Sunteți pe pagina 1din 17

Applied Catalysis A: General 256 (2003) 1935

Review

Acid catalysis by heteropoly acids


M.N. Timofeeva
Institute of Catalysis, Novosibirsk 630090, Russia Received 30 September 2002; received in revised form 23 January 2003; accepted 24 January 2003

Abstract TiIV , The achievements in the eld of acid catalysis by heteropoly acids having different structures Hx PW11 LO40 (L = ZrIV , ThIV ), Hx ZW12 O40 (Z = SiIV , PV , BIII , CoIII , GeIV ), H21 B3 W39 O132 , H6 P2 W18 O62 , H6 P2 W21 O71 , H6 As2 W21 O69 are reviewed. The data on the acidity of HPAs are generalized. The rules of homogeneous and heterogeneous acid catalysis by HPAs are discussed. 2003 Published by Elsevier B.V.

Keywords: Heteropoly acid; Keggin-type structure; Dawson-type structure; Acidity; Hammet acidity function; Acetone dimerization to mesitylene oxide; Acetic acid and n-butyl alcohol esterication; l-Sorbose acetonation

1. Introduction Heteropolyacids due to their unique physicochemical properties are widely used as homogeneous and heterogeneous acid and oxidation catalysts. They are also of great interest as model systems for studying fundamental problems of catalysis [13]. This is stipulated by the possibility of varying their acidity and oxidizing potential as well as by similarity of their catalytic effects in solution and in the solid state [13]. VI O , For catalysis, Keggin-type HPAs (H8x Xx M12 40 VI x V IV IV H8x+n X M12n Vn O40 , where X = Si , Ge , PV , AsV ; M = MoVI , WV ) are of importance (Fig. 1). The considerable number of studies performed during past 2025 years allowed to formulate the selection principles of effective catalysts in the series of Keggin-structure HPAs. Their signicantly higher Brnsted acidity, compared with the acidity of traditional mineral acid catalysts, is of great importance

E-mail address: timofeeva@catalysis.nsk.su (M.N. Timofeeva). 0926-860X/$ see front matter 2003 Published by Elsevier B.V. doi:10.1016/S0926-860X(03)00386-7

for catalysis [14]. Many new catalytic processes for basic and ne organic syntheses based on their employment have been developed [14]. In the future, the number of such processes will undoubtedly increase because HPA-based catalysts have higher activity than known traditional catalysts. Using HPA-based catalysts, it is frequently possible to obtain higher selectivity and successfully solve ecological problems. Nowadays, more than 100 heteropolyacids of different compositions and structures are known [57]. A great body of information concerning their synthesis methods and their structure has been obtained, however, only the Keggin-structure HPAs are well described in respect of their physicochemical and catalytic properties [14]. Probably, this situation is due to several reasons, one of which is a simple synthesis procedure of Keggin-structure HPAs compared with HPAs of other structures. Besides, in contrast to HPAs with Keggin-structure, heteropolyacids of other types are thermally less stable and therefore not capable of dehydrating at 150200 C and thus can not be employed for reactions, which

20

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

Fig. 1. The structures of heteropoly anions (M = MoVI , WVI ).

are conducted at more than 150 C. Nevertheless, in recent years more and more attention is drawn to obtaining new quantitative data for acid-catalytic properties of HPAs having other structures and compositions, mainly, H5 PW11 XO40 (X(IV) = TiIV , ZrIV , ThIV ), H6 As2 W21 O69 (H2 O), -H6 P2 W18 O62 , -H6 P2 Mo18 O62 , H6 P2 W21 O71 (H2 O)3 and H21 B3 W39 O132 (Fig. 1), in order to formulate certain selection rules for effective acid catalysts in the series of HPAs of the above structures. According to a simple electrostatic theory, these HPAs would be expected to be stronger Brnsted acids than the Keggin-type HPAs. The aim of this review is to discuss the scope and limitations of the above mentioned HPAs as the acid catalysts. The review mainly covers the literature published during the last 10 years, although, for a more comprehensive discussion, some earlier works are also included. More detailed information about the preparation of HPAs and their crystallographic data were comprehensively reviewed elsewhere [58].

2. Acidic properties of heteropoly acids 2.1. Acidic properties of heteropoly acids in solutions HPAs are well-known to be strong Brnsted acids but only recently their acidity has been quantitatively characterized and compared with the acidity of usual mineral acids. 2.1.1. The dissociation constants of HPAs A great deal of works is dedicated to the study of HPA acidity in solutions. However, data for the dissociation constants of HPAs are quite limited, mainly because of polyanions are labile in solutions. In aqueous solution, HPAs are completely dissociated at the rst three steps, the consecutive dissociation usually being unnoticeable because of leveling effect of the solvent [8,9]. According to the electroconductivity data, H4 PMo11 VO40 and H5 PMo10 V2 O40 are strong 1-4 and 1-5 electrolytes, respectively, in aqueous solution [10]. Using a potentiometric method

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935 Table 1 Dissociation constants of heteropoly acids and H3 PO4 in aqueous solution at 25 C H5 PMo10 V2 O40 [9] pK4 pK5 pK6 pK7 pK8 1.16 2.14 H6 PMo9 V3 O40 [9] 1.25 1.62 2.00 H7 PV12 O36 [8] 3.4 4.9 6.4 7.9 H8 NbMo12 O42 [10] 3.24 3.43 3.64 4.28 5.73 H8 CeMo12 O42 [10] 2.12 1.98 2.99 4.16 H8 UMo12 O42 [10] 2.13 3.02 4.31 H3 PO4 [10]

21

2.12 (pK1 ) 7.20 (pK2 ) 11.9 (pK3 )

the dissociation constants (starting from K4 ) of H5 PMo10 V2 O40 , H6 PMo9 V3 O40 , H7 PV12 O36 were measured in aqueous solution [8,9]. In the above acids, the rst three protons dissociate completely and the others do stepwise with increasing pH [9]. The dissociation constants are presented in Table 1. Note that the value of dissociation constants of HPAs does not vary considerably, while the distinctions between value of dissociation constants of inorganic acids are more considerable. For comparison, pKi for H3 PO4 are given in Table 1. One can see that all the studied HPAs are much stronger than H3 PO4 . This may be explained by taking into account a large size and low surface density of the polyanion charge as well as by the peculiarity of the HPAs proton structure [9]. Nonaqueous and mixed solvents exhibit differentiating effect on the acid dissociation constants of HPAs. Moreover, HPAs are considerably more stable in
Table 2 Dissociation constants of heteropoly acids in various solvents at 25 C Acid HOAc [12] pK1 H6 P2 W21 O71 (H2 O)3 H6 P2 W18 O62 H6 P2 Mo18 O62 H3 PMo12 O40 H4 SiW12 O40 H3 PW12 O40 H5 PW11 TiO40 H5 PW11 ZrO40 H3 PW11 ThO39 CF3 SO3 H HNO3 HClO4 HBr H2 SO4 HCl 4.66 4.39 4.36 4.68 4.87 4.70 5.32 5.45 5.48 4.97 4.87 5.60 7.00 8.40 CH3 CN [15] pK1 1.8 1.8 2.0 1.9 1.7 2.0 1.8 5.5 pK2 5.6 5.7 6.0 5.9 5.3 6.0 5.5 pK3 7.6 7.7 8.0 7.9 7.2 7.9 7.5

organic solutions [11]. In the works [1215] the acidity of HPAs in organic solutions was systematically studied. In Table 2, the dissociation constants in acetonitrile, acetone, ethanol and acetic acid are given in the form of pKi = log Ki , where Ki is the acid dissociation constant. For comparison, the dissociation constants of some inorganic acids are presented. First of all it is clearly seen that all the HPAs are stronger than the usual inorganic acids (HCl, H2 SO4 , HNO3 , HBr) and even such strong acids as HClO4 and CF3 SO3 H. This is of fundamental importance for the HPAs application in acid catalysis. It is noteworthy that sulphuric acid, which is a traditional acid catalyst, ranks 25 units of pK below HPA. The data given in Table 2 show that acidic properties of HPAs in CH3 CN, C2 H5 OH and (CH3 )2 CO do not vary signicantly depending on the HPAs composition and structure. Acidity series differ for each solvent.

(CH3 )2 CO [13,15] pK1 2.0 2.0 1.6 1.7 2.0 2.7 3.6 4.0 pK2 3.6 3.6 3.0 3.2 3.4 pK3 5.3 5.3 4.1 4.2 5.2

C2 H5 OH [14] pK1 1.8 2.0 1.6 3.6 pK2 3.4 4.0 3.0 pK3 5.3 6.3 4.1

22

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

Thus, in CH3 CN, the acidity of W-containing HPAs changes in the order: H3 PW12 O40 > H5 PW11 ZrO40 > H6 P2 W21 O71 (H2 O)3 > -H6 P2 W18 O62 , H6 As2 W21 O69 (H2 O) > H4 SiW12 O40 > H5 PW11 TiO40 At the same time, in acetone the acidity series falls in the order: H3 PW12 O40 , H5 PW11 TiO40 > H4 PW11 VO40 > H4 SiW12 O40 > H5 PW11 ZrO40 H3 PMo12 O40 > H4 SiMo12 O40 The HPA composition has only a slight effect on their acidity, nevertheless, this effect can be seen. The acidity decreases with the reduction of HPA and replacement of MoVI or WVI atom by VV atom and/or the replacement of the central PV atom by SiIV . In all cases, this decrease in acidity must occur with an increase in basicity of the acid (the number of equivalent protons connected with heteropolyanion). The more pronounced acidity dependence on the composition and structure of HPA is observed in acetic acid solution [12]. For the W-containing HPAs, the acidity changes in the following series in accordance with the anion structure: Dawson > H6 P2 W21 O71 (H2 O)3 > Keggin This correlates well with acidity of reducing forms of the Dawson- and Keggin-type HPAs in solutions [16]. For the Keggin-type HPAs, the dissociation constants in AcOH decrease in the series [12]: H3 PW12 O40 > H4 SiW12 O40 > H5 PW11 TiO40 > H5 PW11 ZrO40 H3 PW11 ThO39 . The replacement of WVI by MoVI does not inuence the value of the HPA dissociation constants, while the replacement of WVI by TiIV , ZrIV or ThIV must result in the decrease of the dissociation constant according to the electrostatic theory. The basicity of the solvent substantially inuences the acidity as well as the extent of acidic dissociation (Table 2). Thus in the less polar HOAc, all the HPAs are comparatively weak monobasic acids. In such polar solvents as CH3 CN, C2 H5 OH and

(CH3 )2 CO, the Keggin-structure HPAs are completely dissociated at the rst step and partially at the second one. Such acids as H3 PW12 O40 and H5 PW11 TiO40 are completely dissociated at the rst, second and partially at the third steps in acetone. For HPAs of other structures data acquisition in some polar solvents is impossible due to an intricate dependence of electroconductivity on acid concentration in solution [15]. However, it can be supposed that their acidity will be close to the acidity of H3 PW12 O40 . 2.1.2. The Hammet acidity function H0 In solution, the acidity of HPAs is characterized quite well by the Hammet acidity function H0 [17,18]. It has been established that the series of the HPA strength in diluted and concentrated aqueous solutions do not coincide with each other. The strength of HPAs in diluted solutions ([HPA] < 0.05 mol/l) decreases in the series [18]: H21 B3 W39 O132 > H5 PW11 TiO40 > H5 PW11 ZrO40 > H6 P2 W21 O71 > H3 PW12 O40 > H4 SiW12 O40 At the same time, the acidity series for concentrated solutions ([HPA] > 0.05 mol/l) decreases in the order [18]: H21 B3 W39 O132 > H6 P2 W21 O71 > H5 PW11 TiO40 > H3 PW12 O40 H4 SiW12 O40 > H5 PW11 ZrO40 This difference is probably due to the fact that acidity of concentrated solutions is determined not only by the values of the acid dissociation constants (as it takes place in the case of diluted acids) but also by the salt effects which depend on the composition and structure of the HPAs [17]. The H0 values obtained in aqueous acid solutions with the same proton concentration differ slightly (Fig. 2, Table 3). This indicates that all the HPAs have similar acid strength, close to the acidity of HClO4 and CF3 SO3 H, and even stronger. The equimolar aqueous solutions of H21 B3 W39 O132 and H6 P2 W21 O71 have higher acidity than Keggin-structure HPAs, HClO4 and CF3 SO3 H. Probably, that is because they have more protons in their composition. The difference in the acidity of aqueous solutions of the HPAs

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

23

Fig. 2. H0 for aqueous solutions of HPAs at 20 C [15].

disappears at the same weight concentration of the HPAs (Table 3). The differences in acidity become more obvious in waterorganic solutions. Thus, in 90% aqueous acetone the following acidity series is observed (acidity is based per proton) [18]: H6 P2 W21 O71 > H21 B3 W39 O132 > H5 PW11 TiO40 , H5 PW11 ZrO40 > H3 PW12 O40 , H4 SiW12 O40

In 90% acetonitrile [18] and 85% HOAc [12], the acidity series are different: H6 P2 W21 O71 > H3 PW12 O40 , H5 PW11 ZrO40 , H5 PW11 TiO40 > H21 B3 W39 O132 > H4 SiW12 O40 and H6 P2 W21 O71 (H2 O)3 > H4 SiW12 O40 H3 PW12 O40 > H5 PW11 TiO40 > H3 PW11 ThO39 H5 PW11 ZrO40 , respectively. It is obvious that the acidic series in waterorganic solvents does not correlate with the thermodynamic dissociation con-

Table 3 Hammet acidity function H0 values of heteropoly acids in different solution Acid H2 O [18] H0 (0.1 mol/l) H21 B3 W39 O132 H6 P2 W21 O71 H5 PW11 TiO40 H3 PW12 O40 H5 PW11 ZrO40 H4 SiW12 O40 HClO4 CF3 SO3 H H2 SO4
a b a

(CH3 )2 CO (90%) [18] H0 (0.3 mol/l) +0.21 +0.15 +0.01 0.05 +0.03 +0.06 +0.36 +0.25
b

CH3 CN (90%) [18] H0 (0.05 mol/l) 0.55 0.58 +0.06 +0.14 0.03 +0.46 +0.86 +0.77
a

HOAc (85%) [12] H0 a (0.05 mol/l) 0.27 +0.19 +0.10 +0.79 0.07 H0 b (0.20 mol/l) 0.20 +0.40 0.17 +1.06 0.04 +0.75

H0 (288.2 g/l) +0.11 +0.04 0.06 0.05 0.05 0.03 0.71 0.83

H0 (0.05 mol/l) +0.21 +0.11 +1.31 +2.17 +1.43 +1.36 +2.32 +2.00

H0 (0.15 mol/l) +1.55 +0.38 +1.77 +2.17 +2.02 +2.13 +1.80 +1.66

H0 (0.15 mol/l) +0.35 0.36 +0.17 +0.14 +0.20 +0.60 +0.66 +0.62

0.52 0.40 0.18 0.05 0.07 0.03 +0.81 +0.56

The molar concentration of acid. The proton concentration of acid. c The weight concentration of acid.

24

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

stants in nonaqueous organic media. This shows that in waterorganic solutions HPAs become polyelectrolytes of a level higher than two and essentially differ from other acids [18]. The acid strength decreases when WVI atom is replaced for MoVI or VV and/or when the central PV atom is replaced for SiIV . The effect of the central atom has been demonstrated for acetonitrile solutions of the Keggin-type heteropolytungstates [19,20]. It has been established that in general the acidity increases with a decrease in the negative charge of the heteropolyanion or an increase in the charge of the central atom: PV > SiIV , GeIV > BIII > CoIII It has been found [18] that the acidity of waterorganic solutions of Keggin HPAs depends inversely on the basicity of the organic solvent. The solvent basicity decreases in the series: acetone > acetonitrile > water. The acidity of the solutions of H3 PW12 O40 , H4 SiW12 O40 , H5 PW11 TiO40 and H5 PW11 ZrO40 changes in the reverse order. This does not occur for the solutions of H6 P2 W21 O71 and H21 B3 W39 O132 most likely due to the inuence of the anion structure of HPAs. The effect of water amount on the acidity of waterorganic solutions has an extreme character, which is a common phenomenon for mixed solutions. The acidity decreases sharply upon addition of water, then goes through a minimum and increases with further increase of water concentration [17]. Such character of dependence is supposed to be due to the inuence of the solvent on the activity coefcient of reagents [14]. 2.1.3. The basicity of heteropolyanions Recently, the acidity of HPAs has been studied by using various physicochemical methods, including comparison of the basicity of heteropolyanons. These data were gained from 1 H NMR study of complexation between the anion and the geminal diol, 1,1-dihydroxy-2,2,2-trichloroethane (chloral hydrate), in organic medium [2123].This reaction is assumed

anion (Fig. 1). The chemical shifts of the hydroxyl protons of chloral hydrate increase in parallel with the increase in the basicity of the anion or, in other words, with the decrease in the acidity of the corresponding HPA. The application of this method for the Keggin HPAs resulted in the following acidity series: H3 PW12 O40 > H3 PMo12 O40 > H4 SiW12 O40 H4 GeW12 O40 > H4 SiMo12 O40 > H4 GeW12 O40 In general, this order is in agreement with that obtained using indicator tests. HPAs of different structural types (Dawson, H6 P2 W21 O71 (H2 O)3 , H6 As2 W21 O69 (H2 O), H21 B3 W39 O132 ) were studied in [23]. Since heteropoly anions of different structures differ substantially in size and are potentially capable of coordinating dissimilar numbers of chloral hydrate molecules, the authors decided that the data presented in Fig. 3, where the 1 HCH values are divided by the number of WVI or MoVI atoms in the anion, are more illustrative. In this respect, heteropoly anions are divided into three groups: 1S2 Mo18 O62 4 , PW12 O40 3 , PMo12 O40 3 ; 2P2 W21 O71 (H2 O)6 , SiW12 O40 4 , As2 W21 O69 (H2 O)6 , B3 W39 O132 21 ; 3P2 Mo18 O62 6 , P2 W18 O62 6 . The basicity of heteropoly anions increases in the following order: S2 Mo18 O62 4 , PW12 O40 3 , PMo12 O40 3 < P2 W21 O71 (H2 O)6 , SiW12 O40 4 , As2 W21 O69 (H2 O)6 , B3 W39 O132 21 < P2 Mo18 O62 6 , P2 W18 O62 6 The specic charge of one atom of WVI or MoVI changes in the same order. Thus, although the surface of the HPA anions is inhomogeneous, the foregoing suggestion generally conrms the correlation found previously for the HPA anions of the Keggin-structure, namely, with a decrease in the anion charge, the anion basicity decreases and, hence, the acidity of the corresponding HPA increases [24]. 2.2. Acidic properties of solid heteropolyacids The strength and the number of acid centers as well as related properties of heteropolyacids can be controlled by the structure and composition of het-

to yield complexes having two hydrogen bonds between bridging (Ob ) oxygen atoms of the heteropoly

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

25

Fig. 3. Variations of 1 HCH related to the number of the WVI or MoVI atoms in the heteropoly anions [22].

eropolyanions, extent of hydration, type of support, thermal pretreatment, etc. Solid heteropolyacids, such as Hx ZM12 O40 (Z = PV , SiIV , GeIV , AlIII ; M = MoVI , WVI ), Dawson H6 PM18 O62 (M = MoVI , WVI ), H6 As2 W21 O69 (H2 O), H6 P2 W21 O71 (H2 O)3 , are pure Brnsted acids [25] and are stronger than conventional solid acids such as SiO2 Al2 O3 . According to the indicator test,

H3 PW12 O40 has a Hammet acidity function less than 8.2 [26] and it has been suggested to be a superacid [27,28]. A superacid is an acid with a strength greater than that of 100% H2 SO4 , i.e. a value of H0 < 12 [29,30]. Thermal desorption of basic molecules also reveals the acidic properties. At the same time, heteropoly acids such as H5 PW11 TiO40 , H3 PW11 ThO39 and H5 PW11 ZrO40

Fig. 4. IR-spectra of pyridine adsorbed on the surfaces of H5 PW11 ZrO40 (1) and H5 PW11 TiO40 (2) [30].

26

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

Table 4 Average PA values for protons of heteropoly acids substituted by the pyridinium ion [25] HPA PA (kJ/mol) n(H+ )a Ib H3 PW12 O40 -H6 P2 Mo18 O62 CF3 SO3 H H6 P2 W21 O71 H4 SiW12 O40 H3 PMo12 O40 H4 SiMo12 O40 H3 PW6 Mo6 O40 H4 PW11 VO40 H21 B3 W39 O132 HClO4 H5 PW11 TiO40 -H6 P2 W18 O62 H6 As2 W21 O69 H5 PW11 ZrO40 H-ZSM-5 CF3 COOH
a b

Keggin HPAs decrease in the series: H3 PW12 O40 > H4 SiW12 O40 , H3 PMo12 O40 > H4 SiMo12 O40 , H3 PW6 Mo6 O40 > H4 PW11 VO40 > H5 PW11 ZrO40 According to this, the acidity of the HPAs decreases with an increase in the anion charge. For the Keggin HPA, W-HPAs are stronger acids than Mo-HPAs, whereas for the Dawson HPA -H6 P2 M18 O62 (M = MoVI , WVI ) this dependence is reverse, which is also conrmed by the method of complex formation with chloral hydrate [23]. The acidic strength of heteropolyacids has been determined more quantitatively by calorimetry of NH3 adsorption [32,35]. It has been determined that after treatment at 423 K the initial heats of NH3 adsorption are as follows: 196, 185, 164 and 156 kJ/mol for H3 PW12 O40 , H4 SiW12 O40 , H6 P2 W21 O71 (H2 O)3 and H6 P2 W18 O62 , respectively. These values show that acidity decreases in the series: H3 PW12 O40 > H4 SiW12 O40 > H6 P2 W21 O71 (H2 O)3 > H6 P2 W18 O62 These data also indicate that the Keggin-type heteropoly acids are much stronger acids than the Dawson-type HPAs. Furthermore, the heats of the NH3 adsorption conrm that HPAs are stronger acids than zeolites or simple metal oxides: 150 and 140 kJ/mol for H-ZSM-5 and -Al2 O3 , respectively. An increased interest in heterogenization of HPA on a support is caused by a superior activity revealed for these systems compared to that of bulk HPAs [3638]. Moreover, the supported systems are of practical importance because the catalytic activity is determined by the catalyst surface area for some HPA-catalyzed reactions. Many porous materials were used as supports for HPAs. A variety of methods has been used to characterize HPAs supported on SiO2 , carbon, Al2 O3 , MgF2 [3742]. 31 P NMR [43] and FT-IR data on SiO2 , SiO2 Al2 O3 and Al2 O3 impregnated with H3 PW12 O40 [44] show that the primary Kegginstructure is preserved but some degradation can be observed with Al2 O3 heated above 200 C. Kapustin et al. [45] have found that the acid strength of supported H3 PW12 O40 decreases accordH5 PW11 TiO40

IIc 3.0 6.0 6.1 4.1 3.0 4.1 3.0 4.1 15 4.0 6.0 6.1 3.5

1070 1110 1120 1130 1140 1140 1150 1150 1170 1180 1180 1190 1200 1200 1210 1200 1350

4.5 4.4 1.1 5.6 3.2 2.4 5.4 3.6 5.5 10 1.07 4.9 6.5 2.4 6.1

Number of H+ ions substituted by PyH+ . IR spectroscopy. c Weight method.

are both Brnsted and Lewis acids [31]. Fig. 4 shows the IR-spectra of pyridine adsorbed on H5 PW11 ZrO40 and H5 PW11 TiO40 . In the spectra of all the HPAs the absorption bands at 1540 cm1 attributed to Brnsted acid sites and the absorption bands at 1450 cm1 attributed to Lewis acid sites are observed. The intensity of this band for H5 PW11 ZrO40 is 2.5 times higher than that for H5 PW11 TiO40 . It is likely that in acidic reactions Lewis sites as well as Brnsted acid sites signicantly contribute to the catalytic activity of H5 PW11 ZO40 (Z = ZrIV , TiIV , ThIV ) [32]. The strengths of heteropolyacids on the proton afnity scale have been determined by IR spectroscopy of the pyridinium salts [25]. Table 4 demonstrates the PA (proton afnity) values of several HPAs. All the HPAs are strong acids like CF3 SO3 H and HClO4 and stronger than zeolites and sulphated alumina. The structure of the HPA anion has almost no effect on the HPA acidity. H3 PW12 O40 is the strongest acid, which agrees well with the data obtained for the solid HPA and their solutions by other methods [12,18,3134]. PA values of the similar

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

27

ing to the series SiO2 > Al2 O3 > carbon, which allowed them to assume a signicant interaction between HPA and carbon. The nature of the interaction between HPA and carbon depends mainly on the concentration of the impregnating solutions, solvent used and pH of the solution. Thus, when H3 PW12 O40 and H4 SiW12 O40 are supported from concentrated aqueous solutions on chemically activated carbon, they retain their Keggin-structure, and the acids are highly dispersed on the support. At the same time, if the HPA content is low, H3 PW12 O40 is partly decomposed [46]. The results of HPA acidity measurement by two different methods (indicator test and the calorimetry of NH3 adsorption) are compared in [47]. The order of the acid strength of unsupported HPAs measured by the indicator test is: H3 PW12 O40 > H4 SiW12 O40 > H4 GeW12 O40 , H6 P2 W18 O62 > H5 BW12 O40 > H6 CoW12 O40 The order obtained by the calorimetry of NH3 adsorption, H3 PW12 O40 > H4 SiW12 O40 > H6 P2 W18 O62 , is consistent with the series above. An interesting fact is that the acid strength of H6 P2 W18 O62 was almost unchanged upon supporting on SiO2 . The acidic strength of 20% H6 P2 W18 O62 /SiO2 and 20% H3 PW12 O40 /SiO2 became closer. This result seems to be quite strange because (see below) upon supporting H6 P2 W18 O62 on SiO2 as well as on the carbon, the strength of acidic centers decreases due to strong interaction with the supporting groups of the carrier [48]. It is likely that such an unusual result is caused by the high temperature of pre-activation of H6 P2 W18 O62 /SiO2 (200 C under vacuum), which results in isomerization and partial destroying of H6 P2 W18 O62 .

3.1. Homogeneous acid-catalyzed reactions The mechanism of homogeneous catalysis by heteropoly acids is in principle similar to the mechanism of catalysis by solutions of inorganic acids. However, heteropolyacids are capable of protonating the substrate and activating it for subsequent chemical reactions more effectively than usual inorganic acids. According to physicochemical data, we may expect that acids with different structures and compositions will differ slightly in the catalytic activity in homogeneous acid-catalyzed reactions in water. Considerable distinctions can be observed in organic solvents. The activity series may differ in different solvents or for different substrates like it was observed for the HPA-catalyzed ether decomposition [24]. Examples of relationship between the catalytic activity of HPA and their dissociation constants (the Brnsted equation) are well precedented [17,19,20,31,49]. It is well known that in concentrated aqueous solutions of heteropolyacids the reaction rate constant is dependent on the Hammett acidity function H0 [49]. Thus the rate constant for hydration of isobutene in aqueous solution in the presence of H3 PW12 O40 and H4 SiW12 O40 obey the equation: log(k) = 1.04 H0 3.46. The same relation takes place for H2 SO4 , HCl, HNO3 and HClO4 . Given this fact in mind, it was suggested that the hydration of isobutylene in the presence of the HPAs and inorganic acids proceeds via the generally accepted mechanism. However, it turned out that not only concentration but also the kind of HPA inuences the catalytic activity of HPA for isobutene hydration, which decreases in the series [50]: H3 PW12 O40 > H3 PMo12 O40 > H4 SiW12 O40 = H4 GeW12 O40 > H4 SiMo12 O40 > H4 GeMo12 O40 This order correlates with the relative basicity of conjugated anions. Judging from the relative basicity of the conjugated anions, the strength of H3 PW12 O40 is almost comparable with that of HClO4 , however, the catalytic activity of the HPA is much higher owing to the nature of the heteropoly anions [50,24]. Thus Izumi et. al. [24] observed that H4 SiW12 O40 is more active for the reaction of dibutyl ether with acetic anhydride than H3 PW12 O40 , despite the fact that H4 SiW12 O40 is a weaker acid compared to

3. Acid catalysis by heteropoly acids In this part of the review the following questions will be discussed: (a) comparison of catalytic activity of HPAs having various composition and structure; (b) relationship between the HPAs activity and their acidity. Table 5 gives examples of reactions catalyzed by both the Keggin HPAs and HPAs of other structures.

28

Table 5 Acid-catalyzed reactions in the presence of heteropoly acids Reaction Homogeneous reactions Condensation: 2CH3 COCH3 CH3 COCH=C(CH3 )2 Acetonation of l-sorbose Catalyst H6 P2 W21 O71 , Hx ZW12 O40 (Z = SiIV , PV ), Hx PW11 LO40 (L = ZrIV , TiIV , ThIV ) H6 P2 W21 O71 , H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV ), Hx PW11 LO40 (L = ZrIV , TiIV , ThIV ) Ref. [18] M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935 [55]

Esterication: BuOH + HOAc BuOAc + H2 O

H21 B3 W39 O132 , H6 P2 W18 O62 , H6 P2 W21 O71 , H6 As2 W21 O69 , Hx ZW12 O40 (Z = SiIV , PV ), Hx PW11 LO40 (L = ZrIV , TiIV , ThIV ) H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV , BIII , CoIII , GeIV ) H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV , GeIV ) Hx ZM12 O40 (Z = SiIV , PV ; M = MoVI , WVI ) H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV , GeIV )

[31]

CH3 CH2 COOH + iso-C4 H9 OH iso-C4 H9 OCOCH3 + H2 O iso-C4 H9 OCOC2 H5 iso-C4 H9 OH + C2 H5 COOH iso-C4 H9 OCOC2 H5 +CH3 COOH iso-C4 H9 OCOCH3 + C2 H5 COOH Higher aliphatic diols cyclic ethers

[52] [52] [74,75] [54]

(CH3 )2 C=CH2 + CH3 OH (CH3 )3 COCH3 Biphasic reactions Condensation:

H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV ) H6 P2 W21 O71 , H6 P2 M18 O62 (M = MoVI , WVI ) Hx ZW12 O40 (Z = Si, P), Hx PW11 LO40 (L = ZrIV , TiIV , ThIV ) Hx ZM12 O40 (Z = SiIV , PV ; M = MoVI , WVI )

[64] [56,57]

[70,71]

Acetoxylation:

Hx ZW12 O40 (Z = SiIV , PV )

[72]

Alkylation

H6 P2 W21 O71 , H6 P2 W18 O62 , H3 PW12 O40

[59]

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

Heterogeneous reactions Alkylation: iso-C4 H10 + iso-C4 H8 fractions C5 C8

H6 P2 W18 O62 , H6 P2 W18 O62 /SiO2 H3 PW12 O40 , -H6 P2 W18 O62 , H6 P2 W21 O71 (H2 O)3 , H6 P2 W18 O62 /SiO2 , H6 P2 W18 O62 /carbon

[60] [48]

Condensation:

H6 P2 W21 O71 , H6 P2 W18 O62 , Hx ZW12 O40 (Z = SiIV , PV ), Hx PW11 LO40 (L = ZrIV , TiIV )

[73]

Higher aliphatic diols cyclic ethers 2-iso-C4 H8 dimer and trimer of isobutylene Esterication: iso-C4 H8 + CH3 OH iso-C4 H9 OCH3 Dehydration: CH3 OH CH4 + H2 O

Hx ZM12 O40 (Z = SiIV , PV ; M = MoVI , WVI ) H6 P2 W21 O71 , H6 P2 W18 O62 , H6 As2 W21 O69 H3 PW12 O40 , Hx PW11 LO40 (L = ZrIV , TiIV ) Hx ZW12 O40 (Z = SiIV , PV , BIII , CoIII ), H6 P2 W18 O62 , H6 P2 W18 O62 /SiO2 Hx ZW12 O40 (Z = SiIV , PV , BIII ), H21 B3 W39 O132

[74,75] [65] [47,61] [62]

29

30

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

H3 PW12 O40 . This difference was explained by stabilization of the cationic intermediate via the formation of an intermediate polyanion complex due to the softness of polyanion, which falls the order: SiW12 O40 4 > PW12 O40 3 [51]. The effect of the softness becomes signicant for reactions in aqueous solutions, in which the difference in the acid strength is less pronounced since most HPAs are completely dissociated. It was shown that W-containing heteropolyacids of the Keggin-type demonstrate remarkably high catalytic activity [52]. Thus for the liquid-phase homogeneous decomposition of isobutyl propionate, HPAs appeared to be 60100 times more active than H2 SO4 and para-toluenesulfonic acid. The catalytic activity for this reaction increases when the central atom charge increases. This can be attributed to an increase in the acid strength of the solution. It rises with a diminution of the negative charge of the polyanions. For reactions involving basic molecules such as alcohols (ester exchange of isobutyl propionate with n-propyl alcohol and esterication of propionic acid with isobutyl alcohol), no signicant differences in the activity among these catalysts, including H2 SO4 and para-toluenesulfonic acid, are observed, suggesting that these catalysts are equally strong in solution due to the leveling effect of the basic reactant [52]. In the reaction of acetone dimerization a correlation between the catalytic activities of HPAs and H0 has been found (Fig. 5) [18]. The catalytic activity de-

creased in the series: H6 P2 W21 O71 > H5 PW11 TiO40 H5 PW11 ZrO40 > H4 SiW12 O40 H3 PW12 O40 The activity of CF3 SO3 H was higher than that of the Keggin-type heteropolyacids but lower than that of H6 P2 W21 O71 . For HPAs and CF3 SO3 H, the H0 value decreased in the similar order. In the esterication reaction of acetic acid and n-butyl alcohol the catalytic activity of HPAs has a good correlation with the dissociation constant of heteropolyacids [31]. The catalytic activity of tungsten-containing heteropolyacids depends on the structural type and decreases in the series: H21 B3 W39 O132 > H6 P2 W18 O62 H6 P2 W21 O71 > H6 As2 W21 O69 > Keggin-type For the heteropoly acids of the Keggin-structure, the activity decreases in the order: H5 PW11 ZrO40 > H4 SiW12 O40 H3 PW12 O40 > H5 PW11 TiO40 H3 PW11 ThO39 The higher catalytic activity of H5 PW11 ZrO40 may be explained by taking into account not only Brnsted acidity but also Lewis acidity. Besides, it has been established that in the presence of H6 P2 W21 O71 (H2 O)3 the reaction rate substantially depends on the order of adding the reagents. HPA added immediately after mixing all the reagents (HOAc, H2 O, n-BuOH) is of higher catalytic activity than HPAs pre-maintained during 14 h in aqueous acetic acid medium. For HPAs of other structures such dependence has not been observed. On the basis of NMR 183 W data (Fig. 6) and kinetic measurements, it could be supposed that in pure acetic acid two of three water molecules, incorporated into heteropolyanion and located on its surface (Fig. 6), are replaced for HOAc molecules to form HPA: H6 P2 W21 O71 (H2 O)2 (AcOH) ( 31 P 12.5 ppm) and H6 P2 W21 O71 (H2 O)(AcOH)2 ( 31 P 11.6 ppm). The third water molecule is located inside the HPA anion and is not capable of such exchange. These reactions are reversible, since the intensity of signals at 12.5 and 11.6 ppm decreases with an increase of water concentration in HOAc. Mono- and diacylated derivatives of the H6 P2 W21 O71 (H2 O)3 HPA can be considered as 7- and 8-basic heteropolyacids. An

Fig. 5. A plot of logarithm of the rate constant of acetone dimerization vs. H0 ([H+ ] = 0.15 mol/l) [15].

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

31

H3 PW12 O40 , H4 SiW12 O40 > H4 GeW12 O40 , H6 P2 W18 O62 In the nucleophilic reaction of the addition of alcohols to cyclohexenone the catalytic activity changes in the series [54]: H4 SiW12 O40 > H3 PW12 O40 > H3 PMo12 O40 > H4 GeW12 O40 > H6 P2 W18 O62 In the reaction of isobutyl propionate esterication by n-propyl alcohol different Keggin-structured HPAs have almost the same activity, the activity of the Dawson-structure HPA (H6 P2 W18 O62 ) being higher [52]. At the same time, the HPA activity (per weight unit of a catalyst) in the synthesis of diacetonesorbose (DAS) decreases in the series [55]: H3 PW12 O40 > H3 PMo12 O40 , H4 SiW12 O40 > H4 SiW12 O40 > -H6 P2 W18 O62 H6 P2 Mo18 O62 > H5 PW11 ZrO40
Fig. 6. 31 P NMR spectra of 5 103 mol/l H6 P2 W21 O71 (H2 O)3 solutions in acetic acid with different water concentrations: (1) no water added, (2) 2.5 vol.%, (3) 10 vol.%, (4) 50 vol.%, and (5) aqueous heteropoly acid solution [30].

> H5 PW11 TiO40 > H6 P2 W21 O71 > -H6 P2 W18 O62 All these suggests that the catalytic action of HPA in acidic-type reactions depends on the HPA acidity, the heteropolyanion structure and the type of reaction (the nature of the reagents). 3.2. Biphasic catalytic systems Misono and co-workers [1] demonstrated that crystalline HPAs in many respects behave like solutions. This is due to the fact that these solids have discrete and mobile ionic structures. Water soluble HPAs have a great afnity for molecules of polar substances (alcohols, ketones, ethers, etc.) These substances can be absorbed in a large amount in the bulk of HPAs to form solvates. At the same time in contrast to polar molecules, non-polar reagents are not capable of absorbing in the bulk of HPA. By virtue of these properties HPAs can be used under the conditions of biphasic catalysis. Up to now, only few catalytic HPA-based systems are known (Table 5). Thus, the employment of the concentrated water solutions of HPAs permitted the polymerization reaction with ring opening of tetrahydrofuran (THF) in the

increase in the HPA basicity leads to a decrease in their acidity [53]. As a result, H6 P2 W21 O71 (H2 O)3 added simultaneously with the other reagents to the reaction mixture exhibits higher catalytic activity compared with H7 P2 W21 O71 (H2 O)2 (AcO) or H8 P2 W21 O71 (H2 O)(AcO)2 . The latter complexes appear after preliminary maintaining of HPA in aqueous HOAc. The catalytic activity series of HPAs having different structures differ depending on the reaction type. Thus, in the reaction of isobutyl propionate decomposition the following activity series is observed [52]: H3 PW12 O40 > H4 SiW12 O40 , H4 GeW12 O40 > H5 BW12 O40 > H6 P2 W18 O62 > H6 CoW12 O40 In the reaction of isobutyl propionate etherication by acetic acid the order is somewhat different [52]:

32

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

presence of water [56,57]. The polymer is formed in the HPA phase and then transferred to the THF phase. The reaction rate increases with decreasing H2 O/HPA ratio, and the molecular weight of the polymer does the same. The process is performed continuously. The polyoxytetramethylene glycol with molecular weight of 5002000 and narrow molecular weight distribution is obtained from the THF phase. According to [57], the reaction depends on the HPA structure and the acidity of HPAs (per catalyst weight) decreases in the series: H6 P2 W21 O71 > H3 PW12 O40 > H4 SiW12 O40 > H3 PMo12 O40 > H5 PW11 TiO40 H6 P2 Mo18 O62 > H5 PW11 ZrO40 > H6 P2 W18 O62 Similarly, the polymerization of cyclic formaldehyde acetal, trimer 1,3-dioxolane and 1,3,5-trioxane are catalyzed by the Keggin-type HPAs such as H3 PW12 O40 , H4 SiW12 O40 , H4 SiMo12 O40 and H3 PMo12 O40 [58]. It has been established that in the polymerization of 1,3,5-trioxane in the presence of the HPA, the reaction rate (based on the catalyst weight) is 25 times higher than that found for BF3 OR2 . The hydroquinone can be alkylated with isobutylene in the presence of the HPAs (H6 P2 W21 O71 , H3 PW12 O40 , H6 P2 W18 O62 ) under phase-transfer conditions in a biphasic system, including toluene and HPA dioxane etherate (HPA-nC4 H8 O2 -mH2 O) [59]. The yield of 2-tert-butyl-hydroquinone decreases in the series: H3 PW12 O40 > H6 P2 W21 O71 > H6 P2 W18 O62 This order correlates well with the relative basicity of conjugated anions (see Section 2.2). Since in these series the anion basicity increases, the acidity of the corresponding HPA falls. H3 PW12 O40 shows higher efciency than H2 SO4 or H3 PO4 . 3.3. Heterogeneous catalytic systems In this section, liquid-phase and gas-phase reactions catalyzed by bulk and supported HPAs are discussed. In heterogeneous as well as homogeneous conditions HPAs are more effective than conventional catalysts, such as SiO2 Al2 O3 , zeolites, etc. The

comparison between the HPA acidity and the acidity of other acidic catalysts was carried out by measuring their catalytic activity in different reactions. To date, a few reactions are known, in which the catalytic activity of bulk and supported H6 P2 W18 O62 , H6 P2 W21 O71 and H21 B3 W39 O13 is compared with that of the Keggin-type heteropolyacids (Hn XW12 O40 ; X = PV , SiIV , GeIV , BIII , CoIII ) [47,48, 6062]. Thus, the study of the isopropyl alcohol dehydration reaction CH3 C(OH)HCH3 CH3 CH=CHCH2 + H2 O, at 150 C showed that the HPAs are stronger than the solid acids widely used in catalysis, such as H3 PO4 /SiO2 and even zeolites [50,63]. In accordance with their composition, the HPA acidity changes in the series [63]: H3 PW12 O40 > H4 SiW12 O40 > H2 SiMo12 O40 It is interesting that the acidity series of solid HPAs agrees with those of acids in solution. However, for reactions where small polar molecules are employed, such relation seems to be an exception rather than a rule, since beside the acidity, the catalytic properties of HPAs may be considerably affected by the substrate absorption in the bulk of HPA with the formation of so called pseudoliquid phase. It was demonstrated for reaction of the addition of methanol to isobutylene in the presence of HPA. In this reaction, the HPA acidity decreases in the series [47,61]: H6 P2 W18 O62 > H4 SiW12 O40 > H4 GeW12 O40 > H3 PW12 O40 > H5 BW12 O40 , H6 CoW12 O40 The high activity of acids with a low acidity (H6 P2 W18 O62 , H4 SiW12 O40 ) was explained by the inuence of the rate of the substrate absorption in the bulk of the HPA. This series differs from the activity series for the esterication reactions in the liquid-phase [64]. The activity (per unit weight) of the bulk H6 P2 W18 O62 is at least 13 times higher than that of H3 PW12 O40 [60]. At the same time, the activity (per HPA weight) of H6 P2 W18 O62 is 2.7 times higher than that of H3 PW12 O40 in the homogeneous H5 PW10 V2 O40

> H3 PMo12 O40 > H5 PW10 V2 O40

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935

33

liquid-phase synthesis of methyl-tert-butyl ether (MTBE) [64]. The reaction of 2,6-di-tert-butyl-4-methylphenol (DMBP) with toluene, which behaves as an acceptor of tert-butyl, has been studied using bulk and supported -H6 P2 W18 O62 , H6 As2 W21 O69 and the Keggin-type HPAs, H3 PW12 O40 , H5 PW11 TiO40 , and H5 PW11 ZrO40 [48]. It has been found that -H6 P2 W18 O62 and H6 As2 W21 O69 are much more active than H3 PW12 O40 . The activity of HPAs considerably depends on the anion structure and composition. For the bulk W-containing HPAs the activity (per number of protons) decreases in the following series [48]: H6 P2 W18 O62 > H6 As2 W21 O69 > Keggin. For the Keggin HPAs, the activity falls in the series: H3 PW12 O40 > H5 PW11 TiO40 > H5 PW11 ZrO40 . Lewis centers in H5 PW11 TiO40 and H5 PW11 ZrO40 do not inuence much the rate of reaction of trans-alkylation, as it takes place in homogeneous reactions of acetone dimerization [18] and n-butyl alcohol etherication of by acetic acid [31]. In the reaction of isobutylene oligomerization in decane solution it was demonstrated [65] that H3 PW12 O40 is less active than the Keggin-structure

HPAs having Lewis centers as well as HPAs of other structures: H6 P2 W18 O62 > H6 P2 W21 O71 > H5 PW11 TiO40 > H5 PW11 ZrO40 > H3 PW12 O40 It seems that in this reaction, Lewis sites along with Brnsted sites signicantly contribute to the catalytic activity of H5 PW11 ZrO40 . Interestingly, the ZrIV -substituted HPA shows higher catalytic activity than TiIV -substituted HPA. The support has a great effect on the catalytic activity of supported HPAs. The activity increases with an increase in the amount of supported HPAs. This occurs for many reactions catalyzed by the Keggin-type HPAs [6669]. Thus, it was shown that in the reaction of 2,6-di-tert-butyl-4-methylphenol with toluene that acts as a tert-butyl acceptor the catalytic activity of HPAs (H6 P2 W18 O62 , H3 PW12 O40 ) supported on SiO2 and lamentous carbon is very close [48]. The activity of the surface protons of supported HPA increases monotonically with increasing acid strength, which in turn rises with the increase in the total HPA loading and reaches a maximum for the bulk HPA (Fig. 7). In the gas-phase synthesis of MTBE [47,61], the catalytic activity of supported Dawson-type acid H6 P2 W18 O62 is compared with that of the Keggin-type

Fig. 7. Dependence of catalytic activity on HPA loading in reaction of 2,6-di-tert-butyl-4-methylphenol with toluene: kb represents the activity based on the total amount of H3 PW12 O40 protons (1H3 PW12 O40 /CFC; 2H3 PW12 O40 /SiO2 ; 3H6 P2 W18 O69 /CFC) [48].

34

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935 [4] J.L. Bonardet, K. Carr, J. Fraissard, G.B. McGarvey, J.B. McMonagle, M. Seay, J.B. Moffat, Microporous metaloxygen cluster compounds (heteropoly oxometalates), in: W.R. Moser (Ed.), Advanced Catalysts and Nanostructured Materials, Modern Synthetic Methods, Academic Press, New York, 1996, p. 395. [5] M.T. Pope, Heteropoly and Isopoly Oxometallates, Berlin, Springer, 1983, 220 pp. [6] A. Muller, F. Peters, M.T. Pope, D. Gatteschi, Chem. Rev. 98 (1998) 239. [7] G.M. Maksimov, Russ. Chem. Rev. 64 (1995) 445. [8] L.D. Kurbatova, A.A. Ivakin, A.M. Voronova, Koord. Khim. 1 (1975) 1481 (in Russian). [9] A.A. Ivakin, L.D. Kurbatova, L.A. Kapustina, Zhurn. Neorgan. Khim. 23 (1978) 2545 (in Russian). [10] I.B. Tatianina, A.P. Borisova, E.A. Torchenkova, V.I. Spitzin, Dokl. Akad. Nauk SSSR 23 (1981) 2545 (in Russian). [11] E.N. Dorohova, I.P. Alimarin, Uspekhi Khim. 48 (1979) 930 (in Russian). [12] M.N. Timofeeva, M.M. Matrosova, G.M. Maksimov, V.A. Likholobov, Kinet. Katal. 42 (2001) 862 (in Russian). [13] S.M. Kulikov, I.V. Kozhevnikov, Izv. Akad. Nauk SSSR, Ser. Khim. (1981) 498 (in Russian). [14] S.M. Kulikov, I.V. Kozhevnikov, Izv. Akad. Nauk SSSR, Ser. Him. (1980) 2213 (in Russian). [15] M.N. Timofeeva, M.M. Matrosova, G.M. Maksimov, V.A. Likholobov, Russ. Chem. Bull. Int. Ed. (in press). [16] M.T. Pope, E. Papaconstantinou, Inorg. Chem. 6 (1967) 1147; G.M. Varga, E. Papaconstantinou, M.T. Pope, Inorg. Chem. 3 (1970) 662. [17] I.V. Kozhevnikov, S.Z. Hanhasaeva, S.M. Kulikov, Kinet. Katal. 29 (1988) 76 (in Russian). [18] M.N. Timofeeva, G.M. Maksimov, V.A. Likholobov, Kinet. Katal. 42 (2001) 37 (in Russian). [19] C. Hu, M. Hashimoto, T. Okuhara, M. Misono, J. Catal. 143 (1993) 437. [20] T. Okuhara, C. Hu, M. Hashimoto, M. Misono, Bull. Chem. Soc. Jpn. 67 (1994) 1186. [21] L. Barcha, M.T. Pope, J. Phys. Chem. 79 (1975) 92. [22] Y. Izumi, K. Matsuo, K. Urabe, J. Mol. Catal. 18 (1983) 299. [23] G.M. Maksimov, M.N. Timofeeva, V.A. Likholobov, Russ. Chem. Bull. Int. Ed. 50 (2001) 1529. [24] Y. Izumi, K. Matsuo, K. Urabe, J. Mol. Catal. 18 (1983) 299. [25] G.M. Maksimov, E.A. Paukshtis, A.A. Budneva, R.I. Maksimovskaya, V.A. Likholobov, Russ. Chem. Bull. Int. Ed. 50 (2001) 587. [26] M. Otake, T. Shokubai, Catalysis 17 (1975) 13. [27] T. Okuhara, T. Nishimura, H. Watanabe, M. Misono, J. Mol. Catal. 74 (1992) 247. [28] M. Misono, T. Okuhara, CHEMTECH 23 (1993) 23. [29] R.J. Gillespie, Acc. Chem. Res. 1 (1968) 202. [30] G.A. Olah, G.K.S. Prakash, J. Sommer, Superacids, Wiley, New York, 1985. [31] M.N. Timofeeva, M.M. Matrosova, G.M. Maksimov, V.A. Likholobov, A.V. Golovin, R.I. Maksimovskaya, E.A. Paukshtis, Kinet. Katal. 42 (2001) 868 (in Russian). [32] F. Lefebre, F.X.L. Cai, A. Auroux, J. Mater. Chem. 4 (1994) 125.

heteropolyacids (Hn XW12 O40 ; X = PV , SiIV , GeIV , BIII , CoIII ). It has been also found that the activity of H6 P2 W18 O62 is considerably enhanced by supporting on SiO2 . However, the activity (per unit weight) of bulk H6 P2 W18 O62 is lower than that of supported H6 P2 W18 O62 . Thus, the available data demonstrate that both the acidity of proton sites and the amount of proton sites are among key factors which determine the catalytic activity in acidic reactions.

4. Conclusion The selected examples reviewed show the broad scope of potentially promising applications of HPAs as acid catalysts in various organic reactions. Due to their unique physicochemical properties, HPAs can be protably used in homogeneous, biphasic and heterogeneous systems. In many cases HPA-based catalysts have higher activity than known traditional catalysts. The examples given allow to conclude that the catalytic effect of HPAs in acidic-type reactions depends mainly on three factors, namely, the acidity, heteropolyanion structure and type of reaction (nature of reagents). The analysis of investigations carried out for HPAs of different structures revealed that HPAs are strong acids and that their acidity can be controlled by the change of the heteropolyanion charge. The catalytic activity is more dependent on the HPA structure rather than its composition. Prediction of catalytic activity is possible only for HPAs having similar structure.

Acknowledgements The author thanks Dr. O.A. Kholdeeva for her help in preparing this manuscript. The author would like to thank Prof. I.V. Kozhevnikov and Dr. G.M. Maksimov for valuable discussion and comments. References
[1] T. Okuhara, N. Mizuno, M. Misono, Adv. Catal. 41 (1996) 113. [2] I.V. Kozhevnikov, Chem. Rev. 98 (1998) 171. [3] I.V. Kozhevnikov, Russ. Chem. Rev. 62 (1993) 473.

M.N. Timofeeva / Applied Catalysis A: General 256 (2003) 1935 [33] D. Farcasiu, J.Q. Li, J. Catal. 152 (1995) 198. [34] J.G. Higheld, J.B. Moffat, J. Catal. 89 (1984) 185. [35] A. Auroux, J.C. Vedrine, Stud. Surf. Sci. Catal. 20 (1985) 311. [36] S. Shikata, S. Nakata, T. Okuhara, M. Misono, J. Catal. 166 (1997) 263. [37] I.V. Kozhevnikov, M.N. Timofeeva, J. Mol. Catal. 75 (1992) 179. [38] S. Shikata, S. Nakata, T. Okuhara, M. Misono, J. Catal. 166 (1997) 263. [39] Y. Wu, X. Ye, X. Yang, X. Wang, W. Chu, Y. Hu, Ind. Eng. Chem. Res. 35 (1996) 2546. [40] S.M. Kulikov, M.N. Timofeeva, I.V. Kozhevnikov, V.I. Zaikovskii, L.I. Plyasova, I.A. Ovsyanikova, Izv. Akad. Nauk SSSR, Ser. Khim. (1989) 763 (in Russian). [41] L.R. Pizzio, C.V. Caceres, M.N. Blanko, Appl. Catal. A 167 (1998) 283. [42] M.E. Chimienti, L.R. Pizzio, M.N. Blanko, C.V. Caceres, Appl. Catal. A 208 (2001) 7. [43] R.I. Maksimovskaya, Kinet. Catal. 36 (1995) 910. [44] K.M. Rao, R. Gobetto, A. Iannibello, A. Zecchina, J. Catal. 119 (1989) 512. [45] G.I. Kapustin, T.R. Brueva, A.L. Klyachko, M.N. Timofeeva, S.M. Kulikov, I.V. Kozhevnikov, Kinet. Catal. 31 (1990) 1017 (in Russian). [46] I.V. Kozhevnikov, A. Sinnema, R.J.J. Jansen, H. van Bekkum, Chem. Lett. 27 (1994) 187. [47] S. Shikata, T. Okuhara, M. Misono, J. Mol. Catal. A: Chem. 100 (1995) 49. [48] M.N. Timofeeva, G.M. Maksimov, T.V. Reshetenko, L.B. Avdeeva, E.A. Paukshtis, A.L. Chuvilin, V.A. Likholobov, 10th International Symposium on Relations between Homogeneous and Heterogeneous Catalysis, 26 July 2001, CPE Lyon, France, p. 160. [49] I.V. Kozhevnikov, S.Z. Hanhasaeva, S.M. Kulikov, Kinet. Katal. 30 (1989) 50 (in Russian). [50] Y. Izumi, Catal. Today 33 (1997) 371. [51] Y. Izumi, K. Urabe, A. Onaka, Zeolite, Clay, and Heteropolyacids in Organic Reactions, Kodansha, Tokio-VCH, Weinheim, 1992. [52] C. Hu, M. Hashimoto, T. Okuhara, M. Misono, J. Catal. 143 (1993) 437. [53] M. Misono, N. Mizuno, T. Okuhara, Adv. Catal. 41 (1996) 113.

35

[54] T. Kengaku, Y. Matsumoto, M. Misono, J. Mol. Catal. A: Chem. 134 (1998) 237. [55] G.M. Maksimov, M.N. Timofeeva, React. Kinet. Catal. Lett. 56 (1995) 191 (in Russian). [56] A. Aoshima, S. Tonomura, S. Yamamatsu, Polym. Adv. Tech. 2 (1990) 127. [57] L.I. Kuznetsova, G.M. Maksimov, V.A. Likholobov, Kinet. Katal. 40 (1999) 688 (in Russian). [58] M. Bednarek, K. Brzezinska, J. Stasinski, P. Kubisa, S. Penczek, Makromol. Chem. 190 (1989) 929. [59] M.N. Timofeeva, I.V. Kozhevnikov, React. Kinet. Catal. Lett. 54 (1995) 413. [60] G. Baronetti, H. Thomas, C.A. Querini, Appl. Catal. A 217 (2001) 131. [61] S. Shikata, T. Okuhara, M. Misono, J. Mol. Catal. A: Chem. 100 (1995) 49. [62] F.X. Lui-Cai, B. Sahut, E. Faydi, A. Auroux, G. Herve, Appl. Catal. A 185 (1999) 75. [63] I.V. Kozhevnikov, K.I. Matveev, Uspehi Khim. 51 (1982) 1875 (in Russian). [64] G.M. Maksimov, I.V. Kozhevnikov, React. Kinet. Catal. Lett. 39 (1989) 317. [65] M.N. Timofeeva, G.M. Maksimov, V.A. Likholobov, Kinet. Katal. (in Russian), in press. [66] I.V. Kozhevnikov, M.N. Timofeeva, J. Mol. Catal. 75 (1992) 179. [67] V.M. Mastikhin, V.V. Terskih, M.N. Timofeeva, O.P. Krivoruchko, J. Mol. Catal. A: Chem. 95 (1995) 135. [68] Y. Izumi, R. Hasebe, K. Urabe, J. Catal. 84 (1983) 403. [69] Y. Izumi, N. Natsume, H. Takamine, I. Tamaoki, K. Urabe, Bull. Chem. Soc. Jpn. 62 (1989) 2159. [70] S. Sato, C. Sakirai, H. Furuta, T. Sodesawa, F. Nozaki, J. Chem. Soc., Perkin Trans. 2 (1993) 385. [71] S. Sato, C. Sakirai, H. Furuta, T. Sodesawa, F. Nozaki, J. Chem. Soc., Chem. Commun. (1991) 1327. [72] I.V. Kozhevnikov, A. Sinnema, A.J. van der Weerdt, H. van Bekkum, J. Mol. Catal. A: Chem. 120 (1997) 63. [73] M.N. Timofeeva, G.M. Maksimov, V.A. Likholobov, Kinet. Katal. 41 (2000) 846 (in Russian). [74] B. Trk, . Molnr, C.R. Acad. Sci. Paris, t.1, Serie II c (1998) 381. [75] B. Trk, I. Bucsi, T. Beregszszi, I. Kapocsi, . Molnr, J. Mol. Catal. A: Chem. 107 (1996) 305.

S-ar putea să vă placă și