.attila@freemail.hu

BUTE, atungler@mail.bme.hu

BUTE, lnagy@mail.bme.hu
∗∗
BUTE, vidal@mail.bme.hu
††
MOL, ikovacs@mol.hu
‡‡
BUTE, kerezsi@vnet.hu

ISSN 1542-6580
Copyright 2004
c by the authors.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system,
or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of the publisher, bepress, which has been given
certain exclusive rights by the author.
A Laboratory Steam-Cracking Reactor to
Characterize Raw Materials
Attila Pinter, Antal Tungler, Lajos Nagy, László Vida, Imre Kovács, and Janos
Kerezsi

Abstract
A laboratory steam-cracking reactor was designed, built and tested in order to
compare yields under different operating conditions. A typical industrial gasoline
was used as feedstock to test the reactor. The design of the reactor was found
appropriate to crack hydrocarbon feedstocks and the olefin yields follow the ex-
pected trends.

KEYWORDS: pyrolysis, steam cracking, gasoline, laboratory reactor, feedstock


characterisation
Pinter et al.: Steam Cracking Lab Reactor 1

1. INTRODUCTION

The worldwide production of ethylene is based on thermal cracking of petroleum hydrocarbons with steam; the
process is called pyrolysis or steam cracking. The hydrocarbon stream is heated, mixed with steam, and further
heated to incipient cracking temperature (500-650°C), and then the mixture enters a fired tubular reactor where the
reactions take place during several hundred milliseconds at higher temperatures (750-875°C). The saturated
hydrocarbons of the feedstock crack into smaller molecules according to free radical reactions. The major products
are hydrogen, methane, ethylene, propylene, other olefins and diolefins. Upon leaving the reactor, the products are
cooled immediately to 550-650°C to prevent secondary reactions, and then the gases are separated. The olefin yield
is influenced by residence time, temperature and hydrocarbon-steam ratio. The increase of residence time or
temperature raises the conversion and the yields; the steam serves to decrease the carbonization of the inside wall of
the reactor and reduces the partial pressure of the hydrocarbons. According to Ullmann’s Encyclopaedia of
Industrial Chemistry (1987), the thermodynamic stability of higher molecular weight hydrocarbons is lower than
that of lighter compounds; therefore gas oil requires lower pyrolysis temperatures than gasoline. It is interesting to
note that, in the US, the light paraffins (ethane, propane) dominate as the steam crackers feedstock of choice,
whereas in Europe gasoline, and to a smaller extent gas-oil, are the main preferred feedstocks.

Pyrolysis of hydrocarbons has been studied for decades. Much effort has been devoted to the development
of mathematical models for pyrolysis reactions to predict the product composition obtained from various feedstocks
under different furnace conditions. Safarik et al. (1996, 1997) and Billaud et al. (1991) stated that researchers have
examined the thermal degradation of hydrocarbons in detail, and that the processes occurring during thermal
cracking of hydrocarbons are well understood. Modelling of industrial hydrocarbon pyrolysis is a powerful tool to
optimize feedstock selection and operating conditions. It has been proved that only models based on free radical
reactions have technical adaptability. An example is the SPYRO model, which was described first by Dente et al.
(1979), and recently by van Goethem et al. (2001). Broad accessibility to this model is limited due to financial
reasons. Another approach is the molecular kinetics model, different versions of which can be found in the paper by
Belohlav et al. (2003), or by Pant and Kunzru (1997, 1998). This model applies formal molecular reactions. Each
feed component (individual hydrocarbon molecule) cracks according to a first order kinetic model; each reaction has
its own basic rate constant and energy of activation. Because the complexity of the cracking furnace modelling, it is
necessary to make simplifications in the:
- pyrolysis reaction scheme,
- modelling of heat transfer and gas flow,
- coke deposit formation,
- characterization of pyrolysed hydrocarbon feedstock.
The reliability of the simplified models is limited; they belong to semi-mechanistic or semi-empirical models. The
models often adjust the kinetics parameters, several experiments are carried out, and thereafter the experimental
results are compared with the predicted yields. Due to the limited amount of reliable experimental data, the
applicability of the models is restricted; the deviation may be significant when applied to another reactor or to a
different type of feedstock. In the absence of an appropriate technical model, the characterisation of different
feedstocks can be made with a laboratory pyrolysis reactor.

The design of the reactor and the experiments reported in this work were initiated and supported by a joint
industrial-state foundation. The primary aim of the experimental work was to characterize and compare the olefin
yields of different commonly used raw materials. In addition, our further goal was to explore the steam-cracking
behaviour of other possible feedstocks, such as high molecular weight mixtures formed in polyolefin production or
recycling which have not been previously studied experimentally in detail.

2. COMPARISON OF DIFFERENT REACTORS FOR HYDROCARBON PYROLYSIS

In the research and development of olefin production processes, the question often arose about what type of
laboratory reactor could simulate the best the commercial reactors with respect to olefin yields. Szepesy (1980)
carried out a comparative study. In three different types of experimental reactors, Romaskino type gasoline was
cracked. The most significant dimensions and parameters of these reactors are included in Table 1.

Published by The Berkeley Electronic Press, 2004


2 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

Table 1. Comparison of different pyrolysis reactors

Parameter Laboratory tube Pilot scale reactor Experimental


reactor commercial reactor

Tube diameter, mm 6 6 19
Tube length, m 1.3 24 80
Length of reaction zone, m 0.7 22.7 50 – 60
Max. feed, kg/s 1.4*10-4 9.7*10-4 1.2*10-2
Temperature profile isothermal controlled controlled
Heater electric electric PB gas burner
Temperature measuring whole length 6 point 7 point
Pressure profile isobaric non-isobaric non-isobaric
Intermediate sampling none 5 point 7 point
Pressure drop 2 –5*102 Pa 2*105 Pa 2.5*105 Pa
Outlet temperature, ˚C 580 – 800 770 - 860 730 - 860
Residence time, s 0.1- 8 0.4 – 2 0.4 – 2
Steam, wt% 0 – 50 40 - 60 40 - 60

The Author presented the change of the yields of hydrogen, methane, ethylene, propylene, butadiene, benzene and
other aromatics with the increasing severity in the three reactors. By using of severity function, instead of
conversion, the yields in non-isothermal, non-isobaric reactors can be calculated on the basis of laboratory reactor
yields (isothermal, isobaric). Linden and Peck (1955) defined the severity function with the following formula:

S = t IJ0.06 (1)

where t is the outlet temperature of the reactor in ºC, and


IJ is the residence time in seconds.

Although the loading of the laboratory and of the commercial reactor is different by two orders of
magnitude, at the same severity, the differences of the yields are only within 4 to 7%. The olefin yields in the
laboratory reactor are typically higher, but the yield curves are completely similar. It can be concluded that, by using
a given feedstock in a laboratory reactor, it is possible to predict the expected value and tendency of the yields of
commercial reactors.

3. ASPECTS OF THE DESIGN OF THE LABORATORY REACTOR


A primary objective was to develop a laboratory-scale cracking reactor that could operate under industrial conditions
(temperature and residence time). Moreover, the reactor had to be sufficiently large to allow the closure of material
balances within a few hours of operation. The most important parameters of the pyrolysis, i.e. the residence time,
temperature, and hydrocarbon/steam ratio for a commercial process using both gasoline and gas-oil are reported
(Albright, 1983). In order to characterize the different feedstocks, it is necessary to crack under a variety of
operating conditions, whose limits are also illustrated in Table 2.

The operation of the reactor beyond these limits is not necessary since cracking of gas-oils below 800 °C is
not economic. On the other hand, the use of high temperature (above 890 °C) under hundreds of milliseconds
contact time is not desirable, because secondary reactions decrease yields of valuable components, the exception
being only the pyrolysis of ethane. Similarly, an increase in steam dilution beyond the limits would be beneficial in
reducing coking, and but it would consume an excessive amount of energy. However, a reduction in steam would
negatively affect the coking of the reactor.
The dimensions of our lab scale reactor were calculated according to the following considerations. During a
two-hours run, we set the target to process 1 to 2 L of raw material in order to achieve a reproducible material
balance and to perform the standard distillation of the condensate.

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 3

Table 2. Reactor parameters

Commercial conditions Limiting values of laboratory conditions


Gasoline Gas-oil Min. Max.
Residence time, s 0.3 0.3 0.2 0.5
Temperature, °C 860 830 800 890
Steam/CH mass-ratio 0.54 0.72 0.33 1.0
Steam fraction 0.35 0.42 0.25 0.5

The parameters used for the calculation are the following:


T=830°C=1103 K, P=105 Pa , mSTEAM/mCH=0.67, feedstock average molecular weight: MCH=100 g/mol, feedstock
mass flow: mCH=2.78*10-4 kg/s (1 kg/h), mSTEAM=1.85*10-4 kg/s (0.667 kg/h), feedstock molecular flow rate:
nCH=2.78*10-3 mol/s (10 mol/h), nSTEAM=1.03*10-2 mol/s (37.0 mol/h), nSUM=1.31*10-2 mol/s (47.2 mol/h).

Since

nRT
w (2)
P
As a result, if the volumetric flowrate is w=1.20*10-3 m3/s and the residence time is IJ = 0.3 s, then the
reactor volume is V=0.36*10-3 m3. In addition, in order to avoid the blocking of the reactor as a result of coke
deposition, the reactor tube should be straight and have a diameter of at least 0.02-0.03 m. Therefore, for the lab
scale tube, a length L=0.85 m and a diameter D=0.024 m were chosen, corresponding to a volume of 0.384 L.

The flow pattern (laminar or turbulent) in the reactor, as characterized by the Reynolds number, has a
significant influence on the operation of the reactor. Typical data relative to a commercial furnace (for one tube) are
(Ullmann’s encyclopaedia, 1987): D=0.18 m, L=100 m, mCH =0.56 kg/s, w=1.44 m3/s (860 °C steam, 2 bar),
P=2*105 Pa, Ș=4.26*10-5 Pa.s, and Re=90000. For the calculations, the data relative to the vapour phase have been
approximated with those of steam, since steam represents the major component. From the commercial values, it is
obvious that commercial reactors operate with gases flowing in turbulent regime. However according to our
calculations, similar flowrates can not be generated in a small scale furnace. As a result, the data in Table 3
demonstrate that the laboratory scale unit would operate in the laminar-flow regime.

Table 3. The Reynolds numbers at the inlet of the reactor tube

Ș, Pa.s 4.26E-05
D, m 0.024
L, m 0.85
IJ, s 0.1 0.3 0.5
v, m/s 8.5 2.8 1.7
Re 958 650 192

4. THE EXPERIMENTAL APPARATUS

The experimental apparatus is illustrated in Figure 1. Two membrane pumps ensure the dosage of the raw material
and the water. The output of the pumps is variable through the piston’s stroke length and the number of strokes per
minute. The rate (ml/min or ml/hour) of the dosage is readable on the available nomogram in the function of these
parameters. In addition to this, the glass-made storage vessels are calibrated to calculate the dosage rate directly
from the changes in level. The distilled water flows from the pump through the vaporizer into the preheater furnace.
The temperature of the preheater and the vaporizer can be set by a PID type controller equipped with a KDCL
microprocessor. In the preheater furnace, the pyrolysis feedstock and the steam are well mixed. The hydrocarbon
vapour and the steam are mixed in a spiral mixer consisting of a steel-rod with a wide external thread fitted inside
the preheater. The main pyrolysis reactor is connected directly to the preheater and equipped with a three-sector

Published by The Berkeley Electronic Press, 2004


4 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

furnace and three individual temperature controllers. The three heating elements are located in the external side of
the Cr-Ni steel reactor tube (0.85 m long, Ø47.7/24 mm) in a 7 mm deep slot. The controller is a KDC 3-4 PID
instrument with LCD display. The furnace can be rapidly heated to the desired temperature; the maximum running
temperature is 950 °C. The cooling system is connected to the outlet of the reactor. First, the high temperature (|800
°C) pyrogas passes through a water cooler and a separator, then through a second separator and a tube-in-tube
cooler. The second separator and the tube-in-tube cooler are held at +5 °C with the mixture of glycol and water as
cooling medium. After the cooling system, the gases are passed through an activated carbon adsorber and into a wet
gas meter, to measure the volume of the produced gases. Finally, the gases are directed into an outdoor flare. A T-
branch is provided between the adsorber and the gas meter for sampling the gases.

The manufacturer of the furnace, Ezila Ltd., recommended the appropriate material for the reactor tube to
achieve long runs at a max. temperature of 1000 °C, i.e. steel grade H9 MSZ 4359 (Hungarian standard) equivalent
to a steel grade AISI 309S/310.

The electric furnace’s three single coil filaments encircle the reactor tube, and are embedded into a ceramic
housing which is thermally insulated by a ceramic body and glass-wool. The distance between the heater filament
and the reactor is 6 mm.

1: water storage
vessel,
2: hydrocarbon
storage vessel,
3: water pump,
4: hydrocarbon
pump,
5: water-vaporizer,
6: hydrocarbon tube
preheater
7: preheater furnace,
8: three sector
furnace,
9: jacketed and coil
water cooler
10: water cooled
separator,
11: cryostat cooled
separator,
12: tube-in-tube gas
cooler
13: adsorber
14: gas meter,
15: T-branch to
sampling

Figure 1.The scheme of the laboratory pyrolysis apparatus

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 5

5. CALCULATION OF THE PYROLYSIS PARAMETERS


The dosage rate of the water and hydrocarbon feed were calculated from the cracking parameters, knowing the size
of the reactor. An equation was developed for this purpose, according to the following considerations found in the
book of Siklos (1988), where T = temperature in K, P = pressure =105 Pa, and IJ = residence time in s.

The volume of 1 mol of gas is:


R ˜T
Vmol (3)
P
Molar flow of the feed is:
wIn P ˜ wIn
n In (4)
Vmol R ˜T
§ s ˜ m SUM sWATER ˜ m SUM ·
1000 ˜ ¨¨ CH  ¸¸ n In (5)
© M CH M WATER ¹
The residence time can be calculated as follows:
VR 2 ˜ VR
W (6)
w wIn  wOut

As a result, the following Equation could be developed from combining Equations (4) to (6):

24 .06 ˜ M CH ˜ M WATER ˜ V R (7)


m SUM
ª w º
T ˜ W ˜ «1  Out » ˜ M WATER ˜ s CH  M CH ˜ sWATER
¬ w In ¼
where wOut/wIn is the ratio of the volumetric gas rates (hydrocarbon + steam) at the beginning and at the end of the
reactor. The value of this ratio is unknown before each experiment, and, therefore, it is usually calculated by the
empirical Equation (8) resulting from the work of Linden (1951) and his co-workers, Pettyjohn (1949) and Reid
(1959):

wOut C  0.00346 S  5.315 (8)


wIn H

From the mSUM value, the discrete dosage rate of the components (hydrocarbon and water) can be calculated on the
basis of mass-ratio and density.

6. EXPERIMENTAL METHODS AND SAMPLING

Special safety measures had to be implemented, particularly during start-up and shut-down of the reactor, because it
is during these operations that it is more probable for the hot hydrocarbon vapours to meet with air in the system.
The specific procedures adopted are:
Start-up:
- start cooling water and cryostat,
- start distilled water pump,
- switch on furnace, preheater, vaporizer heater,
- flush the furnace with steam,
- ignite pilot-flame of the flare,
- after reaching the desired temperature, the feeding of the hydrocarbon can be started
Sampling:
- reading instruments: thermometers, gas-meter, storage vessel-level

Published by The Berkeley Electronic Press, 2004


6 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

- after one hour run begins the sampling through the T-branch
- measuring pyrogas density with Regnault-flask
- five dm3 pyrogas-sample to determine composition
Shut down procedure:
- stop the hydrocarbon pump
- to deflate the compounds from the separators
- switch off the heating
- water dosage while the furnace temperature is above 500 °C.

6.1 Analyses of the feedstock and the pyrolysis gases


The analysis of the feedstock included the gas-chromatographic analysis and the determination of some physical
properties. The analysis provided information about the composition of the sample that was required to calculate the
reaction yields. From the standard distillation curve and density it was then possible to calculate average molecular
boiling point. Diagrams are available in the literature, like in the book of Orlicek and Pöll (1951), and of Berghoff
(1968), which provide the average molar weight and hydrogen content of a hydrocarbon mixture from the above
data.

6.1.1 Analysis of the pyrolysis gases

The gases formed in the pyrolysis experiments were analyzed also by gas-chromatography. This consisted of a two
stage analysis; in the first stage, the hydrogen and methane content of the pyrogas were determined with a thermal
conductivity detector, using a calibration gas made up of 20 v/v% H2; 20 v/v% CH4 in nitrogen and a one point
absolute calibration method. A sample chromatogram is shown in Figure 2.

Type of gas-chromatograph: CHROM 5.


Column: length 1.5 m, internal diameter 4 mm, stainless steel tube filled
with 5 A molecular sieve (60-80 mesh)
Carrier gas: nitrogen at 105 Pa
Injector: 250 µl gas-loop, injection directly to the column
Thermostat: 30 °C
Detector: TCD 180 °C
Bridge current: 60 mA
Sensitivity: 1/20
Integrator: HP 3396A

Analysis conditions for hydrocarbon components of pyrogas (sample shown in Figure 3):

Type of the gas-chromatograph: Carlo Erba Fractovap 2450 split/splitless injector, FID
Column: length 50 m, diameter 0.52 mm, 1.5 µm film thickness, HP
PLOT Al2O3 „S” FS capillary
Carrier gas: helium, 0.31*105 Pa
Temperature program: 35 °C (2 min), 5°C/min to 100°C, 10 °C/min to 180°C, 180 °C
(5 min)
Sample inlet: 250 µl gas-loop
Detector: 250 °C
Data acquisition: HP 35900 AD converter, HP Vectra PC, HP ChemStation

The relative sensitivity of different and discrete hydrocarbons to the FID detector is almost equal and only
methane deviates appreciably. In the second stage of the analysis, a high-performance HP PLOT Al2O3 capillary
column was used to separate the individual saturated and unsaturated compounds. The high capacity of the PLOT
capillary column allowed the analysis of these compounds without preliminary separation. Whilst the same
components are occurring in all measurements, absolute calibration is not necessary. The total hydrocarbon content
can be calculated from the hydrogen and methane content.

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 7

6.1.2 Analysis of the liquid product

The analysis of the liquid products is more complex than that of the gases, because the liquid product – the
pyrocondensate – consisted of an emulsion of water and hydrocarbons, including tars. The emulsion separation was
carried out by centrifugation. Before the gas-chromatographic analysis, it was beneficial to remove the heavier
components from the organic phase with distillation. The residue of the distillation was considered by its mass only.
The clear distillate was analysed by gas-chromatography with the following parameters:

Type of the gas-chromatograph Carlo Erba 8000


EL 980, MFC 800
split/splitless injector, FID
Column 60 m length, 0.32 mm inner diameter, 0.5 µm film thickness,
SE-30 polydimethyl-siloxane on a FS capillary
Carrier gas helium, 91 kPa
Temperature program 35 qC (30 min), 2 qC/min to 67 qC, then 15 qC/min to 250 qC,
250 °C (17 min)
Injector temperature 250 qC
Sample 1 µl split ratio: 1:100
Detector temperature 270 qC
Data acquisition HP 35900 AD converter, HP Vectra PC, HP ChemStation

The gas-chromatogram of a typical feedstock gasoline is illustrated in Figure 4, while the chromatogram of a
pyrocondensate is given in Figure 5.

Figure 2. Chromatogram of hydrogen and methane in nitrogen

Published by The Berkeley Electronic Press, 2004


8 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

Figure 3. Chromatogram of the pyrogas

Figure 4. Chromatogram of gasoline

Figure 5. Chromatogram of pyrocondensate

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 9

7. REACTOR TESTING

A standard gasoline, which is a usual feedstock in a Hungarian olefin plant, was chosen to characterize the
laboratory reactor. In the steam-cracking of gasoline the cracking severity can be changed over a wider range than in
the case of gas-oil, because coke deposition is smaller, less heavy hydrocarbons are formed and more pyrogas are
evolved. However the aim of this work was to prove that the laboratory apparatus was effective in determining the
cracking properties of all kinds of raw materials, including heavier gas-oils, polyolefinic residues, aromatic free
compounds or other hydrocarbon mixtures over a range of optimal cracking conditions.

The standard gasoline used in steam cracking is not typically desulphurized. This resulted in a useful
property, since the sulphur content of the gasoline poisoned the catalytically active sites located on the inner wall of
the steel tubes of the furnace and reduced the catalytically initiated carbonisation. This property is typically utilised
also in industrial practice. The properties of the feed gasoline are provided in Table 4.

Table 5 contains the mass balance calculated from the experimental results. In the top rows there are the
cracking conditions: temperature, residence time, water/hydrocarbon ratio. Table 6 shows the results of the pyrogas
analysis and yields together with the cracking conditions.

Table 4. Parameters of the feed gasoline

Parameter Value Parameter Value

d 15.6 °C 677.3 bpvol 73

Beginning bp 35 bpmol 67

10% 44 MCH 90

20% 48 K 12,2

30% 52 H wt% 15,2

40% 56 C/H ratio 5.6

50% 62 Sulphur wt% 0.026

60% 68 n-paraffin wt% 33.7

70% 77 naphthenic wt% 17.2

80% 92 aromatic wt% 4.0

90% 130

End bp 175

Published by The Berkeley Electronic Press, 2004


10 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

Table 5. Hydrocarbon mass balance

Temperature °C 860 860 860 860 860 830 830 880


Contact time s 0.3 0.3 0.3 0.2 0.4 0.3 0.3 0.3
sWATER/sCH 0.54 0.41 0.85 0.54 0.54 0.54 0.85 0.54
Feedstock g 1075 1320 1103 1588 777 1163 0767 0998
Pyrogas quantity m3 0.934 1.064 0.995 1.261 0.710 0.865 0.630 0.933
density kg/m3 0.974 1.035 0.943 1.055 0.902 1.146 1.029 0.901
weight g 910 1101 938 1330 640 991 648 841
Pyrocondensate weight g 191 203 151 229 139 190 135 170
C5+ -205 °C boiling weight g 161 172 106 196 103 167 117 130
Above 205 °C g 30 31 45 33 36 23 18 40
Pyrogas wt% 84.6 83.4 85.1 83.8 82.4 85.2 84.5 84.2
Pyrocondensate wt% 17.8 15.4 13.7 14.4 17.9 16.3 17.6 17.0
Pyrogasoline wt% 15.0 13.0 9.6 12.3 13.3 14.4 15.3 13.0
Pyrotar wt% 2.8 2.3 4.1 2.1 4.6 2.0 2.3 4.0
Total product wt% 102.4 98.8 98.8 98.2 100.3 101.6 102.1 101.3

Table 6. Ethylene and propylene yields

Temperature °C 860 860 860 860 860 830 830 880


Contact time s 0.3 0.3 0.3 0.2 0.4 0.3 0.3 0.3
sWATER/sCH 0.54 0.41 0.85 0.54 0.54 0.54 0.85 0.54
Feedstock g 1075 1320 1103 1588 777 1163 767 998
Pyrogas quantity m3 0.934 1.064 0.995 1.261 0.710 0.865 0.630 0.933
Ethylene v/v% 38.7 34.6 40.2 29.8 41.1 31 36 40.4
Propylene v/v% 15.8 13.8 8.6 15.2 8.9 16.2 14.2 6.5
Ethylene yield wt% 39.3 32.6 42.4 27.7 43.9 27.0 34.6 44.2
Propylene yield wt% 24.0 19.5 13.6 21.1 14.2 21.1 20.4 10.6
Ethylene+propylene wt% 63.3 52.1 56.0 48.8 58.1 48.1 55.0 54.8

In Table 7 and 8 the olefin yields are listed as a function of temperature and residence time. The effects of
both parameters correspond to the well-known characteristics of steam-cracking furnaces, showing the optimal
yields at intermediate pyrolysis temperatures and residence times.

Table 7. The olefin yields vs. temperature Table 8. The olefin yields vs. residence time
Constants Constants
Residence time, s 0.3 Temperature °C 860
sWATER/sCH 0.54 sWATER/sCH 0.54
Variable Variable
Temperature, °C 830 860 880 Residence time, s 0.2 0.3 0.4
wt% ethylene 27.0 39.3 44.2 wt% ethylene 27.7 39.3 43.9
propylene 21.1 24.0 10.6
propylene 21.1 24.0 14.2
together 48.1 63.3 54.8
together 48.8 63.3 58.1

8. MAIN OBSERVATIONS AND CONCLUSIONS

In general, it was found that the behaviour of the laboratory reactor is very similar to that of commercial furnaces,
although some yield differences were identified. These differences could be attributed to the flow regime: instead of

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 11

the strongly turbulent flow typical of commercial reactors, the laboratory reactor operated in laminar-flow regime.
As a consequence of the laminar flow profile in the lab scale device, the distribution of residence times of the
feedstock is not perfectly uniform. Higher residence times decrease the yields of valuable intermediate products due
to secondary reactions. The coke deposition during several days of operation was negligible according to visual
observation. This finding was attributed to the start-up and shut-down procedures, when steam was fed into the
reactor for 2 to 3 hours, resulting in coke gasification. Although coke gasification is not fast at typical pyrolysis
temperatures, it is certain that significant coke could be removed. The confirmation of this reaction was based on the
observation that, at higher pyrolysis temperatures (860 °C), small amounts of carbon-monoxide were detected with
the TCD gas-chromatograph. Another reason of the low observed coke deposition can be attributed to the sulphur
content of the gasoline used, which poisoned the catalytic sites of the reactor wall and decreased the formation of
coke precursors.

The design and construction of the apparatus was successful: the heat controller was proven to very
accurately maintain the reactor at the desired temperature ( ±1 °C); the selected reactor material was stable and no
deterioration was observed and the pumps worked at every desired feed rate.

The ethylene and propylene yields follow the trends found in the literature. The apparatus has been proven
successful to steam-crack complex mixtures and to make comparative measurements of olefin production capacity
based on different feedstocks. The reproducibility of the measurements was very good, with a maximum deviation
in olefin yields of less than 2%.

ACKNOWLEDGEMENTS
The authors acknowledge the financial and technical support of MOL Rt (Hungarian Oil Company) in building of
the apparatus, and that of TVK Rt. in supplying the hydrocarbon samples, helping with the analysis and with the
evaluation of results, the assistance of András Kristály in the design of the reactor, the financial support of the Co-
operation Research Center of the Institution of Chemical Engineering of Veszprém University and that of the
Ministry of Education.

NOTATION

bp boiling point, qC
d density, kg/m3
C/H carbon and hydrogen mass-ratio of the feedstock hydrocarbon
D intrinsic diameter of the reactor, m
K characterization factor
L reactor length, m
M molecular weight, g/mol
m mass flow, kg/s
n molecular flow rate, mol/s
P pressure, Pa
R ideal gas constant
Re Reynolds number
S severity function
s mass fraction
T absolute temperature, °K
t the outlet temperature of the reactor, ºC
v linear flow velocity, m/s
V volume, m3
w volumetric flow rate, m3/s
w average volumetric flow, m3/s

Greek Letters
Ș viscosity, Pa.s

Published by The Berkeley Electronic Press, 2004


12 International Journal of Chemical Reactor Engineering Vol. 2 [2004], Article A15

IJ residence time, s

Subscripts
CH hydrocarbon feedstock
SUM hydrocarbon feedstock+water
In at the reactor inlet
Out at the reactor outlet
mol molar
vol volumetric
R reactor

REFERENCES

Albright, L. F., “Pyrolysis: Theory and industrial practice”, Academic Press, New York, 277-323 (1983).

Barendregt, S., Valkenburg, P. J. M., Wagner, E. S., Dente, M., Ranzi, E., “History and recent developments in
SPYRO”, a review. Pre-Print Archive - American Institute of Chemical Engineers, [Spring National Meeting], New
Orleans, LA, United States, Mar. 11-14, 2497-2537 (2002).

Belohlav, Z., Zamostny, P., Herink, T., “The kinetic model of thermal cracking for olefins production”, Chem. Eng.
Proc., Vol. 42, No. 6, 461-473 (2003).
Berghoff, W., “Erdölverarbeitung und Petrolchemie”, Leipzig, 110-115 (1968).

Billaud, F., Elyahyaoui, K., Baronnet, F., Kressmann, S., “Chemical kinetic modeling of n-hexane pyrolysis by
ACUCHEM, CHEMKIN and MORSE software packages”, Chem. Eng. Sci., Vol. 46, No. 11, 2941-2946 (1991).

Dente, M., Ranzi, E., Goossens, A. G., “Detailed prediction of olefin yields from hydrocarbon pyrolysis through a
fundamental simulation model (SPYRO)”, Comput. Chem. Eng. , Vol. 3, No. 1-4, 61-75 (1979).

van Goethem, M. W. M., Kleinendorst, F. I., van Leeuwen, C., van Velzen, N., “Equation-based SPYRO (R) model
and solver for the simulation of the steam cracking process”, Comput. Chem. Eng., Vol. 25, No. 4-6, 905-911
(2001).

Linden, H. R., “Thermal cracking”, Petrol. Proc., Vol. 6, 1389-96 (1951).

Linden, H. R., Reid, J. M., “New data for making petrochemicals [petroleum chemicals]”, Petrol.Ref., Vol. 35, No.
6, 189-95 (1956).

Linden, H.R., Peck, R.E., “Severity function in pyrolysis”, Ind. Eng. Chem., Vol. 46, No. 12, 2470 (1955).

Linden, H. R., Reid, J. M., “Ethylene production process variables”, Chem. Eng. Prog., Vol. 55, No. 3, 71-8
(1959).

Orlicek, A.F., Pöll, H., “Hilfsbuch für Mineralöltechniker”, Wien, I. Vol. 34-63, 75-90 (1951).

Pant, K. K., Kunzru, D., “Pyrolysis of methylcyclohexane: kinetics and modeling”, Chem. Eng. J., Vol. 67, No. 2,
123-129 (1997).

Pant, K. K., Kunzru, D., “Catalytic pyrolysis of methylcyclohexane: kinetics and modeling”, Chem. Eng. J.
(Lausanne), Vol. 70, No. 1, 47-54 (1998).

Pettyjohn, E. S., Linden, H. R., “Estimating potential gasification value and yields of oils for gas plants. I.”, Am.
Gas J., Vol. 171, No. 4, 26-31, 57 (1949).

http://www.bepress.com/ijcre/vol2/A15
Pinter et al.: Steam Cracking Lab Reactor 13

Pettyjohn, E. S., Linden, H. R., “Estimating potential gasification value and yields of oils for gas plants. II.”, Am.
Gas J., Vol. 171, No. 5, 31, 34-35 (1949).

Safarik, I., Strausz, O.P., “The thermal decomposition of hydrocarbons. 1. n-alkanes (C>=5)”, Res. Chem.
Intermed., Vol. 22, No. 3, 275–314 (1996).

Safarik, I., Strausz, O.P., “The thermal decomposition of hydrocarbons. 2. Alkylaromatic hydrocarbons:
Alkylbenzenes”, Res. Chem. Intermed., Vol. 23, No. 1, 63–99 (1997).

Safarik, I., Strausz, O.P., “The thermal decomposition of hydrocarbons. 3. Polycyclic n-alkylaromatic compounds”,
Res. Chem. Intermed., Vol. 23, No. 2, 179–195 (1997).

Siklos, P., “Hydrocarbon industrial practices”, Tankönyvkiadó, Budapest, 19-23 (1988).

Szepesy, L., “Feedstock characterisation and prediction of product yields for industrial naphta crackers on the basis
of laboratory and bench scale pyrolysis”, J. Anal. Appl. Pyrol., Vol. 1, No. 3, 243-268 (1980).

Ullmann’s Encyclopaedia of Industrial Chemistry, VCH Veinheim, Vol. A10, 45-93 (1987).

Published by The Berkeley Electronic Press, 2004

Вам также может понравиться