Vous êtes sur la page 1sur 8

Article No : a18_321

Oxo Synthesis
HELMUT BAHRMANN, Celanese GmbH Ruhrchemie, Oberhausen, Federal Republic
of Germany
HANSWILHELM BACH, Celanese GmbH Ruhrchemie, Oberhausen, Federal Republic
of Germany

1. Introduction. . . . . . . . . . ..... .... ..... . 609 3.2. Cobalt Processes . . . . . . . . . . . . . . . . . . . . . 615


2. Theory . . . . . . . . . . . . . . ..... .... ..... . 609 References . . . . . . . . . . . . . . . . . . . . . . . . . . 615
3. Industrial Applications . ..... .... ..... . 612
3.1. Rhodium Processes . . . . ..... .... ..... . 612

1. Introduction Efficiency became extremely important follow-


ing the oil crisis in 1973, which shifted the the
Olefins react with synthesis gas (carbon monox- cost structure in aldehyde production towards the
ide and hydrogen) in the presence of homoge- raw materials. For example, propene and synthe-
neous catalysts to form aldehydes containing an sis gas account for about 75 % of the product
additional carbon atom. This hydroformylation value [9]. With the introduction of more expen-
of olefins, known as oxo synthesis, was discov- sive rhodium catalysts, catalyst reprocessing,
ered in 1938 by ROELEN of Ruhrchemie [1], [28, which can be carried out batchwise or continu-
10]. ously, became increasingly important [1222].
The catalyst lifetime can be severely diminished
by extrinsic poisons such as strong acids,
hydrogen cyanide, sulfur, hydrogen sulfide,
COS, oxygen, and dienes. The feedstocks must
therefore be very pure. Since 1978 [23] increased
The most important oxo products are in the attention has been paid to intrinsic deactiva-
range C3 C19 , [2], [8]; with a share of roughly tion (i.e., internal decomposition) of the rela-
75 % butanal is by far the most significant. In tively expensive catalyst ligands [12] , [18], [24].
1997 the total worldwide oxo production capaci- Next to the efficiency, activity, and selectivity of
ty for aldehydes and alcohols was 6.5  106 t/a. the catalyst, the catalyst lifetime is the most
Table 1 gives an overview of the known facilities industrially important reaction parameter.
and capacities.
Up until the mid-1970s only the cobalt-based
catalysts were important industrially. The situa- 2. Theory
tion changed in 1974 1975 when Union Car-
bide and Celanese, independently of one another, The mechanism developed by HECK and BRESLOW
introduced rhodium-based catalysts on an indus- [2527] for the reaction with cobalt catalysts
trial scale. Since then, cobalt catalysts for the (see ! Organometallic Compounds) can be
hydro formylation of propene have been replaced largely applied to hydroformylation with rhodi-
in nearly all major plants by the more advanta- um. However, the introduction of the triphenyl-
geous rhodium catalysts despite the higher price phosphine (PPh3) ligands in the catalyst system
of the noble metal. This substitution was made brings about critical changes which already
possible because the high price of rhodium was affect the course of the reaction by formation of
offset by cheaper equipment, increased activity, various Rh complexes, which leads to new
and a generally higher selectivity and efficiency. effects that cannot be explained by the older

 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim


DOI: 10.1002/14356007.a18_321
610 Oxo Synthesis Vol. 25

Table 1. Worldwide oxo capacities (1989)

Country Capacity, 103 t/a Products Process type *

Brazil 120 C4 1
Bulgaria 20 C4 5
China 300 C4 1, 5
Czech Republic 40 C4 5
Former Soviet Union 580 C4 5
France 355 C4, C8 C10, C13 1, 5
Germany 1695 C3 C13 15
India 40 C4, C8 C10 1, 4, 5
Italy 85 C12 C15 5
Japan 755 C4, C7 C15 1, 4, 5
Poland 150 C4 1
Romania 135 C4 5
Saudi Arabia 190 C4 1
South Korea 420 C4 1, 2
Spain 70 C4 1
Sweden 220 C4 1
The Netherlands 200 C8 C10 5
United Kingdom 150 C7 C15 4, 5
United States 2500 C3 C15 1, 4, 5
*
See Table 2.

cobalt model. For example, with the cobalt sys- important consequences when the reaction is
tem the fraction of unbranched aldehyde product carried out on an industrial scale.
(or the n/i ratio) increases with increasing partial According to R. L. PRUETT and J. A. SMITH,
pressure of carbon monoxide, while with rhodi- under industrial reaction conditions (i.e., at high-
um no effect or a lower n/i ratio is observed [1]. er temperatures and under CO/H2 pressure) the
Fundamental mechanistic studies of the in- above-mentioned equilibrium is also accompa-
dustrially important triphenylphosphine rhodium nied by the competition of CO and PPh3 for free
catalyst system are based on the work of G. coordination sites on the rhodium atom [29]:
WILKINSONs group and workers at Union Carbide
in the late 1960s [25], [28], [29].
Model studies with the hydride complex
[RhHCO(PPh3)3], the key compound in the hy-
droformylation reaction, were the first to provide
satisfactory explanations for the different reac-
tion behavior compared to cobalt catalysts.
For example, studies on the molecular mass
Each catalytic species is assigned an individ-
of this compound in various organic solvents
ual reaction rate and a characteristic product
(e.g., benzene) indicated extensive dissociation
of this complex. This shows that an equilibrium distribution. High phosphorus/rhodium ratios
and a lower partial pressure of carbon monoxide
exists between the various substituted rhodium
favor the type 1 complex, which is assumed to
complexes before the catalytic cycle occurs
[28]: give a high n/i ratio as a result of steric effects,
since it is more highly substituted with phosphine
ligands.
Rhodium complexes with three phosphine
ligands were also postulated in 1981 by J. D.
UNRUH and O. R. HUGHES of Celanese as the basis
for a high n/i selectivity in the presence of
particular chelating ligands in combination with
certain monophosphine ligands [30]. In contrast,
The predominant catalytic species is thus the two basic dissociative and associative me-
determined by this equilibrium, which has chanisms, proposed by WILKINSON in 1968, are
Vol. 25 Oxo Synthesis 611

Figure 1. Dissociative mechanism after G. WILKINSON [7]

based on rhodium complexes with either one or approach. Additionally, the rotation and sweep-
two phosphines [28]. Both mechanisms start with ing of the phenyl groups of the PPh3 ligands in the
the key complex 4 but differ as regards the transfer step from the p complex (Fig. 2 b) may
primary reaction step, i.e., coordination of the favor formation of the n-alkyl complex (Fig. 2 c)
olefin to the rhodium center. rather than the i-alkyl complex due to minimiza-
In the dissociative mechanism dissociation of tion of steric hindrance, the i-alkyl complex
a phosphine ligand from the bis(phosphine) com- being more crowded than the n-alkyl complex.
plex 4 is followed by addition of the olefin to the Investigations by Exxon [31], [32], Johnson
coordinatively unsaturated square-planar com- Matthey [33], Montedison [24], [34], and J. M.
plex 6 (Fig. 1). There is very little steric hin- BROWN and A. G. KENT [35] clearly support the
drance because the complex only has a single dissociative mechanism. The active catalytic
phosphine ligand. In contrast, in the associative species are probably unsaturated square-planar
mechanism the olefin attaches directly to the bis rhodium complexes containing one or two coor-
(phosphine) complex 4. The associative mech- dinated phosphine ligands (Fig. 3). These com-
anism accounts for the fact that a higher selec- plexes are in equilibrium with one another,
tivity for the n-aldehyde is obtained in the whereby complex 2 produces predominantly the
presence of excess phosphine. However, a high n-aldehyde and complex 6 the i-aldehyde.
n/i ratio can also be explained by the dissocia- Since 1990 an increasing number of investi-
tive mechanism if the coordinatively unsaturat- gations [4649] have been performed on the
ed trans bis(phosphine) complex 2, [RhH(CO)
(PPh3)2] is assumed to be the active catalytic
species. WILKINSON [28] had already taken this
possibility into consideration, but rejected it
(Fig. 2).
As the olefin approaches the trans bis(phos-
phine) complex (Fig. 2 a) and forms an alkene
complex (Fig. 2 b) the substituents on the phos-
phine ligands impinge on the alkene due to
rotation about the Rh P bonds, whereby the Figure 2. Control of the n/i ratio by a trans bis(phosphine)
alkene groups are bent away from the direction of rhodium complex [28]
612 Oxo Synthesis Vol. 25

Figure 3. Modified dissociative mechanism [32]

water-soluble tris(trisodium-3-sulfophenyl) 3. Industrial Applications


phosphine-modified [P(3-C6H4SO3Na)3] rhodi-
um hydroformylation catalyst system used in the The hydroformylation reaction is used on an
Ruhrchemie Rhone-Poulenc (RCH RP) industrial scale to produce aldehydes and alco-
process (Fig. 4). A higher n/i ratio is obtained hols (! Alcohols, Aliphatic; ! Aldehydes,
in the RCH RP process than in the UCC pro- Aliphatic; ! Butanals). These include bulk
cess, which uses triphenylphosphine as the products as well as high-value specialty
ligand. This can be explained by a relative shift products.
of the equilibrium between complexes 6 and 4 The industrially important processes can be
(see Fig. 1) toward complex 2 because of the divided into five main types. Table 2 gives an
higher dissociation energy of the tris(trisodium- overview of their catalyst systems, reaction con-
3-sulfophenyl)phosphine ligand compared to the ditions, and expected products.
triphenylphosphine ligand. The cobalt catalysts are extensively described
in [13] , [8]. In the meantime their decline in
industrial importance has come to an end. Steady
improvements in the processing of branched
medium- to long-chain olefins (Table 2, Type 5
[5052]) and inner long-chain olefins (Table 2,
Type 4) have strengthened the position of these
older cobalt-based processes.

3.1. Rhodium Processes

As shown in Table 2, rhodium-based processes


can be classified into three types. The most
important of these on an industrial scale uses
the so-called phosphine-modified catalyst sys-
Figure 4. Structure of the water-soluble rhodium carbonyl tem. The unmodified rhodium carbonyl complex
hydride complex is used for the reaction of special olefins.
Vol. 25 Oxo Synthesis 613

Table 2. Comparison of the various oxo processes

Catalyst [RhH(CO)(PR3)3] [RhH(CO)4] [CoH(CO)3PR3] [CoH(CO)4]

R C6H5 R 3-C6H4SO3Na
Process type no. 1 2 3 4 5

Hydroformylation conditions
Pressure, bar 15 20 10 100 200 300 50 100 200 350
Temperature,  C 85 115 50 130 100 140 160 200 110 180
Results
Selectivity for aldehydes high high high low medium
n/i Ratio 92 : 8 95 : 5 50 : 50 88 : 12 80 : 20
Hydrogenation low low low high medium

Low-Pressure Oxo (LPO) Process. lyst lost in production are fed in from below. Due
Table 2, Type 1. In the mid-1970s Union Car- to the sensitivity of the rhodium catalyst system
bide and Celanese succeeded in using rhodium toward catalyst poisons, the olefin and synthesis
triphenylphosphine catalysts for the hydrofor- gas or hydrogen used as makeup gas must first be
mylation of olefins on an industrial scale. Since carefully purified (b). The reactor contains the
then some other companies have developed mod- catalyst dissolved in high-boiling reaction by-
ifications of this process. The most important of products. In order to maintain catalyst activity a
these, however, is the low-pressure oxo (LPO) portion of the solution must be continuously
process jointly developed by Union Carbide, removed and reprocessed separately (c), and the
Davy McKee, and Johnson Matthey. A detailed noble metal and the phosphine returned to the
description of this specific low-pressure route process. From time to time, all of the catalyst
follows with the conversion of propene to butyr- must be removed via line (d) and reprocessed
aldehyde as an example. externally [12], [16]. The hydroformylation re-
Figure 5 shows a schematic of hydroformyla- action takes place at < 20 bar and 85 115  C.
tion with the LPO gas-recycle process [8], [36]. The reactor jacket is cooled to remove the heat of
The reaction takes place in a stirred-tank reaction. The reaction products and unreacted
reactor (a) made of stainless steel. The reactants gaseous reactants (conversion of about 30 % per
and supplementary catalyst to make up for cata- pass) are forced out of the reactor by the recycled

Figure 5. LPO process (schematic) a) Stirred-tank reactor; b) Reactant purification system; c) Reprocessing; d) To catalyst
reprocessing; e) Demister; f) Condenser; g) Separator; h) Compressor; i) Stripping column; j) Outlet
614 Oxo Synthesis Vol. 25

gas and pass through the demister (e) and the Type 2) is based on a water-soluble rhodium
condenser (f) into the separator (g). catalyst (see Chap. 2) [3743]. As with the LPO
Unreacted starting materials and part of the process, the RCH RP process has found its
propane formed as byproduct are recycled to the greatest importance in the hydroformylation of
reactor by means of the compressor (h). The level propene.
of propane in the circulating gas is adjusted by The use of a water-soluble catalyst system is
means of the outlet (j). The liquid reaction pro- associated with substantial advantages for indus-
ducts are freed from residual olefin in a stripping trial practice, because the catalyst can be consid-
column (i), and are worked up by multistage ered to be heterogeneous. Since the catalyst is
distillation. Residual olefin from the stripping insoluble in the organic phase formed, separation
column is recycled. To limit the buildup of inerts of the aqueous catalyst phase and the butanal is
in the recycled gas stream and to reduce losses by greatly simplified by phase separation, and losses
venting, other treatment steps may be applied. of the noble metal in the crude aldehyde stream
These include extraction of propene with the are negligible. High-boiling byproducts do not
aldehyde products, stripping the olefin from the dissolve in the aqueous catalyst phase, dispens-
aldehydes with synthesis gas and recycling both ing with the need for continuous catalyst regen-
to the reactor [55], and washing the off-gas with eration. In the LPO process, however, these
stripped catalyst solution [56]. byproducts are retained in the catalyst phase and
The LPO gas-recycle process has been partly lead to catalyst problems, unless the catalyst
replaced by a liquid-recycle variant, in which the system is continuously regenerated.
catalyst solution and the aldehyde products leave The process is explained by means of the
the reactor as a liquid. The catalyst solution is simplified flow sheet (Fig. 6). The reaction takes
separated from the aldehydes in several distilla- place in a stirred-tank reactor (a), which contains
tion steps and recycled. Combinations of gas and the catalyst solution; the reactants are introduced
liquid recycle have also been described and are from below. Before entering the reactor, the
claimed to give increased propene conversion synthesis gas is first passed through a stripping
[53], [54]. column (b) in countercurrent to the crude alde-
hyde stream in order to recover the unreacted
Ruhrchemie Rhone-Poulenc (RCH propene. Furthermore, purification of the reac-
RP) Process. Another industrial low-pressure tants can be avoided by this procedure [42]. The
process for olefin hydroformylation (Table 2, crude aldehyde product leaves the top of the

Figure 6. Schematic of the RCH RP process a) Stirred-tank reactor; b) Stripping column; c) Separator; d) Phase separator;
e) Heat exchanger; f) Water inlet; g) Off-gas outlet
Vol. 25 Oxo Synthesis 615

reactor and passes through the trap (c) into the process was based on an unmodified cobalt car-
phase separator (d) where it is separated from bonyl hydride catalyst (Table 2, Type 5) which
the entrained catalyst solution. The catalyst so- required high pressures and temperatures. How-
lution is returned to the reactor via the heat ever, since the successful introduction of rhodi-
exchanger (e). um catalysts its importance has declined but
Because of the higher temperature compared since stabilized at a lower level.
to the LPO process (see Table 2), the heat of
reaction can be used for steam generation in the
heat exchanger (e). The crude aldehyde from the
phase separator (d) is distilled. Part of the water is References
retained in the crude aldehyde in homogeneous
solution. This loss of water is compensated for General References
via inlet (f). Part of the off-gas which escapes via 1 B. Cornils in J. Falbe (ed.): New Syntheses with Carbon-
the separator (c) can be recirculated; part must be Monoxide, Springer-Verlag, Berlin 1980.
2 Ullmann, 4th ed., 7, 118.
drawn off to maintain a constant propane level in 3 Kirk-Othmer, 3rd ed., 16, 637.
the gas. Depending on the purity of the olefin 4 R. L. Pruett, Adv. Organomet. Chem. 17 (1979) 1.
used, the off-gas (g) may contain considerable 5 Houben-Weyl, 4th ed., vol. E 3, p. 180; Science of
amounts of olefin. The RCH RP process may Synthesis, vol. 25, 2006, p. 1.
then, for example, be combined with a cobalt 6 F. H. Jardine, Polyhedron 1 (1982) 569.
high-pressure process to convert the residual 7 I. Wender, P. Pino (eds.): Organic Syntheses via Metal
olefin [44]. Carbonyls, vol. 2, John Wiley & Sons, New York 1977,
p. 43.
8 Winnacker-K uchler 4th ed., 5, 537.
Rhodium High-Pressure Process. If un- 9 C. D. Frohning, C. W. Kohlpaintner in B. Cornils, W. A.
modified rhodium carbonyl hydride is used as Herrmann (eds.): Applied Homogeneous Catalysis with
a hydroformylation catalyst, the reaction product Organometallic Compounds, vol. 2, VCH, Weinheim,
consists of roughly equal amounts of branched Germany 1996, p. 64.
and straight-chain aldehydes (see Table 2,
Type 3). For this reason this catalyst is only Specific References
applicable if the n/i ratio is not important (i.e., 10 Ruhrchemie AG, DE 849 548, 1938 (O. Roelen).
both aldehydes are valuable products) or if the 11 B. Cornils, R. Payer, K. C. Traenckner, Hydrocarbon
Process. 54 (1975) 89.
formation of a branched aldehyde is impossible
12 P. E. Garrou, Chem. Rev. 85 (1985) no. 3, 171.
(e.g., in the hydroformylation of ethylene to give 13 Ruhrchemie AG, DE 3 235 030 A 1, 1982 (R. Gartner, B.
propanal) [45]. Anhydrous propanal can be ob- Cornils, H. Springer, P. Lappe).
tained by this process. 14 Ruhrchemie AG, EP 0 163 233 A 1, 1985 (H. Bahrmann,
et al.).
15 Ruhrchemie AG, EP 0 175 919 A 1, 1985 (L. Bexten, B.
3.2. Cobalt Processes Cornils, D. Kupies).
16 P. E. Garrou, R. A. Dubois, W. J. Chu, CHEMTECH 1985,
no. 2, 123.
Modified Catalysts (Shell Process). If the
17 Ruhrchemie AG, EP 0 103 845 A 2, 1983 (R. Gartner, B.
corresponding alcohol rather than the aldehyde is Cornils, L. Bexten, D. Kupies).
the desired product, the Shell process can be 18 A. G. Abatjoglou, E. Billig, D. R. Bryant, Organometal-
applied (Table 2, Type 4). A phosphine-modified lics 3 (1984) no. 6, 923.
cobalt catalyst with a strong hydrogenation 19 UCC, EP 0 357 997, 1988 (D. J Miller, D. R. Bryant, E.
activity is used and gives a favorable n/i ratio Billig, L. B. Shaw).
(88 : 12) in the initially formed aldehyde product 20 UCC, EP 0 552 797, 1993 (J. E. Babin, D. R. Bryant, A.
mixture. The catalyst readily isomerizes feed M. Harrison, D. J. Miller).
21 Hoechst, EP 544 091, 1992 (W. Konkol, H. Bahrmann,
olefins and can therefore be used to hydroformy-
W. A. Herrmann, C. Kohlpaintner).
late inner olefins. For a detailed description of 22 UOP, EP 504 814, 1992 (D. R. Bryant, J. E. Babin, J. C.
this process see [2], [8]. Nicolson, D. J. Weintritt).
23 G. Gregorio et al., Symp. Rhodium Homogeneous Catal.
Unmodified Catalysts. The first and, for a [Proc.] 1978, 121.
long time, the most important, industrial oxo 24 G. Gregorio et al., Chim. Ind. (Milan) 62 (1980) 389.
616 Oxo Synthesis Vol. 25

25 D. Evans, J. A. Osborn, G. Wilkinson, J. Chem. Soc. A 50 Exxon Chemical Patents, EP 0 357 405, 1989 (N. A. De
1968, 3133. Munck).
26 D. S. Breslow, R. F. Heck, Chem. Ind. (London) 1960, 51 Exxon Chemical Patents, EP 0 391 650, 1989 (D. M.
467. Olijve, N. A. De Munck).
27 R. F. Heck, D. S. Breslow, J. Am. Chem. Soc. 83 (1961) 52 Exxon Chemical Patents, EP 0 372 925, 1988 (E. van
4023. Driessche et al.).
28 D. Evans, G. Yagupsky, G. Wilkinson, J. Chem. Soc. A 53 Union Carbide, EP 0 188 246, 1986 (D. L. Bunning, M.
1968, 2660. A. Blessing).
29 R. L. Pruett, J. A. Smith, J. Org. Chem. 34 (1969) 327. 54 Davy Powergas, GB 1 387 657, 1973 (R. Fowler).
30 O. R. Hughes, J. D. Unruh, J. Mol. Catal. 12 (1981) 71. 55 Union Carbide Chemicals and Plastics Company, EP 048
31 Exxon Research and Engineering Co., DE 3 034 352, 976, 1991 (K. D. Sorensen).
1980, (A. A. Oswald, T. G. Jermansen, A. A. Westner, I- 56 Union Carbide Chemicals and Plastics Company, EP 0
Der. Huang). 404 193, 1991 (D. L. Bunning).
32 I. T. Horvath, R. V. Kastrup, A. A. Oswald, E. J. Moze-
leski, Catal. Lett. 2 (1989) 85.
33 M. J. H. Russell, Platinum Met. Rev. 32 (1988) no. 4, 185.
34 P. Cavalieri dOro et al., Chim. Ind. (Milan) 62 (1980) Further Reading
572.
35 John M. Brown, Alexander G. Kent, J. Chem. Soc. Perkin D. Astruc: Organometallic Chemistry and Catalysis,
Trans. 2 1987, 1597. Springer, Berlin 2007.
36 F. Heinrich, M. Bernard, 27th DGMK-Haupttagung, R. Ballini (ed.): Eco-friendly Synthesis of Fine Chemicals,
Aachen Oct. 6 8, 1982, Compendium 82/83, p. 189. Royal Society of Chemistry, Cambridge, UK 2009.
37 Rhone-Poulence Industries, DE 2 627 354, 1976 M. Beller (ed.): Catalytic Carbonylation Reactions, Springer,
(E. Kuntz). Berlin 2006.
38 Rhone-Poulenc Chimie de Base, EP 0 104 967 B 1, 1982 E. Billig, D. R. Bryant: Oxo Process, Kirk Othmer Ency-
(J. L. Sabot). clopedia of Chemical Technology, John Wiley & Sons,
39 Rhone-Poulenc Recherches, EP 0 133 410 A 1, 1984 (J. Hoboken, NJ, online DOI: 10.1002/0471238961.
Jenck, D. Morel). 15241502091212.a01.
40 Rhone-Poulenc Recherches, EP 0 158 572 A 1, 1985 (C. A. Borner (ed.): Phosphorus Ligands in Asymmetric Cataly-
Barre, M. Desbois, J. Nouvel). sis, Wiley-VCH, Weinheim 2008.
41 Rhone-Poulenc Industries, FR 8 005 488, 1980 (J. Jenck). G. P. Chiusoli, P. M. Maitlis (eds.): Metal-Catalysis in
42 Ruhrchemie AG, EP 0 103 810, 1983 (B. Cornils Industrial Organic Processes, Royal Society of Chemis-
et al.). try, Cambridge, UK 2006.
43 Ruhrchemie AG, EP 0 158 246 A2, 1985 (B. Cornils B. H. Davis, M. L. Occelli (eds.): Advances in Fischer-
et al.). Tropsch Synthesis, Catalysts, and Catalysis, CRC Press/
44 Ruhrchemie AG, EP 0 111 257 B 1, 1983 (B. Cornils Taylor & Francis, Boca Raton, FL 2010.
et al.). P. A. Evans (ed.): Modern Rhodium-catalyzed Organic Re-
45 Ullmann, 4th ed., 19, 443. actions, Wiley-VCH, Weinheim 2005.
46 B. Cornils, Org. Process Res. Dev. 2 (1998) no. 2, 121. J. Hagen: Industrial Catalysis, 2nd ed., Wiley-VCH,
47 W. A. Herrmann, C. W. Kohlpaintner, Angew, Chem. Int. Weinheim 2006.
Ed. Engl. 32 (1993) 1524. C. Higman, M. d. van Burgt: Gasification, 2nd ed., Elsevier,
48 M. Beller, B. Cornils, C. D. Frohning, C. W. Kohlpaint- Amsterdam 2008.
ner, J. Mol. Cat. A 104 (1995) 17 85. L. Kollar (ed.): Modern Carbonylation Methods, Wiley-
49 F. Joo, A. Katho,, J. Mol. Cat. A 116 (1997) 3 26. VCH, Weinheim 2008.

Vous aimerez peut-être aussi