You are on page 1of 13

Applied Catalysis, 78 (1991) 213-225 Elsevier Science Publishers B.V.

, Amsterdam

213

Methylbutynol: a new and simple diagnostic tool for acidic and basic sites of solids
H. Lauren-Pernot*,F. Luck and J.M. Popa
RMne-Poukmc Recherches, Centre de Recherches dAuberviUiers, 52 rue de la Haie-Coq, 93308AuberviUiers (France), tel. (+33-1)49376262, fax. (+33-I)49376100 (Received 28 February 1991, revised manuscript received 29 July 1991)

Abstract In order to characterize acid-base properties of a material surface under gas-phase reaction conditions, a new test reaction has been investigated. We have shown that 2-methyl-3-butyn-2-olundergoes dehydration over solid acid materiah~and gives acetone and acetylene over solid basic materiak These reactions are very sensitive to the ueual modifications that can occur during cat&et preparation. Production of 3-hydroxy-3-methyl-2-butanone seems to be characteristic of amphoteric catalysta possessing acid-base ion pairs. Keywords: acid-base properties, 2-methyl-3-butyn-2-01 reactivity, catalyst characterization, magnesium oxide, alumina, zirconium oxide, silica-alumina, silica.

INTRODUCTION

The acidity or basicity of catalystsis important in determiningthe catalytic activity and selectivity, not only in typical acid-base reactions, but also in other reactionssuch as oxidation processes.Solid acid catalystsareextensively used for a varietyof industrialprocesses,such as catalyticcracking,alkylation, isomerizationand organic synthesis.By contrast, due to their limited number of commercial applications,basic and amphotericcatalystshave receivedless attention even though their characterizationin an area of great interest [l41. Besides these catalysts, new materialssuch as binary and ternary oxides, doped oxides, perovskites, etc., are studied increasingly,and require characterizationof their surfaceproperties. From a theoreticaland practical viewpoint,the importanceof acid and base heterogeneouscatalysts has led to numerous studies of the acidic and basic propertiesof surfaces.Severalclasses of methods can be used to determinethe type, strengthand amount of both types of sites,the characterization of surface basicity being far less studied than that of the surface acidity. The methods used to characterizeboth types of sites have been recentlyreviewed [ $61, and
0166-9834/91/$03.50 0 1991 Eleevier Science Publiehere B.V. Ail rights reserved.

214

can be classified in titration methods, spectroscopic methods using probe molecules and catalytic tests. Many of these techniques require the use of sophisticated equipment and sample preparation techniques. Some methods, such as vibrational spectroscopy, display only limited sensitivity, and may be unable to detect active surface centres present in trace amounts, often responsible for the catalytic activity of the sample. Others, such as aqueous titrations, may give erroneous results, the state of the surface of a solid catalyst in a water suspension being quite different from its state during a gas-phase reaction, for example. The use of Hammett indicators is highly controversial especially when used to give quantitative results. As there is no ideal universal method, catalytic tests which can be carried out in realistic conditions are of special interest. Isopropanol dehydration and dehydrogenation have been employed to probe respectively the acid and base properties of metal oxides at high temperatures. However, the assignment of the decomposition products to acid and base sites is somewhat ambiguous, since the activity for dehydrogenation to acetone is assumed to be related to the acidity and basicity (even to redox properties) of the solid [7]. We have thus sought a new probe molecule, able to give information about acid and base properties of solid catalysts, in conditions as close as possible to those of the most common catalytic reactions, i.e. sufficiently sensitive and selective in a wide range of temperatures. 2-Methyl-3-butyn-2-01 (MBOH), the simplest tertiary alkynol, is a versatile and readily available chemical, used in the manufacture of vitamins A and E via isomerization to prenal [8-l 11, isoprene, corrosion inhibitors and solvent stabilizers. It is a highly polar compound, soluble in water in all proportions, produced with high yields (ca. 90% ) via the ethynylation of acetone in liquid ammonia or polar solvents with catalytic amounts of alkali metals or hydroxides. A vapour phase synthesis method, based on a NaOH/clay catalyst has been claimed, with a conversion of 23% at 125 C [ 121. The lower yield may be due to the fact that base-catalyzed cleavage of methylbutynol becomes significant above 50 C [ 131. On the other hand, methylbutynol undergoes dehydration to yield 3-methyl-3-buten-1-yne (Mbyne) on acid catalysts such as acidic activated alumina, P205/Si02 and NH4H,P04/C [ 111. By-products such as 3hydroxy-3-methyl-2-butanone (HMB) and the corresponding dehydrated product 3-methyl-3-butene-2-one (MIPK) were found in the reaction of MBOH in concentrated sulfuric acid [lO,ll]. Hence, MBOH is expected to give different products when reacting on acidic or basic materials. This paper describes the overall reactivity of MBOH over solid acids and bases as well as preliminary results obtained over amphoteric materials; this reactivity is shown in Scheme 1.

215

acidic

Y-H

and 0 (Mbyne) 0 and (prenal)

HO -+-T (MBOH) (HMB) (MIPK)

reactivity

K 0

H-H

Scheme 1. Overall reaction scheme. EXPERIMENTAL

Materials
2-Methyl-3-butyn-2-01 ( > 99% ) was obtained from Fluka and used without further purification. Some data on the oxide powders used are summarized in Table 1. Aerosil A 200 silica was mixed with water to form a paste, which was calcined at 400C before grinding and compressing into cylindrical pellets of 3 mm in height and 5 mm in diameter. This treatment did not change the surface area of the silica. Rh8ne-Poulenc alumina SAS 350 is a commercial alumina obtained from gibbsite Al(OH)Q calcined and granulated (beads diameter = 2 mm). NaOH doped SAS 350 samples were prepared from SAS 350 by incipient wetness impregnation of the beads with a sodium hydroxide solution. CC 381 y-alumina is a laboratory sample, obtained by precipitation of sodium aluminate by carbon dioxide leading to an amorphous hydroxycarbonate, subsequently processed into y-alumina [ 141. Other commercial carriers were used as received. Each material was ground to 0.25-0.50 mm.

Pulse reactor system and procedure


The study was carried out in a fixed-bed, automated bench unit. Liquid MBOH was heated to 35 oC in a saturator system and the vapour was carried by a nitrogen flow. The reactant stream thus consisted of 4% MBOH in nitrogen [15]. In a typical experiment, the reactant stream was introduced into the reactor at a total rate of 4 l/h. A catalyst charge of 400 mg was supported in the centre section of a quartz tubular reactor. The catalysts were pretreated above 200 C

TABLE

1 of the oxide samples Supplier Composition (wt.-%) > 99.8% silica BET surface area (m/x) 299 Main impurities (wt.-%)

Characteristics Oxide sample

Aerosil2oO silica Magnesium oxide Zinc oxide Aerosil COK 84 Silica-alumina Aluminium oxide SAS 350 Aluminium oxide CC381 Zirconium oxide

Degussa

A1203, Fez03, TiO, < 0.005 Cl- < 0.025 CaO < 2; Fe,O, < 0.054; Cl- < 0.23 soa < 0.68

Cie des Salins du Midi Prolabo Degussa

>90.7% MgO

170

>99% 84% SiO, 16% Also, > 99.5% AlsO3

3.5 150

RhBne-Poulenc

340

NasO= 2799 ppm

RhBne-Poulenc

> 99% y-A1203

220

NaxO = 250 ppm

Stauffer

> 95% ZrO,

260

Cl- ~0.1; Tit0.S; Hfz 1

Six 1

(see the results tables) under nitrogen (4 l/h) during 2 h. The reactor temperature was then decreased to 130 or 150C and the reactant stream was admitted into the reactor for 5 min. This reaction temperature was adequate for the catalysts reported in this paper, but higher temperatures up to 250C can be used for less active materials. After the first 4 min, the products were analysed by an on-line gas chromatograph. During the analysis (40 min), the catalyst was kept at 130C under nitrogen. This procedurewas repeated twelve times in order to follow the deactivation occurring during the first pulses. Reactions products were analysed using a Hewlett-Packard Model 437 gas chromatograph equipped with a flame ionization detector. The packed column which contained 15% TCEPE on Silocel, effectively separated all the components which were identified by GC-MS and use of pure reference samples.

Physico-chemical analysis
Specific surface areas were measured on a Flowsorb 2300 Micromeritics by the BET method using nitrogen physisorption at 77 K. Samples were pretreated one night at 200C under air. Partial nitrogen pressure in the helium carrier was 20%.

217

Thermal analyses were carried out with a Mettler TA thermograph in a dry air stream. The temperature varied from room temperature to loo0 C with an average heating rate of 10 C/min. Infrared spectra were recorded with a Nicolet 7199 Fourier transform spectrometer. The samples (ca. 20 mg) were pressed into 18 mm diameter pellets and placed into an in situ IR vacuum cell with KBr windows. They were heated at 306 oC under vacuum for one night. IR spectra of adsorbed pyridine were taken after room temperature pyridine adsorption (pressure= 2 Torr, 266.6 Pa) and evacuation for one hour at the desired temperature. Calculation method As usual in catalysis, results are expressed in terms of conversion and selectivity. Ci is the molar concentration of compound i in the vapour phase and ( Cmou)o the initial molar concentration of MBOH. The total resulting products concentration is defined as: C=C CviCi with (Yi= 1 for all products, except for acetone and acetylene for which, in order to normalize conversion to 160%, oJi= l/2. Then, MBOH conversion (Conv% ) and selectivity for product i (Si% ) can be defined as follows: Conv% =
ci si= c -cm,,

(CMBoH)-C&4~OH~100 (CMBOH)

or acetylene and acetone Si%Z ci 2-(1-&B,,) This difference will

In most experiments C is slightly lower than (C,on)O. appear as C balance and will be discussed below. C balance= (C~on)oC Ci are obtained from the GC peaks.
RESULTS AND DISCUSSION

Evidence for acid-base selectivities


!&pical acidic or basic materials Table 2 shows the results obtained by reacting MBOH on typical basic or acidic catalysts and on pure silica. The values correspond to the conversion

218 TABLE 2 MBOH reaction on typical acidic or basic materials Activation temperature = 400C; reaction temperature = 180 C Sample Calc. temp. (C) Surface area (m2/g) Conv. (%) C balance (W) Selectivities (% ) C2H2 Acetone 52 50 0.5 48 50 0.5 Mbyne 0 0 90 Prenal 0 0 9

MgO ZnO Si02-A&O, Si02

400 400 400 400

100 3.5 150 200

70 20 25 0.5

3 1 1.5 -

%
100

.
80

l *ee.~*eeeeeeeee*eeeee~e**e*

60

1
0 0 -I

10

pulse

20 number

30

Fig. 1. Deactivation curve on silica-alumina. Activation temperature = 400C; reaction temperature=18OC; (m) Conv%, (m)Sprenal%, (+)SMbyne%.

reached after twelve pulses. The first pulses are generally accompanied with a non-negligible deactivation, but the selectivity remains constant during the whole series of pulses (Fig. 1) . Thus we can expect that the chemical nature of the active sites is not affected by deactivation. This behaviour is observed on all the materials tested. Deactivation can be due either to strong product adsorption on the catalyst surface or to secondary reactions leading to coke production on acidic sites [ 161. On basic sites, acetone condensation is a well known reaction [ 171 which can lead to heavy materials poisoning the catalyst. Our procedure (5 min pulses of MBOH and 40 min nitrogen steps during the chromatographic analysis) does not accelerate deactivation over silica-

219

alumina: the same result is found with our classical procedure ( 12 pulses) and after 59 min of continuous treatment with MBOH. On the contrary, over MgO deactivation is more rapid with the classical procedure, probably because the deactivation reaction proceeds during the 40 min step under nitrogen. Magnesium oxide is a well documented solid base catalyst. According to Hattori [ 181, evacuation at high temperature (460-1OCKP C ) leads to the appearance of three different types of basic sites, due to the removal of surface hydroxides and carbonates exposing 02- ions with different coordination numbers. The influence of calcination temperature on MgC reactivity towards MBOH is reported in the Section Reaction sensitivity to acid-base properties: effect of MgO calcination. Zinc oxide is presented here to show that a transition metal oxide can give a basic reactivity without any ambiguity related to its oxidative power: MBOH conversion in acetone and acetylene is not an oxidation reaction like isopropan01 conversion in acetone. Tested under the same conditions, silica-alumina COK 64 shows a Mbyne selectivity close to lOO%, acetone and acetylene being formed in minor quantities. Amorphous silica-alumina contains indeed a variety of strong Brransted and Lewis sites, as well as weakly acidic OH groups, and displays a fairly high acidity. Only 0.5% of MBOH is converted or adsorbed at 166C over Aerosil silica. This low reactivity can be directly related to the high purity of this silica, containing as major impurities Cl- (less than 250 ppm), A1203, Fe203 and Ti02 (less than 50 ppm of each). In fact, pure silica is an example of an inert surface. Acidity of silanol sites in silica gel is so weak, with pK, values in the range of 4 to 7, that the usual acid-catalyzed reactions do not proceed on Si02 [ 191. There are neither strong basic sites. In recent studies, it has been shown that silica adsorbs little to no pyridine [ 201, ammonia or carbon dioxide [ 211. These four examples allowed us to confirm the reaction pattern given in Scheme 1 for the conversion of MBOH over acidic or basic catalysts. The mechanisms we can propose to explain these results are given in Scheme 2. The temperature effect of the selectivity for both acidic and basic reactions must be known in order to use this reaction as a catalytic test. We have thus determined in non-diffusion limited conditions (Conv% < 25 obtained by dilution), the total order and apparent activation energy over MgC for basic reaction and silica-alumina for the acidic one. Best fit of the kinetic data was obtained for a total order of 0.3 5 0.1 for the acidic reaction and 0.3 2 0.1 for the basic one. We have found an apparent activation energy of 106 kJ/mol on MgO and of 75 kJ/mol on silica-alumina. Amphoteric materials Alumina is generally considered to be an amphoteric material. Kniizinger et al. [3] proposed that surface oxide ions and acidic OH groups are necessary

220

Basic

sites

Y I
0

H-H

0-

Acidic sites

3-H

+=-H H /

6
t
0 H

HGH I

Scheme 2. Basic and acidic reaction mechanieme. Acidic mechanism can be written either over Bran&d or Lewis acidic site.

221

for alkene production from alcohols, through a one step concerted E 2 elimination, whateverthe structureof the alcohol. The mechanismthus proposed would imply that alcohols ratherprovide a probe for hydroxyl-oxideion pairs on the alumina surface. Two different ahnninas containing respectively250 and 2700 ppm of NazO have thus been tested in MBOH reaction. Results are shown in Table 3. Both ahuninasgive the same new products: HMB and MIPK. Hydroxybutanone (HMB ) can be producedby hydrationof MBOH, waterbeingproduced either by formation of Mbyne or from the material itself. MIPK can be provided by dehydrationof butanone, hydration or Mbyne or by direct isomerization of MBOH [ 111. These reaction mechanismswill be investigatedin further studies.The most basic aluminais SAS 350 probably due to its relatively higher NazO content. A detailed study of the influence of sodium content on the reactivityof aluminawill be exposed in the following section. The same kind of reactivity (see Table 3 ) was obtained on zirconiumoxide for which butanone is the major product. As zirconiumoxide is also known to be amphoteric i.e. to possess acid-base ion pairs [22], butanone formation might reveal acid-base ion pairs on the materialsurface. Reaction sensitivity to acid-base properties We have seen that this reaction was able to determinethe acid-base propertiesof a surfaceand to distinguishbetweenthe major classesof catalysts.On the other hand it is well known that a catalyst surfacecan be slightlymodified by varyingpreparationconditions. These modifications are often of great interest for a certain catalytic application. Our purpose is thus to see whether the MBOH test reaction is sensitiveor not to some kind of modificationsthat can usuallyoccur on a catalyst surface.
TABLE 3 MBOH reaction on amphoteric materials Reaction temperature = 160 C Sample Activation temp. (Cl 250 250 400 Conv.
(56)

C balance
ml

ScIwz
(56)

S-,,m6)

%I@J)

SHMB BY

%lIPK WY-

cc 381 250 ppm Na,O SAS 350 _~ 2700 ppm NqO ZlO,b

7 9.5 8

3 5 3.5

22 41 2

23 43.5 3.5

16 4.5 18

0 4.5 76

39 6.5 0

See text. bFteaction temperature = 206 oC to increaee the conversion level.

222

Effect of MgO calcination MgO has been calcined at different temperatures between 430 and 1030 C and tested with MBOH at a reaction temperature of 150 C. This latter value was chosen to give moderate conversion, providing the opportunity to compare the different products. Table 4 shows texture properties and catalytic results obtained on these products. As expected, all samples provide a base type selectivity. Conversion is stable for a calcination temperature range of 430 to 730C and drops remarkably between 730 and 1030 C. In order to take into account the changes in specific area, the intrinsic activity has been calculated as follows: Activity (mmol/m TT.N, h) = 7 e

where TT is Conv-%, No is 6.9 mmol/h (initial quantity of MBOH) and A, ( m2) is the effective area introduced in the reactor. We can see from Table 4 that the activity increases with rising calcination temperature in the range of 530 to 1030 C. Examination of TGA-TDA results (see Fig. 2) obtained on the initial hydrated MgO sample shows that the decrease of specific surface area between 530 and 1030 C is correlated with the elimination of the most strongly bound carbonates. The basic sites thus revealed remarkably enhanced MgG activity towards MBOH. This effect can be due either to an increase of basic sites number or to an increase of basic sites strength. Further work will have to be done to elucidate this point. 11 $h.ence of sodium content on SAS 350 alumina reactivity We have seen above that MBOH reactivity on alumina seems to be very sensitive to the sodium content of the material. It is well known that sodium hydroxide impregnation of alumina increases its basicity [ 231. In order to investigate the versatility of our test reaction, we have prepared four sodium hydroxide impregnated aluminas from SAS 350. Na20 content, specific area and catalytic results observed with these samples are given in Table 5.
TABLE 4 Effect of MgO calcination temperature on catalytic results Calcination temperature ( C ) 430 530 730 1030 Surface area (d/g) 70 66 25 3 Conv (%) 31 27 32 11 C balance (%) 4 2 5 2
SC1H2 shtine

(%) 50 50 50 50

(%) 50 50 50 50

Activity ( mmol/m2 h ) 0.07 0.07 0.22 0.63

223 200 I physisorbed


4

I
I

300 I I H,O from

400 I
I

500 I
I

600 I I
l 4

700 I + TC

Mg(OH),

1 Ha0 from
Id

W
loss

H,O

I I I I
I
I I I

cap), I c--) co2


I

?I I f-)
I
I I I I
I

A-I I co,
I I I
I

co,
loss
III

CO,

1.95%

3.65%

1%

3.2%

weight loss

Fig. 2. TGA-TDA on MgO eample. TABLE 5 Influence of odium content on SAS 350 alumina reactivity Activation temperature = 250C; reaction temperature = 180C Na,O
(wt.-%)

S.A.
bZ/d

Conv
(%I

Cbahnce
(%a)

SCPH~ (%I

SA,,,, (%I

SM (%I

sHMB (%I

s MIPK (%I

0.27 0.66 1 4.1 8

295 290 274 260 238

9.5 12 14 100 100

5 4.5 5 0.5 0

41 47 40.5 50 50

43.5 47 40.5 50 50

4.5 1.2 0.1 0 0

4.5 4.6 2.8 0 0

6.5 0.2 0.1 0 0

The increase of the sodium content is accompanied by a poisoning of acid sites of alumina. Between 1 and 4% Na,O, the acidic reactivity is totally suppressed and from 4 to 8 wt.-% Na,O content, ahnninas exhibit a typical basic activity. This point was confirmed by pyridine infrared spectroscopy. Only Lewis acid sites are detected on these aluminas. Table 6 gives the intensity of the 1450 cm- band which is characteristic of Lewis bound pyridine for each desorption temperature. The maximum desorption temperature gives an acid strength evaluation. The 1450 cm- band intensity is correlated to the acid sites density From these results it is clear that an increase of sodium content reduces both the strength (maximum desorption temperature of pyridine) and the number (intensity of the 1450 cm- band) of alumina acid sites. A good correlation is found with the MBOH test. Nevertheless, pyridine, as a strongly basic molecule, is able to detect very weak acid sites on SAS 350+ 4% Na,O which is

i241.

224 TABLE 6 Infrared spectroscopy of pyridine adsorbed on ahminas Sample Desorption temperature ( C!) 100 200 300 100 200 100 200 100 200 100 1450 cm- intensity (Optical density/g) 17.1 7.3 2.1 12.9 4.3 15.3 2.7 1.3 0 0

SAS 350

SAS 350+0.66% Na,O SAS 350 + 1% NazO SAS 350+4.1% Na,O SAS 350 + 8% Na,O

inactive for MBOH dehydration. The same observation has been made with MgO for which pyridine is slightly adsorbed until 100C. Thus, these very weak acid sites cannot provide a typical acidic reactivity.
CONCLUSIONS

The preliminary results obtained in this study demonstrate that methylbutynol is a very sensitive probe molecule for simultaneous characterization of acidic and basic properties of solids. It allows a qualitative assay of the effect of the presence of impurities in catalysts or carriers, and can be used in a wide range of temperatures. Further work is necessary to know which kind of acid sites are responsible for the formation of respectively Mbyne, MIPK and HMB. We would expect that butanone is formed on an acid-base ion pair and that acid sites involved in MIPK formation are strong sites, since they are rapidly poisoned by NaOH impregnation. Additional work is in progress to determine whether this method is applicable to other oxide systems and to establish a two dimensional acido-basicity scale.
ACKNOWLEDGEMENTS

The authors acknowledge Rh&ne-Poulenc Recherches for permission to publish this work. We thank S. Morel and M. Arbeille for the test results, 0. Le Part for the BET measurements, B. Amram for the IR spectra and C. Poltz for the thermogravimetric analysis.

225

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

P.E. HathawayandME. Davis, J. Catal., 116 (1989) 263. H. Scheper, J.J. Berg-Slot and W.H.J. Stork, Appl. Catal., 54 (1989) 79. H. Kniizinger, H. BW and K. Kochloefl, J. Catal., 24 (1972) 57. E.J. Rode, P.E. Gee, L.N. Marquez and T. Uemura, Prepr. ACS Symp. Series, 34 (1989) 467. J. Kijenski and A. Baiker, Catal. Today, 5 (1989). K. Tanabe, in J.R. Anderzon and M. Boudart (Editors) Catalysis, Vol. 2, Springer, NewYork, 1981, p. 231. M. Ai, Bull. Chem. Sot. Jpn., 49 (1976) 1328. P. Chabardes, E. Kuntz and J. Varagnat, Tetrahedron, 33 (1977) 1775. P. Chabardes, Tetrahedron L&t., 23 (1988) 6253. A.V. Muzhegyen, R.Kh. Dzhulakyan, A.R. Tzagikyan, Kinet. Catal., 25 (1984) 63. E.D. Bergmann, J. Am. Chem. Sot., 73 (1951) 1218. A.I. Nogaideli, R.Sh. Tkeshelazhvilii and I.S. Khitarizhvilii, Zh. P&l. Khim. (Leningrad), 41 (1968) 1358. R.J. Tadezchi, Acetylene-Based Chemicals From Coal and Other Natural Resources, Chemical Industries Vol. 6, Marcel Dekker, New York & Bazel, 1982, p. 154. E. Trebillon, Fr. Pat., 2 520 722 (1982) to Rhhne-Poulenc. A.Z. Conner, P.J. Elving, J. Benizcheck, P.E. Tobiaz and S. Steingizer, Ind Eng. Chem., 42 (1950) 106. C. Chanmugathas end J. Heicklen, Int. J. Chem. Kinet., 17 (1985) 871. G. Zhang, H. Hattori and K. Tanabe, Appl. Catal., 40 (1988) 183. H. Hattori, in M. Che and G.C. Bond (Editors), Adsorption and Cat&& on Oxide Surfaces, Studies in Surface Science and Catalysis, Vol. 21, Elzevier, Amsterdam, 1985, p. 319. H.A. Benesi and B.H.C. Winquizt, Adv. Catal. Relat. Subjects, 27 (1978) 97. G. Connell and J.A. Dumesic, J. Catal., 105 (1987) 285. Y. Matsumura, K. Hazhimoto and S. Yozhida, J. Catal., 117 (1989) 135. Y. Nakano, T. Iizuka, H. Hattori and K. Tanabe, J. Catal., 57 (1979) 1. S. Siddhan and K. Naranayan, J. Catal., 68 (1981) 383. A.V. Kiselev and V.I. Lygin, Infra-red Spectra of Surface Compounds, Wiley, New York, 1975, Chapter 2.

You might also like