Anda di halaman 1dari 7

Top Catal (2010) 53:1387–1393

DOI 10.1007/s11244-010-9598-1

ORIGINAL PAPER

Enhancing the Production of Light Olefins by Catalytic Cracking


of FCC Naphtha over Mesoporous ZSM-5 Catalyst
M. A. Bari Siddiqui • A. M. Aitani •

M. R. Saeed • S. Al-Khattaf

Published online: 5 October 2010


Ó Springer Science+Business Media, LLC 2010

Abstract The enhanced production of light olefins from 1 Introduction


the catalytic cracking of FCC naphtha was investigated
over a mesoporous ZSM-5 (Meso-Z) catalyst. The effects Light olefins (mainly ethylene and propylene) are impor-
of acidity and pore structure on conversion, yields and tant feedstocks for the production of useful materials, such
selectivity to light olefins were studied in microactivity test as polyethylene and polypropylene, vinyl chloride, ethyl-
(MAT) unit at 600 °C and different catalyst-to-naphtha ene oxide, ethylbenzene and others. Polyolefins remain the
(C/N) ratios. The catalytic performance of Meso-Z catalyst largest sector of light olefins demand showing the highest
was compared with three conventional ZSM-5 catalysts overall growth rate compared with other derivatives.
having different SiO2/Al2O3 (Si/Al) ratios of 22 (Z-22), 27 Annual projected growth rate for ethylene and propylene
(Z-27) and 150 (Z-150). The yields of propylene (16 wt%) demand is estimated at 4.5 and 5.4%, respectively [1]. The
and ethylene (10 wt%) were significantly higher for Meso- major sources of propylene are steam crackers and fluid
Z compared with the conventional ZSM-5 catalysts. catalytic cracking (FCC) units. While ethylene is the main
Almost 90% of the olefins in the FCC naphtha feed were product of steam crackers and gasoline is the also the main
converted to lighter olefins, mostly propylene. The aro- product in FCC units, propylene remains a byproduct in
matics fraction in cracked naphtha almost doubled in all both units. In conventional FCC, typically about 2 wt%
catalysts indicating some level of aromatization activity. ethylene and 3–6 wt% propylene yields are obtained [1].
The enhanced production of light olefins for Meso-Z is About, 30% of world’s propylene is supplied by refinery
attributed to its small crystals that suppressed secondary FCC operations, 64% is co-produced from steam cracking
and hydrogen transfer reactions and to its mesopores that of naphtha or other feedstocks, and the remaining is pro-
offered easier transport and access to active sites. duced on-purpose using metathesis or propane dehydro-
genation processes [1]. Compared with thermal steam
Keywords Cracking  Naphtha  Olefins  ZSM-5  cracking, the catalytic cracking of naphtha is carried out at
Mesoporous  Ethylene  Propylene lower temperatures and consumes 10–20% less energy.
The reduction of carbon dioxide emissions is also another
driving force for the low temperature naphtha catalytic
cracking [2].
Many types of zeolites have been investigated for the
catalytic cracking of naphtha to produce ethylene and
propylene including ZSM-5 zeolites, zeolite A, zeolite X,
zeolite Y, zeolite ZK-5, zeolite ZK-4, synthetic mordenite,
M. A. Bari Siddiqui  A. M. Aitani  M. R. Saeed  dealuminated mordenite, as well as naturally occurring
S. Al-Khattaf (&) zeolites including chabazite, faujasite, mordenite [2–5].
Center of Research Excellence in Petroleum Refining
High selectivity to light olefins depends on the extent of
& Petrochemicals, King Fahd University of Petroleum
& Minerals (KFUPM), Dhahran 31261, Saudi Arabia reaction, reaction path and residence time, which are
e-mail: skhattaf@kfupm.edu.sa determined by the zeolite acidity, pore structure and crystal

123
1388 Top Catal (2010) 53:1387–1393

size [2]. Selectivity to light olefins is also enhanced by Table 1 Composition of FCC naphtha feed (wt%)
lower conversion (weak acidity), shorter residence time Carbon n-Paraffins i-Paraffins Olefins Naphthenes Aromatics
(small crystal size) and increased accessibility (mesopores) no.
which allow full utilization of zeolitic volume by the feed
C4 0.7 1.2 4.0 – –
molecules.
ZSM-5 is considered the catalyst of choice when shape C5 0.9 5.7 9.8 1.0 –
selectivity influences the preferential formation of light C6 0.8 5.5 8.4 1.5 0.5
olefins. Several strategies have been proposed to enhance C7 0.8 5.0 5.3 2.4 2.4
the accessibility to active sites in ZSM-5, such as the use of C8 0.7 3.7 3.2 2.3 6.2
nanocrystals [4], composite materials [5–7] low level of C9 0.4 2.8 0.7 1.8 7.7
acid concentration provided by high Si/Al molar ratio, and C10–C12 1.2 4.1 0.6 0.5 8.2
incorporation of intra-crystalline mesoporosity [8]. The Total 5.5 28.0 32.0 9.5 25.0
latter can be accomplished either by carbon templating
[9–11] or by post synthesis modification by alkali treatment
[12–15]. The generation of mesopores by secondary tem- Cataloid AP-3 (which contains 5.4 wt% alumina,
plating with carbon particles during synthesis increases the 3.4 wt% acetic acid, and water as balance) was used as
accessibility of acid sites in ZSM-5 and diminishes the alumina binder for the four catalysts. The required quantity
role of transport [16]. In such a case, zeolites are able to of alumina was mixed with demineralised water which has
nucleate around mesoporous particles of carbon matrix and been acidified to a pH of 5 with dilute nitric acid. ZSM-5
form typical ZSM-5 crystals. After the removal of carbon sample was added to the alumina-water slurry, while stir-
particles by combustion, ZSM-5 crystals with a controlled ring. The slurry was dried at 120 °C for 2 h followed by
pore size distribution and a highly crystalline micro-mes- calcination at 600 °C for 3 h. The calcined catalyst was
oporous hierarchical structure are formed. Mesoporous pelletized, crushed and sieved to obtain 80–90 microns size
ZSM-5 has shown improved catalytic performance in the catalyst particles.
cracking of model compounds due to the hierarchical FCC naphtha feed was obtained from Saudi Aramco’s
porosity of the pore walls and to the easier transport and Jeddah refinery and was used in all MAT experiments. The
access to the active sites. Hierarchical ZSM-5 combines the composition of the naphtha feedstock comprising paraf-
high activity, shape selectivity, and hydrothermal stability fins (iso and normal), olefins, naphthenes and aromatics
of conventional ZSM-5 [16]. (PIONA), is presented in Table 1.
In this study, the enhancement of light olefins yield from
FCC naphtha cracking over mesoporous ZSM-5 catalyst 2.2 Catalyst Characterization
was investigated. The catalytic performance of Meso-Z
was compared with three conventional ZSM-5 catalysts The surface area of the catalysts were measured in a
having low and high Si/Al ratios. The catalytic activity and Quantochrome NOVA 1200 gas sorption analyzer by the
selectivity were discussed to elucidate the effects of acidity adsorption of nitrogen at 77 K according to standard
and pore structure on the selectivity to light olefins and ASTM D3663 method. Prior to nitrogen sorption, catalyst
aromatics. samples were evacuated for 2 h while being heated up to
350 °C at a steady rate of increase of temperature. Micro
pore volume was determined by t-plot method. Total pore
2 Experimental volume was determined by BJH method. Difference of
total pore volume and micro pore volume gives the mes-
2.1 Catalysts and Naphtha Feed opore volume, Vme.
Bronsted and Lewis acid sites density of zeolites was
The H-form of Z-27 and Z-150 (ZSM-5 zeolites) were determined by adsorption of pyridine as a probe molecule
procured from Catal International, while the H-form of followed by FTIR spectroscopy. Zeolites were activated in
Z-22 zeolite was obtained from Industrial Chemicals, USA. a form of self-supporting wafers at 450 °C overnight. The
The Meso-ZSM-5 zeolite having a Si/Al of 30 was adsorption of pyridine was carried out 170 °C, according to
obtained from the J. Heyrovský Institute of Physical Gil et al. [19].
Chemistry, Czech Republic. It was synthesized using car-
bon black (CBP 2000, Cabot Corporation) as a secondary 2.3 Catalytic Evaluation
template with an average particle diameter of 12 nm as per
the procedure outlined by Pavlačkova et al. [17] and The cracking of FCC naphtha was carried out in a fixed-
Al-Khattaf et al. [18]. bed microactivity test (MAT) unit, manufactured by

123
Top Catal (2010) 53:1387–1393 1389

Sakuragi Rikagaku, Japan. The procedure is carried out Table 2 Textural and acidic properties of catalystsa
according to ASTM D-3907 which is a standard method Item Meso-Z Z-22 Z-27 Z-150
used to determine the activity and selectivities of FCC
catalysts. MAT tests were performed at 600 °C and 30 s SiO2/Al2O3 ratio 30 22 27 150
time-on-stream (TOS). Conversion was varied by chang- Total pore volume, cc/g 0.496 0.344 0.298 0.462
ing the catalyst/naphtha ratio (C/N) in the range of Micro pore (Vmi), cc/g 0.088 0.079 0.075 0.081
1.0–5.0. This variable was changed by keeping constant Meso pore (Vme), cc/g 0.408 0.265 0.233 0.381
the amount of feed (1.0 g) and changing the amount of Surface area, m2/g 370 293 284 320
catalyst. Prior to MAT test, the system was purged with Bronsted acidity, mmol/g 0.08 0.13 0.22 0.10
N2 flow at 30 cc/min of for 30 min. About 1.0 g of Lewis acidity, mmol/g 0.32 0.31 0.28 0.16
naphtha was then fed to the reactor during 30 s. After the Total acidity, mmol/g 0.40 0.44 0.50 0.26
reaction mode, the stripping of the catalyst was carried a
Catalysts were mixed with 33.3 wt% alumina binder
out for 15 min using 30 cc/min of N2. During the reaction
and stripping modes, liquid products were collected in a
glass receiver kept in an ice-bath. Gaseous products (for 380

GC analysis) were collected in a gas burette by water 330


displacement.

Volume adsorbed, cc/g


A thorough gas chromatographic analysis of all MAT 280
products was conducted to provide detailed yield patterns
230
and information on the selectivity of the catalyst being
tested. The gases were analyzed using two Varian gas 180
chromatographs (GC) equipped with FID and TCD detec-
tors. This allowed the quantitative determination of all light 130
hydrocarbons up to C4, hydrogen and fixed gases. Hydro-
80
carbon distribution of feed and cracked naphtha was
determined by a GC PIONA analyzer manufactured by 30
0 0.2 0.4 0.6 0.8 1
Shimadzu, Japan. The GC was equipped with FID detector,
Relative pressure, P/Po
a 100 m capillary column (CP-Sil5) and software that
identifies and quantifies different hydrocarbon compounds Fig. 1 Nitrogen adsorption–desorption isotherms for Meso-Z (filled
in naphtha. Coke on spent catalyst was determined by triangle) and Z-27 (empty circle)
Horiba Carbon–Sulfur Analyzer Model EMIA-220V.
Naphtha conversion is defined as the sum of gaseous Intercrystalline void volume is reflected by a steep rise in
product yields and coke [20]. Mass balance was considered the nitrogen adsorption isotherm at P/Po of 0.9–1.0 [21].
acceptable when comprised in the limits 95–105% of the Surface area, calculated from nitrogen adsorption isotherm,
injected naphtha. was highest for Meso-Z (370 m2/g) compared with the
lowest Z-27 at 284 m2/g. As shown in Table 2, the Bron-
sted acidity of Meso-Z was 0.08 mmol/g compared with
3 Results and Discussion 0.10 mmol/g for Z-150, 0.31 mmol/g for Z-22 and
0.28 mmol/g for Z-27. Total acidity of the catalysts
3.1 Catalyst Characterization decreased in the following order: Z-150 \ Meso-Z \
Z-22 \ Z-27.
The textural properties of the catalysts, presented in
Table 2, were determined from nitrogen adsorption iso- 3.2 Catalytic Evaluation
therms. Meso-Z has significantly higher mesopore volume
(Vme) of 0.41 cc/g compared with Z-27 (0.23 cc/g), Z-22 The catalytic cracking of FCC naphtha was carried out at
(0.26 cc/g) and Z-150 (0.38 cc/g) which includes inter- 600 °C over the four catalysts in MAT unit. Product yields
crystal void volume. The high Vme of Meso-Z catalyst is were mapped over a reasonable range of conversions by
due to actual mesopores and does not represent intracrys- varying the C/N in the range of 1.5–4.0. Conversion and
talline void volume [19]. This is evident from the steep rise product yields were plotted as a function of C/N ratio.
in adsorption isotherm and hysteresis effect in desorption Typical plots for conversion, propylene, ethylene and
curve for Meso-Z in the P/Po range of 0.6–0.9, as illus- C2–C3 olefins yields are shown in Fig. 2.
trated in Fig. 1. The isotherms of conventional Z-27 The results show that naphtha conversion increased
catalyst showed a gradual rise in P/Po of 0.6–0.9. with increased C/N (Fig. 2), as is expected in MAT

123
1390 Top Catal (2010) 53:1387–1393

Fig. 2 Effect of catalyst/ 60 12


naphtha ratio on naphtha Meso-Z Z-22 Z-27 Z-150 Meso-Z Z-22 Z-27 Z-150
conversion, ethylene and

Ethylene Yield, wt.%


propylene yields for Meso-Z 55 10

Conversion, %
(filled square), Z-22 (empty
triangle), and Z-27 (filled
triangle) and Z-150 (filled 50 8
circle) at 600 °C

45 6

40 4
1.0 2.0 3.0 4.0 5.0
1.0 2.0 3.0 4.0 5.0
C/N Ratio C/N Ratio

20 30
Meso-Z Z-22 Z-27 Z-150 Meso-Z Z-22 Z-27 Z-150

C2-C3 Olefins Yield, wt.%


Propylene Yield, wt.%

16 25

12 20

8 15

4 10
1.0 2.0 3.0 4.0 5.0 1.0 2.0 3.0 4.0 5.0
C/N Ratio C/N Ratio

experiments. The highest conversion was obtained for was produced by direct catalytic cracking and by secondary
Z-150, it varied from 49 to 54% as C/N was increased from cracking of propylene [22].
1.5 to 4. Within the same C/N range, Z-27 and Z-22 gave
naphtha conversion of 46–50%. Meso-Z exhibited naphtha 3.3 Catalytic Performance at Constant Catalyst/
conversion of 44–50%. The difference in conversion Naphtha Ratio
between the four catalysts was in the range of 4–5%, not
much significant change. The difference between the catalytic performances of the
The trends of propylene and ethylene yield with four catalysts was elucidated by comparing conversion and
increased C/N ratio are shown in Fig. 2. Propylene yield product yields at constant C/N of 3.0 (the data were cal-
decreased with increased C/N ratio for the four catalysts. culated by interpolation). Naphtha conversion, product
Significant enhancement in propylene yield was achieved yields (gaseous and liquids) and other values determined at
by the use of meso-Z catalyst. Propylene yield was also a constant C/N ratio are presented in Table 3. Meso-Z
enhanced by Z-150 compared with Z-27 and Z-22. Pro- catalyst exhibited a naphtha conversion of 47% compared
pylene yield of 13–16 wt% and 9–12 wt% was observed with 48.4% for Z-22, 49.4% for Z-26 and 50.4% for Z-150.
for Meso-Z and Z-150, respectively compared with A major portion of the FCC naphtha used in this study
5–9 wt% for Z-27 and Z-22. The effect of mesopores on consists of olefins, which are known to be highly reactive
propylene enhancement was clearly evident. The mesop- hydrocarbons. Because of presence of highly reactive
ores allowed fast elution of propylene from the zeolitic olefins in the naphtha feed, no strong correlation between
pores before secondary, olefin consuming oligomerization acidity and conversion was observed.
reactions take place. Besides ethylene and propylene, significant amounts of
Ethylene yield was also higher for Meso-Z and Z-150 methane, butenes, and LPG were detected over all cata-
(8–10 wt%) compared with 6–7 wt% obtained for Z-27 lysts. The highest propylene yield of 14.6 wt% was
and Z-22. The extent of increase (30%) in ethylene yield obtained for Meso-Z, compared with 7.8–8.2 wt% obtained
for Meso-Z was lower compared with the increase (80%) in for Z-22 and Z-27 containing conventional ZSM-5 having
propylene yield. Ethylene yield increased slightly as C/N low S/Al ratio similar to Meso-Z. Propylene selectivity was
ratio increased up to about 3.0 and then remained constant also enhanced by similar margins for Meso-Z. Higher Si/Al
as the C/N was further increased. At lower C/N, ethylene catalyst, Z-150, gave 9.4% propylene yield which was

123
Top Catal (2010) 53:1387–1393 1391

Table 3 Naphtha conversion and product yields at constant catalyst/ Meso-Z Z-22 Z-27 Z-150
naphtha ratio of 3.0 and 600 °C
Catalyst Meso-Z Z-22 Z-27 Z-150

Naphtha conversiona, % 47.4 48.4 49.5 49.6


Gaseous products, wt%
Ethylene 9.4 6.6 7.8 8.2
Propylene 14.6 7.1 8.2 9.4
Butenes 7.5 2.9 3.2 3.9
Methane 1.4 5.5 4.9 3.1
Ethane 1.6 5.1 4.2 3.6
Propane 6.7 15.7 14.9 14.6
i-Butane 3.5 2.5 3.0 3.6
n-Butane 2.2 2.2 2.3 2.6
Fig. 3 Comparison of ethylene and propylene yields for Meso-Z,
Total gas yield 46.9 47.6 48.5 49.0
Z-22, Z-27 and Z-150 at constant catalyst/naphtha ratio of 3.0 and
Coke 0.47 0.84 1.0 0.57 600 °C
HTIb 0.8 2.9 2.5 2.2
Ethylene ? Propylene 24.0 13.7 16.0 17.6 compared with 1.2 for Z-150 and 1.1 each for Z-22 and
Propylene/Ethylene 1.6 1.1 1.1 1.2 Z-27. Coke yield was lowest for Meso-Z (0.47 wt%) and
Selectivity, % highest for Z-27 (1.1 wt%). Figure 3 illustrates these
Ethylene 20.0 13.6 15.8 16.2 differences in product yields among the four catalysts.
Propylene 31.0 14.8 16.5 18.7
Liquid products, wt% 3.4 Hydrogen Transfer Activity
n-Paraffins 0.74 0.1 0.3 0.3
i-Paraffins 6.9 2.1 2.5 2.1 One of the factors for light olefins enhancement by Meso-Z
Olefins 0.9 0.2 0.3 0.5 was its low hydrogen transfer activity. Hydrogen transfer
Naphthenes 2.3 1.0 1.1 0.8 index (HTI) is a measure of hydrogen transfer activity of
Aromatics 42.0 48.0 46.0 46.0 the catalyst during cracking reactions [23]. It is defined as
Benzene 2.2 5.3 5.2 3.1 the ratio of sum of propane and butanes selectivity to
Toluene 11.4 17.3 17.2 15.4 propylene selectivity. Z-22 and Z-27 showed the highest
m-Xylene 7.2 6.8 6.7 8.3 HTI of 2.5–2.9 compared with 0.7 for Meso-Z and 2.2 for
p-Xylene 2.7 2.5 2.6 2.7 Z-150. The values of HTI for Z-27 were similar to those
o-xylene 3.0 2.9 2.7 3.6 found by Zhu et al. [23]. Hydrogen transfer activity for low
C9? 15.5 13.3 12.1 12.8
Si/Al zeolites, Z-22 and Z-27 was high because of their
a
high acidity, as hydrogen transfer reactions require high
Naphtha conversion is the sum of total gas and coke strength of acid sites.
b
HTI (hydrogen transfer index) ratio of propane and butanes selec- Hydrogen transfer reactions occurred at the outer sur-
tivity to propylene selectivity
face of low Si/Al zeolites as higher C1 yield was observed
(Table 3) and steric constraints [24] did not play a role.
slightly higher compared with lower Si/Al ratio catalysts, Reactions on external surface follow a radical mechanism,
Z-22 and Z-27. Highest ethylene yield was also achieved which gives more C1 and C2 hydrocarbons. Corma and
for Meso-Z (9.4 wt%) compared with ethylene yield Orchilles [24] showed that methane and ethane can be
obtained for Z-22 (6.6 wt%) and Z-27 (7.8 wt%). Ethylene formed by a protolytic cracking of branched products. Part
yield was slightly increased for high Si/Al Z-150 (8.2 wt%) of these two gases and most of the ethylene are formed by a
compared with low Si/Al Z-27. Methane yield, which is an radical type of cracking, in which extraframework alu-
indication of the extent of cracking reaction, was lowest for minium plays an important role. Propane yield was sig-
Meso-Z (1.4 wt%) compared with 5.5 wt% for Z-22, nificantly low for Meso-Z because of its significantly low
4.9 wt% for Z-27 and 3.1 wt% for Z-150. hydrogen transfer activity, also evidenced by lower BTX
Enhanced light olefins yield (ethylene and propylene) yields. Propane yield was higher for the other ZSM-5
was achieved for Meso-Z (24.0 wt%) compared with the catalysts because of their higher hydrogen transfer activity.
other three catalysts. For the other catalysts, light olefins There was an increasing trend of propane yield with
yield decreased in the order of Z-150 [ Z-27 [ Z-22. The increasing hydrogen transfer index (Table 3). The highest
propylene/ethylene ratio of 1.6 was highest for Meso-Z yield of light olefins was obtained for Meso-Z because of

123
1392 Top Catal (2010) 53:1387–1393

suppression of olefins consuming, secondary hydrogen


transfer reactions as indicated by its low HTI (Table 3).
These secondary reactions were suppressed predominantly
due to shorter residence time, and rapid elution of primary
reaction products caused by larger intracrystalline spaces
of Meso-Z and not because of its low Bronsted acidity. The
primary products can undergo secondary cracking, aro-
matization and hydride transfer reactions. Z-150 gave
higher light olefins yield compared with Z-27 because of
low density of acid sites and because of its higher meso-
pore volume.

3.5 Hydrocarbon Distribution in Liquid Product


Fig. 4 Comparison of BTX content in cracked naphtha for Meso-Z,
The composition of liquid products in cracked naphtha at Z-22, Z-27 and Z-150 at constant catalyst/naphtha ratio of 3.0 and
600 °C
constant C/N ratio for the four catalysts is given in Table 3.
The results show a significant decrease in iso-paraffins,
n-paraffins, olefins and naphthenes and an increase in linear increase in naphtha conversion with increased tem-
aromatics fraction. BTX aromatics were produced either by perature was observed. At C/N of 4.5, naphtha conversion
cyclization of olefins or by hydrogen transfer or by dehy- increased from 50% at 600 °C to 56% at 650 °C. This is
drogenation of corresponding naphthenes [25] Olefins, typical of acid catalyzed cracking reactions in which
comprising 32% of the naphtha feed, were reduced to cracking is higher at higher temperatures.
0.95 wt% for Meso-Z. For Z-22 and Z-27 catalysts, the At lower C/N ratio light olefins yield was the highest.
olefins were almost completely cracked. Olefins are known Figure 6 shows the variation of propylene and ethylene
to be very reactive compared with paraffins and naphth- yields with temperature at the lowest C/N ratio of 1.5.
enes. Iso-paraffins fraction comprising 28 wt% of naphtha Propylene yield increased from 15.6 wt% at 600 °C to
feed was decreased to 7 wt%, for Meso-Z corresponding to 17.1 wt% at 625 °C, an increase of 10%. On further
a decrease of 75%. For Z-22 and Z-27 and Z-150, iso- increase of temperature to 650 °C, propylene yield
paraffins were decreased to 2–5 wt%. Naphthenes com- increased only by 4% to 17.8 wt%. Ethylene yield
prising 10 wt% of naphtha feed was decreased to 2.3 wt% increased from 8 wt% at 600 °C to 9.2 wt% at 625 °C, an
by Meso-Z and decreased to about 1 wt% by Z-22, Z-27 increase of 14%. It increased further by the same extent to
and Z-150. 10.5 wt% on further increase of temperature to 650 °C.
The aromatics content of increased to 42 wt% for Meso- The extent of increase in ethylene yield at higher tem-
Z compared with 25 wt% in naphtha feed. An increase in perature was higher than the increase in propylene yield.
aromatics content was also observed for Z-22, Z-27 and This trend was due to their different selectivity trends.
Z-150 catalysts. It increased to 46–48 wt% from 25 wt% in
the feed. The distribution of BTX in the liquid fraction of 60
cracked naphtha for the four catalysts is presented in
Table 3 and illustrated in Fig. 4. Benzene content for low
Si/Al catalysts Z-22 and Z-27 was higher (5 wt%) com- 55
Conversion, %

pared with Z-150 (3 wt%). About 17 wt% toluene was


produced by Z-22 and Z-27 compared with 11 wt% for
Meso-Z. Xylenes yield was similar for the four catalysts. 50

Higher aromatics content (C9?) was lower for Z-22, Z27


and Z-150 catalyst compared with Meso-Z. This was due to 45
easier accessibility in large pores of Meso-Z catalyst.

3.6 Effect of Reaction Temperature 40


575 600 625 650 675

MAT experiments were conducted at 600–650 °C with Temperature,°C


Meso-Z catalyst to study the effect of temperature on
Fig. 5 Effect of reaction temperature on naphtha conversion for
conversion and light olefins yield. Figure 5 shows the Meso-Z at catalyst/naphtha ratio of 1.5 (filled square), 3 (filled
effect of temperature on conversion at three C/N ratios. A triangle) and 4.5 (filled circle)

123
Top Catal (2010) 53:1387–1393 1393

Fig. 6 Effect of reaction 30


65
temperature on the yield and
selectivity to propylene (filled 25
square), ethylene (filled 55
triangle) and 20

Selectivity, %
Yield, wt.%
propylene ? ethylene (filled
45
circle) for Meso-Z
15

35
10

5 25

0 15
575 600 625 650 675 575 600 625 650 675
Temperature,°C Temperature,°C

Propylene selectivity was 36% at 600 °C and it did not References


show any increase with increasing temperature (Fig. 6).
However, ethylene selectivity increased with increased 1. Aitani A (2006) In: Lee S (ed) Encyclopedia of chemical pro-
cessing. Taylor & Francis, New York, p 2461
temperature. At higher temperatures, free radical mecha- 2. Jung JS, Park JW, Seo G (2005) Appl Catal A 288:149
nism increased ethylene yield further. Ethylene is the main 3. Wang G, Xu C, Gao J (2008) Fuel Process Tech 89:864
product of thermal cracking. Because of hydrogen transfer 4. Yang X, Feng Y, Tian G, Du Y, Ge X, Di Y, Zhang Y, Sun B,
reactions at higher temperature, propylene was consumed Xiao F (2005) Angew Chem Int Ed 44:2563
5. Wang S, Don T, Zhang Y, Li X, Yan Z (2005) Catal Commun
thereby reducing the increase in propylene yield. 6:87
6. Abrevaya H (2007) Stud Surf Sci Catal 170B:1244
7. Xie Z, Yao H, Yao W, Yang G, Ma J, Xiao L, Chen L (2010) US
4 Conclusions Patent 7686942
8. Hartmann M (2004) Angew Chem Int Ed 43:5880
9. Schmidt I, Boisen A, Gustavsson E, Stahl K, Pehrsons S, Dahl S,
High selectivity to light olefins was achieved over a novel Carlsson A, Jacobsen CJ (2001) Chem Mater 13:4416
mesoporous ZSM-5 with enhanced diffusion rates of FCC 10. Janssen H, Schmidt I, Jacobsen CJH, Koster AJ, de Jong KP
naphtha feed and products, mainly light olefins and aro- (2003) Micro Meso Mater 65:59
11. Christensen CH, Johannsen K, Schmidt I, Christensen CH (2004)
matics. The mesoporosity of ZSM-5 has significantly Catal Commun 5:543
enhanced the production of light olefins and aromatics 12. Motz J, Heinichen H, Holderich WF (1998) J Mol Catal A Chem
showing a two-fold increase in propylene and BTX yields 136:175
compared with conventional ZSM-5 catalyst. Using high 13. Ogura M, Shinimiya S, Tateno J, Nara Y, Kikuchi E, Matsuka M
(2000) Chem Lett 29:882
Si/Al ZSM-5 catalysts enhanced light olefins yields com- 14. Groen JC, Peffer LA, Moulijn JA, Perez-Ramirez J (2004) Micro
pared with low Si/Al ZSM-5 catalysts. However, signifi- Meso Mater 69:29
cant light olefins enhancement was achieved with 15. Musilová-Pavlačková Z, Zones SI, Čejka J (2010) Top Catal
mesoporous ZSM-5. The yield of ethylene and propylene 53:273
16. Čejka J, Mintova S (2007) Catal Rev 49:457
over the mesopore ZSM-5 reached 24.0 wt% compared 17. Pavlačková Z, Košová G, Žilková N, Zukal A, Čejka J (2006)
with 13.7 wt% over conventional ZSM-5. Propylene-to- Stud Surf Sci Catal 162:905
ethylene ratio of about 1.6 was obtained for Meso-Z due to 18. Al-Khattaf S, Pavlačková Z, Ali MA, Čejka J (2009) Top Catal
the relatively large pore volume compared with Z-22 and 52:140
19. Gil B, Zones SI, Hwang SJ, Bejblová M, Čejka J (2008) J Phys
Z-27. Mesopores, low acidity, and small crystals were Chem C 112:2997
important parameters to enhance light olefins yield from 20. Corma A, Melo FV, Sauvanaud L, Ortega F (2005) Catal Today
FCC naphtha cracking. Olefins consuming hydrogen 107–108:69
transfer reactions were suppressed and secondary reactions 21. Viswanadham N, Kamble R, Saxena SK, Singh M (2008) Catal
Comm 9:1894
were reduced due to shorter residence time and rapid 22. Nawaz Z, Tang X, Wei F (2009) Braz J Chem Eng 26:1
elution of primary reaction products. 23. Zhu X, Liu S, Song Y, Xu L (2005) Appl Catal A Gen 288:134
24. Corma A, Orchilles AV (1989) J Catal 115:551
Acknowledgments The authors express their appreciation to the 25. Hollander M, Wissink M, Makkee M, Moulijn J (2002) Appl
support from the Ministry of Higher Education, Saudi Arabia, in Catal A 223:85
establishing the Center of Research Excellence in Petroleum Refining
& Petrochemicals at King Fahd University of Petroleum & Minerals
(KFUPM).

123

Anda mungkin juga menyukai