Вы находитесь на странице: 1из 7

Journal of Natural Gas Chemistry 12(2003)5662

Properties and Characterization of Modified


HZSM-5 Zeolites
Renqing L
u1 ,

Hejin Tangbo1 ,

Qiuying Wang2 ,

Shouhe Xiang1

1. Institute of New Catalytic Materials, Department of Material Chemistry, Nankai University, Tianjin, 300071, China
2. Catalytic Factory, Nankai University, Tianjin, 300071, China
[Manuscript received July 12, 2002; revised November 4, 2002]

Abstract: Physicochemical and catalytic properties of phosphorus and boron modified HZSM-5 zeolites
treated with 100% steam at 673 K were investigated. The acidity and distribution of acidic sites were studied by infrared spectroscopy using pyridine as probe molecule and temperature programmed desorption
(TPD) of ammonia. The structure of the samples was characterized by XRD, and the textural properties
of the catalysts were determined by nitrogen isothermal adsorption-desorption measurements and scanning
electron microscopy (SEM). The XRD results show that the modified samples have no novel crystalline
phase, indicating a high dispersion of phosphorus and boron species. After treatment, the microporous
volume and surface area of the samples markedly decrease, implying the blockage of the channel. The nitrogen adsorption-desorption measurements suggest that the isothermal type of all samples is a combination
of isothermal type I and IV, and all hysteresis loops resemble the H4-type in the IUPAC classification.
The total acidity of the modified samples, determined by pyridine adsorption IR and TPD of ammonia,
decreases in contrast to that of the parent HZSM-5. The conversion of n-heptane over P and B steammodified HZSM-5 is higher than that of P and B-modified HZSM-5 zeolites but lower than that of the
parent HZSM-5.
Key words: HZSM-5 zeolite, steam treatment, phosphorus, boron, secondary pore, texture, cracking
activity

1. Introduction

The properties of catalysts are carefully tuned


for the desired catalytic process before use. Zeolites are crystalline aluminosilicate, and their acidbase properties depend on the aluminum content in
the framework. The adjustment of the acidity may
be realized by proper SiO2 /Al2 O3 molar ratio crystallization, other elements replacing framework constituents, or modification of the zeolite. Dealumination, the removal of framework aluminum from the
zeolite lattice, is a well known procedure for stabilizing zeolites and creating mesopores, which help over-

come diffusional problems in the zeolite micropores


[1]. ZSM-5 is a member of the pentasil family of highsilica zeolites and, due to its unusual properties, has
found a wide range of applications as a catalytic material. Modification of HZSM-5 zeolites by impregnation with phosphoric or boric acid has been investigated because of its promising catalytic properties
in many reactions (e.g. conversion of methanol to hydrocarbons (MTH) [2,7,15,17,22], n-hexane cracking
[3,11], disproportionation of toluene [4,6,8,10], alkylation of toluene with methanol [46,10,13,15,16,20,21],
xylene isomerization [6], alkylation of benzene with
ethanol [9,23], alkylation of ethylbenzene [12], conversion of methyl chloride to ethylene and propylene

Corresponding author. Tel: (022)23509932; E-mail: shxiang@public.tpt.tj.cn.


On leave from Chemistry and Chemical Engineering College, University of Petroleum (East China), Dongying, Shandong
Province

57

Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

[18,19]). To our knowledge, no attempts have been


made to combine phosphoric or boric acid modification with steam treatment over HZSM-5 zeolites.
2. Experimental
2.1. Catalyst preparation
A template-free synthesized commercial HZSM-5
zeolite (SiO2 /Al2 O3 =50), supplied by the Catalytic
Factory of Nankai University, was used as the starting material (denoted parent 50). The zeolite material
was impregnated with an aqueous solution of phosphoric acid (H3 PO4 ) in order to reach a 1%P content.
After being dried in air, the product was heated to
393 K in a muffle furnace for 1 h. Then, the temperature was increased to 823 K and held for 3 h. This
product was designated as P501. Another product,
B501, with a B content of 1% was prepared in a similar manner except for the replacement of H3 PO4 with
H3 BO3 . Some P501 and B501 samples were treated
with 100% steam of 673 K for 4 h and designated as
P5014 and B5014, respectively.

673 K in a special IR cell under vacuum (0.04 Pa) for


1 h. After cooling to room temperature, excess pyridine was adsorbed and outgassed at 433 K to eliminate the physisorbed pyridine. The concentrations
of Br
onsted and Lewis sites able to retain pyridine
at 433 K were determined using the extinction coefficients and the adsorbance surface of the corresponding bands at around 1,540 and 1,450 cm1 , respectively.
TPD patterns of chemisorbed ammonia were
recorded using a DuPont 2000 thermoanalyzer by
means of NH3 adsorption-desorption.
2.3. Catalytic activity measurements
The catalytic activity of samples in n-heptane
cracking was determined in a pulse microreactor (i.d.
4 mm) connected to a gas chromatograph. The reaction was carried out with 0.2 g catalyst, a 30 cm3 /min
N2 flow rate, a 2 l pulse and at 773 K. Before
the activity was measured, the catalyst was activated
in situ at 793 K for 1 h in dry nitrogen stream.
3. Results and discussion

2.2. Catalyst characterization


3.1. Results of catalyst characterization
X-ray diffraction patterns were recorded on
D/max-2500 powder diffractometer using nickelfiltered Cu K radiation (=0.1542 nm) and equipped
with a graphite monochromator. The step scans were
taken over a 2 range from 5 to 50o .
The BET specific surface area and porosity texture of each sample were determined by nitrogen adsorption measurements at liquid nitrogen temperature
with an automatic Micromeritics ASAP 2400 apparatus. The samples were first degassed at 573 K for
approximate 6 h and then studied with a static volumetric technique.
IR measurements were carried out using pyridine
as the probe molecule, and the vibration spectra of
chemisorbed pyridine were recorded between 1,400
and 1,700 cm1 . The samples were pressed into selfsupporting wafers 20 mm in diameter and heated to
Table 1.

3.1.1. The measurement of physicochemical


properties
Specific surface area of the catalysts was computed according to the BET method from the nitrogen
adsorption isotherms obtained at 77 K, taking a value
of 0.162 nm2 for the cross-section of the adsorbed N2
molecule at that temperature. BET areas of the various samples are summarized in Table 1. The BET
surface area, microporous area and microporous volume of all modified samples significantly decreased
compared to the parent zeolite. Among the modified samples, BET surface area, mesoporous area and
mesoporous volume of the steam-treated P5014 and
B5014 are higher than those of P501 and B501, respectively, resulting in diffusion benefits. However,

Physicochemical properties of samples

BET surface area

Micropore area

Mesopore area

Micropore volume

Mesopore volume

(m2 /g)

(m2 /g)

(m2 /g)

(cm3 /g)

(cm3 /g)

Parent 50

388.5

303.1

85.4

0.1209

0.0429

P501

318.9

231.2

87.7

0.0940

0.0532

P5014

324.1

226.6

97.5

0.0925

0.0740

Sample

B501

323.3

251.3

72.0

0.1016

0.0324

B5014

340.2

233.7

106.5

0.0958

0.0731

58

Renqing L
u et al./ Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

the microporous area and microporous volume of


steam-treated P5014 and B5014 are lower than those
of P501 and B501. These results suggest that the
channel of phosphorus and boron-modified samples
is occluded, and steam treatment may result in secondary pore formation.
3.1.2. X-ray powder analysis
The x-ray powder pattern peaks of the five samples, exhibited by all samples in the XRD patterns,
are typically of the MFI topology. The patterns indicate that crystallinity was retained after treatment.
Also, there is no novel crystalline phase, which indicates that phosphorus and boron species are highly
dispersed on the zeolites. The interaction between
ZSM-5 zeolites and phosphorus and boron species may
result in the peak split around 2=23o .
3.1.3. IR measurement of adsorbed pyridine
Information about the type of acid sites and their
distribution in the catalysts could be obtained from
the infrared spectra of pyridine adsorbed on the samples in the 1,4001,700 cm1 spectral region. The
acidity of the five samples is shown in Figure 1, and
a qualitative estimation of the band intensity ratio
representing pyridine adsorbed at Br
onsted acid sites
and pyridine absorbed at Lewis acid sites is illustrated
in Table 2. As shown in Figure 1 and Table 2, the
parent zeolite possesses the largest number of both
Br
onsted and Lewis acid sites of the five samples.
The reason for the difference between the Br
onsted
and Lewis acid number in P501 and P5014 remains
unclear. Steam-treated B5014 and P5014 have a lower
acidity compared to B501 and P501, respectively.

Figure 1. Total acidity of samples measured by FTIR

3.1.4. TPD of ammonia


TPD profiles of the parent HZSM-5 and modified HZSM-5 are shown in Figure 2. The two-peak
pattern is well documented for HZSM-5 [24], indicating the existence of weak and strong acid sites in the
parent ZSM-5 zeolite. The profiles of modified samples reveal similar patterns as the parent sample, but
the peak intensity of weak and strong acid sites in the
modified samples markedly decreased. The significant
decrease in the acidity of the single-phosphorus modified P501 and the single-boron modified B501 may
be explained by the phosphorus and boron species
combining with bridging hydroxyl Al(OH)Si groups
[2]. After steam treatment at 673 K, some framework
aluminum atoms are partially hydrolysed to form nontetrahedrally symmetric aluminum atoms, which act
as a strong electron withdrawal centers for the remaining tetrahedral framework aluminum atoms thus
creating stronger Br
onsted acids [26]. Table 2 shows
the acidity of all samples, strongly suggesting the decrease in acidity of the modified samples in contrast to
Parent50. The acidity of the samples determined by
the pyridine adsorbed IR method is lower than that
of their counterparts determined by TPD of ammonia. This difference may be the results of differences
in the molecular size of these bases, i.e. the smaller
molecules of ammonia may penetrate through more
pores than the larger molecules of pyridine.

Figure 2. Ammonia TPD profiles of samples

Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

Table 2. Ratio of the B to L acid site intensity


(denoted Ratio) and acidity measured by
TPD of NH3 (denoted TPD)
Sample

Ratio

TPD (mmol/g)

Parent 50

1.643

0.79

P501

3.910

0.71

P5014

3.211

0.68

B501

1.321

0.60

B5014

1.144

0.54

3.1.5. Porosity measurement


The porous structure of all samples was determined by N2 adsorption-desorption measurements,
and the nitrogen isotherm for the samples is illustrated in Figure 3. According to IUPAC [25], the
shape of the adsorption isotherm can be classified into
one of six groups. Of these, the most common are type

59

I (Langmuir) isotherms for purely microporous solids,


and type IV for mesoporous goods in which capillary
condensation takes place at higher pressures of adsorbate as well as a hysteresis loop. As is shown in Figure
3, the adsorption volume at very low relative pressures (p/p0 <0.1) is high, indicating the presence of
microporous adsorption. Increasing the relative pressure causes capillary condensation, which illustrates
type IV behavior. All five materials show a hysteresis
loops that resembles the H4 type in the IUPAC classification. This can be attributed to the crystalline
agglomerates that result in the mesoporous structure
formed by the interparticle space and steam treatment that causes the formation of secondary pore. A
distinct increase in the adsorbate volume in the low
p/p0 region and the hysteresis loop in the high p/p0
region indicate the presence of micropores associated
with mesopores. Therefore, the isothermal type of all
samples is a combination of type I and type IV.

Figure 3. N2 adsorption-desorption isotherms of all samples

3.1.6. Scanning electron microscopy


A scanning electron micrograph (SEM, Figure 4)
of all the samples indicates the morphology of the

parent and modified HZSM-5 zeolite crystals. The


SEM photographs reveal a change in the morphology
of HZSM-5 upon steam treatment. A comparison between P501 and P5014 as well as between B501 and

60

Renqing L
u et al./ Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

B5014 reveals some cracks and faults that appeared


on the surface of steam-treated B5014 and P5014.
This shows the formation of secondary pores and this
formation is an important explanation for the higher
heptane cracking activity because cracking is often
limited by diffusion inside the micropores of the zeo-

lite [27]. P501 has few cracks or faults, and this may
be because of the stronger acidity of H3 PO4 than that
of H3 BO3 . This clarified that the mesoporous volume
of P501 is higher than that of B501 and the heptane
cracking activity of P501 is higher than that of B501.

Figure 4. SEM pictures of all samples: (a) parent, (b) enlarged parent, (c) P501, (d) P5014, (e) B501, (f )
B5014

61

Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

3.2. Activity of n-heptane cracking over the


catalyst
To further study the activity of catalysts,
n-heptane cracking was used as a test reaction.
The results of an n-heptane cracking
conversion over the catalysts are presented in
Table 3.
It reveals an order of conversion
(parent50=B5014>P5014>P501>B501) that shows
no correlation to the acidity determined by pyridine
adsorption and TPD-NH3 . The activity of steamtreated phosphorus and boron-modified samples is
higher than that of single-phosphorus and boronmodified samples. This may be the result of steam
enhancement of the BET surface area and mesoporous
Table 3.
Sample
Parent 50

area as well as proper steam treatment enhancement


of acid strength. The selectivity of products is also
shown in Table 3. It can be seen that B5014 shows the
highest C=
3 selectivity, while B501 shows the highest
C=
selectivity.
According to the acidic results calcu4
lated by TPD of ammonia and FT-IR, the acidity of
steam-treated P5014 and B5014 is slightly lower than
that of P501 and B501, respectively. This suggests
a decrease in diffusion constraints brought about by
the creation of mesopores in the steam-treated samples (as seen from the enhancement of the mesoporous
area and mesoporous volume of the steam-treated
samples). In addition, higher acid site strength may
contribute to the activity enhancement.

Product selectivity and conversion of heptane cracking


Product selectivity (%)

C1 -C2

C3

C=
3

C4

C=
4

C+
5

11.3

31.5

8.0

16.2

4.6

28.4

Conversion (%)
100

P501

10.9

30.5

5.7

17.3

4.9

30.7

93

P5014

12.5

36.5

6.1

17.9

4.2

22.9

94

B501

11.3

31.0

4.8

19.0

7.6

26.3

59.2

B5014

9.9

32.5

8.5

13.4

3.7

32.0

100

4. Conclusions

References

The BET surface area, microporous area and microprous volume of modified samples decreased pronouncedly in contrast to Parent50. Phosphorus and
boron species were highly dispersed over the HZSM5 (as suggested by XRD and SEM). The acidity of
treated samples (measured by FT-IR and TPD of ammonia) pronouncedly decreased. The isothermal type
of all samples is a complex of type I and IV, while
hysteresis loops belong to the H4 type. The heptane
cracking activity of a phosphorus or boron-modified
sample is lower than that of the parent zeolite. The
activity of steam-treated P5014 is higher than that
of only phosphorus-modified P501, while the activity
of steam-treated B5014 is remarkably enhanced compared to B501.

[1] Bertea L, Kouwenhoven H W, Prins R. Appl Catal A,


1995, 129(1): 229
[2] Kaeding W W, Butter S A. J Catal, 1980, 61(1): 155
[3] Vinek H, Rumplmayr G, Lercher J A. J Catal, 1989,
115(2): 291
[4] Ashton A G, Batmanian S, Dwyer J et al. J Catal,
1986, 34(1): 73
[5] Nunan J, Cronin J, Cunningham J. J Catal, 1984,
87(1): 77
[6] Young L B, Butter S A, Kaeding W W. J Catal, 1982,
76(2): 418
[7] Vedrine J C, Auroux A, Dejaifve P et al. J Catal,
1982, 73(1): 147
[8] Kaeding W W, Chu C, Ying L B et al. J Catal, 1981,
69(2): 392
[9] Chandawar K H, Kulkarni S B, Ratnasamy P. Appl
Catal, 1982, 4: 287
[10] Chen N Y, Kaeding W W, Dwyer F G. J Am Chem
Soc, 1979, 101(2022): 6783
[11] Li D, Tao L, Zhang Y et al. Shiyou Huagong
(Petrochem Technol), 1990, 19(3): 449
[12] Kaeding W W. J Catal, 1985, 95(2): 512
[13] Kaeding W W, Chu C, Ying L B et al. J Catal, 1981,
67(1): 159

Acknowledgements
Financial support from Catalytic Key Laboratory
of China Petroleum and Natural Gas Group Corporation (University of Petroleum) was greatly appreciated. We thank the National Science Foundation
Committee for Grant NSFC 20233030.

62

Renqing L
u et al./ Journal of Natural Gas Chemistry Vol. 12 No. 1 2003

[14] Kim J H, Namba S, Yashima T. Bull Chem Soc Jpn,


1988, 61(4): 1051
[15] Derouane E G, Dejaifve P, Gabelica Z et al. Chem
Soc, 1981, 72(2): 331
[16] Chen N Y. J Catal, 1988, 114(1): 17
[17] Dehertog W J H, Froment G F. Appl Catal, 1991,
71(1): 153
[18] Sun Y, Campbell S M, Lunsford J H et al. J Catal,
1993, 143(1): 32
[19] Lersch P, Bandermann F. Appl Catal, 1991, 75(1):
133
[20] Cavallaro S, Pino L, Tsiakaras P. Zeolites, 1987, 7(5):
408
[21] Giordano N, Pino L, Cavallaro S et al. Zeolites, 1987,
7(2): 131

[22] Suzuki K, Kiyozumi Y, Matsuzaki K. Appl Catal,


1988, 39(2): 315
[23] Romannikov V N, Tissler A T, Thome R. React Kinet
Catal Lett, 1993, 51(1): 125
[24] Richter M, Janchen J, Jerschkewitz H-G et al. J Chem
Soc Faraday Trans, 1991, 87(9): 1461
[25] Sing K S W, Everett D H, Haul R A W et al. Pure
Appl Chem, 1985, 57(4): 603
[26] Lago R M, Haag W O, Mikovsky R J et al. In: Murakami Y, Lijima A, Ward J W eds. New Developments in Zeolite Science and Technology: Proceedings
of the 7th International Zeolite Conference. Amsterdam: Elsevier, 1986. 677
[27] Sato K, Nishimura Y, Shimada H. Cata Lett, 1999,
60: 83

Вам также может понравиться