Sie sind auf Seite 1von 8

React. Kinet. Catal. Lett., Vol. 53, No.

2, 397-404 (1994)

RKCL2494

DEHYDRATION-DEHYDROGENATION OF 2-PROPANOL AS A MODEL


REACTION FOR ACID-BASE CHARACTERIZATION OF CATALYSTS

M.A. Aramendfa, V. Borau, C. Jim4nez, J.M. Marinas,


A. Porras and F.J. Urbano

Departamento de Qufmica Org~nica, Facultad de Ciencias,


Universidad de C6rdoba, Avda San Alberto Magno s/n,
E-14004 C6rdoba, Spain

Received March 31, 1994


Accepted April 20, 1994

The relationship between the acid-base properties of


various solids, as determined by a spectrophotomet-
ric procedure and the apparent dehydrogenation (KI)
and dehydration rate constant (K2) of 2-propanol
have been studied by using a c0ntinuous-flow reactor.

INTRODUCTION

The acidity and basicity of solid catalysts are two in-


fluential factors on their activity and selectivity, not only
in typical acid-base reactions, but also in many others in-
volving redox transformations. Acidic solids have been widely
used as catalysts in various industrial processes including
cracking, alkylation, isomerization and a variety of organic
syntheses. On the other hand, basic and amphoteric solids have
received comparatively much less attention even though they
have aroused growing interest lately.
There are a number of methods for characterizing acid and
base sites, most of which were recently reviewed [1,2]. In
broad terms, they can be classified as titration, spectroscopic
and reaction test methods. The determination of acid-base pro-

Akad4miai Kiad6, Budapest


ARAMEND~A et al.: 2-PROPANOL

perties of solid catalysts has been addressed by using a vari-


ety of model reactions involving t-butanol [3], methylbutynol
[4] or isopropyl alcohol [5,6] as the starting substrate. The
catalyst acidity is related to its ability to dehydrate iso-
propyl alcohol, whereas dehydrogenation is generally accepted
to require the presence of both acidic and basic sites. Some
authors have found correlations between the dehydrogenation-to-
dehydration rate ratio and the number of basic sites [6].
In this work we studied the potential correlation between
the acidity and basicity of various solid catalysts (Mg02,
Zr02, an 80:20 w/w SiO2/AIPO 4 mixed system called PM2, and se-
piolite, a natural support) as determined by a spectrophoto-
metric procedure previously developed by the authors [7] and
the dehydrogenation (k I) and dehydration constant (k2) of
2-propanol in a continuous-flow reactor over the temperature
range of 200-350 ~

EXPERIMENTAL

Catalysts
M a g n e s i u m oxide was prepared from Mg(OH) 2 (Merck ref.
5870) by calcination in a ceramic crucible (from room tempera-
ture to 600 ~ at 4 ~ and then at 600 ~ for 2 h), after
which the solid was allowed to cool back to room temperature.
The zirconium oxide was synthesized from zirconium oxy-
chloride (Merck ref. 8917) by preparing a zirconium hydroxygel
[8] and calcining it using the same temperature program as for
magnesium oxide.
Catalyst BM50 was obtained by suspending a mixture of 29.0
g of Mg(OH) 2 and 0.35 g of B203 in 200 mL of distilled water
and immersing it in an ultrasonic bath for 1 h. Then, the sus-
pension was dried in a stove at 120 ~ for 2 h and calcined
similarly as the previous two catalysts. The final Mg/B ratio
obtained was 50:1.
The procedure used to synthesize catalyst PM2 was de-
scribed in detail elsewhere [9].
Finally, the sepiolite was supplied by Tolsa S.A. and

398
ARAMEND~A et al.: 2-PROPANOL

drawn from the company's ores in Vallecas (Madrid, Spain). The


chemical analysis of the parent natural sepiolite (85 wt.%) was
as follows: 62.0% SiO2, 23.9% MgO, 1.7% A1203, 0.5% Fe203, 0.5%
CaO, 0.6% K20 and 0.3% Na20. The weight loss from 293 to 1273 K
was 10.5%. The oxides are parts of the clay structure (none is
in free form). The material contains smectite, quartz and dolo-
mite as major impurities. The sepiolite was designated by PS400,
where the subscript denotes the calcination temperature used
(in ~
Table 1 summarizes the chemical composition of the cata-
lysts tested, the precursors used in their synthesis, and their
calcination temperature, specific surface area (SBET) as de-
termined by means of a Micromeritics ASAP 200 instrument, acidi-
ty and basicity. The last two were determined by using a spec-
trophotometric procedure described elsewhere [7], which allows
the amount of irreversibly adsorbed acidic or basic compound
used as titrant for basic and acid sites, respectively, to be
titrated. The equilibrium monolayer coverage at 25 ~ X m (in
~mol g-l) was determined from the Langmuir adsorption isotherm
and was taken as a measure of acidic and basic sites correspond-
ing to the specific pK a of the base or acid used as titrant.

Table 1

Chemical nature, nomenclature, specific surface area,


acidity and basicity of the catalysts tested

Catalyst Precursor Tca I SBE T Acidity Basicity


(0C) (m2g -I (~mol g-l) (~mol g-l)

PM2 AICI3.6H20 650 402 21.7 3.4


BM50 Mg(OH)2/B203 600 104 2.8 98.4
PS400 Sepiolite 400 121 8%9 7.5
MgO Mg(OH) 2 600 119 2.0 172.7
ZrO 2 ZrOCI2.8H20 600 220 3.2 18.4

399
ARAMEND~A et al.: 2-PROPANOL

Experimental set-up

We used a c o n v e n t i o n a l continuous-flow fixed bed reactor


working at a t m o s p h e r i c pressure and constant temperature. The
reactor consisted of a piece of Pyrex glass of ii0 m m length
and 12 m m ID that was placed in a tubular electric furnace.
The reactor temperature was c o n t r o l l e d to w i t h i n • ~ through-
out the e x p e r i m e n t s by means of a t h e r m o c o u p l e attached to its
walls. Fresh c a t a l y s t was used in every experiment. The cata-
lyst bed was held b e t w e e n two layers of glass wool and the rest
of the reactor was p a c k e d w i t h glass beads in order to ensure
homogeneous mixing and complete vaporization prior to contact
with the catalyst bed. The p r o d u c t s emerging from the reactor
were t r a n s f e r r e d to a Sensorlab VG mass spectrometer that al-
lowed s i m u l t a n e o u s monitoring of 16 mass peaks and gave their
changes w i t h time. The selected masses were 45 (isopropyl al-
cohol), 58 (acetone), 41 (propene), 2 (hydrogen), 18 (water)
and 102 (di-isopropyl alcohol).
Catalytic runs were c a r r i e d out in the absence of diffu-
sional effects from the b o u n d a r y layer and internal or exter-
nal mass transfer processes by selecting appropriate operation-
al variable values, particularly as regards the space v e l o c i t y
(controlled through the catalyst w e i g h t and carrier gas flow-
rate) and particle size. The feed flow-rate was 12 mL h -I and
the carrier (nitrogen) gas flow-rate 120 mL min -I (both were
optimized in order to avoid d i f f u s i o n a l control).
In the absence of d i f f u s i o n a l effects, the c o n v e r s i o n of
2-propanol to propene and acetone fulfills the r e q u i r e m e n t of
the B a s s e t t - H a b g o o d kinetic treatment [101 for first-order
kinetic processes, viz.

in ( Ii~ ) = kT (i)

where X denotes the absolute conversion, T temperature (in K )


and k the a p p a r e n t dehydrogenation or d e h y d r a t i o n rate cons-
tant.

400
ARAMEND~A et al.: 2-PROPANOL

RESULTS AND DISCUSSION

Figure 1 shows the variation of the apparent dehydrogena-


tion and dehydration rate constants for isopropyl alcohol as a
function of time (a), as well as several reaction profiles for
catalyst PM2 at various temperatures (b).
As can be seen, the catalyst was deactivated in the pro-
cess, so all subsequent calculations of kinetic parameters were
done for a reaction time of 32 min (the first point plotted),
once the feed and carrier gas flow-rates had stabilized.
Table 2 gives the apparent dehydrogenation and dehydra-
tion rate constants obtained with various catalysts at differ-
ent temperatures.

Table 2

Apparent dehydrogenation (k I) and dehydration (k~) rate


constants (mol atm -I g-I s-l) for 2-propanol obtained with
various catalysts

T Catalyst

(OK) PM2 BM50 PS400 MgO ZrO 2

kI k2 kI k2 kI k2 kI k2 kI k2

473 0.20 3.65 0.21 1.02 0.23 2.17 0.ii 0.13 0.25 3.25
523 0.20 4.41 0.19 1.30 0.21 3.01 0.05 0.23 0.22 3.31
573 0.09 4.92 0.16 1.61 0.17 4.73 0.06 0.44 0.17 4.32
623 0.08 5.11 0.15 1.66 0.13 6.33 0.06 0.58 0.15 4.82

Table 3 shows the results obtained by correlating the a-


bove apparent rate constant values to the acid and basic site
density, and the temperature (fitted to the equation Y = AT +
+ BDac + CDba + E), as well as the corresponding percent over-
all significances and correlation coefficients.
As expected, there was good correlation between k 2 and
the acid site density, but not with the base site density.

401
A R A M E N D ~ A et al.: 2-PROPANOL

i i i i

Ca)

~3
r-
0

00 50 100 150 200 250


Time (rain)

60 I I I I

(b)
,, 250oC
o 300oC
v 350 ~ C
- 400 ~ C
4o
c-
o
_

k-

0
0

20

I I I I
50 100 150 200
~me (rain)

Fig. i. (a) Variation of the apparent dehydrogenation


(k I) and dehydration (k 2) rate constant of iso-
propyl alcohol with time. (b) Deactivation pro-
files for catalyst PM2 at four different temper-
atures

402
ARAMEND~A et al.: 2-PROPANOL

Table 3

Overall significance (Sover), regression coefficient (r) and


correlation coefficients for the apparent dehydrogenation (kI) and
dehydration (k 2) rate constants and the kl/k 2 ratio (R) as
determined from eq. (i), with temperature (T), acidic site
density (Dac) and basic site density (Dba). The figur@s in
parentheses are the corresponding partial significances

Y A B C E r S
over
(x 104 ) (x 102 ) (x 102 ) (x i0)

kI -7.9(21) -0.5(50) 1.3 (91) 0.7 0.89 98


k2 i01.0(50) 38.9(99) 3.3 (12) -4.1 0.96 99
R -8.6 (24) -1.9(30) 4.3 (95) 0.9 0.86 90

Also, the percent dehydration increased with increasing tempe-


rature. As regards kl, it was influenced by both acid and base
sites; however, the latter had a positive effect, whereas the
former influenced this constant adversely. In relation to R,
calculated as the kl/k 2 ratio, this parameter was more signifi-
cantly influenced by basic sites than was kl; therefore, R can
be used instead of k I to obtain a correlation with the catalyst
basicity, even though the resulting correlation coefficient
will be somewhat smaller.
It is interesting to emphasize the adverse influence of
the temperature on k I and R, which results in an increase in
the percent dehydrogenation with decreasing temperature.
The results provided by the catalysts tested allow them
to be used in hydrogen transfer reactions, one of our group's
goals in forthcoming work.

Acknowledgements. The authors gratefully acknowledge financial


support from the Consejerfa de Educaci6n y Ciencia de la Junta
de Andalucfa and the Spanish DGICyT for the realization of this
work in the framework of Project PB92-0816.

403
ARAMEND~A et al.: 2-PROPANOL

REFERENCES

I. J. Kijensi, A. Baiker: Catal. Today, 5 (1989).


2. K. Tanabe: in J.R. Anderson and M. Boudart (Eds), Cata-
lysis, Vol. 2, p. 231. Springer, New York 1981.
3. H. Grisebach, J.B. Moffat: J. Catal., 80, 350 (1983).
4. H. Lauron-Pernot, F. Luck, J.M. Popa: Appl. Catal., 78,
213 (1991).
5. A. Gervasini, A. Aroux: J. Catal., 131, 190 (1991).
6. M. Ai: Bull. Chem. Soc. Jpn., 50, 2579 (1977).
7. J.M. Campelo, A. Garcfa, J.M. Guti4rrez, D. Luna, J.M.
Marinas: Can. J. Chem., 6!i, 544 (1983).
8. K. Tanabe, M. Misono, Y. Ono, N. Naltori (Eds), Studies
in Surface Science and Catalysis, 51, 47 (1989).
9. M.A. Aramendfa, V. Borau, C. Jim4nez, J.M. Marinas, F.
Rodero: Coll. Surf., 12, 227 (1984).
10. D. Bassett, H.W. Habgood: J. Phys. Chem., 64, 769 (1960).

404

Das könnte Ihnen auch gefallen