You are on page 1of 4

Catalysis Communications 5 (2004) 643–646

www.elsevier.com/locate/catcom

Silica immobilized ruthenium catalyst used for carbon


dioxide hydrogenation to formic acid (I): the effect of
functionalizing group and additive on the catalyst performance
a,b
Yiping Zhang , Jinhua Fei a, Yingmin Yu a, Xiaoming Zheng a,*

a
Department of Chemistry, Institute of Catalysis, Xixi Campus, Zhejiang University, Hangzhou 310028, PR China
b
Department of Chemistry, Zhejiang Education Institute, Hangzhou 310012, PR China

Received 6 June 2003; accepted 9 August 2004


Available online 11 September 2004

Abstract

The amine functionalized silica immobilized ruthenium complexes used for CO2 hydrogenation to formic acid were prepared in
an in situ synthesis way. The formic acid turnover frequency of 1384 h 1 at 100% selectivity has been obtained in CO2-expanded
ethanol over Si–(CH2)3–NH(CH2)3CH3–Ru, while NEt3 used as base, which was significantly higher than homogeneous ones.
 2004 Elsevier B.V. All rights reserved.

Keywords: Immobilized; Ruthenium; Silica; Formic acid; CO2-expanded solvent

1. Introduction DMF could be reached 1860 h 1. The catalyst was pre-


pared by the co-condensation of RuCl2[P(CH3)2(CH2)2
Recently, the atom economic reaction of formic acid Si(OC2H5)3]3 with Si(OC2H5)4.
synthesis from carbon dioxide hydrogenation has at- In this paper, the ruthenium immobilized on func-
tracted much attention. The effective catalysts used for tionalized silica was used as a new catalyst precursor,
this reaction were homogeneous ruthenium complexes and then ruthenium complexes catalysts were formed
in published papers [1–3]. It was also reported [4–6] that in situ at the reaction process of formic acid syntheses
the formic acid production rate reaches turnover fre- from CO2 hydrogenation by the catalyst precursor and
quency (TOF) of 1400 h 1 with PH2/CO2(atm) = 80/130 triphenylphosphine ligand in CO2-expanded solvent.
when scCO2 (supercritical CO2) was used as both a reac- This approach offers the following advantages: (1) easy
tant and a solvent. But the separation and recovery of separation from products; (2) reusability; (3) high activ-
the catalyst is a major problem in homogeneous cataly- ity and selectivity; (4) simplification of the process, no
sis. The immobilization of homogeneous catalyst at the need for preparing ruthenium complexes beforehand.
surface of a support may be a favourable way to solve
this problem. The group of Baiker [7–9] employed
RuCl2[P(CH3)3]4 supported on sol–gel derived silica ma- 2. Experimental
trix as catalyst for the synthesis of N,N-dimethylforma-
mide (DMF) from CO2, H2 and NH(CH3)2, the TOF of The preparation of catalysts is as follows: firstly, a
mixture of 6 g silica and 20 mmol functionalizing agents,
* with the general formula (EtO)3Si(CH2)3–X, where X is
Corresponding author. Tel.: +86 571 882 73417; fax: +86 571 882
73283. a Cl atom, or an NH2 group, in refluxing toluene was
E-mail address: xmzheng@dial.zju.edu.cn (X. Zheng). stirred for 8 h, then 25 ml H2O was added, and stirred

1566-7367/$ - see front matter  2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.catcom.2004.08.001
644 Y. Zhang et al. / Catalysis Communications 5 (2004) 643–646

for 4 h. The mixture was filtered and washed several which contained such components: the solid catalyst, a
times with toluene, then dried at 200 C in vacuum. vapor and a liquid phase (see Fig. 1). A large amount
The functionalized solid was denoted as ‘‘Si’’–(CH2)3– of CO2 was present in the lower phase, so the volume
X. Secondly a mixture of ‘‘Si’’–(CH2)3–Cl, amine (n- of the lower phase was larger than the amount of liquid
butylamine, diethylamine, respectively), was refluxed in solvent added [6,11]. The lower phase could be intro-
toluene for 4 h, then filtered, washed and dried at 200 duced as CO2-expanded solvent as B. Subramaniam
C. Three types of amine functionalized support, de- et al. [12–14] term it. It is likely that the large amount
scribed as ‘‘Si’’–(CH2)3–NH2, ‘‘Si’’–(CH2)3–NH(CH2)3 of CO2 in the lower phase increased the solubility of
CH3 and ‘‘Si’’–(CH2)3–N(CH2CH3)2 were obtained. H2 in that phase compared to the solubility of H2 in liq-
Thirdly, an ethanol solution of RuCl3 was mixed with uid organic solvent, Thus making CO2-expanded solvent
‘‘Si’’–(CH2)3–NH2, ‘‘Si’’–(CH2)3–NH(CH2)3CH3 or a good medium for CO2 hydrogenation.
‘‘Si’’–(CH2)3–N(CH2CH3)2 under stirring. Then, fil- From Table 1, HCOOH selectivity of 100% was at-
tered, washed several times with ethanol, and dried at tained in all of the catalysts, which means that no
60 C in vacuum, which was denoted as A–Ru, B–Ru, byproducts formation under the conditions employed.
or C–Ru in turn. The Ru content measured by ICP- Comparison with the homogeneous system (entry 1),
AES was 0.380 · 10 2, 0.258 · 10 2, 0.290 · 10 2 g/g the activity of the immobilized catalyst was greatly en-
catalyst, respectively. hanced (entry 2, 3, 4), which can be ascribed to so-called
The reactions were performed in a stainless-steel 100 site-isolation [15]. Since attaching a catalyst to a support
ml autoclave. In a typical run, 0.6 g A–Ru (or B–Ru, in such a way, catalytic sites were dispersed and pre-
C–Ru), 20 ml solvent, triphenylphosphine ligand, base cluded their dimerization or oligomerization to less reac-
were successively introduced. H2 (4.0 MPa pressure) tive species and can, therefore, endowed them with
and then liquid CO2 was pumped into the autoclave unique activities compared to their homogeneous
using a syringe pump, the reaction was normally pro- counterparts.
ceeded at 80 C for 1 h, the total pressure is 16.0 MPa The best catalytic performance was achieved over B–
at reaction temperature. At the end of the reaction, Ru as catalyst precursor, and the HCOOH TOF reached
the reactor was cooled rapidly at first and then depressu- 1384 h 1, which is almost three times higher than that of
rized very slowly. The reaction solution collected was homogeneous system with RuCl3 as precursor. The
acidified by 10% HCl and the formic acid was esterified activity order of three types of immobilized ruthenium
completely for analysis [10]. The esterified sample was was as follows: A–Ru < C–Ru < B–Ru. The possible
analyzed both by GC–MS equipped with TC-wax col- reason for the different activity of catalyst may be that
umn and GC equipped with poropak QS column and the electron donating character of alkyl group bound
FID detector. The two results were consistent with each to N atom. The ability of electron-donating increased
other. as the sequence of H (in A), –(CH2CH3)2 (in C) and
–(CH2)3CH3 (in B), leading to an increase in the density
of electron cloud around N atom, this helps to stabilize
3. Results and discussion the Ru–N bond in accordance with the above sequence.
Consequently, the activity of three types of immobilized
The activity, expressed as HCOOH TOF of various ruthenium increased in proper sequence.
Ru catalyst precursors is summarized in Table 1. Table 2 shows the comparison of the results on A–Ru
Although the reaction conditions were above critical catalyst precursor in different bases.
point of CO2 (Tc: 31.3 C, Pc: 7.39 MPa), the reaction As shown in Table 2, the production rate of HCOOH
system was a multi-component and multiphase system, was high in the presence of NEt3, while in the absence of
base, no HCOOH was obtained. The reaction
CO2(g) + H2(g) M HCOOH(1) occurs with DG298 =
Table 1
Catalytic results of CO2 hydrogenation with various ruthenium
immobilized catalysts
Entry Catalyst PPh3/Ru HCOOH
(mol ratio)
TOF (h 1)a Selectivity (%)
1 RuCl3 + PPh3 10:1 462 100
2 A–Ru + PPh3 10:1 656 100
3 B–Ru + PPh3 10:1 1384 100
4 C–Ru + PPh3 10:1 868 100
Reaction conditions: reaction temperature, 80 C; H2 pressure, 4.0
MPa; total pressure, 16.0 MPa; reaction time, 1 h; ligand, PPh3; sol-
vent, 20 ml ethanol; base, 5 ml NEt3; stirring rate, 250 r/min. Fig. 1. Schematic diagram of the phase behavior of reaction system
a
TOF: turnover frequency, mol of HCOOH/mol of Ru/h. using immobilized catalyst in CO2-expanded ethanol.
Y. Zhang et al. / Catalysis Communications 5 (2004) 643–646 645

Table 2 Table 3
The catalytic results on A–Ru precursor in different bases The catalytic results on A–Ru precursor in different solvents
Entry Catalyst Base HCOOH Entry Catalyst Solventa HCOOH
1
TOF (h ) Selectivity (%) TOF (h 1) Selectivity (%)
1a A–Ru + PPh3 – 0 0 1 A–Ru + PPh3 Methanol 933 100
2a,b A–Ru + PPh3 NEt3 656 100 2 A–Ru + PPh3 Ethanol 656 100
3c,d A–Ru + PPh3 NH2CONH2 523 100 3 A–Ru + PPh3 Isopropanol 74 100
4c,d A–Ru + PPh3 Na2CO3 227 100 4 A–Ru + PPh3 DMSOb 93 100
5c A–Ru + PPh3 C2H5ONa 68.5 100 5 A–Ru + PPh3 THFc 0 0
6c,d A–Ru + PPh3 CH3COONa 93 100 6 A–Ru + PPh3 toluene 0 0
Reaction conditions: reaction temperature, 80 C; H2 pressure, 4.0 Reaction conditions: reaction temperature, 80 C; H2 pressure, 4.0
MPa; total pressure, 16.0 MPa; reaction time, 1 h; ligand, PPh3; PPh3/ MPa; total pressure, 16.0 MPa; reaction time, 1 h; ligand, PPh3; PPh3/
Ru mol ratio, 10:1; stirring rate, 250 r/min. Ru mol ratio, 10:1; base, 5 ml NEt3; stirring rate, 250 r/min.
a a
Solvent: 20 ml ethanol. Solvent: 20 ml.
b b
NEt3: 5 ml. DMSO: dimethylsulphoxide.
c c
Wbase: 1.8 g. THF: tetrahydrofuran.
d
Solvent: 16.7 ml ethanol + 3.3 ml H2O.

+32.9 KJ mol 1, which is thermodynamically unfavour- NEt3, the most efficient base we tested, may be also used
able, and the equilibrium conversion of CO2 at 3.0 MPa, as reducing agent for the ruthenium complex formed in
100 C and 6.0 MPa, 100 C is 0.72 · 10 5 and situ [17].
0.14 · 10 4, respectively [16]. The equilibrium can be Table 3 shows the comparison of the results on A–Ru
shifted to the right by the addition of a base. The formic catalyst precursor in different solvents.
acid and formate ester could be detected when the reac- As shown in Table 3, no HCOOH could be observed
tion solution collected was acidified and esterified com- in non-protonic solvent, such as THF and toluene (entry
pletely. But neither formic acid nor formate ester could 5, 6). But the hydrogenation of CO2 to formic acid can
be detected in the reaction solution before being acidi- occur in a non-protonic DMSO solution (entry 4). The
fied to all of the reactions in which base was added. It reason may be that the strong polarity of DMSO, these
suggested that the product formic acid was converted solvents dissolved extremely large amounts of CO2 than
to a formate salt in the presence of a base. other apolar solvents [18]. Addition of protonic solvent
The purpose of adding water in entry 3, 4 and 6 was accelerated the reaction. The highest HCOOH TOF (933
to cause the base dissolve in the solution. Table 2 shows h 1) was achieved in methanol, the catalytic activity de-
that the use of organic bases (entry 2) resulted in much creased in the order of methanol, ethanol and isopropa-
higher activity as compared to that of inorganic bases nol (entry 1, 2, 3), consistent with the order of steric
(entry 3, 4, 5, 6). The activity was in the order of hindrance effect of alkyl group in ROH molecular.
NEt3 > NH2CONH2 > Na2CO3 > CH3COONa > C2H5 The smallest steric bulk of methanol made it the most
ONa, which was independent on the order of basicity. efficient alcohol for CO2 hydrogenation, while isopropa-
These results suggest that the presence of base was not nol with large steric bulk resulted in only 93 h 1 of TOF.
only served to neutralize HCOOH, but also played a After the reaction was completed, the catalyst can
crucial role in the catalytic performance of reaction. readily be recovered by filtration, hence, the catalyst

Table 4
The catalytic results of formic acid production with reused catalysts
Entry Catalyst PPh3/Ru (mol ratio) HCOOH
TOF (h 1)a Selectivity (%)
1 A–Ru 0 0 0
2 A–Ru + PPh3 10:1 656 100
3 Recycle-1b 0 612 100
4 Recycle-2b 0 600 100
5 B–Ru 0 0 0
6 B–Ru + PPh3 10:1 1384 100
7 Recycle-1b 0 1263 100
8 Recycle-2b 0 1239 100
Reaction conditions: reaction temperature, 80 C; H2 pressure, 4.0 MPa; total pressure,16.0 MPa; reaction time, 1 h; ligand, PPh3; solvent, 20 ml
ethanol; base, 5 ml NEt3; stirring rate, 250 r/min.
a
TOF: turnover frequency, mol of HCOOH/mol of Ru.
b
The recycled catalyst was the catalyst in the above entry.
646 Y. Zhang et al. / Catalysis Communications 5 (2004) 643–646

can be reused (see Table 4). A worthy noting detail Acknowledgements


was that no formic acid was formed in the absence
of triphenylphosphine ligand (entry 1, 5), while the This work was financially supported by the Natural
synthesis of formic acid only occurred with the addi- Science Foundation of China (No. 20173048) and the
tion of triphenylphosphine (entry 2, 6). The reused specialized research fund for the Doctoral program of
catalysts, formic acid could be obtained while the higher education (No. 20030335068).
recycle was performed in the absence of triphenyl-
phosphine ligand (entry 3, 4, 7, 8). It was attributed
to the formation of the catalytic active species from References
the immobilized ruthenium and triphenylphosphine
ligand under reaction conditions in the first run. [1] D.J. Darensbourg, C. Ovalles, M. Pala, J. Am. Chem. Soc. 105
Therefore, besides the high activity of the catalyst, it (1983) 5937.
[2] D.J. Darensbourg, C. Ovalles, J. Am. Chem. Soc. 106 (1984)
is also a new, clean, simple and highly efficient
3750.
method for the synthesis of complexes catalyst. [3] D.A. Palmer, R.V. Eldik, Chem. Rev. 83 (1983) 651.
From Table 4, it can be found that 600 and 1239 h 1 [4] P.G. Jessop, T. Ikariya, R. Noyori, Nature 368 (1994) 231.
of HCOOH TOF was maintained after catalyst A–Ru [5] P.G. Jessop, T. Ikariya, R. Noyori, Science 269 (1995) 1065.
and B–Ru was used three times, respectively. [6] P.G Jessop, Y. Hsiao, T. Ikariya, R. Noyori, J. Am. Chem. Soc.
118 (1996) 344.
The moderate loss of activity was due to the inevita-
[7] O. Kröcher, R.A. Köppel, A. Baiker, J. Chem. Soc., Chem.
bly loss of catalyst in the process of filtration. Because Commun. (1997) 453.
we used the filtrate in the further reaction as Sheldon [8] O. Kröcher, R.A. Köppel, M. Fröba, A. Baiker, J. Catal. 178
noted [19], and the filtrate exhibited no activity, indicat- (1998) 284.
ing that the leaching of ruthenium from amine-modified [9] O. Kröcher, R.A. Köppel, A. Baiker, J. Mol. Catal. 140 (1999)
185.
silica in each run was almost negligible, hence the recy-
[10] J.Z. Zhang, Z. Li, H. Wang, C.Y. Wang, J. Mol. Catal. 112
cling of the catalyst was possible. (1996) 9.
[11] P.E. Savage, S. Gopalan, T.I. Mizan, C.J. Martino, E.E. Brock,
AIChE J. 41 (1995) 1723.
[12] G. Music, M. Wei, B. Subramaniam, D.H. Busch, Coord. Chem.
4. Conclusion Rev. 219 (2001) 789.
[13] B. Kerler, R.E. Robinson, A.S. Borovik, B. Subramaniam, Appl.
In conclusion, amine functionalized silica immobi- Catal. B: Environ. 49 (2004) 91.
lized ruthenium complexes for the formation of formic [14] B. Subramaniam, C.J. Lyon, V. Arunajatesan, Appl. Catal. B:
Environ. 37 (2002) 279.
acid in CO2-expanded solvent have been developed [15] B. Pugin, M. Müller, Stud. Surf. Sci. Catal. 78 (1993) 107.
with an in situ synthesis way. The functionalizing [16] F.H. Cao, D.H. Liu, Q.S. Hou, D.Y. Fang, J. Nat. Gas. Chem.
group, the additives such as the base, the solvent, 10 (2001) 24.
influenced the performance of the catalyst. And the [17] J.Z. Zhang, Z. Li, H. Wang, C.Y. Wang, J. Mol. Catal. 112
immobilized catalysts were superior to the homogene- (1996) 9.
[18] W. Leitner, E. Dinjus, F. Gaßner, J. Organometal. Chem. 475
ous catalysts, not only exhibiting higher activity, but (1994) 257.
also offer the practical advantages such as ease separa- [19] R.A. Sheldon, M. Wallau, I.W.C.E. Arends, U. Schuchardt, Acc.
tion and recycling. Chem. Res. 31 (1998) 485.

You might also like