You are on page 1of 9

J Porous Mater

DOI 10.1007/s10934-017-0410-5

Catalytic cracking of n-dodecane over hierarchical HZSM-5


zeolites synthesized by double-template method
Hong Liu1 · Zhenheng Diao1 · Guozhu Liu1,2 · Li Wang1,2 · Xiangwen Zhang1,2 

© Springer Science+Business Media New York 2017

Abstract  Hierarchical HZSM-5 zeolites have been pre- Keywords  Catalytic cracking · n-Dodecane · ZSM-5
pared by a double-template method with tetrapropylam- zeolites · Hierarchical zeolites · Mesoporous materials ·
monium hydroxide and polydiallydimethylammonium Zeolite coatings
chloride (PDDA) as micro- and meso-scale template,
respectively, and Al(NO3)3 as aluminum source. The as-
prepared zeolites were characterized by XRD, SEM, TEM, 1 Introduction
­N2-adsorption–desorption, TGA, ICP, in  situ FT-IR and
­NH3-TPD. The hierarchical HZSM-5 zeolite prepared at Recently, there has been a great interest in the catalytic
the PDDA/TEOS weight ratio of 0.2 in the synthesis gel cracking of hydrocarbon fuels due to its application poten-
(HZ-0.2P) had uniform mesopores (11–13  nm) and high tial in thermal management system of hypersonic aircrafts
acid site accessibility. An increase of the PDDA/TEOS where hydrocarbon fuels can serve as not only source
resulted in enlargement of mesopore size and broadening of heat through combustion, but also coolant through the
of size distribution, together with appearance of some non- cracking reaction to remove the waste heat from aircrafts [1,
uniform accumulation pores. Moreover, the accessibility 2]. Some studies have revealed that wall-coated HZSM-5
of acid sites in hierarchical HZSM-5 decreased. Catalytic zeolite catalyst (i.e. structured catalyst) could meet extreme
cracking of n-dodecane carried out at 500 °C and 4 MPa in working conditions of hypersonic flight (above 400 °C and
a wall-coated tubular reactor with HZSM-5 zeolites showed 3.4–6.9 MPa, i.e., the supercritical conditions) [3–5]. How-
that HZ-0.2P exhibited the superior catalytic activity and ever, the sole presence of micropores in HZSM-5 zeolites
stability, with the increase higher than 20% in the conver- induces severe diffusion restrictions, which slows down the
sion of n-dodecane and reduction to ca. 1/2 in the deacti- rate of cracking reaction of fuel [6].
vation rate compared with micro-porous HZSM-5 zeolite. Hierarchical HZSM-5 zeolites have received an ever-
This could be ascribed to its highest acid site accessibility, increasing attention owing to improved catalytic per-
which improves the diffusion of n-dodecane and its crack- formance during recent years [7, 8]. It has been reported
ing product molecules inside pore channels of catalyst. that hierarchical zeolites with uniform size of mesopore
was desired for catalytic cracking. For example, Góra-
Marek et al. [9] synthesized hierarchical HZSM-5 zeolites
through NaOH desilication in the presence of tetrabutyl-
* Li Wang
wlytj@tju.edu.cn amine hydroxide. They found that the existence of organic
amines had the advantages of narrowing the mesopore size
1
Key Laboratory for Green Chemical Technology distribution as well as achieving smaller mesopores, which
of Ministry of Education, School of Chemical Engineering
improved catalytic cracking activity of 1,3,5-tri-isopropylb-
and Technology, Tianjin University, Tianjin 300072, China
2
enzene. Similar results were also observed by Wang et al.
Collaborative Innovation Center of Chemical Science
[10].
and Engineering (Tianjin), Tianjin University, 92
Weijin Road, Nankai District, Tianjin 300072, In the last decades, numerous efforts have been put forth
People’s Republic of China to develop hierarchical zeolites. Top-down techniques such

13
Vol.:(0123456789)
J Porous Mater

as desilication and recrystallization are simple and efficient 2 Experimental


to prepare hierarchical zeolites [11, 12], but their applica-
tions are limited because of significant loss of micropore 2.1 Materials
volume and crystallinity. Bottom-up methods offer the
opportunity for better integrity of HZSM-5 zeolite struc- Tetrapropylammonium hydroxide (TPAOH, 25 wt% in
tures [13, 14]. Both hard-templating and soft-templating water) was purchased from J&K Chemical Ltd (Beijing,
routes are classic to prepare hierarchical zeolites but the China). Tetraethylorthosilicate (TEOS) and aluminum
use of hard-templating could not avoid the phase separation nitrate [Al(NO3)3·9H2O] were purchased from Guangfu
due to the hydrophobicity. Chemical Reagent Company (Tianjin, China). Polydially-
Generally, soft-templating routes use large hydrophilic dimethylammonium chloride solution (PDDA, average Mw
organics as meso-porogens. Ryoo et  al. [15, 16], Koek- 100,000–200,000, 20  wt% in water) and colloidal silica
koek et al. [17] and Zhao et al. [18] prepared hierarchical (40 wt% suspension in water) were purchased from Sigma-
HZSM-5 zeolites with tunable mesoporosity using amphi- Aldrich Co. (St. Louis, MO, USA). n-Dodecane with 99.5%
philic organosilane surfactant as mesopore-directing agent. purity was purchased from Sinopharm Chemical Reagent
The presence of a positive charge and silanols are favorable Co., Ltd (Shanghai, China). All chemicals were used as
to a strong interaction between the template and the silica received without further purification.
species [19]. Additionally, silane-functionalized polymers,
such as silane-functionalized polyethylenimine and poly- 2.2 Synthesis of hierarchical HZSM‑5 zeolites
propylene oxide diamine, are also used as a porogens for
introducing mesopores in HZSM-5 zeolite, as reported by The hierarchical HZSM-5 zeolites were prepared by a dou-
Wang et al. [20] and Park et al. [21]. But caution should be ble-template method according to [25] with some modifica-
paid to that these templates are expensive or always com- tions. Because the hierarchical HZSM-5 zeolites will apply
mercially unavailable [19], which would hinder their indus- to catalytic cracking of supercritical n-dodecane, we used
trial applications. Al(NO3)3 instead of N ­ aAlO2 as aluminum source avoid-
Alternatively, the use of cationic polymers as meso- ing a further transformation of Na-form zeolite to H-form
porogen for the synthesis of hierarchical zeolites is a fac- zeolite by ion exchange and calcination. And polydial-
ile route [22]. Xiao et  al. synthesized Beta zeolite with lydimethylammonium chloride instead of dimethyldiallyl
5–40  nm mesostructures in the presence of TEAOH and ammonium chloride acrylamide copolymer was used as
polydiallyldimethylammonium chloride (PDDA) [23]. meso-template. In addition, some crystallization conditions
These Beta zeolites exhibited high activity and selectivity were also adjusted. In a typical process, a required amount
in alkylation of benzene with isopropanol after a transfor- of TPAOH was added to the solution of Al(NO3)3·9H2O
mation process of Na-form to H-form by ion-exchange and in water, and then TEOS slowly dropped into the mixture
calcination compared with conventional zeolites prepared under stirring. The resulting solution with a molar compo-
in the absence of PDDA. Subsequently, the authors syn- sition of 0.36 TPAOH:1 TEOS:58 H ­ 2O:0.02 Al(NO3)3 was
thesized mesoporous NaX zeolites with preformed FAU stirred for 24  h at room temperature until a homogeneous
zeolite precursors and PDDA, and hierarchical mesoporous aluminosilicate gel was obtained. The gel was transferred
ZSM-5 zeolites with tetrapropylammonium hydroxide and into a Teflon-lined stainless-steel autoclave and pre-crystal-
dimethyldiallyl ammonium chloride acrylamide copolymer, lized at 100 °C for 5 h. A required amount of PDDA solu-
demonstrating universality of this route [24, 25]. tion was added, followed by an additional stirring at room
In our previous work, mesoporous ZSM-5 zeolites were temperature for 24 h. After crystallization at 180 °C for 72 h
prepared through alkali-desilication and used for catalytic in a Teflon-lined stainless steel autoclave, the samples were
cracking of supercritical n-dodecane [26]. It was found that recovered by centrifugation and washed thoroughly with
appropriate desilication via alkali-treatment could improve deionized water, dried overnight at 100 °C and calcined at
the catalytic activity and stability of ZSM-5 zeolites due to 550 °C for 6 h to remove the template. The obtained sam-
the introduction of intracrystal mesopores. The aim of this ples are denoted as HZ-xP where x (=0, 0.2, 0.6 and 0.8)
paper is to study the effects of meso-pore structure of hier- is the weight ratios of PDDA to TEOS in the synthesis gel.
archical HZSM-5 zeolites on their catalytic activity and sta-
bility. Firstly, hierarchical HZSM-5 zeolites with uniform 2.3 Characterization
mesopores were prepared using cationic polymer (PDDA)
as meso-porogen. Then, these zeolites were employed as The powder X-ray diffraction (XRD) measurements
catalysts in the form of coatings on the inside of a tubu- were carried out on a Philips X’Pert MPD diffractometer
lar reactor for cracking of supercritical n-dodecane (500 °C equipped with Cu-Kα radiation source (40  kV, 200  mA).
and 4 MPa). Transmission electron microscope (TEM) images were

13
J Porous Mater

collected using a JEM-2100F microscope. Scanning elec- procedure, the coated reactor was dried overnight at room
tron microscopy (SEM) was obtained using a Nanosem temperature and then calcined at 600 °C for 2 h. The result-
430 scanning electron microscope (FEI-SEM). Nitrogen ing zeolite coatings are denoted as HZC-xP series.
adsorption and desorption processes at 77  K were meas- The tubular reactor was heated by a direct current power
ured using a Micomeritics ASAP 2020 volumetric adsorp- and the outlet temperature was kept at 500 °C (reaction tem-
tion analyzer. The Brunauer–Emmett–Teller (BET) method perature), as described in our previous work [34]. The reac-
was applied to evaluate the total specific surface area (SBET) tion pressure was maintained at 4 MPa by a back-pressure
[27]. The t-plot method was used to calculate the volume of valve. The flow rate of n-dodecane pumped into the reactor
micropores (Vmicro) and external surface area (Sexter) [28]. was 10 mL  min−1. The products were first quenched by a
The total pore volume (Vt) was determined from ­N2 uptake condenser and then flowed into a gas–liquid separator. Gas
at a relative pressure (P/P0) of 0.99 [29]. The mesopore and liquid samples were collected at an interval of 5 min to
size distribution was calculated by the BJH model follow- ensure their weight enough for the material balance.
ing the adsorption branch of the isotherm [30]. Elemental The liquid products were analyzed by a GC (Agilent
analyses were conducted on an inductively coupled plasma 7890A) with a flame ionization detector and a capillary
(ICP) spectroscopy [USA Thermo Jarrell-Ash Corp. column (PONA, 50  m × 0.2  mm × 0.5  µm). The peak area
ICP-9000(N + M)]. normalization method was used for calculating the con-
Ammonia temperature-programmed desorption tents of the liquid products. The gaseous products were
­(NH3-TPD) measurements were carried out on a Quan- analyzed online with an Agilent 3000A Micro gas chroma-
tachrome CHEMBET 3000. The samples of 0.05  g were tograph equipped with three thermal conductivity detec-
charged in a quartz tubular reactor and pretreated at 600 °C tors and multichannel analytical columns, i.e., molecular
with an Ar flow of 30 mL  min−1 for 1  h and then cooled sieve (10 m × 12 µm), Plot U (10 m × 30 µm) and alumina
to 50 °C. Ammonia (20% N ­ H3 in He) was introduced at a (10  m × 8  µm). External standard method was used for
flow rate of 30 mL min−1 for 0.5 h at 50 °C and then a He quantitation of the gaseous products (standard gas mixture
stream was fed in until a constant TCD signal was obtained. was purchased from Tianjin Lianbo Chemical Co.).
The physisorbed ammonia was removed by flowing He for
60  min at 100 °C. The chemically adsorbed ammonia was
determined by rising the temperature up to 600 °C with a 3 Results and discussion
heating rate of 10 °C ­min−1.
Pyridine-adsorbed infrared spectroscopy (Py-IR) was 3.1 Structures characterization
recorded on a FT-IR Bruker Equinox spectrometer (Vertex
70). The samples were pressed in self-supported wafers (ca. 3.1.1 Crystal phase and morphologies
10  mg  cm−2) and activated in an in  situ IR cell at 400 °C
and vacuum of ­10−3 Pa prior to the adsorption of pyridine. As illustrated in Fig.  1, for all the samples there were
Adsorption was performed at 50 °C for 30 min. IR spectra five characteristic diffraction peaks between 2θ = 7° and
with a resolution of 2 cm−1 by collecting 45 scans for a sin-
gle spectrum were recorded after evacuation desorption for
20  min at 150 °C. The adsorption of 2,6-di-tert-butylpyri-
101 501
dine (DTBPy) was proceeded at 50 °C and IR spectra were 020 151
recorded after evacuation desorption for 20 min at 150 °C. 303
For the calculation of the accessibility index (ACI) of acid HZ-0.8P
sites, extinction coefficients for pyridine were used [31].
Intensity

HZ-0.6P
2.4 Catalytic cracking of n‑dodecane

The catalytic cracking of supercritical n-dodecane was HZ-0.2P


carried out in a tubular reactor with 300  mm in length
and 2 mm id. Prior to the catalytic test, the as-synthesized
HZSM-5 zeolites (HZ-xP series) were coated on the inner HZ-0P
surface of the reactor by a washcoating method using a
freshly prepared coating slurry (by mixing zeolite, colloi- 10 20 30 40 50
dal silica and ethanol and then ball-milling for 3  h) with 2θ (degrees)
the mass composition of 1.0 zeolite:2.5 colloidal silica:4.5
ethanol [32, 33]. After completion of the washcoating Fig. 1  XRD patterns of HZ-xPs

13
J Porous Mater

25°, attributed to the (110), (020), (501), (151) and (303) HZ-0P and appearance of some mesopores in HZ-0.2P,
crystal faces of the MFI zeolite. This indicated that the HZ-0.6P and HZ-0.8P.
as-synthesized zeolites have topology of MFI zeolite.
Compared with that synthesized without the addition of 3.1.2 Textural structure
PDDA (HZ-0P), all the HZSM-5 zeolites synthesized
with the addition of PDDA into the synthesis gels (HZ- Figure  4 gives nitrogen adsorption isotherms and BJH
0.2P, HZ-0.6P and HZ-0.8P) showed decreased crystal- adsorption pore size distribution curves. HZ-0P showed
linity (Table  1), indicating a negative effect of the addi- type I isotherm typical for microporous materials. The oth-
tion of PDDA on crystallinity of MFI zeolite. But, three ers displayed type I isotherms at the relative pressure below
zeolites with the addition of PDDA have almost the same 0.01 and type IV with a capillary condensation loop at the
crystallinity of ∼90%. Xiao et  al. [25] synthesized hier- relative pressure higher than 0.4 which is indicative of the
archical mesoporous ZSM-5 zeolites with dimethyldiallyl presence of mesopores. The hysteresis loop of HZ-0.6P and
ammonium chloride acrylamide copolymer (PDD-AM) HZ-0.8P extends up to the relative pressure higher than
as meso-scale template and observed that the relative 0.85, indicating the presence of non-uniform accumulation
crystallinity increased with increasing the amount of pores. In Fig.  4b, HZ-0P did not present any distribution
PDD-AM in the solution gels. However, the contradic- peaks in the range of 2–50  nm. HZ-0.2P displayed a nar-
tory results that the relative crystallinity of ZSM-5 zeo- row and intensive peak at 11–13 nm, implying the presence
lites decreased with the amount of PDD-AM were also a large amount of uniform mesopores. Both HZ-0.6P and
reported by Yang et al. [35] and Wang et al. [36]. It can HZ-0.8P showed the broad pore size distributions in the
be explained by the different amounts of cationic poly- range of ca. 11–34  nm. Overall, the size distributions of
mers, i.e., PDD-AM/TEOS weight ratio lower than 0.092 mesopores obtained in this work are narrower than those
was used by three research groups, whereas PDDA/TEOS reported by Xiao et  al. [23, 25], which is possibly caused
weight ratio higher than 0.2 was used in this work. In by the difference in the distribution range of molecular
addition, the difference in aluminum sources might be weight of used cationic polymer. In their works, PDDA
another reason [37]. with the molecular weight of 1 × 105 to 1 × 106 was used
SEM images (Fig. 2) revealed that HZ-0P has hexago- for the crystallization of Beta zeolites and PDD-AM with
nal pseudoprism morphology with smooth crystal-faces the molecular weight of ∼2.6 × 106 for the crystallization of
and an average crystal size of approximately 500  nm. ZSM-5 zeolites. The authors found that both the hierarchi-
After adding PDDA into the synthesis gels, the crystal cal mesoporous Beta and ZSM-5 zeolites obtained showed
size of zeolites increased remarkably and the morphology mesopores at 5–40  nm, which is in good agreement with
became ellipsoidal shape. Moreover, the crystal surface the size of molecular and aggregated cationic polymer. In
roughness increased with the amount of PDDA. TEM this work, the molecular weight of used PDDA is 1–2 × 105,
images (Fig.  3) showed the sole presence of micropores which is lower and narrower than those used by Xiao et al.
with the pore diameter of around 0.6  nm in crystals of Table  1 summarizes the values of total specific surface
area (SBET) and volume (Vt), as well as that of external

Table 1  Textural and acid properties of HZ-xPs


Samples Rca SBET ­(m2 g−1) Sexter ­(m2 g−1) Vmesob ­(cm3 g−1) Vmicro ­(cm3 g−1) Si/Alc Acid ­densityd ACIe
(μmol NH3 g−1)
B L T

HZ-0P 100 398 112 0.070 0.132 64 120 148 268 0.09
HZ-0.2P 91 490 244 0.164 0.113 66 111 156 267 0.36
HZ-0.6P 90 479 235 0.196 0.112 65 101 159 260 0.29
HZ-0.8P 90 471 229 0.209 0.111 66 101 158 259 0.25
a
 XRD relative crystallinity calculated according the sum of the peak intensities in the 22° < 2θ < 25° region based on that synthesized without
addition of PDDA (HZ-0P)
b
 Vmeso = Vt − Vmicro
c
 Si/Al ratio measured by ICP
d
 Acid density calculated by combing the ­NH3-TPD and Py-FTIR methods. B, L and T represent Brønsted, Lewis and total acid site
e
 Accessibility index determined as the ratio of acid sites accessible to DTBPy (the band at 1617  cm−1) to that of pyridine (the band at
1543 cm−1)

13
J Porous Mater

Fig. 2  SEM images of HZ-0P


(a), HZ-0.2P (b), HZ-0.6P (c)
and HZ-0.8P (d)

Fig. 3  TEM images of HZ-0P


(a), HZ-0.2P (b), HZ-0.6P (c)
and HZ-0.8P (d)

13
J Porous Mater

Quantity Adsorbed (cm3/g STP)


(a) HZ-0P
HZ-0.2P
HZ-0.6P
HZ-0.8P

TCD single (a.u)


HZ-0P
HZ-0.8P
HZ-0.2P

HZ-0.6P

0.0 0.2 0.4 0.6 0.8 1.0


Relative Pressure (P/P0) 100 200 300 400 500
o
Tempreture ( C)
0.10
HZ-0P
(b)
Fig. 5  NH3-TPD profiles of HZ-xPs
dV/dlogD (cm3 g-1 nm-1)

HZ-0.2P
0.08 HZ-0.6P
HZ-0.8P
0.06
Figure 5 shows the N­ H3-TPD results of HZ-Ps. There are
two peaks at around 185–200 and 370–400 °C, assignable
0.04 to the weak and strong acid sites, respectively, on profiles
of all zeolites. The slightly lower intensities of strong acid
0.02 sites for hierarchical zeolites than for HZ-0P, together with
similar maximum temperature, suggested that the amount
0.00 of strong acid sites decreased slightly and the acid strength
10 20 30 40 was unchanged after introducing mesopores. Figure  6a
Pore Diameter (nm) illustrates pyridine-adsorbed FT-IR spectra of HZs. All
zeolites showed characteristic bands at around 1452, 1543,
Fig. 4  N2 adsorption–desorption isotherms (a) and BJH adsorption and 1489  cm−1, assignable to pyridine interacting with
pore size distribution (b) of HZ-xPs Lewis (L), Brønsted (B), and both acid sites. By combining
­NH3-TPD and pyridine-adsorbed FT-IR, the amount of B
and L acid sites was calculated and the results are listed in
specific area (Smeso) and microporous volume (Vmicro). SBET Table 1. Overall, the differences in both B and L acid sites
of HZ-0.2P increased significantly compared with that of
HZ-0P. But, HZ-0.6P and HZ-0.8P gave slightly lower SBET
than HZ-0.2P. This suggests that uniform mesopores ben- (a) (b)
efit SBET of hierarchical zeolites. Vmeso of zeolites gradually
increased with increasing the PDDA amount and Vmicro was
HZ-0.8P HZ-0.8P
opposite, which is coincident with the results reported by
Xiao et al. [25]. A careful comparison shows that SBET and
Absorbance %

Vmicro obtained in this work are higher and Vmeso is lower HZ-0.6P HZ-0.6P
than those by Xiao et  al. This might be ascribed to the
lower molecular weight and its narrower distribution range
HZ-0.2P HZ-0.2P
of PDDA used in this work, as well as the higher addition
of PDDA in the solution gels.
HZ-0P HZ-0P
3.2 Compositions and acid properties
1450 1500 1550 1450 1500 1550 1600 1650
The Si/Al ratios of HZ-xPs were measured by ICP and the Wavenumber (cm-1)
results are shown in Table 1. Obviously, very similar Si/Al
ratios were obtained, indicating the addition of PDDA and Fig. 6  Pyridine (a) and 2,6-di-tert-butylpyridine (b) adsorbed FT-IR
its amount had negligible effects on incorporation of Al. spectra of HZ-xPs

13
J Porous Mater

were very small among zeolites, especially among three 55


hierarchical zeolites. HZC-0P

Conversion of n-dodecane(%)
The adsorption of 2,6-di-tert-butylpyridine (DTBPy) 50 HZC-0.2P
HZC-0.6P
is used commonly to probe the accessibility of the Brøn- HZC-0.8P
45
sted acid sites in zeolites because the kinetic diameter of
DTBPy (1.05  nm) exceeds the size of HZSM-5 micropo- 40
res (0.53 × 0.56  nm and 0.51 × 0.55  nm) [38, 39]. From
Fig.  6b, the strong band at 1617  cm−1 and weak band at 35
1530  cm−1 were observed after the adsorption of DTBPy,
which would be attributed to the DTBPy bonded to Brøn- 30
sted sites [38]. The absence of the band at 1543 cm−1 evi-
25
denced that dealkylation of DTBPy did not occur. More-
over, there was no band at 1452  cm−1 being observed 0 10 20 30 40 50 60
because the tert-butyl groups prevent the interaction with Time on stream (min)
Lewis acid sites [39]. Among HZ-Ps, an evident difference
in the peak areas at 1617  cm−1 could be found, revealing Fig. 7  Conversion of n-dodecane over HZSM-5 zeolite coatings
the difference in accessibility of acid sites. In this work, the (reaction conditions: 500 °C, 4 MPa, 10 mL min−1)
accessibility index (ACI) was calculated and is presented
in Table  1. ACI was only 0.09 for HZ-0P, corresponding
to the sole presence of micropores. While those for hierar- indicating the quick deactivation of the catalyst with a
chical zeolites were higher than 0.25 due to the formation deactivation rate (rd) of 39.9% (Table  2). All hierarchi-
of mesopores. Although the ­NH3-TPD and Py-IR results of cal zeolites showed improved stability and increased con-
hierarchical zeolites are similar, their ACI show great dif- version compared to HZC-0P. Among them, HZC-0.2P
ference, from 0.25 for HZ-0.8P to 0.29 for HZ-0.6P, and presented the lowest rd with a value of 21.1% (Table  2)
further to 0.36 for HZ-0.2P. This implies that HZ-xPs are which was only ca. half of that over HZC-0P and the
significantly different in pore diffusion. highest conversion of n-dodecane with the improvement
higher than 20% compared with HZC-0P. This could be
3.3 Catalytic cracking performances ascribed to the higher ACI of hierarchical zeolites. Stud-
ies on catalytic cracking demonstrated that catalytic
3.3.1 Catalytic cracking activities cracking activities of zeolites did not vary monotonously
with the amount of acid sites, instead diffusive property
Catalytic cracking of supercritical n-dodecane were carried also played an important role [39–42]. HZ-xPs have the
out in a flowing reactor with zeolite coatings (HZC-xPs) at similar acid density but significantly different ACI. The
500 °C and 4 MPa. In the test of catalytic cracking, the sim- higher ACI implies the enhanced diffusion of reactant
ilar loading amount of zeolite (ca. 5.56 mg cm−2) and mean and product molecules inside pore channels of catalyst.
thickness of coatings (ca. 40.0 μm) were used (as shown in HZ-0P has extremely low ACI, which is bad to the diffu-
Table 2). sion of n-dodecane molecules and cracked product mol-
Figure  7 depicted the conversion of the n-dodecane ecules. As a result, HZC-0P exhibits the lowest initial
catalytic cracking as a function of the time on stream conversion and worst stability. HZ-0.2P has the highest
(TOS) over HZC-xPs. The conversion of n-dodecane over ACI, resulting in the highest conversion of n-dodecane
HZC-0P dropped rapidly with the going on of reaction, and best stability.

Table 2  Physical properties Coatings Samples Loading amounts Mean ­thicknessa (μm) rdb (%) Average
and catalytic performances of (mg cm−2) conversion
zeolite coatings (%)

HZC-0P HZ-0P 5.61 ± 0.14 40.4 ± 2.3 39.9 28.68


HZC-0.2P HZ-0.2P 5.56 ± 0.20 39.5 ± 2.4 21.1 44.41
HZC-0.6P HZ-0.6P 5.61 ± 0.17 40.7 ± 2.8 21.7 42.62
HZC-0.8P HZ-0.8P 5.51 ± 0.18 38.9 ± 1.6 28.5 40.05
a
 Mean thickness determined by SEM
b
 Deactivation rate (rd) is defined as: rd = (Xt = 5 − Xt = 45)/Xt = 5 × 100%, wherein Xt = 5 and Xt = 45 are n-dode-
cane conversions at time on steam of 5 and 60 min, respectively

13
J Porous Mater

3.3.2 Product distribution Table 3  Mass selectivity and conversion of n-dodecane catalytic


cracking over zeolite coatings
It is well known that n-dodecane is cracked firstly into pri- Products HZC-0P HZC-0.2P HZC-0.6P HZC-0.8P
mary paraffin and olefin, and the formed molecules would
Hydrogen 0.08 0.05 0.07 0.07
then be deeply cracked into lower carbon number hydro-
Methane 4.47 3.06 3.14 3.26
carbons [41, 43]. Therefore, the high selectivity of liquid
Ethylene 8.72 9.25 8.92 8.84
products (C5-C11 alkenes/alkanes and C12 alkene) implies
Ethane 8.35 6.31 6.56 7.19
a low degree of deep-cracking. It can be seen in Table  3
Propane 9.04 6.87 7.52 7.78
that the obviously high selectivity of liquid products was
Propylene 14.12 13.15 13.83 14.2
obtained over the coatings of hierarchical zeolites, suggest-
iso-Butane 0.98 1.12 1.08 1.23
ing the inhibition of deep-cracking. This supports the spec-
n-Butane 4.25 3.75 3.88 3.90
ulation of that the high ACI is favorable to the diffusion of
trans-2-Butene 2.67 1.68 0.99 1.14
cracked product molecules.
iso-Butene 2.59 3.91 3.36 3.6
Hydrogen transfer is an important secondary reaction
1-Butene 6.05 6.75 7.08 6.45
during the catalytic cracking of fuels and its extension can
cis-2-Butene 1.34 2.07 2.82 3.19
be quantitatively described using the olefin/paraffin ratio
Isopentene 0.24 0.33 0.36 0.32
(o/p) in the C3 and C4 products and even better in the isob-
Isopentane 0.57 0.46 0.52 0.45
utene/isobutane ratio. The higher the o/p ratio, the lower
Pentene 5.97 6.44 5.78 5.41
the extension to that hydrogen transfer reactions occur [44].
n-Pentane 6.12 5.37 4.77 4.65
It can be seen in Table 3 that C­ 3=/C3, ­C4=/C4 and i-C4=/i-
Hexene 3.95 5.23 4.81 4.93
C4 ratios are in good agreement with ACI, indicating
n-Hexane 2.07 2.69 2.24 2.58
high ACI could reduce the extension of hydrogen transfer
Benzene 0.46 0.31 0.34 0.35
because of the improvement of diffusion.
Heptene 3.14 3.55 3.90 3.30
Ishihara et  al. reported that mesopores of the matrix
n-Heptane 2.17 1.28 1.73 1.08
surrounding microporous zeolite crystals in the mixed
Toluene 0.69 0.62 0.60 0.64
catalysts of zeolite (β and Y) and silica-alumina are very
Octene 2.71 3.13 3.49 3.27
favorable for the improvement of catalytic activity, forma-
n-Octane 0.93 1.46 1.02 1.65
tion of branched products and inhibition of over-cracking
Nonene 2.64 2.97 3.60 3.56
of n-dodecane [41, 45]. In this work, zeolite coatings were
n-Nonane 0.89 1.51 1.85 1.63
prepared for the catalytic cracking tests and a similar struc-
Decene 2.15 3.69 2.82 2.86
ture that S­ iO2 in catalytic coatings which is formed from
n-Decane 0.58 0.35 0.65 0.46
colloidal silica as binder by calcination surrounds ZSM-5
Undecene 0.93 1.58 1.37 1.27
zeolite crystals might be created. To understand the pore
n-Undecane 0.86 0.64 0.52 0.45
structure of ­SiO2, colloidal silica was dried at 100 °C and
Dodecene 0.27 0.42 0.38 0.29
calcined at 600 °C for 2 h, and then nitrogen adsorption and
Gaseous products 62.66 57.97 59.25 60.85
desorption measurement of the as-obtained powder was
Liquid products 37.34 42.03 40.75 39.15
performed. The results showed that the S ­ iO2 powder exhib-
C3=/C3 1.56 1.91 1.85 1.83
ited mesopores with the pore size centered on ca. 5.5 nm.
C4=/C4 2.37 2.84 2.81 2.76
Therefore, the presence of mesopores derived from ­SiO2
i-C4=/i-C4 2.64 3.49 3.11 2.93
could improve the intercrystalline diffusion, which ben-
Conversion (%) 43.39 52.34 51.12 51.15
efits catalytic performance. Moreover, the positive effects
of ­SiO2 on catalytic performance universally existed in all Reaction condition: 4 MPa and 500 °C, TOS = 5 min
coatings because the same mass ratio of colloidal silica to
zeolite was used for the preparation of zeolite coatings.
and high acid site accessibility could be obtained at a
PDDA/TEOS weight ratio of 0.2 in the synthesis gel. The
4 Conclusion coating of this catalyst exhibited the great improvement in
the conversion of n-dodecane cracking and stability, as well
Hierarchical HZSM-5 zeolites were prepared using cationic as high selectivity of liquid product and o/p ratio, indicat-
polymer of PDDA as the mesoscale template. The amount ing the inhibition of deep cracking and hydrogen transfer
of PDDA in synthesis gel showed to be a crucial parameter reactions compared with the micro-HZSM-5 zeolite coat-
controlling mesopore size and distribution. The hierarchi- ing. However, an increase in the amount of PDDA induced
cal HZSM-5 zeolite with uniform mesopores (11–13  nm) the formation of some non-uniform accumulation pores,

13
J Porous Mater

which results in no further improvement in the conversion 21. D.H. Park, S.S. Kim, H. Wang, T.J. Pinnavaia, M.C. Papapetrou,
of n-dodecane cracking and stability being achieved. A.A. Lappas, S.K. Triantafyllidis, Angew. Chem. 121, 7781
(2009)
22. K. Zhang, M.L. Ostraat, Catal. Today 264, 3 (2016)
Acknowledgements  The authors gratefully acknowledge financial 23. F. Xiao, L. Wang, C. Yin, K. Lin, Y. Di, J. Li, R. Xu, D.S. Su, R.
support from National Natural Science Foundation of China (Grant Schlögl, T. Yokoi, T. Tatsumi, Angew. Chem. 118, 3162 (2006)
No. 21522605). 24. S. Liu, X. Cao, L. Li, C. Li, Y. Ji, F. Xiao, Colloid. Surf. A-Phys-
icochem. Eng. Asp. 318, 269 (2008)
25. L. Wang, Z. Zhang, C. Yin, Z. Shan, F. Xiao, Microporous

References Mesoporous Mater. 131, 58 (2010)
26. L. Zhang, S. Qu, L. Wang, X. Zhang, G. Liu, Catal. Today 216,
64 (2013)
1. L. Yue, X. Lu, H. Chi, Y. Guo, L. Xu, W. Fang, Y. Li, S. Hu, 27. S. Brunauer, P.H. Emmett, E. Teller, J. Am. Chem. Soc. 60, 309
Fuel 121, 149 (2014) (1938)
2. D.R. Sobel, L.J. Spadaccini, J. Eng. Gas Turbines Power 119, 28. B.C. Lippens, J.H. de Boer, J. Catal. 4, 319 (1965)
344 (1997) 29. J.C. Groen, J.C. Jansen, J.A. Moulijn, J. Perez-Ramirez, J. Phys.
3. H. Wu, G. Li, Appl. Catal. A 423–424, 108 (2012) Chem. B 108, 13062 (2004)
4. X. Fan, F. Zhong, G. Yu, J. Li, C. Sung, J. Propul. Power 25, 30. E.P. Barrett, L.G. Joyner, P.P. Halenda, J. Am. Chem. Soc. 73,
1226 (2009) 373 (1951)
5. G. Liu, G. Zhao, F. Meng, S. Qu, L. Wang, X. Zhang, Energy 31. T.M. Davis, C. Chen, N. Žilková, D. Vitvarová-Procházková, J.
Fuels 26, 1220 (2012) Čejka, S.I. Zones, J. Catal. 298, 84 (2013)
6. D.P. Serrano, J.M. Escola, P. Pizarro, Chem. Soc. Rev. 42, 4004 32. B. Mitra, D. Kunzru, J. Am. Ceram. Soc. 91, 64 (2008)
(2013) 33. S. Qu, G. Liu, F. Meng, L. Wang, X. Zhang, Energy Fuels 25,
7. N. Linares, A.M. Silvestre-Albero, E. Serrano, J. Silvestre- 2808 (2011)
Albero, J. Garcia-Martinez, Chem. Soc. Rev. 43, 7681 (2014) 34. R. Jiang, G. Liu, X. He, C. Yang, L. Wang, X. Zhang, Z. Mi, J.
8. L. Chen, X. Li, J.C. Rooke, Y. Zhang, X. Yang, Y. Tang, F. Xiao, Anal. Appl. Pyrolysis 92, 292 (2011)
B. Su, J. Mater. Chem. 22, 17381 (2012) 35. Q. Yang, H. Zhang, M. Kong, X. Bao, J. Fei, X. Zheng, Chin. J.
9. K. Sadowska, K. Góra-Marek, M. Drozdek, P. Kuśtrowski, J. Catal. 34, 1576 (2013)
Datka, J. Martinez Triguero, F. Rey, Microporous Mesoporous 36. X. Wang, X. Gao, M. Dong, H. Zhao, W. Huang, J. Energy
Mater. 168, 195 (2013) Chem. 24, 490 (2015)
10. D. Wang, L. Zhang, L. Chen, H. Wu, P. Wu, J. Mater. Chem. A 37. S.M. Alipour, R. Halladj, S. Askari, Rev. Chem. Eng. 30, 289
3, 3511 (2015) (2014)
11. I.I. Ivanova, E.E. Knyazeva, Chem. Soc. Rev. 42, 3671 (2013) 38. S. Zheng, H.R. Heydenrych, A. Jentys, J.A. Lercher, J. Phys.
12. D. Verboekend, J. Pérez-Ramírez, Catal. Sci. Technol. 1, 879 Chem. B 106, 9552 (2002)
(2011) 39. J.B. Koo, N. Jiang, S. Saravanamurugan, M. Bejblová, Z. Musi-
13. M. Hartmann, Angew. Chem. Int. Ed. Engl. 43, 5880 (2004) lová, J. Čejka, S.E. Park, J. Catal. 276, 327 (2010)
14. H. Zhu, Z. Liu, Y. Wang, D. Kong, X. Yuan, Z. Xie, Chem. 40. Y. Goto, Y. Fukushima, P. Ratu, Y. Imada, Y. Kubota, Y. Sugi,
Mater. 20, 1134 (2008) M. Ogura, M. Matsukata, J. Porous. Mater. 9, 43 (2002)
15. M. Choi, H.S. Cho, R. Srivastava, C. Venkatesan, D.H. Choi, R. 41. A. Ishihara, K. Inui, T. Hashimoto, H. Nasu, J. Catal. 295, 81
Ryoo, Nat. Mater. 5, 718 (2006) (2012)
16. V.N. Shetti, J. Kim, R. Srivastava, M. Choi, R. Ryoo, J. Catal. 42. J. Na, G. Liu, T. Zhou, G. Ding, S. Hu, L. Wang, Catal. Lett.
254, 296 (2008) 143, 267 (2013)
17. A.J.J. Koekkoek, C.H.L. Tempelman, V. Degirmenci, M. Guo, Z. 43. X. Qian, B. Li, Y. Hu, G. Niu, D. Zhang, R. Che, Y. Tang, D. Su,
Feng, C. Li, E.J.M. Hensen, Catal. Today 168, 96 (2011) A.M. Asiri, D. Zhao, Chem. Eur. J. 18, 931 (2012)
18. Z. Zhao, Y. Liu, H. Wu, X. Li, M. He, P. Wu, Microporous 44. A. Corma, V. Fornes, J. Martínez-Triguero, S.B. Pergher, J.

Mesoporous Mater. 123, 324 (2009) Catal. 186, 57 (1999)
19. H. Xin, X. Li, L. Chen, Y. Huang, G. Zhu, X. Li, Energy Envi- 45. A. Ishihara, H. Negura, T. Hashimoto, H. Nasu, Appl. Catal. A
ron. Focus 2, 18 (2013) 388, 68 (2010)
20. H. Wang, T.J. Pinnavaia, Angew. Chem. Int. Ed. 45, 7603 (2006)

13

You might also like