Rhodium(I) and rhodium(III)–heteropolyacids supported on MCM-41 for the catalytic hydroformylation...

12
Rhodium(I) and rhodium(III)–heteropolyacids supported on MCM-41 for the catalytic hydroformylation of styrene derivatives B. El Ali * , J. Tijani, M. Fettouhi, M. El-Faer, A. Al-Arfaj Chemistry Department, King Fahd University of Petroleum and Minerals (KFUPM), P.O. Box 202, 31261 Dhahran, Saudi Arabia Received 28 August 2004; received in revised form 18 December 2004; accepted 5 January 2005 Available online 1 February 2005 Abstract Rh supported catalysts were prepared by impregnating rhodium(I) and rhodium(III) complexes with and without heteropolyacids for the hydroformylation of styrene derivatives. The effect of the pore size of MCM-41 was considered. The addition of water showed a promoting effect with Rh(III) based catalysts. The amount of water showed also a big effect on the catalytic activity of the Rh(III) supported catalyst. The change of the temperature affects the selectivity of the reaction time of the reaction. Different heteropolyacids such as H 3 PMo 12 O 40 xH 2 O (HPA-Mo 12 ) and H 3 PW 12 O 40 yH 2 O (HPA-W 12 ) were used as co-catalysts that were impregnated along with rhodium complexes on the inorganic supports. The results showed a clear effect of the heteropolyacid H 3 PW 12 O 40 yH 2 O (HPA-W 12 ) in increasing the catalytic activity of the rhodium supported catalyst. # 2005 Elsevier B.V. All rights reserved. Keywords: Hydroformylation; MCM-41; Zeolites; Clay; Rhodium; Heteropolyacid; Styrene; Phosphine ligands; Syngas 1. Introduction The hydroformylation (Oxo syntheses) is a process of production of aldehydes from alkenes and syngas. This process is mainly catalyzed by Co and Rh complexes. While the industrially relevant product range of hydroformylation has remained more or less unchanged over the past 20 years, the research work has continued actively in this area [1–3]. The immobilization or heterogenization of homogeneous catalysts on inorganic support materials can be achieved by various methods. These methods can be classified by the interaction between catalytically active component and support [4–6]. Hydroformylation of styrene over Rh/SiO 2 Al 2 O 3 has been studied in the presence of the chiral diphosphines ()-Chiraphos and ()-DIOP. Leaching of rhodium was greatly reduced by controlling the water content of the solvent and the degree of hydration of the catalyst surface. Catalytic activity was significantly decreased in the presence of the diphosphines although high levels of both chemo- and regioselectivity towards the chiral aldehyde were maintained [7]. In the two-phase hydro- formylation of propylene, some of the novel catalysts showed activities and regioselectivities similar to those of Rh/TPPTS. Amphiphilic Rh/phosphonate–phosphine cat- alysts were found widely superior to Rh/TPPTS in the hydroformylation of 1-octene [8]. A series of polyamides have been synthesized by low-temperature interfacial condensation of 2,5- and 2,6-pyridinedicarboxylic chlor- ides and aliphatic diamines H 2 N(CH 2 ) n NH 2 , where n =2 and 6, and used as the model supports for immobilization of the Rh(I) complex catalyst. It was found that the catalyst activity decreased with increasing polymer crystallinity. A strong impact of support structure on the catalyst selectivity was also revealed. A high regioselectivity of the catalyst in the hydrosilylation of alkenes toward formation of the linear products was achieved due to the favorable microporous structure of the polyamide supports with pore sizes between 10 and 20 A ˚ . The results of the 6–9 times repeated catalytic runs indicated high stability of the polyamide-supported catalysts [9]. The www.elsevier.com/locate/apcata Applied Catalysis A: General 283 (2005) 185–196 * Corresponding author. Tel.: +966 3 8604491; fax: +966 3 8604277. E-mail address: [email protected] (B. El Ali). 0926-860X/$ – see front matter # 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.apcata.2005.01.005

Transcript of Rhodium(I) and rhodium(III)–heteropolyacids supported on MCM-41 for the catalytic hydroformylation...

www.elsevier.com/locate/apcata

Applied Catalysis A: General 283 (2005) 185–196

Rhodium(I) and rhodium(III)–heteropolyacids supported on MCM-41

for the catalytic hydroformylation of styrene derivatives

B. El Ali*, J. Tijani, M. Fettouhi, M. El-Faer, A. Al-Arfaj

Chemistry Department, King Fahd University of Petroleum and Minerals (KFUPM), P.O. Box 202, 31261 Dhahran, Saudi Arabia

Received 28 August 2004; received in revised form 18 December 2004; accepted 5 January 2005

Available online 1 February 2005

Abstract

Rh supported catalysts were prepared by impregnating rhodium(I) and rhodium(III) complexes with and without heteropolyacids for the

hydroformylation of styrene derivatives. The effect of the pore size of MCM-41 was considered. The addition of water showed a promoting

effect with Rh(III) based catalysts. The amount of water showed also a big effect on the catalytic activity of the Rh(III) supported catalyst. The

change of the temperature affects the selectivity of the reaction time of the reaction. Different heteropolyacids such as H3PMo12O40�xH2O

(HPA-Mo12) and H3PW12O40�yH2O (HPA-W12) were used as co-catalysts that were impregnated along with rhodium complexes on the

inorganic supports. The results showed a clear effect of the heteropolyacid H3PW12O40�yH2O (HPA-W12) in increasing the catalytic activity of

the rhodium supported catalyst.

# 2005 Elsevier B.V. All rights reserved.

Keywords: Hydroformylation; MCM-41; Zeolites; Clay; Rhodium; Heteropolyacid; Styrene; Phosphine ligands; Syngas

1. Introduction

The hydroformylation (Oxo syntheses) is a process of

production of aldehydes from alkenes and syngas. This

process is mainly catalyzed by Co and Rh complexes. While

the industrially relevant product range of hydroformylation

has remained more or less unchanged over the past 20 years,

the research work has continued actively in this area [1–3].

The immobilization or heterogenization of homogeneous

catalysts on inorganic support materials can be achieved by

various methods. These methods can be classified by the

interaction between catalytically active component and

support [4–6].

Hydroformylation of styrene over Rh/SiO2 Al2O3 has

been studied in the presence of the chiral diphosphines

(�)-Chiraphos and (�)-DIOP. Leaching of rhodium was

greatly reduced by controlling the water content of the

solvent and the degree of hydration of the catalyst surface.

Catalytic activity was significantly decreased in the

* Corresponding author. Tel.: +966 3 8604491; fax: +966 3 8604277.

E-mail address: [email protected] (B. El Ali).

0926-860X/$ – see front matter # 2005 Elsevier B.V. All rights reserved.

doi:10.1016/j.apcata.2005.01.005

presence of the diphosphines although high levels of

both chemo- and regioselectivity towards the chiral

aldehyde were maintained [7]. In the two-phase hydro-

formylation of propylene, some of the novel catalysts

showed activities and regioselectivities similar to those of

Rh/TPPTS. Amphiphilic Rh/phosphonate–phosphine cat-

alysts were found widely superior to Rh/TPPTS in the

hydroformylation of 1-octene [8]. A series of polyamides

have been synthesized by low-temperature interfacial

condensation of 2,5- and 2,6-pyridinedicarboxylic chlor-

ides and aliphatic diamines H2N(CH2)nNH2, where n = 2

and 6, and used as the model supports for immobilization

of the Rh(I) complex catalyst. It was found that the catalyst

activity decreased with increasing polymer crystallinity.

A strong impact of support structure on the catalyst

selectivity was also revealed. A high regioselectivity of

the catalyst in the hydrosilylation of alkenes toward

formation of the linear products was achieved due to the

favorable microporous structure of the polyamide supports

with pore sizes between 10 and 20 A. The results of

the 6–9 times repeated catalytic runs indicated high

stability of the polyamide-supported catalysts [9]. The

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196186

[Rh(m-Cl)(COD)]2 complex has been impregnated on

activated carbon and used as catalysts for the hydro-

formylation of 1-octene. The catalyst prepared with the

functionalized activated carbon, where the anchorage

takes place by ion exchange, is more stable and active in

further catalytic runs using the proper recovering method

[10].

Silica-tethered rhodium thiolate catalysts prepared by

the condensation of Rh2[m-S(CH2)3Si(OCH3)3]2(CO)4 or

Rh2[m-S(CH2)3Si(OCH3)3]2[Ph2P(CH2)3Si(OC2H5)3]2 (CO)2

on silica gel were found to exhibit high activity for

hydroformylation of 1-octene in the presence of phosphine

or phosphite ligands under the mild conditions of 60 8C and

1 bar total pressure of H2 and CO. The catalysts are easily

separated from the reaction mixtures [11]. Immobilization of

a dinuclear rhodium(II) acetato complex with bridging

ortho-metalated phosphine ligands on amorphous silica

supports as well as a mesoporous MCM-41 support resulted

in useful catalysts for hydroformylation of styrene and 1-

decene. The effect of recycling on the conversion of the

olefins and on the branched-normal selectivity of the

hydroformylation products was studied [12]. A polysilicate

immobilized rhodium complex with ligand was found

to perform as a selective hydroformylation catalyst

showing an overall selectivity for the linear aldehyde of

94.6% in the case of 1-octene hydroformylation at 80 8Cand 50 bar CO/H2 1:1. In addition the formation of 1-

nonanol (3.6% at 20% conversion) was observed via the

hydrogenation of the corresponding aldehyde [13]. Func-

tionalization of MCM-41 with 3-aminopropyltrimethoxye-

siline (APTS), and subsequent encapsulation and anchoring

of HRh(CO)(PPh3)3 in the mesoporous material was

recently reported [4]. The unique approach of passivation

of the external surface silanol groups (Si–OH) the

mesoporous materials by reacting with controlled amount

of Ph2SiCl2 and subsequent treatment with APTS for

selective functionalization of the inner-surface. Thus, the

initial treatment mesoporous with Ph2SiCl2 compelled

anchoring of APTS as well as Rh complex selectively

inside the mesopore channels [4]. Rh(I) catalyst bearing

the diphosphine 1,10-bis(diisopropylphosphino)-ferrocene

(DiPFc) was immobilized on phosphotungtic acid treated

alumina or silica, the catalyst used for efficient and selective

hydrogenation of a range of functionalized aldehydes, as

well as alkenes and alkynes, under mild conditions. The

catalyst was also recycled several times [14]. Another

approach to immobilizing homogeneous catalysts on solid

supports has been reported, in which the Wilkinson’s

hydroformylation catalyst complex was tethered through

heteropolyacid to zeolites [15] and MCM-41 [4], this result

to an excellent stability, reusability and even improved

activity.

The new heterogenized homogenized catalysts are being

at least as active as the homogeneous species and capable of

being re-used many times with little lost in activity and

selectivity.

2. Experimental

2.1. General

Alkenes were highly pure (>99%) and were purchased

from Sigma-Aldrich Company. The solvents used in the

experiments, such as THF, hexane, CH2Cl2, benzene, 1,2-

dimethoxyethane, 1,4-dioxane, methyl ethyl ketone, acetone,

methanol,ethanol,andothersolventswereallHPLCgradeand

werestoredundernitrogenoveractivated3 Amolecular sieves

and also purchased from Sigma-Aldrich. The 3 A molecular

sieves (productofAldrich)wereactivated for 3 hat300 8C and

used with the solvent to adsorb moistures or water. Anisole,

tetraetoxysilane (TEOS), n-hexadecyltrimethyl ammonium

(or cetyltrimethyl ammonium) bromide, ammonium hydro-

xide, tetramethyl ammonium bromide (25% by weight

solution in water), Ludox AS40 (containing 40% by weight

SiO2) were also purchased from Sigma-Aldrich and used

without any furtherpurification. RhCl3�3H2O, [Rh(COD)Cl]2,

[Rh(CO)2Cl]2, RhHCO(PPh3)3,RhCl(PPh3)3, Rh6(CO)16, and

other catalysts were purchased from Strem and used

without purification. Triphenylphosphine (PPh3), 1,4-bis-

(diphenylphosphino)butane (dppb), H3PW12O40�xH2O,

H3PMo12O40�yH2O and H3SiMo12O40�zH2O were purchased

from Sigma-Aldrich.

2.2. Preparation of catalysts

2.2.1. Preparation of heteropolyacids and MCM-41

The heteropolyacids of general formula H3+nPMo12�nV-

nO40�xH2O (HPA-n; n = 1–5) were prepared according to the

procedure described in the literature [16,17].

The mesoporous materials MCM-41 used as supports for

the impregnation of the catalysts were synthesized accord-

ing to the two methods described in literature [18,19]. Two

types of MCM-41 having different pore sizes, 30 A [18] and

50 A [19], were prepared. These supports have also different

surface areas and showed different adsorption–desorption

properties. However, the method for the synthesis of MCM-

41 (30 A) [18] represents a fast and a simple route for the

preparation of the mesoporous material.

The MCM-41 prepared was analyzed by X-ray diffrac-

tion to determine the X-ray patterns and also to calculate the

pore diameter. X-ray diffraction patterns were recorded on a

JOEL JDX-3530 diffractometer using Cu Ka radiation

(l = 0.15405 nm). Diffraction data were recorded between

18 and 308 2u in steps of 0.028with a count time of 1 s at each

point. The Bragg peaks can be indexed assuming a

hexagonal symmetry. We assume that the pore wall

thickness is 10 A.

2.2.2. Preparation and characterization of rhodium(I) and

rhodium(III) impregnated with and without heteropolyacids

on MCM-41

The MCM-41materials with pore size of 30 A were

chosen as the inorganic support of choice in the impreg-

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196 187

nation processes due to their large pore size and high specific

surface area. In addition, the method of synthesis is easy,

simple and fast compared to MCM-41 (50 A) with

equivalent catalytic activity as a support in the hydro-

formylation tests (vide supra):

RhnmHpSS

where Rh is the rhodium complex; n the oxidation state; m

the % rhodium impregnated; H the type of the heteropo-

lyacid or ligand; p the % impregnation of HPA-n; S the

support; S the solvent of impregnation.

2.2.2.1. Synthesis of 1–2% of Rh(I) or Rh(III)/HPA/MCM-

41 in methanol (M) or water (W) [Rh32HPA10MCM-M

(W)] and [Rh12HPA10MCM-M (W)]. RhCl3�3H2O [or

Rh(I) complexes] alone or with the co-catalyst was

impregnated on the MCM-41 using methanol (M) or water

(W) as a solvent. For example, a mixture solution of

2% (11.4 mg) RhCl3�3H2O and 10% (56.8 mg) of

H3PW12O40�25H2O (HPA-W12) were stirred for 30 min at

room temperature under nitrogen atmosphere. Five hundred

milligrams of MCM-41 were than added to the homo-

geneous solution and mixture was stirred for 24 h. The

solvent was removed under vacuum and the solid sample

was dried at 90 8C for 2 h and stored in the refrigerator. The

catalysts prepared are summarized in Table 1.

2.2.2.2. Synthesis of 5% of Rh(I) or Rh(III)/HPA/MCM-41

i n e t h a n o l ( E ) [ R h 3 5 H PA 1 0 M C M - E ] a n d

[Rh15HPA10MCM-E]. The tethered catalyst was prepared

following the procedure reported by Augustine and co-

workers [8]. Thirty three micromoles of HPA dissolved in

20 ml of ethanol and 500 mg of MCM-41, the solution was

stirred at room temperature for 16 h, the resulting solid was

filtered and washed with ethanol to removed HPA and finally

dried at 100 8C for 6 h.

The solid was suspended in 20 ml dry ethanol and

30 mmol HRhCO(PPh3)3 was added and refluxed (at 80 8C)

under stirring for 18 h. The light gray solid product was then

washed with ethanol repeatedly to remove unanchored to

HPA, dried at 80 8C and used as such for the reaction. The

catalysts prepared are summarized in Tables 1 and 2.

Table 1

Supported catalysts obtained by impregnation of rhodium(III) with and without

Catalyst no. Catalyst Rhodium (%) H

Rh-1 Rh32MCM-M RhIIICl3�3H2O, 2% –

RhW-2 Rh32W5MCM-M RhIIICl3�3H2O, 2% H

RhW-3 Rh32W10MCM-M RhIIICl3�3H2O, 2% H

RhW-4 Rh32W20MCM-M RhIIICl3�3H2O, 2% H

Rh-5 Rh32MCM-W RhIIICl3�3H2O, 2% –

RhW-6 Rh32W10MCM-W RhIIICl3�3H2O, 2% H

RhW-7 Rh32W10MCM0-M RhIIICl3�3H2O, 2% H

RhMo-8 Rh32Mo10MCM-M RhIIICl3�3H2O, 2% H

Rh-9 Rh35MCM-E RhIIICl3�3H2O, 5% –

RhW-10 Rh35W10MCM-E RhIIICl3�3H2O, 5% H

RhMo-11 Rh35Mo10MCM-E RhIIICl3�3H2O, 5% H

2.2.2.3. Impregnation of rhodium(I) or rhodium(III) with

and without heteropolyacid on MCM-41. The impregnation

of the rhodium complexes was realized on MCM-41 (30 A)

and MCM-41 (50 A). Tables 1 and 2 summarize the most

important supported catalysts prepared by impregnation of

2% of RhCl3�3H2O, with and without 10% of the

heteropolyacid H3PW12O40 (HPA-W12) in methanol (cata-

lyst nos. Rh-1 and RhW-4) or in water as a solvent (catalyst

nos. Rh-5 and RhW-6) on MCM-41 (30 A). In addition,

RhCl3�3H2O (2%) was impregnated with HPA-W12 (10%)

on MCM-41 (50 A) in dry methanol as a solvent (catalyst no.

Rh-7). Another heteropolyacid such as H3PMo12O40 was

considered with RhCl3�3H2O for the impregnation on

MCM-41 (30 A) (catalyst no. RhMo-8). 5% of RhCl3�3H2O

3H2O were also impregnated with and without HPA-W12 or

HPA-Mo12 on MCM-41 (30 A) in dry ethanol as a solvent of

impregnation (catalyst nos. Rh-9 and RhMo-11). Similarly,

the rhodium(I) complex [Rh(COD)Cl]2 (RhA) (2%) was

impregnated on MCM-41 (30 A) and MCM-41 (50 A) with

and without HPA-W12 (10%) in dry methanol as a solvent

(catalyst nos. Rh-12 and RhW-14). Other rhodium(I)

complexes such as [RhI(CO)2Cl]2 (RhB) and HRhCO(PPh3)3

(RhC) were impregnated on MCM-41 (30 A) in a similar

manner (catalyst nos. Rh-15 and RhW-18) in dry methanol.

The other heteropolyacid H3PMo12O40 was also considered

with [Rh(COD)Cl]2 (RhA) for the impregnation on MCM-41

(30 A) (catalyst no. RhMo-19). It is important to note that

Rh6(CO)16 was not considered for impregnation due to very

low solubility in methanol or water at room temperature and

even at 60 8C. Five percent of RhHCO(PPh3)3 were also

impregnated with and without HPA-W12 or HPA-Mo12 on

MCM-41 (30 A) in dry ethanol as a solvent of impregnation

(catalyst nos. Rh-20 and RhMo-22).

2.2.3. Characterization of supported catalyst

2.2.3.1. Infrared. The IR spectrum of the rhodium complex

HRh(CO)(PPh)3 showed bands on 1921 and 512 cm�1 for

Rh–CO, and Rh–P, respectively. However, after its impreg-

nation on MCM-41 giving Rh12MCM-M, the Rh–CO

shifted to 1942 cm�1. When the same catalyst was treated

with 600 psi of CO/H2 for 2 h at 110 8C, two peaks appeared

heteropolyacid (HPA-n)

PA-n (H)/co-catalysts (%) MCM-41 (A) Solvent (S)

30 Methanol (M)

3PW12O40, 5% 30 Methanol (M)

3PW12O40, 10% 30 Methanol (M)

3PW12O40, 20% 30 Methanol (M)

30 Water (W)

3PW12O40, 10% 30 Water (W)

3PW12O40, 10% 50 Methanol (M)

3PMo12O40, 10% 30 Methanol (M)

30 Ethanol (E)

3PW12O40, 10% 30 Ethanol (E)

3PMo12O40, 10% 30 Ethanol (E)

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196188

Table 2

Supported catalysts obtained by impregnation of Rhodium(I) with and without Heteropolyacid (HPA-n)

Catalyst no. Catalyst Rhodium (%) HPA-n (H)/co-catalysts (%) MCM-41 (A) Solvent (S)

Rh-12 RhA12MCM-M [RhI(COD)Cl]2, 2% – 30 Methanol (M)

RhW-13 RhA12W10MCM-M [RhI(COD)Cl]2, 2% H3PW12O40, 10% 30 Methanol (M)

RhW-14 RhA12W10MCM0-M [RhI(COD)Cl]2, 2% H3PW12O40, 10% 50 Methanol (M)

Rh-15 RhB12MCM-M [RhI(CO)2Cl]2, 2% – 30 Methanol (M)

RhW-16 RhB12W10MCM-M [RhI(CO)2Cl]2, 2% H3PW12O40, 10% 30 Methanol (M)

Rh-17 RhC12MCM-M RhICl(PPh3)3, 2% – 30 Methanol (M)

RhW-18 RhC12W10MCM-M RhICl(PPh3)3, 2% H3PW12O40, 10% 30 Methanol (M)

RhMo-19 RhA12Mo10MCM-M [RhI(COD)Cl]2, 2% H3PMo12O40, 10% 30 Methanol (M)

Rh-20 RhC15MCM-E HRhCO(PPh3)3, 5% – 30 Ethanol (E)

RhW-21 RhC15W10MCM-E HRhCO(PPh3)3, 5% H3PW12O40, 10% 30 Ethanol (E)

RhMo-22 RhC15Mo10MCM-E HRhCO(PPh3)3, 5% H3PMo12O40, 10% 30 Ethanol (E)

Table 3

Determination of the pore diameters of the supported catalysts by XRD

techniques

Entry Catalysts Estimated pore diameters (A)

1 MCM-41 30.0

2 Mo10MCM-41-E 33.9

3 Rh15MCM-E 32.1

4 Rh15Mo10MCM-E 33.9

5 Rh15W10MCM-E 32.2

6 Rh15W10MCM-E cycle I 34.3

7 Rh15W10MCM-E cycle II 33.9

8 Rh15W10MCM-E cycle III 36.3

9 Rh15W10MCM-E cycle IV 34.3

10 Rh35MCM-E 34.3

11 Rh35W10MCM-E 32.4

12 Rh35Mo10MCM-E 29.4

at 2013 and 2083 cm�1 which probably correspond to two

gem carbonyl functions, which confirms the replacement of

PPh3 ligand by CO. Similar assignments were also made in

previous workers [15,20,21].

After the hydroformylation reaction of styrene, two new

bands were observed to developed, 2040 and 1661 cm�1.

The two peaks are tentatively attributed to Z–O–

RhCO(PhCHCH2) and Z–O–Rh(Ph(CH3)COCH3) (Z = sup-

port), respectively [22].

However, all the catalysts at the end of several cycles has

an additional new absorption, at 1712 cm�1. It occurs in a

frequency range characteristic of Rh–CO–Rh bridging with

C–O stretching. This datum logical requires that, in the

species giving rise to the 1712 cm�1 band, two Rh centers be

less than �3 A apart [23].

2.2.3.2. 29Si MAS NMR. Spectra of W10MCM-41 and

Rh15W10MCM-E at 99.44 MHz as shown, revealed almost

no variation in the Q3 ((SiO3)3BBSi–OH) and Q4

((SiO)3BBSi–O–SiBB) chemical shifts (d = �100 ppm and

d = �110 ppm, respectively) although, there was a sub-

stantial change observed in the Q3/Q4 ratio. A possible

reason for the lower value of Q3/Q4 might be the decrease in

the population of free silanol group due to the linkage arising

from anchoring with HPA [4].

2.2.3.3. 31P CP-MAS NMR. The catalyst was characterized

using 31P CP-MAS NMR. The spectra of HRh(CO)(Ph3)3

complex have earlier been reported [4,15] to have two major

peaks (diso 34.4, 45.4 ppm). Each split by J-coupling to other31P and 103Rh (natural abundance of 100%) nuclei. The

spectra pattern observed in HRh(CO)(PPh3)3 thus elucidated

the trigonal bipyramidal of the complex. The results of

the 31P spectra of HPA-W12 and rhodium complex,

Rh15W10MCM-E showed a totally different spectrum with

peaks (diso 31.23 and 45.36 ppm). This shift in the NMR

spectra of Rh15W10MCM-E catalyst may be due to the

interaction of HPA-W12 and HRh(CO)(PPh3)3 tethered to

MCM-41 matrix, where Rh–O–W–O–Si type of bond

formation may have taken place.

2.2.3.4. XRD. The catalysts materials (mesoporous phases

of MCM-41, HPA–MCM-41, Rh–HPA–MCM-41 and Rh–

HPA–MCM-41 catalysts after various recycles) were

characterized and analyzed by powder XRD patterns and

also calculate the pore diameters (Table 3). Bragg peaks

were indexed assuming a hexagonal symmetry and pore wall

thickness of 10 A. The pore diameters can be approximated

by the formula:

ð2=ð3Þ1=2Þd1 0 0 � 10 A

The distinct reflection and crystallinity or morphology

(determined by the peak positions) remains unaltered from

one another, as observed by the powder XRD. This shows

that there is no change in crystallinity or morphology of

MCM-41 embedded with HPA compared to that of the neat

MCM-41 and the porous framework the Rh–HPA–MCM-41

catalyst was not affected or damage during the complex

formation.

2.3. General procedure for the heterogeneous

hydroformylation of styrene

The following general procedure was considered for

styrene adopted as model substrates for aryl alkenes in the

hydroformylation reaction. Fifty or 20 mg of the supported

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196 189

catalyst were weighted into a glass liner equipped with

magnetic bar using weighing paper. Five milliliters of dry

solvent added to the liner. Five millimoles (=0.5208 g) of

styrene or 5.0 mmol was added, and the liner placed into

45 ml high-pressure Parr autoclave. The autoclave was

flushed thoroughly three times with carbon monoxide,

subsequently pressurized with 300 psi of CO and 300 psi of

hydrogen. The autoclaves were heated on hotplates

controlled by temperature sensor to maintain the tempera-

ture constant (�0.5 8C). The rate of stirring was maintained

at 750 rpm. After the reaction time elapses, the autoclave

was cooled to room temperature and the gas carefully

released under fume hood. The reaction mixture was

centrifuged twice to separate the solid catalyst from the

liquid products. The solid catalyst was dried in oven at 60 8Cand used again in another catalytic cycle. One hundred

micrograms of anisole were added to mixture as internal

standard to determine the conversion of styrene. The

products were identified using GC, GC–MS, 1H NMR and13C NMR techniques. The ratio of branched (B) to linear

aldehydes (L) was determined by GC and 1H NMR.

3. Results and discussion

The heterogeneous hydroformylation reaction represents

an attractive industrial route for the production of aldehydes

due to the relatively easy separation of products and high

turnover of the rhodium catalysts. Many heterogeneous

catalytic systems have been described in the literature, but to

our knowledge, the combination of the heteropolyacids as

co-catalysts with the rhodium(I) or rhodium(III) complexes

Table 4

Rhodium(III) supported catalysts in the heterogeneous hydroformylation of styre

Entry Catalyst Time (h)

1 Rh32MCM-M 6

2 Rh32MCM-M 16

3d Rh32MCM-M 16

4 Rh32W10MCM-M 6

5 Rh32W10MCM-M 16

6d Rh32W10MCM-M 16

7 Rh32MCM-W 16

8 Rh32W10MCM-W 16

9 Rh32Mo10MCM-M 6

10 Rh32Mo10MCM-M 16

11e Rh32W10MCM-M 6

12e Rh32W10MCM-M 16

13 Rh32MCM0-M 16

14 Rh32W10MCM0-M 16

a Reaction conditions: catalyst (10 mg), styrene (0.5208 g = 5.0 mmol), T

MCM = MCM-41 (30 A), MCM0 = MCM-41 (50 A).b Determined by GC.c Determined by GC and 1H NMR.d No water was used.e PPh3 (0.04 mmol) was added into the liquid phase.

as catalysts impregnated on MCM-41 has not been yet well

explored in the hydroformylation of alkenes (Eq. (1)):

(1)

3.1. Rhodium(III) supported catalysts in the heterogeneous

hydroformylation of styrene

We have chosen the MCM-41 as inorganic support due to

their high surface area and large pore size. RhCl3�3H2O (1–

2%) impregnated alone or with H3PW12O40 (HPA-W12) (5–

20%) on MCM-41 have been studied in the hydroformyla-

tion of styrene (Table 4). The conversions and the product

distribution were monitored at 6 and 16 h of the reaction.

Rh32MCM-M and Rh32W10MCM-M gave no reaction in

the absence of water (Table 4, entries 3 and 6). However,

Rh32MCM-M catalyzed successfully the hydroformylation

of styrene in the presence of 25 ml of water. The conversions

were 36% at 6 h and 76% at 16 h of the reaction (Table 4,

entries 1–2). The catalyst Rh32W10MCM-M gave excellent

conversion equals to 98% at 16 h compared to 76%

conversion at 6 h but only in the presence of 25 ml of

water (Table 4, entries 4 and 5). It was clear at this stage that

the presence of water was essential for the occurrence of the

reaction at 80 8C. The catalysts Rh32MCM-W and

Rh32W10MCM-W, formed by the impregnation of rhodiu-

m(III) and the heteropolyacid in water, gave slightly lower

conversions of styrene (Table 4, entries 7–8). Interestingly,

when H3PMo12O40 replaced the HPA-W12 in the impreg-

nation process with rhodium(III), we were surprised to

nea

Conversionb (%) Product distributionc (%

B L

36 71 29

76 70 30

0 – –

74 71 29

98 69 31

0 – –

65 73 27

92 71 29

Traces – –

11 74 26

16 80 20

46 78 22

28 72 28

77 76 24

HF (5 ml), 600 psi (CO/H2 = 1/1), H2O (25 ml = 1.39 mmol), 80 8C

)

.

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196190

Fig. 1. Hydroformylation of styrene: effect of % loading of HPA-W12.

Reaction conditions: % loading of rhodium (2%), styrene (0.5208 g

= 5.0 mmol), THF (5 ml), H2O (25 ml = 1.39 mmol), CO (300 psi), H2

(300 psi), 80 8C, 16 h.

observe that the supported catalyst seriously inhibited the

reaction of hydroformylation of styrene (Table 4, entries 9–

10); only 11% of conversion was obtained after 16 h. On the

other hand, we have made serious attempts to improve the

selectivity of the reaction. For these reasons, we have used

triphenyl phosphine (PPh3) in the reaction catalyzed by

Rh32W10MCM-M (Table 4, entries 11–12). The selectivity

was slightly improved (78–80%) towards the branched

aldehyde B but the conversions dropped in a significant

manner to 16% at 6 h and 46% at 16 h of the reaction. The

use of the larger pore size support such as MCM-41 (50 A)

showed again the same advantages of MCM-41 (30 A) of the

Table 5

Rhodium(I) supported catalysts in the heterogeneous hydroformylation of styren

Entry Catalyst Times (h)

1 RhA12MCM-M 6

2 RhA12MCM-M 16

3 RhA12W10MCM-M 6

4 RhA12W10MCM-M 16

5 RhB12MCM-M 6

6 RhB12MCM-M 16

7 RhB12W10MCM-M 6

8 RhB12W10MCM-M 16

9 RhC12MCM-M 6

10 RhC12MCM-M 16

11 RhC12W10MCM-M 6

12 RhC12W10MCM-M 16

13d RhA12MCM-M 16

14e RhA12MCM-M 16

15 RhA12W10MCM0-M 16

16 RhC15W10MCM-E 16

a Reaction conditions: catalyst (10 mg), styrene (0.5208 g = 5.0 mmol),

MCM0 = MCM-41 (50 A).b Determined by GC.c Determined by GC and 1H NMR.d PPh3 (0.04 mmol) was initially added into the liquid phase.e PPh3 (0.08 mmol) was initially added into the liquid phase.

impregnation of the rhodium(III) complex combined with

HPA-W12 compared to the rhodium complex impregnated

alone on the support (Table 4, entries 13–14). Rh32MCM0-M (2% of RhCl3�3H2O supported on MCM-41) gave 28% of

conversion after 16 h compared to 77% with the catalysts

Rh32W10MCM0-M prepared by supporting 2% of

RhCl3�3H2O and 10% of HPA-W12 on MCM-41 (50 A).

When the catalyst Rh32W5MCM-M formed of 2% of

RhCl3�3H2O and impregnated with 5% of HPA-W12 on

MCM-41 in methanol, was used in the hydroformylation of

styrene, the conversion was improved to 48% at 6 h and 85%

at 16 h of the reaction (Fig. 1) The increase of the percentage

of loading of HPA-W12 from 10 to 25% (15, 20, and 25%)

and fixing the percentage of rhodium at 2% decreased the

conversions of the reaction from 98 to 27% (82, 45, and

27%, respectively) (Fig. 1). The selectivity toward the

branched aldehyde was independent to the % loading and

was obtained almost the same in all experiments.

The results summarized in Table 4 and Fig. 1 showed the

advantages of the use of the mesoporous materials, such as

MCM-41, as supports for the impregnation of rhodium

catalysts for the hydroformylation reactions. In addition, the

combination of both RhCl3�3H2O (2%) and HPA-W12 (10%)

in the presence of water strongly enhanced the catalytic

activity of the rhodium catalyst in the hydroformylation of

styrene.

3.2. Rhodium(I) supported catalysts in the heterogeneous

hydroformylation of styrene

Similar to the previous study, we have explored the

supported catalysts obtained by the impregnation of

ea

Conversionb % Product distributionc (%)

B L

43 73 27

76 72 28

70 72 28

97 71 29

51 71 29

82 70 30

71 70 30

95 71 29

Traces – –

20 80 20

45 75 25

85 72 28

53 74 26

45 76 24

74 72 28

95 78 22

THF (5 ml), 600 psi (CO/H2 = 1/1), 80 8C, MCM = MCM-41 (30 A),

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196 191

rhodium(I) complexes combined to HPA-W12 and phos-

phine ligands on MCM-41 (30 A) as a support. The results

are summarized in Table 5. Among rhodium(I) complexes,

[Rh(COD)Cl]2 (denoted as RhA) was supported with and

without HPA-W12 on MCM-41. The catalyst RhA12MCM-

M, obtained by the impregnation of 2% of the rhodium(I)

complex RhA alone on the mesoporous support, catalyzed

the hydroformylation of styrene with conversions of 43% at

6 h and 76% at 16 h (Table 5, entries 1–2). A significant

improvement in the conversion of styrene with the catalyst

RhA12W10MCM-M, formed by the impregnation 2% of

RhA and 10% of HPA-W12 on MCM-41 was noticed. The

conversion increased to 70% at 6 h and to 97% at 16 h with

the selectivity toward branched aldehyde B maintained

almost the same (71–72%) (Table 5, entries 3–4). Another

rhodium(I) complex [Rh(CO)2Cl]2 (denoted as RhB) was

also considered in the impregnation process with and

without HPA-W12 (2% rhodium complex and 10% HPA-

W12) to give the catalysts RhB12MCM-M and

RhB12W10MCM-M. The results of the catalytic hydro-

formylation tests with these two catalysts (Table 5, entries 5–

8) were similar to those obtained with RhA based catalysts,

but the catalytic activity of the catalysts obtained with RhA

was slightly higher. Surprisingly, the rhodium catalyst

HRhCO(PPh3)3 (denoted RhC) impregnated alone on MCM-

41 in methanol gave a supported catalyst that was not very

active in the hydroformylation of styrene; only traces of

products were obtained after 6 h of reaction and the

conversion increased to 20% after 16 h (Table 5, entries 9–

10). It is clear that the reaction was very slow and went

through a prolongeted induction period. However, the

activity of the catalyst obtained from the impregnation of 2%

of RhC with 10% of HPA-W12 was much higher; the

conversion reached 71% after 6 h and 95% after 16 h

(Table 5, entries 11–12). It seems that a very active

intermediate complex was formed between RhC and HPA-

W12 at the surface of the support. However, the small amount

of the impregnated catalyst and probably the instability of

the active rhodium intermediate species made the identifica-

tion and the characterization of these species difficult. In

order to improve the selectivity of the reaction, we have

Table 6

Heterogeneous hydroformylation of styrene: effect of the type of catalyst based

Run Catalyst Time (h)

1 Rh15Mo10MCM-E (50 mg) 3

2 Rh15W10MCM-E (50 mg) 3

3 Rh15MCM-E (50 mg) 3

4 Rh35Mo10MCM-E (50 mg) 6

5 Rh15Mo10MCM-E (20 mg) 6

6 Rh15MCM-E (20 mg) 6

7 Rh15Mo10MCM-E (20 mg) 6

8 Rh15MCM-E (20 mg) 6

a Styrene (0.5208 g = 5.0 mmol), THF (5 ml), 600 psi (CO/H2 = 1/1), cycloheb Determined by GC.c Determined by GC and 1H NMR.

designed some experiments for which 10 mg of the catalyst

RhA12MCM-M and 0.02–0.08 mmol PPh3 were added in

THF (Table 5, entries 13–14). The conversions were low

(45–53%). The use of another mesoporous material such

MCM-41 (50 A) as a support did not show a better catalytic

activity for the catalyst RhA impregnated with HPA-W12

(Table 5, entry 15); the conversion of the reaction was

74%. In order to study the recycling process and the

characterization of the catalyst, another catalyst has been

prepared by impregnating 5% of HRhCO(PPh3)3 (denoted

RhC) with 10% of HPA-W12 on MCM-41 in ethanol as a

solvent of impregnation. The resulted supported catalyst

RhC15W10MCM-E showed high catalytic activity (Table 5,

entry 21). The study of the hydroformylation reaction

catalyzed by supported rhodium(I) catalysts on MCM-41

showed high sensitivity towards the water where no

conversion of styrene was observed in the presence of water.

Another class of rhodium catalysts with higher load (5%)

of rhodium have been prepared and tested in the

hydroformylation of styrene. For example, Rh15MCM-E,

and Rh15W10MCM-E and Rh15Mo10MCM-E, which was

prepared by refluxing 5% of HRhCO(PPh3)3 in MCM-41

functionalized by HPA-W12 and HPA-Mo12 in ethanol, were

evaluated for their activity and selectivity in hydroformyla-

tion of styrene. Different amount of catalyst (20–50 mg)

were also used the catalytic run. The catalyst based on

rhodium(III), Rh35Mo10MCM-E, was also considered for

comparison. The results are summarized in Table 6. High

conversions (94–99%) were obtained at 110 8C after 3–6 h

in cyclohexane as a solvent. However, the selectivity of the

reaction was relatively lower (B/L = 1.3). The low

selectivity is related to the high temperature of the reaction,

which is well known phenomena in the chemistry of the

hydroformylation.

3.3. Hydroformylation of styrene. effect of the amount of

water on Rh(III) based catalysts

The addition of water into the liquid phase of the reaction

catalyzed by Rh32W10MCM-M, obtained by the impreg-

nation of RhCl3�3H2O with HPA-W12 on MCM-41, has an

on rhodium(I) or rhodium(III)a

Conversionb (%) Product distributionc (%)

B L

97 57 43

97 59 41

99 54 46

98 58 42

98 63 37

99 86 14

94 61 39

97 60 40

xane (5.0 ml), 110 8C. MCM = MCM-41 (30 A).

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196192

Fig. 3. Hydroformylation of styrene: effect of the temperature. Reaction

conditions: Rh32W10MCM-M (10 mg), styrene (0.5208 g = 5.0 mmol),

water (25 ml = 1.39 mmol), THF (5 ml), CO (300 psi), H2 (300 psi), 6 h.Fig. 2. Hydroformylation of styrene: effect of the volume of water (ml;

5 ml = 0.278 mmol). Reaction conditions: Rh32W10MCM-M (10 mg),

styrene (0.5208 g = 5.0 mmol), THF (5 ml), CO (300 psi), H2 (300 psi),

80 8C, 6 h.

enormous influence on the hydroformylation of styrene

(Fig. 2). The amount of water added to the mixture was also

critical to the occurrence of the reaction. Only 15% of

conversion was achieved in the absence of water after 6 h of

reaction. The conversion increased to 60 and 70% with the

addition of 5 and 10 ml of water, respectively. Fig. 2 has a

maximum conversion of 74% obtained with 25 ml of water.

Interestingly, the increase in the amount of water led to a

serious decrease of the conversion of styrene to reach 40%

with 100 ml of water.

The possible role of water in the reaction was to allow the

reduction of Rh(III) into rhodium(I) in the presence of

carbon monoxide (Eq. (2)). This effect was described by

Primet et al. [24] and later by Davis and coworkers [25] and

it takes place at the surface of the silica as support:

RhðIIIÞ þ CO!RhðIIIÞ�CO �!2COþH2ORhþðIÞðCOÞ2

þ CO2 þ 2Hþ (2)

We also support the idea that rhodium(III) carbonyl complex

was formed on the surface of the MCM-41, which was

reduced in the presence of water to rhodium(I) carbonyl

intermediate. This later represents the most probable active

catalytic species in the hydroformylation reaction of styrene.

This reduction was produced on the surface and in the pores

of the mesoporous materials. The use of high concentration

of water inhibited the catalytic activity of the supported

catalyst. It seems that the use of the excess of water enhances

the conversion of Rh(I)(CO)2 into Rh0(CO)m and subse-

quently into rhodium cluster Rh6(CO)16 via Rh4(CO)12

(Eq. (3)) as reported previously [25,26]. These rhodium

clusters are not active catalytic complexes under the present

experimental conditions:

2RhðIÞðCOÞ2 �!COþH2O2Rh0ðCOÞm �!

CO=highH2O conc:

Rh4ðCOÞ12 �!2Rh0ðCOÞm

Rh6ðCOÞ16 (3)

On the other hand, we have found that part of the rhodium

catalyst was leaching into the liquid phase at the end of the

reaction, and the main catalytic hydroformylation might

take place in the homogenous solution. In addition, we also

believe that the reaction was also partially catalyzed by the

heterogeneous system.

3.4. Hydroformylation of styrene: effect of the temperature

with Rh(III) based catalyst

The effects of the temperature on the conversion and the

selectivity of the hydroformylation of styrene have been

studied with the catalyst Rh32W10MCM-M formed by the

impregnation of RhCl3�3H2O with HPA-W12 on MCM-41 in

methanol (M) (Fig. 3). The time of the reaction was fixed as

6 h. The results showed that the hydroformylation of styrene

catalyzed by Rh32W10MCM-M (Fig. 3) was relatively fast

and there was no induction period observed. With

Rh32W10MCM-M no reaction took place at 50 8C, and

the conversion at 60 8C was 8% with 90% selectivity toward

branched aldehyde. The conversion of the reaction increased

and the selectivity of the branched aldehyde dropped with

the increase of the temperature. At 100 and 120 8C the

conversions were 94 and 96% but the selectivities in

branched aldehyde decreased to 54 and 47%, respectively. It

is clear that the rate of the reaction and the selectivity are a

function of temperature; the selectivity deteriorated at the

end of the reaction.

3.5. Hydroformylation of styrene: effect of the type of

solvent

The catalyst Rh32W10MCM-M in the hydroformylation

of styrene showed higher catalytic activity mostly in polar

oxygenated solvent having ether function such as tetra-

hydrofurane (THF), 1,2-dimethoxyethane (DME), and

methyl tertiary butyl ether (MTBE) (Fig. 4). The reactions

were carried out at 80 8C for 6 h in the presence of 25 ml of

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196 193

Fig. 4. Hydroformylation of styrene by Rh32W10MCM-M: effect of the

type of solvent. Reaction conditions: Rh32W10MCM-M (10 mg), styrene

(0.5208 g = 5.0 mmol), water (25 ml = 1.39 mmol), CO (300 psi), H2

(300 psi), 80 8C, 6h.

water. The significant conversions were obtained using THF,

DME, and MTBE (74, 62, and 34%, respectively). On the

other hand, very low conversions of styrene (1–9%) were

detected with acetone, methyl ethyl ketone, hexane and 1,4-

dioxane as a solvent, and no reaction was observed in

dichloromethane (CH2Cl2), benzene, and toluene.

Similar solvent effect has been also observed with the

reaction catalyzed by the rhodium(I) RhA12W10MCM-M

(Fig. 5). The same solvents have been considered for the

reaction with RhA12W10MCM-M. Again, the highest

conversion of styrene was obtained with THF followed by

DME and MTBE as a solvent. Small or no conversions were

obtained with the other considered solvents (1,4-dioxane,

acetone, MEK, hexane, CH2Cl2, benzene, toluene). The

effect of the type of solvent in Figs. 4 and 5 can be also

explained in terms of dielectrical constant. The two solvents

that led to high conversions were THF and DME having

dielectrical constant of 7.52 and 7.20, respectively. The

Fig. 5. Hydroformylation of styrene by RhA12W10MCM-M: effect of the

type of solvent on the conversion and the selectivity. Reaction conditions:

catalyst RhA12W10MCM-M (10 mg), styrene (0.5208 g = 5.0 mmol), CO

(300 psi), H2 (300 psi), 80 8C, 6 h.

solvent of low (2.02–2.8) and high (9.08–20.70) dielectrical

constants gave very low conversion of styrene (2–28%). It is

very clear that the type of solvent has an enormous effect on

the occurrence of the reaction. THF appeared to increase the

catalytic activity and the stability of the key intermediate

species [Rh(I)–HPA-W12].

3.6. Hydroformylation of styrene derivatives by supported

Rh(I) and Rh(III) catalysts

The reactions of the hydroformylation of styrene

derivatives were carried out in the presence of the catalytic

system RhA12W10MCM-M/CO/H2 in THF as a solvent

(Table 7, Eq. (4)). Excellent conversions (95–99%) were

obtained in general after 16 h with mono-, di- and

trisubstituted benzene ring of the styrene with no major

change in the selectivity in the corresponding branched

aldehyde (69–72%) (Table 7, entries 1–12). The styrene

derivatives with electron withdrawing groups (chloro, or

nitro) in ortho, para, or meta position, were also very

reactive in the hydroformylation reaction. The conversions

were very high 88–96% with the selectivities maintained

again at the same level (68–73%) (Table 7, entries 13–18).

Interestingly, 2-vinylnaphthalene was almost totally

converted (99%) with higher selectivity (79%) toward the

branched aldehyde B. It is important to mention that no

hydrogenated by-products were detected with all previous

substrates.

(4)

3.7. Recycling of the supported catalysts in the

hydroformylation of styrene

Various heterogeneous catalysts were synthesized and

tested in successive hydroformylation reactions of styrene to

investigate the effect of the support and anchoring agents on

the recycling ability of the catalysts.

Hydroformylation of styrene proceeded in the presence

of the catalytic amount of Rh15Mo10MCM-E affording a

good yield and selectivity for the branched and linear

aldehydes, even in the absence of free PPh3 or any other

ligand often used in homogeneous or immobilized systems

[26]. The results are presented in Fig. 6. The conversions of

over 95% were achieved in 3 h for five cycles with branched

to linear ratio remaining fairly constant (55/45), while

cycles 6 and 7 required a little more time (9 and 12 h,

respectively) for the same conversion, while the selectivity

remain the same. It is very clear that the catalyst system is

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196194

Table 7

Hydroformylation of various aryl alkenes by RhA12W10MCM-M or Rh32W10MCM-M with CO/H2 in THFa

No. Substrate 1 Catalyst Conversionb (%) Product distributionc (%), B/L

1 Rh(I) 97 69/31

2 Rh(III) 98 69/31

3 Rh(I) 94 74/26

4 Rh(III) 99 72/28

5 Rh(I) 95 73/27

6 Rh(III) 95 70/30

7 Rh(I) 95 75/25

8 Rh(III) 95 72/28

9 Rh(I) 97 74/26

10 Rh(III) 97 70/30

11 Rh(I) 95 70/30

12 Rh(III) 95 71/29

13 Rh(I) 98 71/29

14 Rh(III) 96 73/27

15 Rh(I) 92 70/30

16 Rh(III) 96 73/27

17 Rh(I) 88 69/31

18 Rh(III) 64 68/32

19 Rh(I) 98 72/28

20 Rh(III) 64 68/32

a Reaction conditions: catalyst (10 mg), styrene (5.0 mmol), THF (5 ml), 600 psi (CO/H2 = 1/1), 80 8C, 16 h. Catalyst: RhA12W10MCM-M [Rh(I)] or

Rh32W10MCM-M [Rh(III)].b Determined by GC.c Determined by GC and 1H NMR.

not leached out in the first cycle, and hence can be recycled

at least seven times.

To investigate the role of heteropolyacid as tethered

agent, Rh15MCM-E was prepared by refluxing the same 5%

of HRhCO(PPh3)3, and MCM-41 in dry ethanol, but without

Fig. 6. Hydroformylation of styrene: the recycling of the catalyst

Rh15Mo10MCM-E. Reaction conditions: Rh15Mo10MCM-E (50 mg),

styrene (0.5208 g = 5.0 mmol), cyclohexane (5.0 ml), CO (300 psi), H2

(300 psi), 110 8C.

heteropolyacid. The catalyst was tested under the same

above reaction conditions with Rh15MCM-E and the results

are shown in Fig. 7. The conversion was again very high

(>95%) and was achieved in 3 h for the first three cycles, but

for the cycles 4 and 5, the reaction has to be run for 12 and

Fig. 7. Hydroformylation of styrene: the recycling of the catalyst

Rh15MCM-E. Reaction conditions: Rh15MCM-E (50 mg), styrene

(0.5208 g = 5.0 mmol), cyclohexane (5.0 ml), 110 8C, CO (300 psi), H2

(300 psi).

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196 195

Fig. 9. Hydroformylation of styrene: the recycling of the catalyst

Rh15SiW10MCM-E. Reaction conditions: Rh15SiW10MCM-E (50 mg),

styrene (0.5208 g = 5.0 mmol), cyclohexane (5.0 ml), 110 8C, CO (300 psi),

H2 (300 psi).

Fig. 10. Hydroformylation of styrene: the recycling of the catalyst

Rh35W10MCM-E. Reaction conditions: Rh35W10MCM-E (50 mg), styr-

ene (0.5208 g = 5.0 mmol), cyclohexane (5.0 ml), 110 8C, CO (300 psi), H2

(300 psi).

Fig. 8. Hydroformylation of styrene: the recycling of the catalyst

Rh15W10MCM-E. Reaction conditions: Rh15W10MCM-E (50 mg), styr-

ene (0.5208 g = 5.0 mmol), cyclohexane (5.0 ml), 110 8C, CO (300 psi), H2

(300 psi).

20 h to achieve similar conversion. A clear indication faster

leaching out of this catalyst could be seen in the fifth cycle,

were less than 40% conversion was achieved after 20 h of

reaction. The almost was almost deactivated after the fifth

cycle. The branched to linear ratio was almost the same as in

the previous experiment. These results indicate clearly the

effect of the heteropolyacid in the formation of stable and

active supported catalysts.

The catalysts prepared by impregnation rhodium com-

plex RhA with different heteropolyacids of Keggin structure

such as H3PW12O40 (Fig. 8) and H4SiMo12O40 (Fig. 9) were

studied. The catalyst that incorporates phosphotungstic acid,

H3PW12O40, was found to be as active as the catalyst made

of H3PMo12O40. The supported catalyst Rh15W10MCM-E

was still very active even after running the catalyst for seven

cycles. Earlier reports have shown that phosphotungstic acid

to be very effective tethering agent in hydroformylation

[15,24] and hydrogenation of functionalized and unfunctio-

nalized alkene, alkynes and aldehydes [14].

Si was used as central atom in place of P in HPA-W12 to

produce the supported catalyst Rh15SiW10MCM-E. Sur-

prisingly, the use of Si reduced the recycling ability of the

catalyst even when the reaction is ran at 110 8C. Cycle 4

gave less than 60% even after running the experiment for

40 h, but the branched-linear ratios (61/39) were slightly

higher than for Rh15Mo10MCM-E (55/45) ran under the

same conditions (110 8C) (Fig. 9). The lower recycling

potential of the Rh15SiW10MCM-E is probably related to

the Si–O–Si interaction between HPA and MCM-41 or the

possible decomposition of HPA-W12 on the surface of the

catalyst subsequently lowering the catalytic activity of the

catalyst. We believe that the HPA–MCM-41 interaction goes

through W–Si bond formation, which probably affects the

recycling ability of the catalyst.

The reaction of hydroformylation of styrene was also

studied using the supported system obtained from Rh(III)–

HPA-W12–MCM-41 (Fig. 10). It is important to note again

that 5% (27.1 mg) of RhCl3�3H2O were refluxed with

500 mg MCM-41 already functionalized with 10%

(56.8 mg) of H3PW12O40.

The activity of the catalysts based on Rh(III), as expected,

is generally lower than Rh(I) based catalyst. For example,

cycle 1 with Rh35W10MCM-E gave 99% conversion with

61/39 branched to linear ratio after 20 h reaction time

compared to 3 h in the case of Rh15W10MCM-E. In the

same way cycles 2–6 was ran for 20, 20, 30, 40 and 40 h,

respectively, to obtained a significant amount of yield

compared to 4, 6, 6, 20 and 20 h for Rh15W10MCM-E

catalyst.

4. Conclusion

The study of the hydroformylation of styrene and styrene

derivatives by the heterogeneous catalysts led to interesting

results. The impregnated catalysts based on rhodium(III)

showed high catalytic activity in the presence of water.

B. El Ali et al. / Applied Catalysis A: General 283 (2005) 185–196196

Rh32W10MCM-M was among the catalysts that gave the

highest conversion at 80 8C. The amount of water added

should be controlled because the use of excess of water did

inhibit the reaction. The supported catalysts based on

rhodium(I) were also very active and produce high

conversions of styrene and styrene derivatives. The detailed

study of the heterogeneous catalyst recycling showed clearly

the role of the heteropolyacid in the supported catalyst. It is

important to note that the influence of the heteropolyacid

was in both catalyst systems that involve rhodium(I)

and rhodium(III). The choice of the solvent of impregnation

was very important in order to minimize the leaching.

Cyclohexane as a solvent does not enhance the leach-

ing of Rh(III) and rhodium(I) supported catalysts.

RhA15W10MCM-E catalyzed efficiently the hydroformyla-

tion of styrene with excellent conversion. It was very clear

that the presence of HPA-W12 or HPA-Mo12 with Rh(I) or

Rh(III) on the support increased the catalytic activity of

rhodium catalysts by improving the conversions of styrene

and the recycling of the supported catalyst.

Acknowledgment

We gratefully acknowledge King Abdulaziz City for

Science and Technology (KACST-Saudi Arabia) for the

financial support under the project KACST 18–15. We

equally thank King Fahd University of Petroleum and

Minerals (KFUPM-Saudi Arabia) for providing all support

to this project.

References

[1] P.W.N.M. Van Leeuwen, C. Claver, P.W. Van Leeuwen, Rhodium

Catalyzed Hydroformylation, Kluwer Academic Publishers, New

York, 2000;

B. Cornils, W.A. Herrmann (Eds.), Applied Homogeneous Catalysis

by Organometallic Complexes, VCH Verlag, Chernie, Weinheim,

1996.

[2] L. Marko, J. Organomet. Chem. 245 (1983) 309;

L. Marko, J. Organomet. Chem. 261 (1984) 485;

L. Marko, J. Organomet. Chem. 283 (1985) 221;

L. Marko, J. Organomet. Chem. 305 (1986) 333;

L. Marko, J. Organomet. Chem. 357 (1989) 481.

[3] P. Pino, J. Organomet. Chem. 200 (1980) 223;

L. Marko, F. Ungvary, J. Organomet. Chem. 432 (1992) 1;

F. Ungvary, J. Organomet. Chem. 457 (1993) 273;

P.W.N.M. van Leeuwen, G. van Koten, Stud. Surf. Sci. Catal. 79

(1993) 199.

[4] K. Mukhopadhyay, A.B. Mandale, R.V. Chaudhari, Chem. Mater. 15

(2003) 1766.

[5] W.A. Herrmann, B. Cornils, Angew. Chem. Int. Ed. 36 (1997)

1048.

[6] M.H. Valkenberg, W.F. Holdrich, Catal. Rev. 44 (2) (2002) 321.

[7] J.M. Coronado, F. Coloma, J.A. Anderson, J. Mol. Catal. 154 (2000)

143.

[8] S. Bischoff, M. Kant, Catal. Today 66 (2001) 183.

[9] Z.M. Michalska, K. Strzelec, J. Mol. Catal. 177 (2001) 89.

[10] J.A. Dıaz-Aunon, M.C. Roman-Martınez, C. Salinas-Martınez, C. de

Lecea, J. Mol. Catal. 170 (2001) 81.

[11] H. Gao, R.J. Angelici, Organometallics 17 (1998) 3063.

[12] Y. Wang, J. Jiang, Q. Miao, X. Wu, Z. Jin, Catal. Today 74 (2002)

85.

[13] O. Stenzel, H.G. Raubenheimer, C. Esterhuysen, J. Chem. Soc., Dalton

Trans. (2002) 1132.

[14] M.J. Burk, A. Gerlach, D. Semmerji, J. Org. Chem. 65 (2000) 8933.

[15] K. Mukhopadhyay, R.V. Chaudhari, J. Catal. 213 (2003) 73;

K. Mukhopadhyay, B.R. Sarkar, R.V. Chaudhari, J. Am. Chem. Soc.

124 (2002) 9692.

[16] G.A. Tsigdinos, C. Hallada, Inorg. Chem. 7 (1968) 437.

[17] A. Atlamsani, M. Ziad, J.-M. Bregeault, J. Chim. Phys. (1995) 1344.

[18] M. Grun, K. Unger, A. Matsumoto, K. Tsutsumi, Micropor. Mesopor.

Mater. 27 (1999) 207;

P. Jalil, M. Al-Daous, A. Al-Arfaj, A. Al-Amer, J. Beltramini, S. Barri,

Appl. Catal. 207 (2001) 159.

[19] Z. Seddegi, U. Budrthumal, A. Al-Arfaj, A. Al-Amer, S. Barri, Appl.

Catal. 225 (2002) 167.

[20] A.M. Trzeciak, J.J. Ziolkowski, Z. Jaworska-Galas, W. Mista, J. Mol.

Catal. 189 (2002) 203.

[21] B.E. Hanson, M.E. Davis, D. Taylor, E. Rode, Inorg. Chem. 23 (1984)

52.

[22] M. Lenarda, L. Storaro, R. Ganzerla, J. Mol. Catal. 111 (1996)

203.

[23] A.R. Siedle, W.B. Geaspm, R.A. Newmark, R.P. Skarjune, P.A. Lyon,

C.G. Markell, K.O. Hodgson, A.L. Roe, Inorg. Chem. 29 (1990)

1667.

[24] M. Primet, J.C. Vedrine, C. Naccache, J. Mol. Catal. 4 (1978) 411.

[25] E.J. Rode, M.E. Davis, B.E. Hanson, J. Catal. 96 (1985) 574.

[26] A.N. Ajjou, H. Alper, Mol. Online 2 (1998) 53.