You are on page 1of 5

Fuel Processing Technology 116 (2013) 217–221

Contents lists available at SciVerse ScienceDirect

Fuel Processing Technology


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

Catalytic conversion of C4 fraction for the production of light olefins


and aromatics
Xianghai Meng, Zhixi Wang, Rui Zhang, Chunming Xu, Zhichang Liu ⁎, Yadong Wang, Qiang Guo
State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing 102249, China

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

Article history: The catalytic conversion of the C4 fraction from a fluid catalytic cracking (FCC) unit over a commercial FCC equi-
Received 11 March 2013 librium catalyst was investigated using a confined fluidized bed reactor system. Butenes were easier to convert
Received in revised form 10 June 2013 than butanes, and 1-butene was the easiest to convert among the butene isomers. The ethene and propene yields
Accepted 11 June 2013
increased with increased reaction temperature and decreased with increased weight hourly space velocity
Available online xxxx
(WHSV). The formation of propene involved two successive steps: butene dimerization and large hydrocarbon
Keywords:
cracking. Aromatics were formed by the aromatization of the intermediate large olefins. A mechanism parameter
Catalytic conversion RCA was proposed to describe the relative function of the cracking reaction to the aromatization reaction. RCA
Butene increased with increased reaction temperature and decreased with increased WHSV. The cracking reaction
Butane predominates on the aromatization reaction at high reaction temperatures.
Cracking © 2013 Elsevier B.V. All rights reserved.
Propene
Aromatics

1. Introduction than butane [6,7]. Catalysts play an important role in converting C4


fractions to propene. The porous structure and acidity of zeolites are
Light olefins (ethene and propene) and light aromatics (benzene, key to their catalytic performance during the butene cracking process.
toluene, and xylene) are basic materials in the petrochemical industry. Small pores and high Si/Al ratio of zeolites benefit the production of
Light olefins are traditionally produced by the steam cracking/pyrolysis propene and ethene [8]. HZSM-48 zeolites, particularly those with low
of ethane, propane, butane, naphtha, light diesel, or heavier hydrocar- Si/Al ratios, exhibit better selectivity for propene than HZSM-5 zeolites
bons [1,2]. However, the steam pyrolysis process requires high reaction [9]. P-modified HZSM-5 zeolites exhibit high propene selectivity and
temperatures and consumes large amounts of energy. Light aromatics excellent anti-coking ability in the catalytic conversion of butene [10].
are traditionally produced by the catalytic reforming of naphtha. How- Yang et al. reported the catalytic cracking of 1-butene over HMCM-49
ever, naphtha is in short supply in some regions, such as China, Korea, zeolite. At a reaction temperature of 580 °C and a weight hourly space
and Japan. velocity (WHSV) of 9.4 h−1, 1-butene conversion reached 90.82%, and
Fluid catalytic cracking (FCC) is another important process for the total selectivity of propene plus ethene reached 51.38% [11].
propene production. The introduction of co-catalysts such as conven- For the catalytic cracking of n-butane, the yields of ethene and
tional ZSM-5, mesoporous ZSM-5, TNU-9, and SSZ-33 zeolite to a com- propene over alkaline-earth-modified HZSM-5 were higher than those
mercial equilibrium FCC catalyst can enhance the propene yield [3,4]. over non-modified HZSM-5 [12]. ZSM-23 zeolite is an efficient and
High yields of ethene and propene (>60 wt.%) have been achieved stable catalyst for the catalytic cracking of butanes, resulting in ethene
when C4+ olefins were cracked over hybrid catalysts at 610 °C to and propene yields exceeding 50 wt.%, and about 90% conversion at a
640 °C [5]. reaction temperature of 600 °C [13]. P-modified HZSM-5 zeolite also
The C4 fraction is an important fraction of FCC. C4 hydrocarbons shows positive catalytic performance in the cracking of butanes, with
can potentially be used for producing light olefins and light aromatics. propene and ethene yields reaching 25.6 wt.% and 33.9 wt.%, respectively,
The catalytic conversion of C4 hydrocarbons offers an alternative to at a reaction temperature of 650 °C [14].
steam cracking to produce light olefins and aromatics. The aromatization of butane and butene is also a research hotspot.
The conversion of C4 hydrocarbons to light olefins is attracting n-Butane is generally cracked over HZSM-5, whereas aromatic produc-
considerable research interest. Butene is relatively easier to convert tion is significantly enhanced when Mo2C is loaded onto HZSM-5 [15].
The introduction of Mo2C into ZSM-5 remarkably enhances the aromati-
zation of n-butane, and the selectivity of the aromatics reaches 34.5% at a
⁎ Corresponding author at: State Key Laboratory of Heavy Oil Processing, China University
of Petroleum, Changping District, Beijing 102249, China. Tel.: +86 10 8973 1252 (office);
conversion rate of 68.9% [16]. Mo2C/ZSM-5 is also an efficient catalyst
fax: +86 10 6972 4721. for the aromatization of 1-butene, and the yield of aromatics reaches
E-mail address: lzch@cup.edu.cn (Z. Liu). 40.6 wt.% at a conversion rate of 98.0% [17]. Methane, ethane, ethene,

0378-3820/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.fuproc.2013.06.002
218 X. Meng et al. / Fuel Processing Technology 116 (2013) 217–221

propene, and hydrogen are the primary products, and aromatics are Table 2
formed during a secondary process, namely, the oligomerization and Properties of the commercial FCC equilibrium catalyst.

aromatization of butenes [17]. Item Value


The C4 fractions are traditionally used as liquefied petroleum gas
Micro-activity index 60
(LPG). The market demand for LPG becomes weak as natural gas plays Pore volume (cm3·g−1) 0.22
more and more important role in the domestic fuel. Therefore, efficiently Surface area (m2·g−1) 88
converting the C4 fraction has become an urgent problem. Lu et al. Packing density (g·cm−3) 0.91
Coke content (wt.%) 0.04
proposed a method of converting n-butene into propene, i-butene,
Particle size distribution (wt.%)
i-butane, and gasoline on spent FCC catalyst in the stripper part. As 0–20 μm 0.8
shown in simulation tests, high temperature and low WHSV favor the 20–40 μm 16.2
formation of propene, i-butene, and i-butane, and the propene yield 40–80 μm 54.6
reaches 10.22 wt.% at a reaction temperature of 500 °C and a WHSV of 80–105 μm 15.8
>105 μm 12.6
1.44 h−1 [18]. Another possible method of efficient conversion of C4
fraction is to recirculate the C4 fraction to the FCC riser, where catalytic
conversion reactions occur. This process has the advantages of low
cost, easy operation, and high propene production. was used to measure the mass of the C4 fraction fed into the reactor.
This study investigated the catalytic conversion of the C4 fraction The C4 feed was then pumped and mixed with the steam. The mixture
from an FCC unit, as well as the influence of reaction temperature was heated to approximately 350 °C in a preheater and then fed into
and WHSV on conversion and product yield. The main reactions and the reactor. The reactions occurred as the feed made contact with the flu-
reaction pathways of C4 fraction catalytic conversion were discussed idized catalyst. After reaction, the oil gas was cooled and separated into
based on the above factors. gas and liquid samples. The spent catalyst was drawn out of the reactor
using a vacuum pump.
The experiments were conducted at various reaction tempera-
2. Experimental
tures between 400 and 700 °C, various WHSVs from 1 h−1 to 9 h−1,
and a steam flow rate of about 10 wt.% of the C4 fraction.
2.1. Feedstock and catalyst

The C4 fraction collected from the FCC unit of a petrochemical com-


2.3. Analytical methods
pany was used as the feed. The components of the C4 fraction are listed
in Table 1. The contents of butenes (1-butene, 2-butene, and i-butene)
A refinery gas analyzer (an Agilent 6890 gas chromatograph with
and butanes (n-butane and i-butane) were 53.62 wt.% and 45.10 wt.%,
a hydrogen flame ionization detector and a thermal conductivity
respectively. A commercial FCC equilibrium catalyst was used. The
detector) was used to measure the volume percentage of the compo-
physical properties and particle size distribution of the catalyst are
nents in the gas sample. The equation of state for an ideal gas was
listed in Table 2.
used to convert the data into mass percentages. A gasoline component
analyzer (another Agilent 6890 gas chromatograph) was used to deter-
2.2. Apparatus mine the weight percentages of the components in the liquid sample.
The coke content of the spent catalyst was measured using a self-
The experiment was conducted in a fluidized bed reactor with a fil- made coke analyzer, which included a combustion chamber, a thermal
ter on the bed top to prevent the catalyst from escaping. A diagram of conductivity detector, and a signal transfer system. Coke was burned in
the apparatus is shown in Fig. 1. The apparatus consists of five sections: an oxygen atmosphere inside a combustion chamber to form carbon
oil and steam input mechanisms, a reaction zone, a temperature control dioxide, which was detected using a thermal conductivity detector. Soft-
system, and a product separation and collection system. ware was used to calculate the coke content on the spent catalyst sample
The experiments were conducted in batch mode. For each experi- by comparing the peak area of carbon dioxide inside the combustion
ment, 60 g of catalyst was loaded into the reactor with an effective vol- chamber and the mass of the spent catalyst sample with those of a cata-
ume of approximately 580 mL. Distilled water was then pumped into lyst sample with known coke content.
the steam furnace to form the steam used to fluidize the catalyst in the
reactor. High-pressure nitrogen was injected into the C4 feed tank and
C4 buffer tank to keep the feedstock in a liquid state. A specific amount
of C4 feed was transferred from the C4 feed tank to the C4 buffer tank
by a pressure difference. An electronic balance under the C4 buffer tank 4 10
N2
8 19
15
Table 1 9
Components in the C4 feed (wt.%). 7

Component Content 6 11 = =
3 2
Propane 0.35 12
Propene 0.39 16
i-Butane 35.84 2 13 = =
n-Butane 9.26
2 5
2-Butene 22.28 14
1-Butene 12.86
1 17 18
i-Butene 18.48
C5+ 0.54
Fig. 1. Experimental setup (1 — water tank, 2 — filter, 3 — water pump, 4 — steam furnace, 5
Total 100
— C4 feed tank, 6 — electronic balance, 7 — C4 buffer tank, 8 — constant-pressure nitrogen, 9
Butanes 45.10
— C4 feed pump, 10 — counterbalance valve, 11 — preheater, 12 — thermocouple, 13
Butenes 53.62
— reactor, 14 — heater, 15 — entrance and exit of catalyst, 16 — condenser, 17 — liquid
C4 hydrocarbons 98.72
product sampler, 18 — wet gas flow meter, and 19 — gas sample bag).
X. Meng et al. / Fuel Processing Technology 116 (2013) 217–221 219

80 25

Yields of olefins and aromatics (wt%)


i-butane n-butane WHSV: 3 h-1
Conversion of C4 components (%)

2-butene 1-butene ethene


60 i-butene C4 20
propene
aromatics
40 15

20 10

0 5

-20 0
400 450 500 550 600 650 700 400 450 500 550 600 650 700
Reaction temperature (oC) Reaction temperature (oC)

Fig. 2. Conversion of C4 components as a function of reaction temperature at a WHSV Fig. 3. Yields of ethene, propene and aromatics as a function of reaction temperature.
of 3 h−1.
and aromatics as a function of reaction temperature. The yields of
3. Reaction performance of the catalytic conversion of the ethene and propene slightly increased below 600 °C, and then signifi-
C4 fraction cantly increased above 600 °C. The yield of the aromatics slightly in-
creased with increased reaction temperature from 400 °C to 550 °C,
3.1. Effects of reaction temperature and then varied slightly with increased reaction temperature from
550 °C to 700 °C.
The effects of reaction temperature on the conversion and product
yield were investigated by fixing the WHSV at 3 h−1. Fig. 2 shows the 3.2. Effects of WHSV
conversion of C4 components as a function of reaction temperature. The
conversion of the C4 components slightly increased with increased reac- The effects of WHSV on the conversion and product yields were in-
tion temperature from 400 °C to 600 °C, and it remarkably increased vestigated by fixing the reaction temperature at 700 °C. Table 4 lists the
with increased reaction temperature from 600 °C to 700 °C. Below product yields as a function of WHSV. With increased WHSV from
600 °C, 1-butene had the highest conversion among the butene isomers, 1 h− 1 to 9 h− 1, the yields of dry gas, C3, liquid, and coke decreased.
showing that 1-butene was easier to convert than i-butene and 2-butene. A low WHSV results in a long reaction time, which leads to a high
Butane conversion below 550 °C was negative, indicating that butane conversion of the C4 feed. Hence, low WHSV favors the conversion
was formed under the experimental conditions. Butane may be formed of the C4 fraction. Fig. 4 shows the yields of ethene, propene and aro-
in two ways, i.e., through the hydrogen transfer of butenes or the cracking matics as a function of WHSV. The yields of ethene and propene de-
of paraffins in liquid components. creased with increased WHSV, showing that low WHSV favors the
The product yields are listed in Table 3. By increasing the reaction production of ethene and propene. The aromatic yield was nearly con-
temperature from 400 °C to 700 °C, the yields of dry gas, C3 stant with increased WHSV at a reaction temperature of 700 °C.
(propene and propane), and coke increased, whereas those of the
liquid (hydrocarbons with carbon number higher than four) de- 4. Reaction pathways of the catalytic conversion of the C4 fraction
creased. The butenes underwent dimerization and formed liquid
components on the FCC equilibrium catalyst. The so-formed liquid Reactions such as dimerization, hydrogen transfer, isomerization,
components further underwent cracking, cyclization, hydrogen cyclization, aromatization, cracking, and condensation occur during
transfer, aromatization, and condensation reactions [6]. The cracking the catalytic conversion of the C4 fraction. Fig. 2 shows that the conver-
reaction is endothermic, whereas cyclization, hydrogen transfer, and sion of butanes was almost zero or negative at low reaction tempera-
aromatization are exothermic. Thus, high reaction temperatures tures (b 600 °C), indicating that butanes were produced under the
enhance the cracking reactions of the formed liquid components. experimental conditions. The cracking reaction of butanes occurred at
Accordingly, the yield of liquid decreased with increased reaction high reaction temperatures (>600 °C).
temperature, whereas the dry gas and C3 yields increased. High reaction Butenes are the predominant reactants during catalytic conversion.
temperature also accelerates the rate of condensation reactions by Fig. 5 illustrates the reaction network of the C4 fraction catalytic conver-
which coke is generated. Thus, the coke yield increased with increased sion. Butenes can be converted into butanes through hydrogen transfer,
reaction temperature. into octenes through dimerization, or into other large olefins through
Light olefins and aromatics are the desired products in the catalytic polymerization with propene or ethene. Octenes can undergo further
conversion of the C4 fraction. Fig. 3 shows the yields of ethene, propene isomerization to produce multi-branched octenes, hydrogen transfer

Table 3 Table 4
Product yield as a function of reaction temperature at a WHSV of 3 h−1 (wt.%). Product yield as a function of WHSV at a reaction temperature of 700 °C (wt.%).

Reaction temperature (°C) 400 450 500 550 600 650 700 WHSV (h−1) 1 3 5 7 9

Dry gas 0.03 0.21 0.45 0.63 1.29 4.62 12.58 Dry gas 14.05 12.58 11.62 11.34 10.97
C3 2.80 3.81 4.64 6.12 8.22 15.86 26.64 C3 28.39 26.64 26.07 25.38 24.84
C4 84.01 83.76 82.81 81.91 79.70 68.47 48.66 C4 45.10 48.66 50.16 51.24 52.06
Liquid 12.47 11.37 10.98 10.06 9.43 9.16 8.84 Liquid 8.85 8.84 9.06 9.24 9.45
Coke 0.69 0.85 1.12 1.28 1.36 1.89 3.28 Coke 3.61 3.28 3.09 2.80 2.68
220 X. Meng et al. / Fuel Processing Technology 116 (2013) 217–221

25 0.5
Yields of olefins and aromatics (wt%)

Yield ratio of dry gas to C3


20 0.4
WHSV: 3 h-1
700 oC
15 ethene 0.3
propene
aromatics
10 0.2

5 0.1

0 0.0
0 2 4 6 8 10 400 450 500 550 600 650 700
Weight hourly space velocity (h-1) Reaction temperature (oC)

Fig. 4. Yields of ethene, propene and aromatics as a function of WHSV. Fig. 6. Yield ratio of dry gas to C3 as a function of reaction temperature.

to produce octanes, aromatization to produce aromatics, cyclization to Cracking and aromatization reactions simultaneously occur during
produce naphthenes, cracking to produce propene or butenes, or poly- the catalytic conversion of the C4 fraction. Light products (dry gas and
merization with olefins to produce larger olefins. Octanes can undergo C3) are formed by cracking reactions, and aromatics are formed by
cracking to produce light olefins and light alkanes. Aromatics can aromatization reactions. A yield ratio of dry gas plus C3 to aromatics
undergo condensation to produce polycyclic aromatics. Intermediate [RCA, shown in Eq. (1)] was proposed to describe the relative function
large olefins can undergo aromatization to produce aromatics, and po- of the cracking reaction to the aromatization reaction.
lymerization to produce polymers. Polymers and polycyclic aromatics
are finally converted into coke. ydry gas þ yC3
The above analysis shows that the formation of propene involves RCA ¼ : ð1Þ
yaromatics
two successive steps: the dimerization of butenes and the cracking of
large hydrocarbons [6]. The formed olefins can polymerize with butene
to produce other larger olefins that can undergo all of the aforemen- A large RCA indicates that the cracking reaction is important in the
tioned reactions, resulting in a wide variety of end products. catalytic conversion of the C4 fraction, whereas a small RCA indicates
Both catalytic and thermal functions play roles in the catalytic that the aromatization reaction is significant. Fig. 7 shows RCA as a
conversion of the C4 fraction. Dry gas is generally formed through function of reaction temperature at a WHSV of 3 h−1. With increased
thermal cracking. C3 is generally formed through catalytic cracking reaction temperature from 400 °C to 600 °C, RCA gradually increased
at low temperatures, while it is formed through both catalytic crack- from 0.73 to 1.33, indicating the importance of both cracking and aro-
ing and thermal cracking at high temperatures [19,20]. The yield ratio matization reactions. RCA reached 2.77 at 650 °C and 5.23 at 700 °C,
of dry gas to C3 was proposed to describe the relative function of indicating that cracking predominates on aromatization at high reac-
thermal cracking to catalytic cracking. Fig. 6 shows the yield ratio of tion temperatures. Fig. 7 also shows RCA as a function of WHSV at a re-
dry gas to C3 as a function of reaction temperature. The yield ratio action temperature of 700 °C. RCA decreased with increased WHSV,
of dry gas to C3 is below 0.11 when the reaction temperature is indicating that low WHSV favors cracking. RCA at a WHSV of 1 h−1
below 550 °C. This result shows that catalytic cracking plays an im- was about 0.16 times higher than that at a WHSV of 9 h−1, indicating
portant role at low reaction temperatures (b550 °C). The yield ratio that low WHSV strengthens the cracking function.
of dry gas to C3 increases from 0.16 at 600 °C to 0.47 at 700 °C,
indicating that both catalytic and thermal cracking play important
roles at high reaction temperatures (> 600 °C). Weight hourly space velocity (h-1)
1 3 5 7 9
6
3
C2=, C3= C8o C2=, C3=, C4=, C5=, C6=...
5
3 3 3 3 700 oC
1
4
1 2 2
C4o C4= C8= C12=
RCA

2 2 3

5 oligomers 2
WHSV: 3 h-1
4 4 6 4
1
6 coke
1
naphthenes aromatics 0
400 450 500 550 600 650 700
Fig. 5. Reaction network of C4 fraction catalytic conversion (1 — hydrogen transfer re- Reaction temperature (oC)
action, 2 — polymerization reaction, 3 — cracking reaction, 4 — aromatization reaction,
5 — cyclization reaction, 6 — condensation reaction). Fig. 7. RCA as a function of reaction temperature and WHSV.
X. Meng et al. / Fuel Processing Technology 116 (2013) 217–221 221

5. Conclusions [5] R. Le Van Mao, A. Muntasar, H.T. Yan, Q. Zhao, Catalytic cracking of heavy olefins
into propylene, ethylene and other light olefins, Catalysis Letters 130 (2009)
86–92.
Butenes were easier to convert than butanes on commercial FCC [6] L. Li, J.S. Gao, C.M. Xu, X.H. Meng, Reaction behaviors and mechanisms of catalytic
equilibrium catalysts. High reaction temperature and low WHSV favored pyrolysis of C4 hydrocarbons, Chemical Engineering Journal 116 (2006) 155–161.
[7] X.H. Meng, C.M. Xu, J.S. Gao, Secondary cracking of C4 hydrocarbons from heavy
the production of ethene, propene, and aromatics. Butenes initially oil catalytic pyrolysis, Canadian Journal of Chemical Engineering 84 (2006)
underwent dimerization into large olefins, which then underwent crack- 322–327.
ing to produce propene and aromatization to produce aromatics. A yield [8] X.X. Zhu, S.L. Liu, Y.Q. Song, L.Y. Xu, Catalytic cracking of C4 alkenes to propene and
ethene: influences of zeolites pore structures and Si/Al2 ratios, Applied Catalysis A:
ratio of dry gas to C3 was proposed to describe the relative function of General 288 (2005) 134–142.
thermal cracking to catalytic cracking. This ratio increased with in- [9] G.L. Zhao, J.W. Teng, Y.H. Zhang, Z.K. Xie, Y.H. Yue, Q.L. Chen, Y. Tang, Synthesis of
creased reaction temperature, and the importance of thermal cracking ZSM-48 zeolites and their catalytic performance in C4-olefin cracking reactions,
Applied Catalysis A: General 299 (2006) 167–174.
increased at high temperatures. A yield ratio of dry gas plus C3 to aro-
[10] G.L. Zhao, J.W. Teng, Z.K. Xie, W.Q. Jin, W.M. Yang, Q.L. Chen, Y. Tang, Effect of phos-
matics (RCA) was proposed to describe the relative function of cracking phorus on HZSM-5 catalyst for C4-olefin cracking reactions to produce propylene,
reaction to aromatization reaction. RCA increased with increased reaction Journal of Catalysis 248 (2007) 29–37.
temperature and decreased with increased WHSV. High reaction tem- [11] X.C. Yang, Y.C. Shang, P.P. Yang, Catalytic cracking of 1-butene to propene and
ethene on HMCM-49 zeolite, Reaction Kinetics, Mechanisms and Catalysis 100
perature and low WHSV favor cracking reaction. (2010) 399–405.
[12] K. Wakui, K. Satoh, G. Sawada, K. Shiozawa, K. Matano, K. Suzuki, T. Hayakawa, Y.
Yoshimura, K. Murata, F. Mizukami, Cracking of n-butane over alkaline earth-
Acknowledgment containing HZSM-5 catalysts, Catalysis Letters 84 (2002) 259–264.
[13] D. Ji, B. Wang, G. Qian, Q. Gao, G.M. Lu, L. Yan, J.S. Suo, A highly efficient catalytic
C4 alkane cracking over zeolite ZSM-23, Catalysis Communications 6 (2005)
Financial support was provided by the National Basic Research 297–300.
Program of China (973 Program, No. 2012CB215001), the Program for [14] G.Y. Jiang, L. Zhang, Z. Zhao, X.Y. Zhou, A.J. Duan, C.M. Xu, J.S. Gao, Highly effective
New Century Excellent Talents in University of China (NCET-12-0970), P-modified HZSM-5 catalyst for the cracking of C4 alkanes to produce light olefins,
Applied Catalysis A: General 340 (2008) 176–182.
and the Science Foundation of China University of Petroleum, Beijing
[15] S. Yuan, S. Hamid, Y. Li, P. Ying, Q. Xin, E.G. Derouane, C. Li, Preparation of Mo2-
(Nos. KYJJ2012-03-23 and KYJJ2012-03-25). C/HZSM-5 and its catalytic performance for the conversion of n-butane into
aromatics, Journal of Molecular Catalysis A: Chemical 184 (2002) 257–266.
[16] F. Solymosi, R. Nemeth, A. Szechenyi, Aromatization of n-butane over supported
References Mo2C catalysts, Catalysis Letters 82 (2002) 213–216.
[17] F. Solymosi, A. Szechenyi, Aromatization of n-butane and 1-butene over supported
[1] B.Z. Qian, World ethylene industry and its advance, Petrochemical Technology & Mo2C catalyst, Journal of Catalysis 223 (2004) 221–231.
Application 21 (2003) 37–45. [18] Y. Lu, M.Y. He, X.T. Shu, B.N. Zong, Exploratory study on upgrading 1-butene using
[2] X.H. Meng, J.S. Gao, L. Li, C.M. Xu, Advances in catalytic pyrolysis of hydrocarbons, spent FCC catalyst/additive under simulated conditions of FCCU's stripper,
Petroleum Science and Technology 22 (2004) 1327–1341. Applied Catalysis A: General 255 (2003) 345–347.
[3] M.A.B. Siddiqui, A.M. Aitani, M.R. Saeed, N. Al-Yassir, S. Al-Khattaf, Enhancing propyl- [19] X.H. Meng, C.M. Xu, J.S. Gao, L. Li, Studies on catalytic pyrolysis of heavy oils: reaction
ene production from catalytic cracking of Arabian light VGO over novel zeolites as behaviors and mechanistic pathways, Applied Catalysis A: General 294 (2005)
FCC catalyst additives, Fuel 90 (2011) 459–466. 168–176.
[4] M.A.B. Siddiqui, A.M. Aitani, M.R. Saeed, S. Al-Khattaf, Enhancing the production [20] X.H. Meng, C.M. Xu, J.S. Gao, L. Li, Z.C. Liu, Catalytic and thermal pyrolysis of
of light olefins by catalytic cracking of FCC naphtha over mesoporous ZSM-5 atmospheric residue, Energy & Fuels 23 (2009) 65–69.
catalyst, Topics in Catalysis 53 (2010) 1387–1393.

You might also like