Mechanisms in Homogeneous Catalysis A Spectroscopi PDF
Mechanisms in Homogeneous Catalysis A Spectroscopi PDF
Mechanisms in Homogeneous Catalysis A Spectroscopi PDF
Volume 22
Editors:
B. R. James, The University of British Columbia, Vancouver, Canada
P. W. N. M. van Leeuwen, University of Amsterdam, The Netherlands
Advisory Board:
I. Horvth, Exxon Corporate Research Laboratory, Annandale, NJ, U.S.A.
S. D. Ittel, E. I. du Pont de Nemours Co., Inc., Wilmington, Del., U.S.A.
A. Nakamura, Osaka University, Osaka, Japan
W. H. Orme-Johnson, M.I.T, Cambridge, Mass., U.S.A.
R. L. Richards, John Innes Centre, Norwich, U.K.
A. Yamamoto, Waseda University, Tokyo, Japan
The titles published in this series are listed at the end of this volume.
RHODIUM CATALYZED
HYDROFORMYLATION
Edited by
and
CARMEN CLAVER
Department de Quimica Fsica i Inorgnica,
Universitat Rovira i Virgili,
Tarragona, Spain
No part of this eBook may be reproduced or transmitted in any form or by any means, electronic,
mechanical, recording, or otherwise, without written consent from the Publisher
xi
xii Preface
in fine and bulk chemistry. The reader will notice the importance of one
discipline for the other. In many cases these relationships have already been
established, but for other cases the book might assist future developments.
The key roles that ligands may play in selectivity may be an eye-opener for
organic chemists and it will further enhance the large number of new
applications and reactions that are being discovered. The comments in
several chapters on catalyst preparation and feed purification may be useful
for scientists who are not specialized in homogeneous catalysis using
transition metal complexes.
Hydroformylation is also a model reaction system in homogeneous
catalysis as it contains so many aspects such as ligand effects (electronic,
steric, bite angle), in situ studies, complicated kinetics, and effects of
conditions and impurities. All this, combined with its practical value, makes
it an ideal topic in education.
The editors are very grateful to the authors for the good work they did
and the prompt responses. The writing took only a few months, as did the
production by the publisher. Writing the book has been rewarding, because
we learnt many things. Most of all perhaps, we obtained a clearer view on
what we still dont fully understand.
Amsterdam, Tarragona
Piet van Leeuwen, Carmen Claver
TABLE OF CONTENTS
Preface xi
1 Introduction to hydroformylation 1
Piet W. N. M. van Leeuwen
1.1 History of phosphorus ligand effects 1
1.2 Hydroformylation 6
1.3 Ligand parameters 8
3.2 Monophosphites 37
3.2.1 Catalysis 37
3.2.2 Mechanistic and kinetic studies 40
3.3 Diphosphites 44
3.3.1 Catalysis 44
3.3.2 Mechanistic and kinetic studies 48
3.4 Hydroformylation of internal alkenes 55
3.4.1 Hydroformylation of less reactive internal and
functionalized alkenes 55
3.4.2 Formation of linear aldehydes starting from
internal alkenes 57
3.5 Calixarene based phosphites 59
4 Phosphines as ligands 63
Piet W. N. M. van Leeuwen, Charles P. Casey, and Gregory T. Whiteker
4.1 Monophosphines as ligands 63
4.1.1 Introduction 63
4.1.2 The mechanism 64
4.1.3 Ligand effects 66
4.1.4 In situ studies 68
4.1.5 Kinetics 69
4.1.6 Regioselectivity 72
4.1.7 Conclusion 75
4.2 Diphosphines as ligands 76
4.2.1 Introduction 76
4.2.2 Ferrocene based diphosphine ligands 78
4.2.3 BISBI ligands and the natural bite angle 82
4.2.4 Xantphos ligands: tunable bite angles 87
4.2.5 The mechanism, regioselectivity, and
the bite angle. Concluding remarks 96
Index 281
Chapter 1
Introduction to hydroformylation
Phosphorus ligands in homogeneous catalysis
[16] reported that HCo(CO)4 gave unidentifiable complexes with dppe. The
use of dppe in cobalt catalyzed hydroformylation was reported by Slaugh
[17], but compared to PBu3 it had little effect on the rate and the selectivity
of the cobalt carbonyl catalyst. Copolymerization of butadiene and
propylene oxide using nickel bromide and dppe was published in 1965 [18].
In the late sixties at Shell Development, Keim and coworkers discovered
that certain bidentates containing an oxygen and a phosphorus donor atom
formed excellent catalysts with nickel for the oligomerization of ethene [19].
Typical ligands are diphenylphosphinoacetic acid or 2-diphenyl-
phosphinobenzoic acid (SHOP ligand, Figure 1). The ligand required a
relatively laborious ligand synthesis for those days. In addition it was the
first process utilizing the concept of two-phase catalysis. This discovery led
to the Shell Higher Olefins Process that came on stream in 1977.
Hata [20] reported a phosphine-free iron catalyst for the codimerization
of butadiene and ethene in 1964. A year later this was followed by
phosphine-free rhodium catalysts [21]. The oldest publication describing
advantageous results for diphosphines we found is by Iwamoto and Yuguchi
(1 966) who studied the same reaction using iron catalysts containing a range
of diphosphines varying in bridge lengths [22]. In many instances the
activity of catalysts containing dppe instead of PPh3 is lower. For example,
the hydrogenation of styrene using rhodium(I) chloride and dppe is 70 times
slower as compared to the PPh3 based system [23]. The strong chelating
power of the diphosphine was held responsible for this. Thus, initially the
use of dppe and other bidentate phosphines in catalysis found little support
as they were supposed to lead mostly to more stable complexes, rather than
more active or selective catalysts.
Theoretical work of Thorn and Hoffmann [24] explained why migration
reactions in complexes containing for instance dppe were slow. The
constrained P-M-P angle would slow down the migration reaction, since
ideally the phosphine ligand coordinated in the position cis to the migrating
group, would have a tendency to widen the P-M-P angle in the process to
pursue the migrating group.
1.2 Hydroformylation
Figure 5. Structures of ttpms, van Leeuwen's "bulky phosphite", and a highly stable, bulky
phosphite from UCC
The second generation processes use rhodium as the metal and the first
ligand-modified process came on stream in 1974 (Celanese) and more were
to follow in 1976 (Union Carbide Corporation) and in 1978 (Mitsubishi
Chemical Corporation), all using triphenylphosphine (tpp). The UCC
process has been licensed to many other users and it is often referred to as
the LPO process. Not only are rhodium catalysts much faster - which is
translated into milder reaction conditions -, but also their feedstock
utilization is much better than that of cobalt catalysts. For example, the
cobalt-alkylphosphine catalyst may give as much as 10% of alkane
1. Introduction to hydroformylation 7
Figure 6. Eastmans BISBI and typical diphosphites from Union Carbide Corporation
promising. Platinum has been known for many years to have a high
preference for the formation of linear products, but ligand decomposition
hampers applications [57]. Palladium and platinum will not be discussed, but
recent advances for rhodium have been collected in Chapter 10.
A ligand with great potential for hydroformylation of higher alkenes in a
one-phase system that is worked up by adding water to separate catalyst and
product afterwards is the monosulfonated triphenylphosphine, tppms, that
was studied by Abatjoglou, also at Union Carbide [58] (Chapter 8).
The fourth generation process for large-scale application still has to be
selected from the potential processes that have been nominated. In the
chapters to follow several of these candidates will be discussed. The fourth
generation will concern higher alkenes only, since for propene
hydroformylation there are hardly wishes, if any, left [59] (a cheaper catalyst
would be on my shopping list!). Many new phosphite-based catalysts have
been reported that will convert internal alkenes to terminal products and
recently also a new diphosphines have been reported that will do this [60].
The most interesting ligand discovered for asymmetric hydroformylation
is undoubtedly BINAPHOS, introduced by Takaya [30]. Diphosphites have
also been studied to this end by UUC [61] and by us [62]; Babin and
Whiteker reported the first successful ones [61]. Asymmetric rhodium
catalysts are discussed in Chapter 5.
Ligand design for fine chemical applications has been very limited and
usually the ligands designed for large-scale applications are also tested for
more complicated organic molecules. Tpp has been the workhorse in fine
chemicals hydroformylation since Wilkinsons first examples [63, 64], but
also bulky phosphite [49], tppts and tppms [41-43] turned out to be very
useful, and recently diphosphites have been studied [65] (see Chapter 6).
Tolman reviewed ligand effects for the first time [5]. Prior to his studies
[66] the effects of phosphorus ligands on reactions or properties of metal
complexes were rationalized mainly in terms of electronic effects.
Systematic studies had shown, however, that steric effects are at least as
important as electronic effects, and in terms of stability of complexes can
even be dominant. Since then, numerous studies have appeared using both
the electronic parameter and the steric parameter for the cone angle, .
The electronic parameter is a measure for the overall effect of electron
donating and accepting properties of the phosphorus ligand L. It is measured
as the symmetric stretching frequency of the carbonyls in Ni(CO)3L, similar
to the methods proposed by Strohmeier and Horrocks [67]. High -values
stand for strong -acceptors and low -values stand for strong -donor
1. Introduction to hydroformylation 9
Figure 7. A simple picture of Tolmans -value. On the left strong back donation to CO
leading to a low stretching frequency. On the right weak back-donation to CO leading to a
high stretching frequency [5]
Figure 8. Tolmans cone angle [5]. The distance M-P amounts to 2.24 . Caseys natural
bite angle n as calculated by Molecular Mechanics [79, 80].
References
1 van Leeuwen, P. W. N. M.; Kamer, P. C. J.; Reek, J. N. H.; Dierkes, P. Chem. Rev. 2000,
100, in press.
2 Reppe, W.; Schweckendiek, W. J. Annalen 1948, 104, 560.
3 Slaugh, L. H.; Mullineaux, R. D. U.S. Pat. 3,239,569 and 3,239,570, 1966 (to Shell);
Chem. Abstr. 1964, 64, 15745 and 19420. J. Organornet. Chem. 1968,13,469.
4 (a) Brown, E. S. Aspects of Homogeneous Catalysis; Ugo, R., Ed.; Reidel: 1974; Vol. 2;
pp 57-78. (b) Drinkard, W. C.; Lindsey, R. V. U. S. Pat. 3,655,723, 1970 (to Du Pont);
Chem. Abstr. 1972, 77, 4986.
5 Tolman, C. A. Chem. Rev. 1977, 77, 313.
6 Halpem, J.; James, B. R. J. Am. Chem. Soc. 1961, 83, 753.
7 Cramer, R. D.; Jenner, E. L.; Lindsey, R. Y.; Stolberg, U. G.; J. Am. Chem. Soc. 1963,
85, 1691.
8 Sloan, M. F.; Matlack, A. S.; Breslow, D. S. J. Am. Chem. SOC. 1963, 85,4015.
9 Young, J. F.; Osbom, J. A,; Jardine, F. A.; Wilkinson, G. J. Chem. SOC. , Chem. Comm.
1965, 131.
10 O'Connor, C.; Wilkinson, G., Tetrahedron Lett. 1969,18, 1375.
11 Yaska, L.; Rhodes, R. E. J. Am. Chem. SOC. 1965, 87, 4970.
12 (a) Jardine, F. H.; Osborn, J. A.; Wilkinson, G.; Young, J. F. Chem. and Ind. (London)
1965, 560. (b) Evans, D.; Osborn, J. A.; Wilkinson, G. J. Chem. Soc. A. 1968, 3133.
13 Pruett, R. L.; Smith, J. A. J. Org. Chem. 1969, 34, 327.
14 Issleib, K.; Muller, D. W. Chem. Ber. 1959, 92, 3175.
15 (a) Chatt J.; Hart, W. J. Chem. Soc. 1960, 1378. (b) Hieber, W.; Freyer, W. Chem.Ber.
1960, 93, 462.
16 Heck, R. F.; Breslow, D. S. J. Am. Chem. SOC. 1962, 84, 2499.
17 Cannel, L. G.; Slaugh, L. H.; Mullineaux, R. D. Ger. Pat. 1,186,455 (priority date 1960,
to Shell); Chem. Abstr. 1965, 62, 16054.
18 United States Rubber Co, Neth. Appl. 6,507,223; Chem. Abstr. 1965, 62, 16013.
19 Bauer, R. S.; Glockner, P. W.; Keim, W.; van Zwet, H.; Chung, H. U.S. Pat. 3,644,563,
1972.
20 Hata, G. J. Am. Chem. Soc. 1964, 86 ,3903.
21 Cramer, R. J. Am. Chem. SOC. 1965, 87 , 4717 and 1967, 89, 1633.
22 Iwamoto, M.; Yuguchi, S. J. Org. Chem. 1966, 31, 4290.
23 Dang, T. P.; Kagan, H. B. Chem. Commun. 1971, 481 and J. Am. Chem. Soc. 1972, 92,
6429.
24 Thorn D. L.; Hoffmann R. J. Am. Chem. SOC. 1978, 100, 2079.
25 Poulin, J. C.; Dang, T. P.; Kagan, H. B. J. Organometal. Chem. 1975, 84, 87.
26 (a) Knowles, W. S.; Sabacky, M. J. J. Chem. SOC., Chem. Commun. 1971, 481 (b)
Knowles, W. S.; Sabacky, M. J.; Vineyard, B. D., and Weinkauff, J. J. Am. Chem. SOC.
1975 97, 2567.
27 Selke, R. and Pracejus, H.; J. Mol. Catal., 1986, 37, 213.
28 Miyashita, A.; Yasuda, A.; Tanaka, H.; Toriumi, K.; Ito, T.; Souchi, T.; Noyori, R. J. Am.
Chem. Soc., 1980, 102,7932,
29 Burk, M. J. J. Am. Chem. SOC. 1991, 113, 8518.
30 Sakai, N.; Mano, S.; Nozaki, K.; Takaya, H. J. Am. Chem. Soc. 1993, 15, 7033.
31 Togni, A. Angew. Chem. 1996, 108, 1581.
32 Transition Metals for Organic Synthesis; Building Blocks and Fine Chemicals, ed. M.
Beller and C. Bolm, Wiley-VCH, 1998, Vol 1 and 2.
12 Chapter 1
59 Wiebus, E.; Cornils, E. Chem. Ing. Tech. 1994, 66, 916; Comils, B.; Wiebus, E.
ChemTech, 1995, 25, 33,; Comils, B.; Wiebus, E. Recl. Trav. Chim. Pays-Bas, 1996,
115,211.
60 van der Veen. L. A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Angew. Chem. Int. Ed.
Engl. 1999, 38, 336.
61 Babin, J. E.; Whiteker, G. T. U.S. Pat. 5,360,938 1994 (to Union Carbide Corp.); Chem.
Abstr. 1995, 122, 186609.
62 Buisman, G. J. H.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Tetrahedron:
Asymmetry, 1993, 4, 1625.
63 Brown, C. K.; Wilkinson, G. .J. Chem. Soc. (A) 1970, 2753.
64 Eilbracht, P.; Brfacker, L.; Buss, C.; Hollmann, C.; Kitsos-Rzychon, B. E.; Kranemann,
C. L.; Rische, T.; Roggenbuck, R.; Schmidt A. Chem. Rev. 1999, 99, 3329.
65 Cuny, G. D.; Buchwald, S. L. J. Am. Chem. Soc. 1993, 115, 2066.
66 Tolman, C. A. J. Am. Chem. Soc. 1970, 92, 2953 and 2956.
67 (a) Strohmeier, W.; Muller, F. J. Chem. Ber. 1967, 100, 2812. (b) Horrocks Jr., W. D.;
Taylor, R. C. Inorg. Chem. 1963, 2,723.
68 Bartik, T.; Himler, T.; Schulte, H.-G.; Seevogel, K. J. Organomet. Chem. 1984, 272, 29.
69 Grim, S.O.; Yankowsky, A. W. J. Org. Chem. 1977, 42, 716.
70 Trogler, W. C.; Marzilli, J. G. J. Am. Chem. Soc. 1974, 96, 7589.
71 Heimbach, P.; Kluth, J.; Schlenkluhn, H.; Weimann, B. Angew. Chem. 1980, 92, 567.
72 Femandez A.; Reyes C.; Wilson M. R.; Woska D. C.; Prock A.; Giering, M. P.
Organometallics, 1997, 16, 342.
73 Joerg S.; Drago R. S.; Sales J. Organometallics, 1998, 17, 589.
74 Haar, C. M.; Nolan, S. P.; Marshall, W. J.; Moloy, K. G.; Prock, A.; Giering, W. P.
Oganometallics, 1999, 18,474.
75 (a) Hirota M.; Sakakibara K.; Komatsuzaki T.; Akai, I. Comp. Chem. 1991, 15, 241. (b)
White D.; Taverner, B. C.; Coville, N. J.; Wade, P. W. J. Organomet. Chem. 1995, 495,
41. (c) White D.; Taverner, B. C.; Leach, P. G. L.; Coville N. J. J. Organomet. Chem.
1994, 478, 205.
76 Koide, Y.; Bott, S. G . ; Barron, A. R. Organometallics, 1996, 15, 2213.
77 a) Brown, T. L. Inorg. Chem. 1992, 31, 1286. b) Choi, M.-G.; White, D.; Brown, T. L.
Inorg. Chem. 1993, 32, 5591.
78 Angemund, K.; Baumann, W.; Dinjus, E.; Fornika, R.; Gorls, H.; Kessler, M.; Krger,
C.; Leitner, W.; Lutz, F. Chem. Eur. J. 1997, 3, 755.
79 Dierkes, P.; van Leeuwen, P. W. N. M. J. Chem. Soc. Dalton Trans. 1999, 1519.
80 Casey, C. P.; Whiteker, G. T. Isr. J. Chem. 1990, 30, 299.
81 Kranenburg, M.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Vogt, D.; Keim, W. J. Chem.
Soc. Chem. Commun. 1995, 2 177; Marcone, J. E.; Moloy, K. G. J. Am. Chem. Soc. 1998,
120,8527.
Chapter 2
2.1 Introduction
a) Vinyl substrates
b) Allyl substrates
2. Hydroformylation with unmodified rhodium catalysts 19
c) Vinylidenic alkenes
C
Catalyst T P Reaction times b:1 Reference
precursor (C) (bar)b (h)
Rh4(CO)12 60 150 5 95/5 12c
Rh4(CO)12 25 150 9.5 98/2 12c
[Rh(COD)(OAc)]2 25 50 16 96/4 15
a
At complete substrate conversion
b
CO/H2= 1:l
c
d
Regioselectivity
Zwitterionic rhodium complex [BPh4]-[Rh(COD)]+
a
At complete substrate conversion
b
CO/H2= 1:l
c
Regioselectivity
2. Hydroformylation with unmodified rhodium catalysts 21
In the case of styrene the CO and H2 partial pressures affect the reaction
regioselectivity only when the reaction is carried out at high temperatures. In
particular it has been observed that a decrease of carbon monoxide or
hydrogen partial pressure causes an increase of linear aldehydic isomer, this
effect being more evident at higher temperatures (100 C). So in the case of
styrene hydroformylation at 100 C the b:l ratio ranges from 80/20 at 170
bar of CO/H2 (1:1) to 56/44 at PH2 = 6 bar, PCO = 85 bar or to 60/40 at PH2 =
85 bar, PCo = 6 bar [12c]. In the case of 1-hexene gas pressure does not
affect the regioselectivity of the reaction either at room temperature or at
high temperature. By contrast the chemoselectivity to aldehydes increases
with increasing temperature, (b+l)/E-2 being 44/56 at 40 bar and 77/23 at
140 bar [22b].
As described above, both the nature of the substrate and the reaction
conditions strongly influence the regioselectivity in the hydroformylation of
vinyl substrates. The above results clearly demonstrate that, by raising the
reaction temperature, and decreasing the CO and H2 partial pressures, the
amount of linear aldehydes increases. Indeed, this is a general trend in the
hydroformylation of different substrates and constitutes a fundamental
starting point for a rationalization of the influence of experimental
parameters on the reaction selectivity.
2. Hydroformylation with unmodified rhodium catalysts 23
In this context we are going to examine the above results in the light of
the generally accepted mechanism for hydroformylation, taking into account
the more recent findings on the behavior under reaction conditions of the
main intermediate species, namely alkyl- and acyl-rhodium complexes. A
simplified scheme for the hydroformylation of a typical vinyl substrate is
shown in Figure 3.
The rhodium hydride tricarbonyl species easily coordinates the vinyl
substrate generating the complex (1), which is converted into the alkyl-
rhodium intermediates (2) through insertion of the alkene into the Rh-H
bond. Migratory insertion of the alkyl moiety on to a CO molecule
coordinated to the metal center provides the acyl-rhodium species 3, which,
at the end of the catalytic cycle, interacts with hydrogen via an oxidative
addition, giving rise to aldehydic products and regenerating the rhodium-
hydride species.
R = OEt
Figure 6. H-NMR spectrum (46 MHz, 25 C, C6D6 as external standard) of the crude
mixture resulting from deuterioformylation of (ethyl)vinyl ether at (a) 20 C and (b) 100 C
28 Chapter 2
rate=k0 [RCORh(CO)4]ss1.0[CO]-1.0[H2]1.0[C8H8]0.0
Table 3. Selected values of kinetic constants and regioisomeric ratios for styrene
hydroformylation in the presence of Rh4(CO)12 as catalyst precursor, at 25 C and 40 C.
Values Temperature
25 C 40 C
kb 0.93 4.18
kn 1.27 10.3
3b:31a 97.5/2.5 87.51/12.5
b:1 96.7/3.3 66/34
a
Regioisomeric ratio between branched (3b) and linear (31) acyl-rhodium intermediates.
this is due to the higher stabilization induced by the two phenyl groups
adjacent to the carbon-rhodium bond. However the migratory insertion on to
the CO coordinated to the metal in the case of tertiary alkyls is prevented by
steric reasons. Thus it seems evident that the behavior of the two isomeric
alkyl-rhodium intermediates is completely different: while the primary one is
converted into the linear aldehyde, the tertiary one exclusively undergoes -
hydride elimination, regenerating the starting alkene [1le].
2
In conclusion, the H-NMR analysis of crude deuterioformylation
products derived from vinyl or vinylidenic aromatic substrates is a direct and
simple way to detect the different behavior of a primary, secondary and
tertiary alkyl-metal intermediate, related to the -hydride elimination process
under typical hydroformylation conditions.
experiments carried out by Casey [27] and Takaya [7a], thus accounting for
the low variation of regioselectivity with temperature obtained in the
presence of phosphine-modified precursors [20b].
It is to remark that in the hydroformylation of styrene, the most
investigated vinyl aromatic substrate, the predominance of the branched
aldehyde at room temperature is higher with unmodified rhodium precursors
than with phosphine-modified ones [4, 7c, 28]. In this context, when
hydroformylation of styrene with chiral phosphines occurs without
asymmetric induction and with a large prevalence of the branched aldehyde
(> 96%), it is likely that unmodified rhodium-catalysts are also present in the
reaction mixture [ 14, 29, 30].
References
1 (a) Schiller, G. Ger. Pat. 965,605 1956 (to Chem. Verwertungsges. Oberhausen); Chem
Abstr. 1959, 53, 11226. (b) Hughes, V. L. Br. Pat. 801,734 1958 (to Esso Res. Eng
Comp.); Chem Abstr. 1959, 53, 7014.
2 Roelen, O .Ger. Pat. 849,548 1938; Chem. Zentr. 1953, 927.
3 Organometallic Chemistry of Transition Elements, ed. F. P. Pruchnik, Plenum Press,
New York, 1990, p 691.
4 Evans, D.; Osborn, J. A.; Wilkinson, G. J. Chem. Soc., A 1968. 3133.
5 (a) Organic Syntheses via Metal Carbonyls, eds. I. Wender and P. Pino, Wiley-
Interscience, New York, 1977, Vol. 2. (b) Homogeneous Catalysis with Metal
Phosphine Complexes, ed. L. H. Pignolet, Plenum Press, New York, 1983. (c) Applied
Homogeneous Catalysis with Organometallic Compounds, eds. B. Cornils and W. A.
Herrmann, VCH, Weinheim, 1996.
6 (a) Casey, C. P.; Paulsen, E. L.; Beuttenmueller, E. W.; Proft, B. R.; Matter, B. A.;
Powell, D. R. J. Am. Chem. Soc. 1999, 121, 63. (b) Casey, C. P.; Paulsen, E. L. ;
Beuttenmueller, E. W.; Proft, B. R.; Petrovich, L. M.; Matter, B. A.; Powell, D. R. J.
Am. Chem. Soc. 1997, 119, 11817. (c) van der Van, L. A.; Kamer, P. C. J.; Van
Leeuwen, P. W. N. M. Organometallics 1999, 18, 3765. (d) van Rooy, A.; Kamer, P. C.
J.; Van Leeuwen, P. W. N. M.; Goubitz, K.; Fraanje, J.; Veldman, N.; Spek, A.;
Organometallics 1996, 15, 835.
7 (a) Nozaki, K.; Sakai, N.; Nanno, T.; Higashijima, T.; Mano,, S.; Horiuchi, T.; Takaya,
H. J. Am. Chem. Soc. 1997, 119, 4413. (b) Nozaki, K.; Nanno, T.; Takaya, H. J.
Organomet. Chem. 1997, 527, 103. (c) Gladiali, S.; Bayn, J. C.; Claver, C.
Tetrahedron: Asymm. 1995, 6, 1453. (d) Bayn, J. C.; Claver, C.; Masdeu-Bulto, A. M.
Coord. Chem. Rev. 1999, 195, 73. (e) Miquel-Serrano, M. D.; Masdeu-Bulto, A. M.;
Claver, C; Sinou, D. J. Mol. Cut., A: Chemical 1999, 143, 49.
8 (a) Thatchenko, I. Comprehensive Organometallic Chemisty, eds. G . Wilkinson, F. G.
A. Stone and E. W. Abel, Pergamon, Oxford, 1982, Vol 8, pp. 101. (b) Transition
metals for Organic Synthesis, eds. M. Beller and C. Bolm, Wiley VCH, 1999.
9 Beller, M.; Cornils, B.; Frohning, C. D.; Kohlpaintner, C. W. J. Mol. Cat. A: Chem.
1995, 104, 17.
10 (a) Garland, M.; Pino, P. Organometallics 1991, 10, 1693. (b) Garland, M.
Organometallics 1993, 12, 535. (c) Fyhr, C.; Garland, M. Organometallics 1993, 12,
1753. (d) Feng, J.; Garland, M. Organometallics 1999, 18 ,417. (e) Guowei, L.; Volken,
R.; Garland, M Organometallics 1999, 18, 3429.
2. Hydroformylation with unmodified rhodium catalysts 33
11 (a) Lazzaroni, R.; Uccello-Barretta, G.; Benetti, M. Organometallics 1989, 8, 2323. (b)
Raffaelli, A.; Pucci, S.; Settambolo, R.; Uccello-Barretta, G.; Lazzaroni, R.
Organometallics 1991, 10, 3892. (c) Uccello-Barretta, G.; Lazzaroni, R.; Settambolo,
R.; Salvadori, P. J. Organomet. Chem. 1991, 417, 111. (d) Lazzaroni, R.; Settambolo,
R.; Uccello-Barretta, G. Organometallics 1995, 14, 4644. (e) Lazzaroni, R.; Uccello-
Barretta, G.; Scamuzzi, S.; Settambolo, R.; Caiazzo, A. Organometallics 1996, 15,
4657. (f) Lazzaroni, R.; Settambolo, R.; Uccello-Barretta, G.; Caiazzo, A.; Scamuzzi, S.
J. Mol. Cat., A: Chemical 1999, 143, 123.
12 (a) Pino, P.; Oldani, F.; Consiglio, G. J. Organomet. Chem. 1983, 250, 491, (b) Ojima,
I. Chem. Rev. 1988, 88, 1011. (c) Lazzaroni, R.; Raffaelli, R.; Settambolo, R.; Bertozzi,
S.; Vitulli, G. J. Mol. Cat. 1989, 50, 1.
I3 Botteghi, C.; Paganelli, S.; Bigini, L.; Marchetti, M. J. Mol. Cat. 1994, 93, 279.
14 Basoli, C.; Botteghi, C.; Cabras, M. A.; Chelucci, G.; Marchetti, M. J. Organomet.
Chem. 1995, 488, C20.
15 Doyle, M. P.; Shanklin, M. S.; Zlokazov, M. V. Synlett 1994, 615.
16 Amer, I.; Alper, H. J. Am. Chem. Soc. 1990, 112, 3674.
17 (a) Kalck, P.; Serein-Spiran, F. New J. Chem. 1989, 13, 515. (b) Lapidus, A. L.; Rodin,
A. P.; Pruidze, I. G.; Ugrak, B. I. Izv. Akad. Nauk SSSR, Ser. Khim. 1990, 7, 1661.
18 Browning, A. F.; Bacon, A. D.; White, C.; Milner, D. J. J. Mol. Catal. 1993, 83, L11.
19 (a) Caiazzo, A.; Settambolo, R.; Uccello-Barretta, G.; Lazzaroni, R. J. Organomet.
Chem. 1997, 548, 279. (b) Lazzaroni, R.; Settambolo, R.; Mariani, M.; Caiazzo, A. J.
Orgamomet. Chem. 1999, 592, 69.
20 (a) Settambolo, R.; Scamuzzi, S.; Caiazzo, A.; Lazzaroni, R. Organometallics 1998, 17,
2127. (b) Caiazzo, A.; Settambolo, R.; Pontorno, L.; Lazzaroni, R. J. Organomet.
Chem. 2000, in press.
21 Lazzaroni, R.; Bertozzi, S.; Pocai, P.; Troiani, F.; Salvadori, P. J. Organomet. Chem.
1985, 295, 371.
22 (a) Hanson, B. E.; Davis, N. E. J. Chem. Educ. 1987, 64, 928. (b) Lazzaroni, R.; Pertici,
P.; Bertozzi, S.; Fabrizi, G. J. Mol. Catal. 1990, 58, 75.
23 Botteghi, C.; Cazzolato, L.; Marchetti, M.; Paganelli, S. J. Org. Chem. 1995, 60, 6612.
24 Botteghi, C.; Marchetti, M.; Paganelli, S.; Sechi, B. J. Mol. Catal. A: Chem. 1997, 118,
173.
25 (a) Vidal. J. L.; Walker, W. E. Inorg. Chem. 1981, 20, 249 (b) Oldani, F.; Bor, G. J.
Organomet. Chem. 1983, 246, 309.
26 Evans, J.; Schwartz, J.; Urquhart, P. W. J. Organomet. Chem. 1974, 81, C37.
27 Casey, C. P.; Petrovich, L. M. J. Am. Chem. Soc. 1995, 117, 6007.
28 Sakai, N; Mano, S.; Nozaki, K.; Takaya, H. J. Am. Chem. Soc. 1993, 115, 7033.
29 Brown, J. M.; Cook, S. J.; Khan, R. Tetrahedron 1986,42, 5105.
30 Eckl, R. W.; Priermeier, T.; Herrmann, W. A. J. Organornet. Chem. 1997, 532, 243.
Chapter 3
3.1 Introduction
Figure 1. Actual rhodium catalyst containing various phosphorus and carbonyl ligands
3.2 Monophosphites
3.2.1 Catalysis
Figure 2. Structure of bulky phosphites (1, 3) and electron poor phosphite (2)
complex G was the only observable species. When all alkene was consumed
the rhodium hydride A was formed back. Both the spectroscopic and the
kinetic studies show that the rate limiting step for the hydroformylation of 1-
alkenes using the bulky phosphite modified rhodium catalyst is the
hydrogenolysis of the rhodium acyl intermediate.
wavenumber (cm-1)
Figure 4. Metal-CO region of IR spectra recorded with rapid-scan method after addition of 1-
octene to RhH(CO)31b
phosphite ligand. Probably, the phosphite ligand can exchange with carbon
monoxide under higher CO pressure as is also observed in the formation of
rhodium complexes of bulky diphosphites (see section 3.3.2).
Catalyst and ligand stability are important features in catalysis (for more
details see chapter 9). The sensitivity towards hydrolytic reactions and/or
solvolysis is strongly dependent on the structure of the phosphites. For
instance, UCC reported that the hydrolytic stability of bulky phosphites
could be improved by the use of bulky bisphenols in the ligand structure
(structure 3, figure 2). Ligand hydrolysis and other destructive side reactions
like Michaelis-Arbuzov rearrangement [62] are among the problems that can
be encountered when employing phosphites in catalysis (see figure 5). Many
examples of metal catalyzed Arbuzov-type decomposition reactions have
been reported [48]. The Arbuzov reaction is restricted to alkyl phosphites,
which is probably the reason that almost exclusively aryl phosphite ligands
are used in catalysis.
3. 3 Diphosphites
3.3.1 Catalysis
Under syn gas pressure the rhodium acac precursors were converted to
the catalytically active hydride complexes HRh(CO)2(L-L). The complexes
are generally assumed to have a trigonal bipyramidal structure and two
isomeric structures of these complexes are possible, containing the
diphosphite coordinated in a bisequatorial (ee) or an equatorial-apical (ea)
fashion. The structure of the complexes can be elucidated by (high pressure)
IR and NMR data (Table 3). In the carbonyl region of the infrared spectrum
the vibrations of the ee and ea complex can be easily distinguished. The ee
complexes typically show absorptions around 20 15 and 2075 cm-1 [24, 26,
3 1, 32], whereas the ea complexes exhibit carbonyl vibrations around 1990
and 2030 cm-1 [26,32]. The rhodium hydride vibration lies in the same
region as the carbonyl ligands but is often very weak. In fact additional
carbonyl signals in case of mixtures of complexes are often erroneously
assigned to the rhodium hydride stretching frequencies. Van Leeuwen et al.
were able to isolate some complexes as powders and the IR spectra were
measured as nujol mulls (see for IR data Table 3). For complex
50 Chapter 3
HKh(CO)2 (4b) it was found that three absorptions were present in the CO-
stretching region instead of two observed with IR in solution. Upon
measuring DRh(CO)2(4b), only two absorptions remained (2058, 201 2 cm-
1), which were shifted compared to HRh(CO)2(4b) (2035, 1998, 1990 cm-1).
This implies that the three frequencies in the hydrido complex are a
combination of two CO stretching frequencies and one rhodium hydrido
stretch. The rhodium-hydride vibration disappears upon deuteration of the
complex as the rhodium-deuteride vibration is situated in the fingerprint
region. The large frequency shift of the highest energy absorption is
indicative of a trans hydrido-CO relation [46]. In solution IR, the rhodium
hydride vibration and the lowest energy CO vibration overlap, which results
in two absorptions.
Additional information can be obtained from NMR data. Phosphorus
donors coordinated in the equatorial plane have generally small coupling
constants of phosphorus to hydrogen. The coupling constant of a trans
coordinated phosphite shows a large phosphorus to hydrogen coupling
constant of 180-200 Hz. The phosphorus to phosphorus coupling constant is
much larger for the ee (around 250 Hz) than for the ea complex (usually <70
Hz). Some representative NMR and IR data of rhodium-diphosphite
complexes are given in Table 3.
phosphorus to the hydride was observed, whereas the cis coupling constant
remained very small and unresolved [32]. By variable temperature NMR
studies the thermodynamic activation parameters could be determined by
comparison of the calculated and experimental NMR spectra (see figure 10).
The H, calculated from the Eyring equation, was 35 kJ.mol and the S
-1
1
Figure 10. Calculated and observed variable-temperature H (300 MHz) NMR spectra for
HRh(CO)213
3. Rhodium phosphite catalysts 53
Chan (see figure 12) [58]. They also observed the formation of red dimers
after losing two carbon monoxide molecules.
structural proof for the orthometallated species has not always been
convincing. Parshall reported the ortho-deuterium incorporation in phosphite
ligands [56]. In agreement with this, Coolen showed that the hydrogenation
catalyst HRh[P(OPh)3]4 provided ortho-deuterated product when stirred
under deuterium atmosphere [55]. Both researchers suggested that the
deuterium exchange proceeded via orthometallation. The intermediate
orthometallated complex was characterized by 31P NMR [55].
without any problem at relatively high rates. Next to formation of the desired
product methyl formylstearate (as a mixture of isomers) the rapid formation
of the trans isomer of methyl oleate, methyl elaidate, was observed. As
expected the hydroformylation of the trans isomer was slower than that of
the cis isomer (see figure 14). Comparison of the bulky phosphite modified
system with the rhodium triphenylphosphine catalyst revealed that the
reaction rates of the former were much higher, as expected for internal
alkene substrates.
time/h
Figure 14. Hydroformylation of pure methyl oleate (3.64 mmol), Pinitially = 20 bar, CO/H2 =
1:1, T=100 C, Rh=4.10-3 mmol, Ligand:Rh =25:1, methyl oleate,
methyl elaidate,
methyl fomylstearate, 0 methyl formylstearate (isomers)
Technical grade methyl oleate, however, contains about 14% of the diene
methyl linoleate that easily isomerizes to a conjugated diene. These
conjugated dienes strongly retard the hydrofomylation reaction, probably by
the formation of stable -allyl type rhodium complexes [43]. The initial rate
of the hydroformylation reaction is much lower than for the pure methyl
oleate substrate that does not contain significant amounts of dienes. Only
after the methyl linoleate has been converted to other hydrogenation and
hydroformylation products the rate of the reaction increases. Again the
reaction mixture shows isomerization of methyl oleate to the trans product
methyl elaidate during the hydroformylation process (see figure 15). Since
hydrogenation of the diene methyl linoleate to mono alkenes was required
prior to hydroformylation, the reaction rate increased significantly when
higher hydrogen pressures were applied.
3. Rhodium phosphite catalysts 57
Figure 15. Hydroformylation of technical grade methyl oleate (3.64 mmol), Pinitially = 20 bar,
CO/H2 = I: 1, T=100 "C, Rh = 4.10-3 mmol, Ligand : Rh =25: 1, A methyl oleate, methyl
elaidate, methyl linoleate, o methyl linoleate (isomers), methyl formylstearate, methyl
formylstearate (isomers)
formed it will react with the rhodium complex because the alkene addition is
much faster than for internal alkenes. If the catalyst is very selective and
gives a high linear to branched ratio in combination with fast isomerization,
only the linear rhodium alkyl intermediate will show preferential
carbonylation. As a consequence the linear aldehyde will be the main
product, even using internal alkenes as starting material. So far only few
catalysts are known that show these properties.
The strongly electron withdrawing monophosphite 2 was one of the first
examples that provided rhodium catalyst that gave good selectivity for the
formation of linear aldehyde starting from internal alkenes (see table 1) [8].
The high -value of the phosphite induces enhanced selectivity for the
formation of the rhodium alkyl complex and high isomerization rates,
resulting in a high overall selectivity for the linear aldehyde.
For lower alkenes such as 2-butenes UCC has achieved high contents of
linear products (see Table 2). Bryant reported 74% selectivity for the
formation of linear pentanal by hydroformylation of 2-butene using the
rhodium bulky diphosphite catalyst [22, 23].
Du Pont and DSM have patented the use of ligand 16 (Figure 16) for the
hydroformylation of methyl 3-pentenoate to linear product [33]. Instead of t-
butyl groups on the bridge they use electron-withdrawing ester groups, while
the remaining substituents are monophenols [34, 35], rather than diols [36,
37] or bisphenols. The necessary isomerization to the linear alkene prior to
hydroformylation reaction is probably promoted by the electron withdrawing
ester groups. They were able to obtain high selectivity for the linear
aldehyde, providing a useful intermediate to nylon-6 feedstocks. Comparable
selectivity up to 97% was reported for the hydroformylation of 2-hexene.
The hydroformylation of butadiene to the linear 1,6-hexanedial proved to
proceed less satisfactory; the main products were pentanal and 3-pentenal
and only small amounts of the desired 1,6-hexanedial were formed.
Similar results were obtained using the UCC systems; in several patents
the reported yields for dialdehyde in the hydroformylation of butadiene were
3. Rhodium phosphite catalysts 59
relatively low [60]. Moderate yields for the linear aldehyde were generally
obtained when methyl 3-pentenoate was hydroformylated with the bulky
diphosphite modified UCC catalyst [60]
The efficacy of monophenols containing bulky substituents is described
in the patents from Mitsubishi Chemical Corporation [34,35] (17, Figure 16).
They report high yields of aldehyde (> 90%) for the hydroformylation of 1-
alkenes with high linearities. The 1:b ratios are generally above twenty using
this ligand.
Figure 17. Calix[4]arene based diphosphite by BASF [44] and a monophosphite employed in
hydroformylation [45]
References
1 Young, J. F.; Osbom, J. A.; Jardine, F. A.; Wilkinson, G. J. Chem. Soc., Chem. Comm.
1965, 131. (b) Evans, D.; Osborn, J. A.; Wilkinson, G . J. Chem. Soc. (A) , 1968, 3133. (c)
Evans, D.; Yagupsky, G.; Wilkinson, G. J. Chem. Soc. A 1968, 2660. (d) Brown, C. K.;
Wilkinson, G. J. Chem. Soc. A 1970, 2753.
2 Heck, R. F; Breslow, D. S. J. Am. Chem. Soc. 1961, 83, 4023.
3 Cavalieri d'Oro, P.; Raimondt, L.; Pagani, G.; Montrasi, G.; Gregorio, G.; Oliveri del
Castillo, G. F.; Andreeta, A. Symposium on rhodium in homogeneous catalysis,
Veszprem, 1978, pp 76-83; b) Gregorio, G.; Montrasi, G.; Tampieri, M.; Cavalieri d'Oro:
P.; Pagani, G.; Andreetta, A. Chim. Ind. 1980, 62(5), 389; Cavalieri d'Oro, P.; Raimondi,
L.; Pagani, G.; Montrasi, G.; Gregorio, G.; Andreetta, A. Chim. Ind. 1980, 62, 572
4 Van Leeuwen, P. W. N. M.; Van Koten, G. Catalysis. An Integrated approach to
Homogeneous, Heterogeneous and Industrial Catalysis, Moulijn, J. A.; Van Leeuwen, P.
W. N. M.; Van Santen, R. A. Eds., Elsevier: Amsterdam. London, New York, Tokyo,
2nd ed. 1995, pp 199-222.
5 Unruh, J. D.; Christenson, J. R. J. Mol. Catal. 1982, 14, 19.
6 L. A. van der Veen, M. D. K. Boele, F. R. Bregman, P. C. J. Kamer, P. W. N. M. van
Leeuwen, K. Goubitz, J. Fraanje, H. Schenk and C. Bo, J. Am. Chem. Soc., 1998, 120,
11616.
7 Pruett, R. L.; Smith, J. A. J. Org. Chem. 1969, 34, 327. (b) Pruett, R. L.; Smith, J. A. S.
African Pat. 6804937, 1968 (to Union Carbide Coorporation); Chem. Abstr. 1969, 71,
90819.
8 (a) Van Leeuwen, P. W. N. M.; Roobeek, C. F. J. Organomet. Chem. 1983,258, 343. (b)
Van Leeuwen, P. W. N. M.; Roobeek, C. F. Brit. Pat. 2 068 377, 1980 (to Shell); Chem.
Abstr-. 1984, 101, 191 142.
9 Jongsma, T.; Challa, G.; van Leeuwen, P. W. N. M. J. Organonmetal. Chem. 1991, 421,
121.
10 Tolman. C. A. Chem. Rev. 1977, 77, 313.
11 Van Rooy, A.; Orij, E. N., Kamer, P. C. J.; Van den Aardweg, F.; Van Leeuwen, P. W.
N. M. J. Chem. Soc., Chem. Commun. 1991, 1096.
12 Van Rooy, A.; Orij, E. N.; Kanier, P. C. J.; Van Leeuwen, P. W. N. M. Organometallics,
1995, 14, 34.
13 Brown, J. M.; Kent, A. G. J. Chem. Soc. Perkin Trans. II 1987, 1597.
14 Van Rooy, A. PhD thesis, University of Amsterdam 1995.
15 Collman, J. P.; Hegedus, L. S.; Norton, J. R.: Finke, R. G. Principles and Applications of
Organotransition Metal Chemistry University Science Books: Mill Valley, CA 2nd ed.
1987.
16 Trzeciak, A. M.; Zilkowski, J. J.; Aygen, S.; Van Eldik, R. J. Mol. Cutal. 1986, 34, 337.
(b) Trzeciak, A. M.; Zilkowski, J. J. Trans. Met. Chem. 1987, 12 , 408. (c) Trzeciak, A.
M.; Zilkowski, J. J. Inorg. Chim. Acta Lett. 1982, 64, L267. (d) Trzeciak, A. M.;
Zilkowski, J. J. J. Mol. Catal 1986,34, 213. (e) Janecko, H.; Trzeciak, A. M.;
Zilkowski, J. J. J. Mol. Catal. 1984, 26, 355. (f) Trzeciak, A. M.; Zilkowski, J. J. J.
Mol. Cutal. 1987, 43, 13.
17 Yoshinura, N.; Tokito, Y. Eur. Pat. 223 103, 1987 (to Kuraray); Chem. Abstr 1987, 107,
154896. (b) Omatsu, T. Eur. Pat. 0 303 060, 1989 (to Kuraray); Chem. Abstr. 1989, 111,
38870.
18 Polo, A.; Real, J.; Claver, C.; Castilln, S.; Bayn, J. C. J. Chem. Soc. Chem. Commnun.
1990, 600.
3. Rhodium phosphite catalysts 61
19 (a) Polo, A.; Femandez, E.; Claver, C.; Castilln, S. J. Chem. Soc. Chem. Commun.
1992, 639. (b) Fernandez, E.; Claver, C.; Castilln, S.; Polo, A.; Piniella, J. F.; Alvarez-
Larena, A. Organometallics 1998, 17, 2857.
20 Muilwijk, K. F.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Am. Oil Chem. Soc. 1997,
74, 223.
21 Jongsma, T.; Fossen, M.; Challa, G.; van Leeuwen, P. W. N. M. J. Mol. Catal. 1993, 83,
17. (b) Jongsma, T.; van Aert, H; Fossen, M.; Challa, G.; van Leeuwen, P. W. N. M. J.
Mol. Catal. 1993,83, 37.
22 Billig, E.; Abatjoglou, A. G.;. Bryant, D. R. U.S. Pat. 4,668,651; Eur. Pat. Appl.
213,639, 1987 (to Union Carbide); Chem. Abstr. 1987, 107, 7392.
23 Billig. E.; Abatjoglou, A. G.; Bryant, D. R.; Murray, R. E.; Maher, J. M. U.S. Pat.
4,599,206, 1986 (to Union carbide); Chem. Abstr. 1988, 109, 233177.
24 van Rooy, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Goubitz, K.; Fraanje, J.;
Veldman, N.; Spek, A. L. Organometallics, 1996, 15, 535.
25 Babin, J. E.; Whiteker, G. T. W.O. 93 03839, US. Pat. 91 1 518, 1992 (to Union Carbide
Corporation); Chem. Abstr. 1993, 119, 159872.
26 Buisman, G. J. H.; Vos, E.; Kamer, P. C. J.; Van Leeuwen, P. W. N. M. J. Chem. Soc.,
Dalton Trans., 1995, 409.
27 Casey, C. P.; Whiteker, G. T.; Melville, M. G.; Petrovich, L. M.; Gavney Jr., J. A.;
Powell, D. R. J. Am. Chem. Soc. 1992, 114, 5535.
28 Kranenburg, M.; van der Burgt, Y. E. M.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.
Organometallics. 1995, 14, 3081.
29 van der Veen, L. A.; Keeven, P. H.; Schoemaker, G. C.; Reek, J. N. H.; Kamer, P. C. J.;
van Leeuwen, P. W. N. M.; Lutz, M.; Spek, A. L. Organometallics, 2000, 19, 872.
30 Tolman, C. A.; Faller, J. W. Homogeneous Catalysisis with Metal Phosphine Complexes,
Pignolet, L. H. Ed., Plenum Press: New York and London, 1983; Chapter 2, pp 88-89.
31 Moasser, B.; Gladfelter, W. L.; Roe, C. D. Organometallics 1995, 14, 3832.
32 Buisman, G. J. H.; van der Veen, L. A.; Kanier, P. C. J.; van Leeuwen, P. W. N. M.
Organometallics 1997, 16, 5681.
33 Burke, P. M.; Garner, J. M.; Tam, W.; Kreutzer, K. A.; Teunissen, A. J. J. M. WO
97133854, 1997, (to DSM/Du Pont); Chem. Abstr. 1997, 127, 294939.
33 Sato, K.; Karawagi, J.; Tanihari, Y. Jpn. Kokai Tokkyo Koho JP 07,278,040 (to
Mitsubishi); Chem. Abstr. 1996, 124, 231851.
35 Sato, K.; Kawaragi, Y.; Takai, M.; Ookoshi, T. US Pat. 5,235,113, EP 518241 (to
Mitsubishi); Chem. Abstr. 1993, 118, 191183.
36 Rper, M.; Lorz, P. M.; Koeffer, D. Ger. Offen. DE 4,204,808 (to BASF); Chem. Abstr.
1994,120, 133862.
37 Lorz, P. M.; Bertleff, W.; Roper, M.; Koeffer, D. EP 472,071 (to BASF); Chem. Abstr.
1992,117,34513.
38 Cuny, G. D.; Buchwald, S. L. J. Am. Chem. Soc. 1993, 115, 2066.
39 Johnson, J. R.; Cuny, G. D.; Buchwald, S. L. Angew. Chem. Int. Ed. Engl. 1995, 34,
1760.
40 Van den Hoven, B. G.; Alper, H. J. Org. Chem. 1999, 64, 3964.
41 Van den Hoven, B. G.; Alper, H. J. Org. Chem. 1999, 64, 9640.
42 van Rooy, A.; de Bruijn, J. N. H.; Roobeek, K. F.; Kanier, P. C. J.; van Leeuwen , P. W.
N. M. J. Organomet. Chem. 1995, 507, 69.
43 (a) Van Leeuwen, P. W. N. M.; Roobeek, C. F. Brit. Patent Appl. 33,554, 1981 (to Shell
Resareh); Chem. Abstr. 1982, 96, 6174. (b) Van Leeuwen. P. W. N. M.; Roobeek, C. F.
J. Mol. Catal., 1985, 31, 345.
62 Chapter 3
44 Paciello, R.; Rper, M.; Kneuper, H.-J.; Langguth, E.; Peter, M. DE 4321194, 1995,
(BASF AG); Chem. Abstr. 1995, 122, 160937. (b) Paciello, R.; Siggel, L.;Rper, M.
Angew. Chem. Int. Ed. Engl. 1999, 38, 1920. (c) Paciello, R.; Siggel, L.; Walker, N.;
Rper, M. J. Mol. Catal. A 1999, 143, 85.
45 Parlevliet, F. J.; Kiener, C.; Fraanje, J.; Goubitz, K.; Lutz, M.; Spek, A. L.; Kamer, P. C.
J.; van Leeuwen, P. W. N. M. J. Chem. Soc. Dalton Trans. 2000, 1113.
46 Vaska, L. J. Am. Chem. Soc. 1966, 88 , 4100.
47 Meakin, P.; Jesson, J. P.; Tebbe, F. N.; Muetterties, E. L., J. Am. Chem. Soc., 1971, 93,
1797. (b) Meakin, P.; Muetterties, E. L.; Jesson, J. P., J. Am. Chem. Soc., 1972, 94,
5271
48 Brill, T. B.; Landon, S. J. Chem. Rev. 1984, 84,577. (b) Werener, H.; Feser, R. Z Anorg.
Allg. Chem. 1979, 458, 30 1.
49 Whyman, R., J. Chem. Soc.. Chem. Commun., 1970, 230-231.
50 Booth, B. L.; Else, M. J.; Fields, R.; Haszeldine, R. N., J. Organomet. Chem., 1971, 27,
119-131.
51 Cattermole, P. E.; Osborne, A. G., Inorg. Synth., 1977, 17, 115.
52 Castellanos-Pez, A.; Castilln, S.; Claver, C.; van Leeuwen, P. W. N. M.; de Lange, W.
G. J. Organometallics, 1998, 17, 2543.
53 James, B. R.; Mahajan, D.; Rettig, S.J .; Williams, G. M. Organometallics, 1983, 2, 1452
54 Moasser, B.; Gladfelter, W. L. Inorg. Chim. Acta 1996, 242, 125.
55 Coolen, H. K. A. C.; van Leeuwen, P. W. N. M.; Nolte, R. J. M. J. Organornet. Chem.
1995, 496, 159.
56 Parshall, G. W.; Knoth, W. H.; Schunn, R. A. J. Am. Chem. Soc. 1969, 91,4990.
57 Lazzaroni, R.; Rafaelli, A.; Settambolo, R.; Bertozzi, S.; Vitulli, G. J. Mol. Catal., 1989,
50, 1. (b) Lazzaroni, R.; Settambolo, R.; Rafaelli, A.; Pucci, S.; Vitulli, G. J. Organomet.
Chem. 1988, 339, 357.
58 Chan, A. S. C.; Shieh, H. S.; Hill, J. R. J. Chem. Soc., Chem. Commun. 1983, 688.
59 Bryant, D. R; Leung, T. W. US patent 5,741,943, 1998, (to UCC); Chem. Abstr. 1998,
128, 283881.
60 Packett, D. L.; Briggs, J. R.; Bryant, D. R.; Phillips, A. G. U.S Patent 5,886,237 (to
UCC); Chem. Abstr. 1998, 128, 14354.
61 Buhling, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Mol. Catal. A: Chemical,
1995, 98, 69.
62 Bhattacharya, A. K.; Thyagarjan, G. Chem. Rev. 1981, 81, 415.
Chapter 4
Phosphines as ligands
Bite angle effects for diphosphines
4.1.1 Introduction
63
P.W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 63-105.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
64 Chapter 4
isomeric forms ae and ee, having a hydride in an apical position and ethene
coordinating in the equatorial plane. While four-coordinate species rhodium
3c (cis) have never been observed, the analogous four-coordinate complex
for 3t trans-Rh(PCy3)2(CO)H has been isolated and structurally
characterized [56]. In addition, magnetization transfer experiments indicate
that both 1 and 2 are in equilibrium with free PPh3 [8]. These results are
highly suggestive of the formation of the four-coordinate complex 3 by
phosphine dissociation from 1 and 2. The complexes never observed
experimentally are shown in brackets in Figure 1.
Although the complexes shown in Figure 1 contain at least two
coordinated PPh3 ligands, a large body of indirect evidence suggests the
influence of equilibria involving catalytically active species containing a
single coordinated phosphine ligand. Such monophosphine species were
initially invoked to explain the positive effect on hydroformylation
regioselectivity that is typically seen with increasing P:Rh ratios (see 4.1.6).
The 31P NMR magnetization transfer experiments described by Brown [8]
also indicate that PPh3 dissociation from the RhL2 complex 2 occurs, albeit
at a significantly slower rate than the corresponding PPh3 dissociation from
tris-phosphine complex, 1. In addition, IR spectra have revealed the
presence of Rh[P(OAr)3](CO)3H with a very bulky coordinated phosphite
ligand [48].
It has not been established experimentally whether alkene complexation
is reversible or not; in the scheme we have drawn all steps but the
hydrogenolysis at the end as reversible. Brown observed that complex 1
catalyzes the isomerization of cis-1,2-dideuteriostyrene under nitrogen [8].
Although under these conditions the isomerization of this labeled alkene is
slower than the rate of hydroformylation using 1, this result suggests that
alkene coordination and insertion into the Rh-H bond can be reversible
processes. Complex 4 undergoes a migratory insertion to give square-planar
alkyl complexes 5c and 5t, which are respectively cis or trans. In the
absence of CO, the 16-electron complexes of this type can be isolated as
mixtures of cis and trans isomers for aryl groups or alkyl groups not
containing -hydrogen atoms, [9]. Complex 5 can undergo -hydride
elimination, thus leading to isomerization when higher alkenes are used, or it
can react with CO to form trigonal bipyramidal complexes 6. Thus, under
low pressure of CO more isomerization may be expected. At low
temperatures (< 70 C) and a sufficiently high pressure of CO (> 10 bar) the
insertion reaction is usually irreversible and thus at this point also the
regioselectivity of hydroformylation of -alkenes is determined.
Complexes 6 undergo the second migratory insertion in this scheme to
form the acyl complexes 7. Complexes 7 can react either with CO to give the
saturated acyl intermediates 8, which have been observed spectroscopically,
66 Chapter 4
4.1.5 Kinetics
(conditions 90-1 10 C, p(CO) = 1-25 bar, p(H2) = 1-45 bar, [PPh3] = 0.05-5 M,
[Rh] = (0.5-7). 10-3 M, PPh3/Rh = 300:1 to 7:1, [propene]t=0 = 2-7 M)
They reported an overall activation energy of 84 kJ.mol-1 for the process.
The important features of this kinetic study are the zero order dependence in
dihydrogen, the negative order in PPh3 ligand (and CO), and the positive
70 Chapter 4
Under standard conditions, we propose that the best starting point for
the kinetics is an equation of the type:
A [alkene][Rh]
Rate(type - I ) =
B +[L ]
Two possible scenarios exist which are consistent with this rate equation.
Rate determining alkene coordination by 3c and 3t followed by rapid alkene
insertion into the Rh-H bond is one possibility. An equally valid alternative
explanation for the observed kinetics involves rate determining migratory
insertion of alkene into Rh-H preceded by fast, reversible alkene
coordination. In both cases, the concentrations of coordinatively unsaturated
3c and 3t are influenced by PPh3 and CO concentrations.
The kinetics at moderate PPh3 concentrations are in agreement with a
catalyst resting state of RhH(PPh3)2(CO)2 and a transition state of
composition RhH(PPh3)2(CO)(alkene). At very high PPh3 concentrations, the
likely resting state is RhH(PPh3),(CO) and a transition state of composition
RhH(PPh3)2(CO)(alkene) is again likely. The set of rate-limiting reactions
are dissociation of CO (or PPh3), complexation of alkene, and migratory
insertion. Dissociation/association of CO is reversible and faster than
hydroformylation for arylphosphines (see 4.2). Complexation of alkene is
most likely reversible, although there are no experimental data on this
process. Theoretical studies [26] indicate that complexation of CO or ethene
to species 3 (Figure 1) has a low barrier. For migratory insertion of ethene
into the metal-hydride an early transition state was found, involving a re-
organization of the complex. This means that steric hindrance will play a
crucial role and especially the rotation of the alkene from the in-plane
coordination to a perpendicular coordination mode contributes to the barrier.
There is a strong body of evidence in the literature in disagreement with
the kinetics presented above [2-4, 27, 28]. Ever since the first statements of
the Wilkinson group [2a, not repeated though in 2c!] the oxidative addition
of hydrogen has been considered as the rate-determining step [5, 6]. The first
suggestion was based on an analogy with the hydrogenation reaction
employing Wilkinsons catalyst RhC1(PPh3)3. In the textbooks [6] and
reviews [5] oxidative addition of dihydrogen is still mentioned as the most
likely rate-determining step. The theoretical work cited shows the highest
barriers for the migration reactions [21, 26], but yet supports the common
point of view!
72 Chapter 4
4.1.6 Regioselectivity
the aldehyde product is high, but the overall selectivity to linear aldehyde
may be low. This will be discussed further in 4.2.
Clearly, the number of phosphines coordinated to rhodium, along with
the stereochemistry at Rh, determines the regioselectivity. At high PPh3
concentrations, the resting state of the catalyst is 1 (Figure 1, 2), which
undergoes PPh3 dissociation to form selectively 4-coordinate intermediate
3t. At lower PPh3 concentrations, the resting state is an equilibrating mixture
of 2ae and 2ee which undergo either CO dissociation to form 3c and 3t or
PPh3 dissociation to form 10c and 10t. According to NMR studies,
conversion between axial and equatorial phosphines in 2ae and 2ee is many
orders of magnitude faster than the rate of hydroformylation [8]. The rate of
hydroformylation is only two orders of magnitude slower than that of carbon
monoxide exchange of species of type 2 [31]. The relative concentrations of
these four-coordinate diphosphine (3) and monophosphine (10)
intermediates are controlled by the PPh3 and CO concentrations. Wilkinson
[2] and Andreetta [24b] suggested that species 3, formed at high PPh3
concentration, lead to higher linear aldehyde selectivity (1:b = 20: 1], and that
species 10, containing only one phosphine, lead to a lower selectivity for
linear aldehyde (1:b = 4: 1) (see Figure 2).
Dissociation of PPh3 from 1 gives 3t, dissociation of CO from 2ee also
gives 3t, while dissociation of CO from 2ae will give 3c. Dissociation of
PPh3 from 2 gives the isomers 10t and 10c. The use of diphosphines such as
dppe and dppp gives modest linear-branched ratios and their putative
intermediates 3 must have cis structure 3c. Therefore, 3c is not a likely
intermediate for the selective hydroformylation. For linear-branched ratios of
4:1 we don't have to invoke mono-phosphine species of type 10, as already
structure 3c suffices for the production of moderate I:b ratios (3:1).
Hydroformylation regioselectivity can be established by either four-
coordinate or five-coordinate intermediates. In one possible mechanism, the
regioselectivity is controlled by the relative concentrations of four-
coordinate intermediates, 3c, 3t, 10c and 10t, which undergo irreversible
formation of alkene hydride complexes, 4. Accordingly, this mechanism
requires that interconversion of isomeric pairs 4ee/4ae and 4e/4a be slow
relative to irreversible migratory insertion of alkene.
Alternatively, the regioselectivity can be determined during migratory
insertion of alkene into the Rh-H bond in five-coordinate intermediates, 4.
The concentrations of these five-coordinate alkene hydride complexes are
controlled by the relative concentrations of PPh3 and CO. In addition, these
stereoisomers (4ee/4ae, 4e/4a), which differ in the nature of the apical
ligand, presumably undergo rapid interconversion via a pseudorotation
mechanism, as has been observed for 2ee/2ae [8].
In both scenarios, the dependence of I:b on PPh3 concentration is
attributed to changes in the extent of phosphine ligation, and the
stereochemistry at Rh is also proposed to influence regioselectivity. Studies
using diphosphine ligands with well-defined coordination number and
geometry (section 4.2) have provided much insight into these factors that
control hydro formylation regioselectivity.
Substituted alkenes will be discussed in Chapters 5 and 6. From a
mechanistic viewpoint the results obtained with substituted 1 -alkenes are
worth mentioning as they stress the importance of steric interactions. Simple
1-alkenes carrying methyl substituents at the carbon atoms 3 or 4 give
progressively more linear product for a higher degree of substitution closer
to the alkene bond, without losing much activity as is shown in Table 1[ 18].
The table shows that in most instances the rate of hydroformylation to the
terminal product is retained and that a remarkably high linearity can be
obtained when steric hindrance is provided at carbon-3. Substitution at
carbon-2 under these conditions gives rates of only 100 mol.mol-1.h-1, and
4. Phosphines as ligands 75
only linear product is formed. Internal alkenes also show rates that are one to
two orders of magnitude lower than those of 1-alkenes [2].
Cis-2-alkenes react a few times faster than trans-2-alkenes. Steric
influences, both from the complex part and the alkene, play a dominant role
in determining the fate of hydroformylation of alkenes not containing
functional groups. For alkenes such as styrene, 3,3,3-trifluoropropene,
acrylates, and vinyl esters, the electronic properties of the alkene determine
to a large extent the regioselectivity.
4.1.7 Conclusion
4.2.1 Introduction
apparently not attractive for the formation of such complexes. Bayn and
coworkers observed that constrained I ,2-diphosphines form dimeric
complexes in addition to the expected monomeric ones [36].
Studies using dppe or dppp under standard conditions, in the absence
of PPh3, have not been published in the seventies, but it would seem likely
that several industries have conducted research in this field. As described in
4.1.1, two explanations for the low regioselectivities observed with such
diphosphines can be brought forward. Resting state 11a can lead to 3c by
dissociation of CO. Since the small bite angle imposed by chelates such as
dppe and dppp cannot accommodate the trans orientation in 3t, the resultant
cis complex 3c leads to low regioselectivity. Alternatively, 3c can reversibly
add alkene to form the five-coordinate alkene hydride complex, which
contains an aical-equatorial diphosphine chelate. This ae isomer,
analogously to 4ae, leads to low regioselectivity. In addition to these two
possibilities, 11a may also undergo dissociation of one phosphine donor
giving 12t, the analog of unselective 10t. The kinetics would be different for
the two possibilities, but unfortunately these have not been reported.
In the following we will present the results obtained with ferrocene based
ligands and a variety of bidentate phosphine ligands containing bite angles
larger than 100. Since the catalysis for these ligands has been carried out at
somewhat different conditions we will present them separately and discuss
the results in three parts. Finally we will speculate about a general view in an
attempt to find a global explanation.
The table shows that with increasing accepting ability, i.e. increasing -
value (the authors used Hammett constants), the rate of the reaction
increases and so does the 1:b ratio. For phosphites a similar trend had been
observed, see Chapter 3. The rate increase can be easily understood, as
dissociation of CO will be more facile when -back-donation decreases.
Assuming that steric properties are the same coordination of alkene may also
be enhanced by a more electrophilic rhodium center. Rates of migration
reactions are usually not affected by electronic changes according to MO
studies [21], thus we observe an overall rate increase with electron
withdrawing phosphines. Steric properties of the ligands are the same,
except that the electronic changes do have an influence on the mode of
coordination as we shall see later, but this was not yet known at the time.
Substituting trifluoromethyl groups for hydrogens in other arylphosphines
(trans-2,3-dppm-nor) also has a favorable effect on rate and selectivity
[15c].
The I:b ratio of the product aldehyde increases even though electron-
withdrawing phosphines give rise to increasing amounts of alkene
isomerization. The total preference for linear alkyl formation still increases
when the 2-hexene is included in the total amount of intermediary 2-
hexylrhodium that must have been formed. Thus, we can say that the
intrinsic preference for 1 -alkyl formation increases with increasing -values
of the ligand. Figure 6 explains the influence of 2-hexene formation on the
observed 1:b ratio in the aldehyde product in a system in which the internal
alkene is not converted into (branched) aldehyde. The higher I:b ratio for
more positively charged metal centers was supposed to originate from a
higher preference for 1 -alkyl formation for such centers. An increasing rate
of isomerization is a general observation, also for phosphite calalysts. Note
that none of the species in Figure 6 has been observed experimentally.
The authors note that under their conditions, involving relatively high
temperatures and low pressures, many steps of the cycle have similar rates,
as the kinetics are rather complex. The best results were obtained when an
excess of phosphine was employed. A ligand to metal ratio of 1.5 was
preferred, as at higher ratios the selectivity did not improve further.
80 Chapter 4
This led Unruh and Christenson to the conclusion that the most selective
species was one containing more than two phosphorus ligands, while the
current view was, Wilkinsons view, that complexes containing two
phosphines were the ones leading to high 1:b ratios of 20. They propose that
the most selective species contains three phosphorus atoms, one dppf as a
bidentate on one metal and one dppf bridging between two metals. The two
metals are to function independently in the catalysis.
Complexes of this type have indeed been identified [41] in 31P NMR
studies. A whole range of diphosphines was studied including those of
Figure 4, Figure 5, and the common diphosphines dppe and dppp. The study
involved the addition of various amounts of the diphosphine to complex 1.
The spectra were all interpreted in terms of the species shown in Figure 7.
13 14 15
Figure 7. Hughess phosphine complexes 13, 14, and 15
4. Phosphines as ligands 81
typically ranging from 100 to 130 Hz, clearly show that in most cases tris-
phosphine complexes are present having all phosphorus atoms in the
equatorial plane. In a few instances the coupling constants are smaller (50 -
100 Hz) indicating that the phosphines occupy in part an apical position,
probably involving a fast exchange. These results support the catalytic
results suggesting that three phosphine ligands are coordinated to the catalyst
in its resting state. However, it should be borne in mind that the
concentrations for the catalysis experiments and the NMR experiments differ
considerably. In catalysis the free ligand concentration is at most 1.2 mM
([Rh] = 0.6 mM) and p(CO) = 4 bar, while in the NMR experiments the
concentrations are one order of magnitude, maybe two, higher (27 mM for
[Rh] and 130 mM for phosphine in one example) and no CO is added. Thus
one might wonder whether indeed tris-phosphine species are present under
the reaction conditions of the catalytic experiments.
Re-examination of the dppf-Rh system has shown that the coordination
behavior under syn-gas is different [42]. Both NMR spectra and in situ IR
([Rh] = 8 mM, [dppf] = 17 mM, p(CO) = 20 bar, 40 C) spectra show that
the only species under these conditions is RhH(dppf)(CO)2 being a mixture
of rapidly equilibrating 2ae (approximately 80 %) and 2ee (20 %). The 1:b
Figure 9. BISBI, 17, and its complexes 18, and 19, and ligands 20 and 21
bite angle ligands, it is suggested that the extra strain imposed by chelation
be calculated to determine whether the new ligand will from a stable chelate.
Hydroformylation of 1-hexene under mild conditions (34 C, 6 bar of syn
gas, 4mM solution of 19 )gave a 1:b ratio of 66.5 for BISBI and for trans-
dppm-cyp it was 12. DIOP and dppe afforded values of 8.5 and 2.1. No
isomerization was observed. It was concluded that apparently a bis-
equatorial coordination of a diphosphine lead to high 1:b ratios, although the
reason for this was not understood in detail.
The correlation between natural bite angle and 1:b ratio is very interesting
and was an important finding for future work. Thus, complexes 2ee (Figure
1) lead to higher 1:b values than complexes having structure 2ae, or
alternatively ligands with larger bite angles lead to more effective steric bulk
of the diphosphine and a relative destabilization of branched alkyl
intermediates.
Before developing detailed explanations of the control of regiochemistry
based on steric or electronic differences between bis-equatorial and
equatorial-apical diphosphine complexes it was considered crucial to test
whether alkene insertion to produce 1 -alkyl and 2-alkyl rhodium species was
irreversible [47c]. For instance, previous work on the deuterioformylation of
cyclohexene using bulky phosphite had shown that H-D exchange was at
least ten times faster than formylation (40-80 C, < 10 bar) [48], but for 1-
alkenes insertion is irreversible, provided that the CO concentration is
maintained at a sufficiently high level [49]. Similar results have been
reported for unmodified carbonyl catalysts (see Chapter 2). In the alternative
case of fast alkene exchange, the regioselectivity would depend upon the
equilibrium of 1- and 2-alkylrhodium species and their respective rates of
CO insertion. Deuterioformylation of 1-hexene at 6 bar and relatively low
temperature (32 C) has shown that insertion of the 1-alkene is an
irreversible process as very little deuterium was incorporated in recovered
alkenes. The small amounts of secondary alkyl chains that are formed lead
mainly to (deuterated) 1- or 2-hexene and only 25 % proceeds to branched
4. Phosphines asligands 85
a
Ratio of diequatorial:apical-equatorial chelates of (diphosphine)Ir(CO)2H at room
c
temperature.b Mol heptanal:mol 2-methylhexanal. Turnover rate = [mol aldehyde] x [mol
-1 -1
Rh] h- .
energies of the ee and ae isomer of the complex modified with PH3 as model
-1
ligand for a basic phosphine were the same within 0.1 keal mol . Uisng PF3
as a model ligand for a nonbasic phosphine the ee isomer was found to be
favored by 2.4 keal mol-1. This corresponds to an ee:ae equilibrium ratio of
approximately 50:1 at room temperature. This trend is consistent with the
experiments.
Hydroformylation results in Table 6 show that, with the exception of
ligands 39 and 40, the rate of the reaction increases with decreasing
phosphine basicity. An explanation for the deviant behavior of 39 and 40 can
be incomplete catalyst formation or deactivation of the catalyst. Decreasing
phosphine basicity facilitates CO dissociation from the
(diphosphine)Rh(CO)2H complex and enhances alkene coordination to form
the (diphosphine)Rh(CO)H(alkene) complex, and therefore, the reaction rate
increases.
While the 1:b ratio increases with the -value, no electronic effect ofthe
ligands on the selectivity for hear aldehyde is observed. The selectivities
for linear aldehyde are all between 92 and 93%. The increase in 1:b ratio with
decreasing phosphine basicity can be attributed completely to an increased
tendency of the branched alkyl rhodium species to form 2-octene instead of
branched aldehyde (Figure 6). The increasing electrophilicity of the rhodium
center leads to a higher reactivity of the rhodium alkyl species toward CO
dissociation and -hydrogen elimination [ 15a]. The total amount of all other
products (branched aldehyde and 2-octene) is 7-8% for all ligands.
The constant selectivity for linear aldehyde in the hydroformylation of 1-
octene implies that for the basic ligands the 1:b ratio reflects the
regioselectivity of the formation of the rhodium alkyl species. For the less
basic ligands the increase in 1:b ratio results from the different behavior of
branched and linear rhodium alkyls toward -hydrogen elimination, as was
already reported by Lazzaroni and co-workers for the deuteriofomylation of
1-hexene [57]. The linear alkyl is mainly converted to linear aldehyde, while
4. Phosphines as ligands 93
the branched alkyl partially generates 2-octene. Since 2-octene is far less
reactive in the hydroformylation, its formation is irreversible in these
experiments. We conclude that for this catalytic system the ratio of linear to
branched rhodium alkyl species formed is determined not by phosphine
basicity but by steric constraints only.
13
Kinetics of CO dissociation from (Diph osphine)Rh ( 13 CO) 2H Complexes.
The rate determining step in the hydroformylation of 1-octene with
xantphos-type ligands is in an early stage in the catalytic cycle. Dissociation
of CO, alkene coordination, and migratory insertion are relevant for the
determination of the rate. So far, only the first step has been studied
separately [54]. An experimental set-up was introduced enabling the
measurement of the rate constants for CO dissociation from the
(diphosphine)Rh(CO)2H complexes. By exchanging the 12CO ligands for
isotopically pure 13CO ligands, the carbonyl absorptions in the IR spectra of
the (diphosphine)Rh(CO)2H complexes shift 30-40 cm-1 to lower energy.
a 13 13
Reaction conditions: [(diphosphine)Rh( CO)2H] = 2.00 mM in cyclohexane, P( CO) = 1
bar, P(H2) = 4 bar, T = 40 C, diphosphine/Rh = 5. b In mol.mol-1.h-1 at 80 C and 20 bar
c
CO/H2 [(diphosphine)Rh(13CO)2H] = 3.00 mM.d See Figure 17 below.
steric congestion around the rhodium center and consequently in more steric
hindrance for the alkene entering the coordination sphere. What kind of
electronic effect the widening of the bite angle has on the activation energy
for alkene coordination is unclear, since it depends on the bonding mode of
the alkene. Rhodium to alkene back-donation is promoted by narrow bite
angles, while alkene to rhodium donation is enhanced by wide bite angles
[62].
Widening the bite angle of a cis bidentate in a square planar complex
would certainly accelerate a migration reaction, but is is not clear how this
would work out in trigonal bipyramid having the diphosphine as a bis-
equatorial ligand. We will return to this in section 4.2.5.
They constitute the first rhodium phosphine modified catalysts for such a
selective linear hydroformylation of internal alkenes. The extraordinary high
activity of 43 even places it among the most active diphosphines known.
Since large steric differences in the catalyst complexes of these two ligands
are not anticipated, the higher activity of 43 compared to 42 might be
ascribed to very subtle bite angle effects or electronic characteristics of the
phosphorus heterocycles.
The high 1:b ratios of ligands 42 and 43 are in good agreement with
earlier observations that diphosphines with natural bite angles close to 120
give very selective catalysts. The calculated bite angles are largely due to the
interference of the rigid substituent rings at phosphorus. Table 8 shows the
acitivities and selectivities achieved for internal alkenes. To enhance the rate
of isomerization and to prevent hydroformylation of the internal alkenes, a
relatively high temperature and low pressure were used. The dissociation
rate of CO from the resting state (43)RhH(Co)2 was very fast, several times
faster than that found for diphenylphosphino Xantphos ligands (see Table 7).
Thus larger (natural) bite angles promote the formation of four-coordinate
species, as may be expected. For the same reason they will enhance
isomerization, since this involves a competition between CO association and
-hydrogen abstraction by the vacant site.
Summary of ligand effects. A wider bite angle leads to higher 1:b ratios
and to a higher proportion of bis-equatorial hydride resting states. Above a
certain value of n or the ee:ae ratio the effect on the 1:b ratio levels off. It
might be that in a region of 104 to 113 a change in mechanism occurs.
When the bite angle of the diphosphine is changed with large increments
from 78 (dppe), to 107 (trans-dppm-cyp), to 112 (BISBI or Xantphos),
the trend is clear; a higher 1:b ratio is found. The same applies to electronic
effects in symmetrically substituted diphosphines; a higher -value leads to
higher 1:b ratios. However, the latter was not true for the Xantphos series
shown in Figure 14 and Table 6 when the selectivity is considered instead of
the 1:b ratio. Wider bite angles and higher -values both lead to faster
catalysts and a relatively higher rate of isomerization. For propene the
isomerization is immaterial and as a consequence very high linearities can be
obtained using ligands displaying these properties. When isomerization
becomes sufficiently fast, even internal alkenes can be converted to linear
products, albeit at moderate rates (ligands 42 and 43).
Bis-equatorial coordination as in 2ee is not a prerequisite for high
linearity, as selective catalysts show a substantial amount of 2ae. On the
4. Phosphines as ligands 97
Figure 18. Migratory insertion in tbp-like structures for mono and bidentate dppp
Bidentate ligands having wide bite angles will lead to species similar to
3t and 4ee. Ligands with wide bite, such as BISBI and Xantphos, should
impose little deformation of the trans P-Rh-P angle of 4t. Indeed, in
RhH(CO)(PCy3)2, the P-Rh-P angle is 164 [56], which seems well within
reach of the ligands having wide bite angles such as BISBI and Xantphos
[45, 46]. This species 3wide-t associates with the alkene to give 4ee. The
overall large steric hindrance in such a complex could be sufficient
98 Chapter 4
Gleich also assumes that the migratory insertion of alkene into rhodium
hydride determines both the rate and the regioselectivity. In their
calculations they start from a common alkene complex that leads to both
linear and branched alkyl rhodium. The 1:b ratios calculated for BISBI, dppe
and PPh3 agree quite well with those obtained from the experiment.
For PPh3 the outcome seems surprising, because the bis-equatorial
complex was used for the calculation for which one would expect a high 1:b
ratio (~12) according to experiments at high phosphine concentrations. For
equatorial-apical coordination modes for both bidentates and PPh3 they also
calculate very higher 1:b ratios, but indeed the transition state energies are
100 Chapter 4
In this scheme we have assumed that insertion to give the linear alkyl
rhodium species is irreversible and rate-determining and rapid pre-equilibria
exist for CO and propene association and dissociation. From deuteration
studies we know that the conditions can be chosen in such a way that this is
valid. At high temperatures and low pressures this is no longer true. The
free-energy diagram for the kinetics showing irreversible insertion have been
drawn in Figure 22, reactions (1), (2), and (3-linear).
For the formation of branched alkyls either from I-alkenes or internal
alkenes this scheme must be modified, because now often the insertion is
reversible, dependent on conditions and concentrations. A potential scheme
showing the competition between the backward reaction (3-branched) and
the complexation of CO reaction (4) is shown in Figure 23. At low
temperatures and sufficiently high pressures the formation of 2-alkyl species
from 1-alkenes can also be irreversible. We have drawn the barriers for the
backwards reaction (-elimination) and forward reaction (CO complexation)
at about the same height to indicate the competition between the two steps.
The 2-alkyl species shown in Figure 23 can also be formed from 2-
alkenes and this species should have the same energy. We know that in most
catalytic systems the hydroformylation of internal alkenes is one or two
orders of magnitude slower. From the viewpoint of microscopic reversibility
this leads to an interesting point. It would seem therefore that in most
systems a relatively fast hydrogen exchange must take place for internal
alkenes while no or little hydroformylation occurs.
4. Phosphines as ligands 101
Figure 22. Energy diagram Type-I kinetics in which hydride migration is rate-determining for
1 -alkenes
less stable complex than 1 -butene (not shown). Clearly, energy differences
are very small and entropy and concentrations will have large influence on
the actual kinetics.
The few data known on CO dissociation from the starting
pentacoordinate rhodium hydrides show that the rate of these processes are
only two orders of magnitude higher than those of the overall
hydroformylation reaction. For this reason this barrier was drawn as it is in
Figures 22 and 23. One might speculate that coordination of the alkene is
rate-determining. A simplified scheme is shown in Figure 24.
References
20 (a) Hayashi, T.; Tanaka, M.; Ogata, I. J. Mol. Catal., 1979, 6, 1. (b) Hjortkjaer, J.;
Toromanova-Petrova, P. J. Mol. Cat. 1989, 50, 203. (c) Neibecker, D.; Rau, R. J. Mol.
Catal., 1989, 53, 219.
21 Torrent, M.; Sol, M.; Frenking G. Chem. Rev. 2000, 100, 439.
22 Van Leeuwen, P. W. N. M.; Roobeek, C. F. J. Organomet. Chem. 1983, 258, 343.
23 Diguez, M.; Claver, C.; Masdeu-Bult, A. M.; Ruiz, A.; van Leeuwen, P. W. N. M.;
Schoemaker, G. C. Organometallics, 1999, 18, 2107
24 (a) Cavalieri dOro, P.; Raimondo, L.; Pagani, G.; Montrasi, G.; Gregorio, G.; Andreetta,
A. Chim. Ind. (Milan), 1980, 62, 572. (b) Gregorio, G.; Montrasi, G.; Tanipieri, M.;
Cavalieri dOro, P.; Pagani, G.; Andreetta, A. Chim. Ind. (Milan), 1980, 62, 389.
25 (a) Strohnieier, W.; Michel, M. Z. Phys. Chem.Neue Folge, 1981, 124, 23 (in German).
(b) Hjortkjaer, J. J. Mol. Cat. 1979, 5, 377.
26 (a) Musaev, D. C.; Morokuma, K. Adv. Chem. Phys. 1996, 95, 61. (b) Musaev, D. G.;
Matsubara, T.; Mebel, A. M.; Koga, N.; Morokuma, K. Pure Appl. Chem. 1995, 67, 257.
27 (a) Nair, V. S.; Mathew, S. P.; Chaudhari, R. V. J. Mol. Catal., A 1999, 143, 99. (b)
Divekar, S. S.; Deshpande, R. M.; Chaudhari, R. V. Catal. Lett.. 1993, 21, 191. (c)
Deshpande, R. M.; Chaudhari, R. V. Ind. Eng. Chem. Res. 1988, 27, 1996.
28 Palo, D. R.; Erkey, C. Ind. Eng. Chem. Res. 2000, 38, 3786.
29 (a) Heil, B.; Marko, L. Chem. Ber. 1968, 101, 2209. (b) Feng, J.; Garland, M.
Organometallics 1999, 18, 417.
30 Gerritsen, L. A.; Klut, W.; Vreugdenhil, M. H.; Scholten, J. J. F. J. Mol. Catal. 1980, 9,
265.
31 Van der Veen, L. A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Organometallics,
2000, 19, ***.
32 lwamoto, M.; Yuguchi, S. J. Org. Chem. 1966, 31, 4290.
33 (a) Dang, T. P.; Kagan, H. B. J. Chem Soc., Chem. Commun. 1971, 481. (b) Knowles,
W. S.; Sabacky, M. J.; Vineyard, B. D.; Weinkauff, D.J. J. Am. Chem. Soc.. 1975, 97,
2567.
34 (a) Sanger, A. R. J. Mol. Catal. 1977/8, 3, 221. (b) Sanger, A. R.; Schallig, L. R. J. Mol.
Catal. 1977/8, 3, 10 1.
35 Castellanos-Pez, A.; Castilln, S.; Claver, C.; van Leeuwen P. W. N. M.; de Lange, W.
G. J. Organometallics, 17, 1998, 2543.
36 Freixa, Z.; Pereira. M. M.; Pais, A. A. C. C.; Bayn, J. C. J. Chem. Soc., Dalton Trans.
1999, 3245.
37 (a) Consiglio, G.; Botteghi, C.; Salomon, C.; Pino, P. Angew. Chem. 1973, 85, 665. (b)
Pittman Jr., C. U.; Hirao, A. J. Org. Chem. 1978, 43, 640.
38 Kawabata, Y.;Hayashi, T.; and Ogata I. J. Chem. SOC., Chem. Commun., 1979, 462.
39 Van Leeuwen. P. W. N. M.; Roobeek, C. F. Brit. Patent Appl. 33,554, 1981 (to Shell
Resarch); Chem. Abstr. 1982, 96, 6174. Van Leeuwen, P. W. N. M.; Roobeek, C. F. J.
Mol. Catal., 1985, 31, 345-53.
40 Hughes, O. R.; Unruh, J. D. J. Mol. Catal. 1981, 12, 71.
41 Hughes, 0. R.; Young, D. J. Am. Chem. Soc. 1981, 103, 6636.
42 (a) Nettekoven, U.; Kamer, P. C. J.; Widhalm, M.; van Leeuwen, P. W. N. M.
Organometallics, 2000, 19, ***. (b) The first reexamination is believed to be in error,
since it places hydride in an equatorial position: Fung, D. C. M.;James, B. R. Gazz.
Chem. Ital. 1992, 122, 329.
43 (a) Devon, T. J.; Phillips, G. W.; Puckette, T. A.; Stavinoha, J. L.; Vanderbilt, J. J. (to
Texas Eastman) U.S. Patent 4,694,109, 1987; Chem. Abstr. 1988, 108, 7890. (b) Devon,
T. J.; Phillips, G. W.; Puckette, T. A.; Stavinoha, J. L.; Vanderbilt, J. J. (to Texas
Eastman) US. Patent 5,332,846, 1994; Chem. Abstr. 1994, 121, 280,879.
4. Phosphines as ligands 105
44 Casey, C. P.; Whiteker, G. T.; Melville: M. G.; Petrovich, L. M.; Gavney, J. A., Jr.;
Powell, D. R. J. Am. Chem. Soc. 1992, 114, 5535.
45 Casey, C. P.; Whiteker, C. T. Isr. J. Chem 1990, 30, 299.
46 Dierkes, P.; van Leeuwen, P. W. N. M. J. Chem. Soc. Dalton Transnctions, 1999, 1519.
47 (a) Yamamoto, K.; Momose, S.; Funahashi. M.; Miyazawa, M. Synlett. 1990, 711. (b)
Casey, C. P.; Whiteker, G. T. J. Org. Chem. 1990, 55, 1394. (c) Casey, C. P.; Petrovich,
L. M. J. Am. Chem. Soc., 1995, 117, 6007.
48 Jongsma, T.; Challa, G.; van Leeuwen, P. W. N. M. J. Organometal. Chem. 1991, 421,
121.
49 van Rooy, A.; Orij, E. N.; Kamer, P. C. J.; van Leeuwen. P. W. N. M. Organometallics,
1995, 14, 34 and unpublished.
50 Casey, C. P.; Paulsen, E. L.; Beuttenmueller, E. W.; Proft, B. R.: Petrovich, L. M.;
Matter, B. A.; Powell, D. R. J. Am. Chem. Soc. 1997, 119, 11817.
51 Casey, C. P.; Paulsen, E. L.; Beuttenmueller, E. W.; Proft, B. R.; Matter, B. A.; Powell,
D. R. J. Am. Chem. Soc. 1999, 21, 63.
52 Van der Made, A. W.; van Leeuwen, P. W. N. M., unpublished.
53 Kranenburg, M.; van der Burgt, Y. E. M.; Kamer, P. C. J,; van Leeuwen, P. W. N. M.
Organometallics, 1995, 14, 3081.
54 van der Veen, L. A.; Keeven, P. H.; Schoemaker, C. C.; Reek, J. N. H.; Kamer, P. C. J.;
van Leeuwen, P. W. N. M.; Lutz, M.; Spek A. L. Organometallics, 2000, 19, 872.
55 Van der Veen, L. A.; Boele, M. D. K.; Bregman, F. R.; Kamer, P. C. J.; van Leeuwen, P.
W. N. M.; Goubitz, K.; Fraanje, J.; Schenk, H.; Bo, C. J. Am. Chem. Soc. 1998, 120,
11616.
56 Freeman, M. A.; Young, D. A. Inorg. Chem. 1986, 25, 1556.
57 Lazzaroni, R.; Raffaelli, A.; Settambolo, R.; Bertozzi, S.; Vitulli, G. J. Mol. Catal. 1989,
50, 1.
58 van der Veen, L. A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Orgainometallics, 1999,
18, 4765.
59 van der Veen, L. A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Angew. Chem. Int. Ed.
Engl. 1999, 38, 336.
60 (a) Koga, N.; Qian Jin, S.; Morokuma, K. J. Am. Chem. Soc. 1988, 110, 3417. (b)
Matsubara, T.; Koga, N.; Ding, Y.; Musaev, D. G.; Morokuma, K. Organometallics,
1997. 16, 1065.
61 Gleich, D.; Schmid, R.; Hemnann,W. A. Qrganometallics, 1998, 17, 4828.
62 Tatsumi, K.; Hoffmann, R.; Yamamoto, A.; Stille, J. K. Bull. Chem. Soc. Jpn. 1981, 54,
1857.
Chapter 5
Asymmetric hydroformylation
5.1 Introduction
107
P.W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 107-144.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
108 Chapter 5
Optically active diols are useful building blocks for the synthesis of chiral
diphosphite ligands. Chiral diphosphites based on commercialIy available
optically active 1,2 and I ,4-diols, I ,2:5,5-diisopropylidene-D-maiinitol, L-
,,,-tetramethyl-1,3-dioxalan-4,5-dimethanol and L-diethyl tartrate, 1-3,
were first used in the asymmetric hydroformylation of styrene [ 17].
2 (2R, 3R) R= H
ortho and para positions of the biaryl moieties to vary the properties of the
ligands [13].
Diphosphite ligands containing biaryl moieties have low energy barrier
for interconversion in atropisomers, which could lead to the formation of
several diastereomers (see Figure 5). However, the rotation around the biaryl
axis of the diphosphite ligands containing bulky substituents in the biaryl
moieties is hindered. The advantage of using bisnaphthol derivatives is that
interconversion around the bisnaphthol bond is energetically highly
unfavorable, so stable diastereomeric diphosphites were obtained in a pure
form. Because bulky substituents in the aromatic bisphenol positions
significantly affect catalyst performance, derivatives based on 2,4-
pentanediol with ortho substituents, having increasing steric bulk trimethyl-
triethyl and tert-butyl dimethylsilyl have also been used, 8 and 9 [ 18].
The chiral diphosphites 4-9 have been used in the rhodium catalyzed
asymmetric hydroformylation of styrene. The catalysts are prepared in situ
by adding the diphosphite L to [Rh(acac)(CO)2] (acac = acetylacetonate) as a
catalyst precursor. Under typical hydroformylation conditions the active
catalyst precursor [RhH(diphosphite)(CO)2] is formed (See 5.2.3). Two
important features have been found for the preparation of efficient catalyst
using these diphosphites: (1) An excess of diphosphite is used to exclude the
112 Chapter 5
dont improve the enantiomeric excess. The steric bulk in the ortho positions
is optimal with the trimethylsilyl substituent, 9a [ 18].
An interesting connection is found between the enantiomeric excess and
the structure of the diphosphite. Enantiomeric excesses are highest for the
backbones of the diphosphites based on (2R,4R)-pentane-2,4-diol, 5 and 6,
which form eight-membered rings in the catalyst. Ees are low for the
diphosphites based on (2S,5S)-hexane-2,5-diol 7 and moderate for the
backbones of the diphosphites based on (2R,3R)-butane-2,3-diol, 4 (Table
1). In most cases when the ligands are based on (R,R) diols the (S)-aldehyde
is formed predominantly. If the configuration is inverted at the chiral carbon
atom C-2 and C-5 in the (2S,5S)-hexane-diol (R)-aldehyde is predominantly
formed [13,18].
The mode of coordination depends on the length of the bridge [13, 18,
19]. For ligands 5-7 derived from 2,4-pentanediol and 2,5-hexanediol, which
1
form an eight- or nine-membered ring with the metal, the JRh-p coupling
constants are in the range 231-237 Hz. Stable hydride rhodium diphosphite
dicarbonyl complexes are formed. Typically, coupling constants of rhodium
to an equatorial phosphorus are in the range of 220-246Hz [ 13,19,25]. The
2J
P-H coupling constants are < 3Hz, which is also indicative of a cis
disposition between the phosphorus and the hydrogen atom bonded to the
rhodium (Figure 7b). The two phosphorus atoms are diastereotopic in the ee
complexes but at room temperature they are in fast exchange. Evidence for
the rapid exchange process was obtained from variable temperature NMR
experiments between 293 and 203 K. At temperatures between 293 and 253
line broadening occurred. On further cooling to 223K an ABX double
doublet appeared, indicative of two chemically inequivalent phosphorus
atoms P1 and P2 in the C1-symmetric complex.
All complexes of these diphosphite ligands showed fluxional behavior on
the NMR time scale, although the exchange parameters depended on the
mode of coordination (ea or ee). Ligand exchange in trigonal bipyramidal
HML4 complexes has been explained by so-called Berry-type rotations [26],
in one step two apical ligands exchange positions with two equatorial
ligands. However, a Berry mechanism is unlikely to be operative in RhH(L-
L)(CO)2 complexes containing bidentate diphosphite, since phosphorus
exchange requires two successive Berry-type interconversions via a high-
energy intermediate which has a hydride ligand equatorially coordinated to
rhodium. There are no known examples of trigonal bipyramidal HML4
complexes, which have an equatorial coordinated hydride ligand. X-ray and
infrared studies on such complexes always show the hydride in an apical
position. It is also unlikely that phosphorus-rhodium-phosphorus bite angles
can vary freely between 90 and 180 without considerable energy strain.
The low-energy turn-stile rearrangement postulated by Meakin et al. [22,
23] for monophosphites has been proposed to explain the fluxional behavior
of these complexes containing flexible diphosphite. It is believed that the
phosphorus atoms can easily exchange without changing the phosphorus-
rhodium-phosphorus bite angle appreciably.
Figure 8 shows a simultaneous bending motion of the hydride and the
carbon monoxide ligands in the hydride rhodium phosphite dicarbonyl
complexes containing equatorial-apical coordinating phosphite, (A) and the
5. Asymmetric hydroformylation 115
This seems logical since this process involves the movement of two
carbonyl ligands, which should be easier than the movement of the two
phosphite moieties that is required for equatorial-equatorial exchange.
Furthermore, the energy barrier for phosphorus-phosphorus exchange is
lower for the nine-membered bis-equatorial chelate rings than for the eight-
membered ones, as the former are more flexible owing to the reduced steric
congestion.
The fact that the diphosphite ligands behave differently in the formation
of the RhH(diphosphite)(CO)2 complexes demonstrates that both the
absolute configuration of the 2,4-pentanediol ligand backbone and the chiral
bisnaphthol substituents determine the stability and catalytic of the rhodium
complexes. The ligands with configuration (S,2S,4S,S) 11 and (R,2R,4R,R)
12, lead to undefined mixtures of complexes and ligand decomposition.
However the ligands (S,2R,4R,S) 8, 10, and (R,2R,4R,S) 13 form well
defined, stable complexes. When there was an excess of the statistical
mixture of diastereomers 9, only the complexes derived from 10 and 13
formed. When both the pentane backbone and the bisnaphthol substituents
have all S (11) or all R (12) configurations the ee is low. In situ NMR studies
show that the [RhH(CO)2(L-L)] complexes of these ligands cannot be
synthesized selectively.
The low enantioselectivity and relatively high reaction rates observed
have been attributed to the coexistence of highly active rhodium species in
which the ligands coordinate as monodentates. In contrast, the ligands with
different configurations at the pentane bridge backbone and such bisnaphthol
substituents, as ligands 10 and 13, both form stable rhodium complexes, but
only diastereomer 10 gives a high ee (Table 2). These results have been
118 Chapter 5
Figure 11. Diphosphite ligands containing chiral chelate backbones and chiral
dioxaphosphorinane moieties
Figure 12. Chiral diphosphites forming nine membered ring coordinated to the metal
Systematically varying the chirality at both the chelate backbone and the
terminal groups affected the enantioselectivity of the catalyst. The axial
chirality on the bridge determines the product configuration with a
cooperative effect from the terminal groups. When both the bisnaphthol
backbone and the terminal dioxaphosphepine units have the same chiral
denominator, ee's are lower than when they have different configurations, as
can be seen from the systems discussed above [18, 28].
Chiral diphosphite ligands containing a spiro backbone in the bridge and
forming a complex with an eight-membered ring have also been used as
catalysts in the asymmetric hydroformylation of styrene and derivatives
(Figure 13) [30]. At 10 bar of CO/H2 and 40 C regioselectivity to 2-
phenylpropanal is 97% and enantioselectivity 69% using the (1 S,5S,6R)-
(cis,trans)-diol phosphite shown.
5. Asymmetric hydroformylation 121
5.3.1 Introduction
The results obtained for ligands 26 and 28, which contain only one fixed
stereocenter, are interesting and very infomative about the system. Ligand
26-(R,--), in which only the binaphthyl bridge has a predetermined absolute
configuration R, leads to an ee of 83% (R-aldehyde), which is quite close to
the value of 94% for (R,S)-BINAPHOS. This suggests that in the formation
of the complex the binaphthyl bridge controls the conformation of the
bisphenol moiety so that the final conformation is (R,S). Likewise, 28 -(--,R)
gives 69% ee of the S-aldehyde, which suggests that the complex formed
assumes an (S,R) conformation, since now the R-bisnaphthol induces the
formation of an S-diphenyl bridge. The control by the binaphthyl bridge in
26 is somewhat more efficient than that of the bisnaphthol moiety in 28.
In the free ligand 28 we saw a slight preference for the S,R diastereomer,
but in the complexes there is a strong preference for the S,R or R,S
conformation. The NMR spectra of the complexes of the two non-rigid
ligands 26 and 28 led to this conclusion but other data also support this
observation. In the free ligands with R,S configurations the through-space
coupling constants between the two phosphorus nuclei are always larger than
those in R,R ligands. This shows that in R,S ligands the ligand has a natural
tendency to bring the donor atoms into close proximity. Secondly, in
competition experiments between R,R and R,S ligands the latter give the
most stable complexes and are prevalently formed in the mixtures. In
diphosphites, the bisphenol configurations were also observed to be
controlled by the backbone and the most stable configurations lead to the
highest ee's in hydroformylation catalysis [4, 14, 18, 38].
and that more advanced studies are needed. Gleich et al used a semi-
quantitative theoretical model that combined quantum mechanics and
molecular mechanics. Their findings matched previously reported
experimental data; the aldehyde found experimentally had the lowest barrier
of formation and the (R,R)-BINAPHOS complex should give much lower
ees than the (R,S)-BINAPHOS complex [14, 15].
5.4.1 Introduction
[Rh]
CO2Me diphosphine
NHCOMe CO/H2
when the temperature decreases. The same behavior had previously been
observed in the hydroformylation of styrene with the unmodified rhodium
catalyst (chapter 2) [45, 56].
31
The P NMR spectrum reveals the presence of two species in
equilibrium, the [RhM(BDPP)(CO)2] hydride complex and the dimer
[Rh(BDPP)(-CO)(CO)]2. The mononuclear [RhH(BDPP)(CO)2] complex
1
shows only one doublet at 29.8 ppm ( JRh-P)=112 Hz and the 1H NMR
1 2
spectrum revealed a double triplet in the hydride region JRh-H= 11, JH-P= 57
Hz, indicative of a coupling of the hydride with the rhodium nucleus and two
degenerated phosphorus nuclei. Small cis phosphorus-hydride coupling
constants are reported for complexes with ee coordinating diphosphine
ligands [62] and relatively large phosphorus-hydride coupling constants are
characteristic of complexes containing trans P-Rh-H arrangements [22, 24].
These studies revealed that the diphosphine BBPP is equatorially-axially
coordinated, in accord with the calculated bite angle of 91 [54], and that a
fast exchange of the two phosphine groups occurs. The turn-stile
mechanism is the same as that described for diphosphites (5.2) [21].
The monomer-dimer equilibrium. Since the work of Wilkinson it has been
known that complexes of the type [RhH(CO)2(L-L)] and [Rh(L-L)(-
CO)(CO)]2 are in equilibrium with one another. The equilibrium constant
depends on the ligand used and the concentrations of rhodium and hydrogen.
Many authors have studied on the formation of these dinuclear species, but
no quantitative data had been reported before the study involving BDPP [50,
62]. In the Rh-BDPP system, both species can be observed and the
equilibrium constant is measured by 31P NMR. The data indicate that at
room temperature, relatively high catalyst concentrations, and partial H2
pressures of below 5 bar, rhodium will be present predominantly as the
dimeric complex. More monomer is formed when the H2 pressure is raised,
and hence, the rate of hydroformylation increases with the pressure,
regardless the mechanism of the rhodium hydride catalyst. The equilibrium,
therefore, is of great importance to the overall kinetics observed during
hydro formy la tion.
In addition the 31P NMR data show the presence of other species
containing more than two phosphorus atoms, in which the diphosphine acts
as monodentate. Hughes et al. [63] have described related complexes (see
chapter 4).
138 Chapter 5
5.5.1 Regioselectivity
alkyl group [65]. These are the electronic considerations. For steric reasons
one would expect a preference for the linear alkyl.
Ethyl acrylate, the hydroformylation of which has been studied many
times, is an interesting case. Tanaka et al. [66] found that the ratio of 1:b
could be varied from 100:0 (80 C and 1 bar) to 1:99 (40 C and 30 bar) by
using a rhodium catalyst and triphenyl phosphite as the ligand. Electronic
arguments lead to the same result as for trifluoropropene because the frontier
orbitals on ethyl acrylate are the same. Thus, high pressures lead to the
expected result.
Migration of the stabilized ester enolate, however, is expected to be much
1
slower than that of the linear alkyl, irrespective of whether it is 3 or
bonded. At low pressure the catalyst rests in the enolate state of the cycle
for the branched intermediate and the formation of the enolate can be
reversed. The product-forming route is now via the linear alkyl, which
undergoes a relatively rapid migration to form the acyl species (see Figure
28). The role of CO may be either to form a species containing more CO
molecules or to trap the acyl species after migration. In the absence of more
data this remains speculation.
Steric control in the figure is an oversimplification, because this route
is also taken when migration of the ester enolate is too slow, irrespective of
steric control as outlined above.
In a broad range of conditions the rate laws for the rate of formation of
the linear and the branched product are different, which explains the
enormous influence that these have on the product distribution. At low CO
pressures the enolate-rhodium species may be the resting state of the
catalyst. Species of this type often undergo hydrogenation instead of
hydro formy lation.
For styrene the reactivity pattern is less obvious. The assumption that
there is an early transition state does not lead to a simple conclusion, because
the coefficients of both the LUMO and the HOMO of styrene are higher on
the terminal carbon atom. If the interaction of the HOMO on the Rh-H
140 Chapter 5
located on the hydrogen with the LUMO of styrene were to dominate the
branched alkyl would be formed. Clearly the resonance stabilized branched
alkyl is more stable than the linear phenylethyl species. Thus, electronically
the branched species is preferred as most authors have argued (see also
Chapter 2). In addition the branched alkyl species may be stabilized by 3
coordination, but the homo-allylic linear species is also stabilized by alkene-
rhodium interaction.
Insertion of styrene does not always lead to branched alkyl species.
Insertion of styrene into a cationic palladium hydride species may give
purely linear alkyl species, but perhaps not for steric reasons [67]. An early
transition state for this process may involve the interaction of the LUMO for
PdH+, which has the highest coefficient on palladium, with the HOMO of
styrene, which has the highest coefficient on the terminal carbon atom.
References
1 (a) Sheldon, R. A. Chemistry and Industry, 1990, 212. (b) Botteghi, C.; Paganelli, S;.
Schionato, A; Marchetti, M. Chirality, 1991, 3, 355.
2 Consiglio, G. Catalytic Asymmetric Syntheses, I. Ojima editor VCH Publishers, New
York, 1993.
3 (a) Agbossou, F.; Catpentier, J. F.; Mortreux, A. Chem Rev. 1995, 95, 2485. (b)
Gladiali, S.; Bayn, J .C.; Claver. C. Tetrahedron Asymmetry, 1995, 7, 1453.
4 Nozaki, K. Comprehensive Asymmetric Catalysis ; Jacobson E. N.; Pfaltz, A.,
Yamamoto, H. Ed.; Springer Heidelberg, 1999; Vol.1 , pp 381-409.
5 Stille, J. K.; Su H.; Brechot, P.; Paninello, G.; Hegedus, L. S. Organometallics, 1991,
10, 1183.
6 a) Consiglio, G., Nefkens, S. C. A, Tetrahedron Asymmetry, 1990, I 417. b) Consiglio,
G.; Nefkens, S. C. A, Borer, A. Organometallics, 1991, 10, 2046.
7 Cserpi-Szcs, S. Bakos, J. Chem. Commun. 1997, 635.
142 Chapter 5
36 Pmies, O.; Net, G.; Ruiz, A.; Claver, C. Tetrahedron Asymmetry, 2000, 11, 1097.
37 Kadyrov R.; Heller. D.; Selke, R. Tetrahedron Asymmetry. 1998, 9, 329.
38 Sakai, N.; Mano, S.; Nozaki, K.; Takaya, H.; J. Am. Chem. SOC.; 1993, 115, 7033.
39 Sakai, N.: Nozaki, K.; Takaya, H. J. Chem Soc.Chem. Comm. 1994, 395.
40 Higashijima, T.; Sakai, N.; Nozaki, K.; Takaya, H. Tetruhedron Letters, 1994, 35, 2023.
41 a) Horiuchi, T.; Ohta, T.; Nozaki, K.; Takaya, H. Chem Comm. 1996, 155. (b) Horiuchi
T.; Ohta, T.; Shirakawa, E.; Nozaki, K.; Takaya, H. J. Org. Chem.; 1997, 62, 4285.
42 (a)Naieli ,S.; Carpentier, J-F.; Agbossou, F.; Mortreux, A.; Nowogrocki ,G.; Wignacout
J. P. Organometallics, 1995, 14, 401 (b) Kless, A.; Holz, J.; Heller, D.; Kadyrov, R.;
Selke, R.; Fischer, C.; Bmer, A.; Tetrahedron Asymmetry, l996, 7, 33. (c) Basoli, C.;
Botteghi, C.; Cabras, M. A.; Chelucci, G.; Marchetti, M. J. Organomet. Chem. 1995,
488, C20. (d).; Chelucci, G.; Cabras, M. A.; Botteghi, C., Basoli, C., Marchetti, M.
Tetrahedron Aymmetry, 1996, 7, 33.
43 Vaska, L. J. J. Am. Chem.Soc. 1996, 88, 4100.
44 Horiuchi, T.; Ohta, T.; Shirakawa, E.; Nozaki. K., Takaya, H. Organometallics, 1997,
16, 2981.
45 (a) Lazzaroni, R.; Raffaelli, A.; Settambolo, R.; Bertozzi, S.; Vitulli ,G.; J. Mol. Cat.
1989, 50, 1(b) ) Lazzaroni, R; Ucello Barretta, G.; Bennetti, M.; Organometallics, 1989,
8, 2323 (c) Raffaelli A.; Pucci, S.; Settambolo, R.; Ucello Barretta, G.; Lazzaroni, R.;
Organometallics, 1991, 10, 3892, (d) Lazzaroni, R.; Settambolo, R.; Ucello Barretta,
G.; Organometallics, 1995, 14, 4644 (e) Lazzaroni, R.; Ucello Barretta G., Scamuzzi S.;
Settambolo, R.; Caiazzo, A.; Organometallics, 1996, 15, 4657.
46 Consiglio, G.; Pino, P.; Top Curr. Chem., 1982, 105, 77.
47 Nozaki, K.; Itoi, Y.; Sibahara, F.; Shirakawa, E.; Ohta, T.; Takaya, H.; Hiyama, T. J.
Am. Chem. Soc. 1998, 120, 4051.
48 Francio, G.; Leitner, W. Chem. Commun. 1999, 1663.
49 Deerenberg, S.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. Organometallics, 2000, 19,
2065.
50 Pottier, Y.; Mortreux, A.; Petit, F.; J. Organometal. Chem. 1989, 370, 333.
51 Salomon, C.; Consiglio, G.; Botteghi, C.; Pino P. Chimia, 1973, 27, 215.
52 Gladiali, S.; Pinna, L. Tetrahedron: Asymmetry, 1991, 2, 263.
53 Masdeu-Bult, A. M.; Orejon. A.; Castellanos, A.; Castillon, S.; Claver, C. Tetrahedron
Asymmetry, 1996, 7, 1829.
54 Dierkes, P.; van Leeuwen, P. W. N. M. J. Chem. SOC. Dalton Trans. 1999, 1519.
55 Castellanos-Pez, A.; Castillon, S.; Claver, C.; van Leeuwen, P. W. N. M.; de Lange,
W. G. J. Organometallics, 1998, 17, 2523.
56 Diguez, M.; Pereira, M. M.; Masdeu-Bult, A. M.; Claver, C.; Bayn, J. C. J. Mol.
Cat. 1999, 143, I11.
57 Freixa, Z.; Pereira, M. M.; Pais, A. A. C. C.; Bayon, J. C. J. Chem. Soc. Dalton Trans.
1999, 3245.
58 Miquel-Serrano, M. D., Masdeu-Bult, A. M.; Claver, C.; Sinou, D. J. Mol. Cat. 1999,
143, 49.
59 Eckl, R. W.: Primereir, T.; Herrmann, W. A. J. Organometal. Chem. 1997, 532, 243.
60 Rampf, F. A.; Splieger, M.; Herrmann, W. A. J. Organometal. Chem. 1999, 582, 204
61 Cornils, B.;. Herrmann, W. A (eds) Applied Homogeneous Catalysis with
Organometallic Complexes, VCH, Weinheim, 1996. Chap 4.1.
62 Casey, C. P.; Whiteker, G. T.; Melville, M. G.; Petrovich, L. M.; Gavney, J. A.;Jr.;
Powell, D. R. J. Am. Chem. Soc. 1992, 114, 5535.
144 Chapter 5
Sergio Castilln and Elena Fernndez
6.1 Introduction
The discovery of new catalytic systems that enabled the regio- and
steroselectivity of hydroformylation to be controlled has meant that this
reaction has emerged as a flexible and important tool in organic synthesis.
Some interesting reviews on asymmetric hydroformylation (see chapter 5 in
this book) [ 1a-h], hydroformylation tandem reactions, [2,3] other general
aspects of the reaction and how it can be applied in the synthesis of-fine
chemicals [4a-g] give an account of the impressive progress of
hydroformylation over the last decade. Rhodium-catalyzed hydroformylation
takes place under mild conditions and it is compatible with the most
common functional groups present in an olefinic substrate [4a], and it is,
therefore, a synthetically useful tool for the preparation of organic
compounds [2].
This chapter discusses some general aspects of the hydroformylation of
alkenes in organic synthesis. It focuses mainly on regio- and stereoselective
processes, and analyzes the influence of the substrate and the catalysts.
Practical methods which provide high yields and selectivities, and short-cuts
compared to classical organic routes will be described. Particular attention
will be paid to recent advances that have helped to enlarge the synthetic
application of this reaction. Section 6.7 deals with the hydroformylation of
alkynes, and such key aspects as hydroformylation in water-gas shift
conditions and silylformylation, particularly efficient catalytic systems and
the application of hydroformylation in organic synthesis.
145
P.W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 145-187.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
146 Chapter 6
Formyl group
introduction
Reactivity
5 BiPHEPHOS 6 BlSBl 7
8 9 10 11
a c
[Rh4(CO)12, 217 bar, CO/H2= 1,60C b [RhHCO(PPh3]3 100 bar CO/H2= 1, 100C Rh6(CO)16, 120
d e
bar, CO/H2= 1,70C Rh 4(CO)12, 110 bar, CO/H2=1,8OC Rh4(CO)12, 100 bar,CO/H2 (1:1), 20C
product, (Figure 9). Dihydropyrans [27] behave similarly but are much less
reactive.
As has been seen above, ligands have a considerable influence on the
control of regioselectivity because they stabilize a preferred M-alkyl
intermediate or accelerate or avoid the -elimination process. An example of
this is the hydroformylation of 2,3- and 2,5-dihydrofuran [27]. Thus, the
t
hydroformylation of both dihydrofurans 22 and 23, using P(O- o- BuC6H4)3
gives practically the same ratio of products 24:25 (Figure 9) (entries 2 and 4
Table 2).
The evolution of the reaction shows that 23 is isomerized into 22 before
the hydroformylation reaction starts, because the M--alkyl intermediate is
b-eliminated faster than CO is inserted. However, if the reaction is started
from 23 and the ligands have small cone angles, such as P(OMe)3,
practically the only aldehyde detected is the 3-formyl derivative (entry 3). If
it is started from 22, the 3-formyl derivative is the main aldehyde detected
(entry 1). This suggests that under these conditions there is no -elimination
process or that it is very slow.
These results show that the regioselectivity of the process can be
controlled if the ligand and reaction conditions are selected. Thus, PPh3 can
be used to quantitatively convert 23 into 25. The results are similar to when
P(OMe)3 is used in a high P/Rh ratio, at moderate temperatures and high CO
pressures but the catalyst is more active using PPh3. When P(O-o-tBuC6H4)3
is used at high temperatures and low CO pressure, however, the main
aldehyde 24 is obtained.
However, 3,4-di hydro-2H-pyran and 5,6-dihydro-2H-pyran required
more drastic conditions to be hydroformylated and when only P(O-o-
t
BuC6H4)3 was used as the ligand conversions were as high as 80% and the
selectivity in 2-/3-formyl derivatives was 68/32 [27].
a
Table 2. Hydroformylation of dihydrofuran 24 and 25 using different phosphorus ligands.
The hydroformylation of vinyl arenes 26 has been widely studied and the
branched aldehyde 28 is the main product of the reaction (Figure 10). When
rhodium catalysts are used and pressures are high, regioselectivities are
usually between 80 and 98%. The major regioisomer is the inverse of the
one for alkyl monosubstituted olefins. The formation of -complex 27
3
explains the regioselectivity and this has already been discussed in the
literature (Figure 10) [24, 28, 29]. The regioselectivity can vary according to
the substituents in the ring [24b, 30].
36 39
a C + 6 -
[Rh6(CO)16], CO/H 2 =1, 110 bar, 80C. b [Rh], 200 bar CO/H2=1, 100C, [Rh ( C6H5B Ph3)(cod)]
/dppb, CO/H2= 1,40br, 75C.
branched aldehyde is obtained (Figure 15, a). This is the case when R is a
phenyl group or a heteroatom. The regioselectivity in 1,1-disubstituted
alkenes gives the linear aldehyde but a stererocenter is generated in the
1
branched aldehyde when R is different from R2 (Figure 15, b). Two possible
regioisomers can be achieved in the hydroformylation of 1,2-disubstituted
alkenes so the formation of a stereocenter is only interesting in symmetric
alkenes or when exist elements controlling the regioselectivity (Figure 15,
c). And in trisubstituted alkenes, two stereocenters are generated as can be
seen in Figure 15, d.
-3
i) [Rh]= 6.10 M, CO/H2=1, 300 bar, 70C. ii) [Co]= 51.10-2 M, CO/H2=1, 200 bar, 120 C.
- -
iii) [Rh]= 5.10 3 M, PPh3/Rh =100, CO/H2=1, 60 bar, 100 C. iv) [M]= 5.10 3 M, CO/H 2=1,
60 bar, 100 C.
i) [Rh(acac)(CO)2] / PCy3 (from 4), PPh3 (from 5), ratio Rh/P= 1.4, H2/CO=1, 60 bar, 120C,THF
i) [Rh(OAc)2]2, alkene/Rh= 50:1, PhH, CO/H2= 1, 37bar, 100C, benzene, 44h. ii) LiA]H4, Et2O.
i) [Rh(OAc)2]2, alkene/Rh= 50:1, PhH. CO/H2= 1, 37 bar, 50C. benzene, 5-22h. ii) LiAH4 Et2O.
(o-DPPB)= ortho-diphenylphosphinobenzoate
i) 0.7 mol/% [Rh(CO)2(acac)]/4 P(OPh)3, CO/H2= 1, 20 bar, 50C, toluene, 72 h.
i) 1 mol/% [Rh(-OMe)(cod)]2/ 10 t
PPh3/PPTS, CO/H 2= 1. 50 bar, 60C 2,2-dimethoxypropane, 24 h.
ii) Identical to i but using P(O-o - BUC6H3)3.
As has been shown above (see section 6.4), the hydroformylation of 3,4,6-tri-
O-acetylglucal gives considerable amounts of the elimination product 121,
which must be obtained by eliminating of acetic acid from 120 (see also Figure
19) [49]. The hydroformylation of allylic esters to give ,-unsaturated
aldehydes by hydroformylation and acid elimination is a well documented
process [4e]. In situ acetal formation partially avoids this process. Thus, using
[Rh(-OMe)(cod)]2/P(O-o-tBuC6H4)3/PPTS as hydroformylation-acetalyzation
catalysts, 3,4,6-tri- O -acetyl-D-glucal 119 is converted into the dimethoxy acetal
168 Chapter 6
120 (Figure 34). Only a small amount of the elimination product 121 is formed
178].
Unexpectedly, the 3,4,6-tri-O-benzyl-D-glucal gave only the methyl -
glycoside 121 under the hydroformylation-acetalization conditions. In fact, there
are two electrophilic reagents in competition for the nucleophilic alkene, the
rhodium complex and the proton. When the alkene is deactivated (R=Ac) the
coordination of rhodium is preferred and the hydroformylation-acetalization
takes place. But when it is not deactivated (R=Bn) the acidic proton reacts faster
than rhodium and methanol is added.
t
i) 1 mol/% [Rh(-OMe)(cod)]2/ 10P(O-o- BuC6H4)3 / PPTS, CO/H2=1, 50 bar, 100C,
2,2-dimethoxypropane, 48h.
When an amino group is present in the substrate (123, 125, 127), various
processes can take place consecutively under hydroformylation conditions to
afford, cyclic N,O- (124, 126) [79, 80] or N,N-acetals (128) [81] (Figure 34).
Imines and enamines can also be formed. The formation of acetals, imines or
enamines depends on alcohols being present in the substrate or the solvent,
additional amino groups being present in the substrate, the substitution of the
amino group, and the reaction conditions. Moreover the presence of coordinative
atoms such as nitrogen allows a chelate to be formed which control the
regioselectivity of the process.
Because of the chelate control of the process, allylamides such as 129 react
under hydroformylation conditions to give mainly the branched aldehyde 131,
together with cyclic derivatives 132 and 133 (Table 5) [82, 83]. Products 132
and 133 are formed from the linear aldehyde through a sequence of reactions
involving cyclization to give the enamide 134, followed by hydrogenation or
hydroformylation, respectively [84].
6. Hydroformylation in organic synthesis 169
The bulkiness of the substituent on the amide nitrogen virtually does not
have any effect on the regioselectivity, but it exerts a marked effect on the
cyclic/acyclic ratio of the products. The effect of the bulky N-substituents is
particularly pronounced in the trityl derivative (R=Tr, n=l), since no formation
of pyridone 136 was observed. A mixture of open chain linear aldehyde and
pyrrolin-2-one 137 was obtained.
R=Hexyl
t
[Rh(-S-Bu )2(CO)2(PPh3)2]/PPh3,8bar,80C 42 13%
R2=Ph. The reaction provides better yields with cyclic secondary amines than
with acyclic ones [ 107].
Silylenol ethers such as 184 also undergo the hydroformylation-aldol reaction
to give the silylated aldol adducts 185 in good yields through a sequence of
reactions involving the hydroformylation of the alkene and the intramolecular
Mukaiyama type aldol reaction. [108]. Best results were achieved using the
trimethylsilyl group.
Different reaction products are obtained if the hydroformylation of the
unsturated ketone 183 is carried out in the presence of amines. With secondary
amines the hydroaminomethylation of the double bond is observed, leaving the
carbonyl group unaffected.
In the presence of benzylamine, ketone 183 is converted into the
aminoketone 186 by alkene hydroformylation, imine formation and aldol
reaction. Under more drastic hydroformylation conditions the reaction of 183
with benzylamine leads to the amine 187, which results from a mechanism
similar to those above including reductive amination of the ketone moiety.
6. Hydroformylation in organic synthesis 177
191 192
Terminal acetylenes give complex mixtures. The reaction also takes place
in the acetylene that shows that for these substrates the triple bond is more
reactive than the double bond.
The hyfroformylation of propargylamines 202 in the presence of the
classical catalytic system [Rh(OAc)2]2/PPh3 gives 2,4-disubstituted pyrroles
204 in excellent yields, through consecutive hydroformylation, cyclization
and double bond isomerization [ 116, 117].
180 Chapter 6
The reaction gives compounds 206, 209, and 210, and proceeds through
the ,-unsaturated lactones which undergoes an in situ hydrogenation to
give the saturated lactones.
6. Hydroformylation in organic synthesis 181
References
40 Kollr, L.; Consiglio, G.; Pino, P. J. Organomet. Chem. 1990, 386, 389.
41 Stefani, A.; Consiglio, G.; Botteghi, C.; Pino, P. J. Am. Chem. Soc. 1973, 95, 6504.
42 Himmele, W.; Siegel, H. Tetrahedron Lett. 1976 ,907.
43 Siegel, H.; Himmele, W. Angew. Chem. Int. Ed. Engl. 1980, 19, 178.
44 Azzaroni, F.; Biscarini, P.; Bordoni, S.; Longoni, G.; Venturini, E. J. Organomet. Chem.
1996, 508, 5958.
45 Leighton, J. L.; ONeil, D. N. J. Am. Chem. Soc. 1997, 119, 11118.
46 Sarraf, S. T.; Leighton, J. L. Tetrahedron Lett. 1998, 39, 6423.
47 Rosenthal, A. Adv. Carbohydr. Chem. 1969, 24, 59.
48 Rosenthal, A.; Abson, D.; Field, T. D.; Koch, H. J.; Mitchel, R. E. Can. J. Chem. 1967,
45, 1525.
49 Ferndez, E.; Ruiz, A.; Claver, C.; Castilln, S.; Polo, A.; Piniella, J. F.; Alvarez-
Larena, A. Organometallics 1998, 17, 2857.
50 Takahashi, T.; Ebata, S; Yamada, H. Synlett. 1998, 38 1.
51 Breit, B.; Zahn, S. K. Tetrahedron Lett. 1998, 39, 1901.
52 a) Kraft, E. M.; Wilson, L. J. Organometallics 1988, 7, 2528. b) Kraft, E. M.
Tetrahedron Lett. 1989, 301, 539.
53 Jackson, W. R.; Perlmutter, P.; Tasdelen, E. E. J. Chem. Soc., Chem. Commun. 1990,
763
54 Jackson, W. R.; Perlmutter, P.; Tasdelen, E. E. Tetrahedron Lett. 1990, 31, 2461.
55 Jackson, W. R.; Perlmutter, P.; Suh, G. H. J. Chem. Soc., Chem. Commun. 1987, 724.
56 Burke, S. D.; Cobb, J. E. Tetrahedron Lett. 1986, 27,4237.
57 Burke, S. D.; Cobb, J. E.; Takeuchi, K. J. Org. Chem. 1990, 55,2138.
58 Breit, B. Angew. Chem. Int. Ed. Engl. 1996, 35, 2835.
59 Breit, B. Liehigs Ann. Recueil 1997, 1841.
60 Breit, B.; Dauber, M.; Harms, K. Chem. Eur. J. 1999, 5, 2819.
61 Breit, B. Chem. Commun. 1997, 591.
62 Breit, B. Eur. J. Org. Chem. 1998, 1123.
63 Pittman, C. U. Jr.; Honnick, W. D. J. Org. Chem 1980, 45, 2132.
64 Wuts, P. G. M.; Obrzut, M. L.; Thompson, P. A. Tetrahedron Lett. 1984, 25, 4051.
65 Smith, W. E.; Chambers, G. R.; Lindberg, R. G.; Cawse, J. N.; Dennis, A. J.; Harrison,
G. E.; Bryant, D. R. In Catalysis of Organic Reactions; (Ed. Augustine, R. L.) Dekker,
New York, 1985, p.151.
66 Nozaki, K.; Li. W.; Horiuchi, T.; Takaya, H. Tetrahedron Lett. 1997, 38, 4611.
67 Trzeciak, A. M.; Wolszcak, E.; Ziolkwsky, J. New J. Chem. 1996, 20, 365.
68 Anastasiou, D.; Jackson, W. R. Aust. J. Chem. 1992, 45, 21.
69 Anastasiou, D.; Jackson, W. R.; McCubbin, Q. J.; Tmacek, A. E. Aust. J. Chem. 1993,
46, 1623.
70 Baker, R.; Herbert, R. H. Nat. Prod. Rep. 1984, 43, 3309.
71 Jaramillo, C.; Knapp, S. Synthesis 1994, 1.
72 Sirol, S.; Kalck, Ph. New. J. Chem. 1997, 21, 1129.
73 Kitsos-Rzychon, B.; Eilbracht, P. Tetrahedron 1998, 54, 10721.
74 Parriello, G.; Stille, J. K. J. Am. Chem. Soc. 1987, 109, 7122.
75 Fernndez, E.; Castilln, S. Tetrahedron Lett. 1994, 35, 2361.
76 Balu, J.; Bayn, J. C. J. Mol. Catal. 1999, 137, 193.
77 Soulantica, K.; Sirol, S.; Koinis, S.; Pneumatikakis, G.; Kalck, Ph. J. Organomet. Chem.
1995, 498, C 10.
78 Fernndez, E.; Polo, A.; Ruiz, A.; Claver, C.; Castilln, S. Chem. Commun. 1998, 1803.
186 Chapter 6
79 Ojima, I.; Tzamarioudaki, M.; Eguchi, M. J. Org. Chem. 1995, 60, 7078.
80 Campi, E. N.; Jackson, W. R.; McCubbin, Q. J.; Tmacek, A. E. Aust. J. Chem. 1996, 49,
219.
81 Campi, E. N.; Habsuda, J.; Jackson, W. R.; Jonasson, C. A. M.; McCubbin, Q. Aust. J.
Chem. 1995, 48, 2023.
82 Ojima, I.; Zhang, Z. J. Org. Chem. 1988, 53, 4425.
83 Ojima, I.; Zhang, Z. J. Organomet. Chem. 1991, 417, 253.
84 Ojima, 1.; Korda, A. Tetrahedron Lett. 1989, 30, 6283.
85 Ojima, I.; Korda, A.; Shay, W. R. J. Org. Chem. 1991, 56, 2024.
86 Ojima, 1.; Vidal, E. S. J. Org. Chem. 1998, 63, 7999.
87 Ojima, I.; Iula, D. M.; Tzamarioudaki, M. Tetrahedron Lett. 1998. 39, 4599.
88 Anastasiou, D.; Jackson, W. R. J. Chem. Soc., Chem. Commun. 1990, 1205.
89 Anastasiou, D.; Chaouk, H.; Jackson, W. R. Tetrahedron Left. 1991, 32, 2499.
90 Anastasiou, D.; Campi, E. M.; Chaouk, H.; Jackson, W. R. Tetrahedron 1992, 48, 7467.
91 Bergmann, D. J.; Campi, E. M.; Jackson, W. R.; Cubin, Q. J.; Patti, A. F. Tetrahedron Lett.
1997, 38, 4315.
92 Bergmann, D. J.; Campi, E. M.; Jackson, W. R.; Patti, .A. F. Sylik, D. Tetrahedron Lett.
1999, 40 5597.
93 Bergmann, D. J.; Campi, E. M.; Roy Jackson, W.; Patti, A. F. Chem Commun. 1999. 1279.
94 a) Reppe, W. Experientia 1949, 5, 93. b) Reppe, W.; Vetter, H. Liebigs Ann. Chem. 1953,
582, 133.
95 Rische, T.; Eilbracht, P. Synthesis 1997, 1331.
96 Baig, T.; Kalc, Ph. J. Chem. Soc., Chem. Commun. 1992, 1373.
97 Baig, T.; Molinier, J.; Kalc, Ph. J. Organomet. Chem. 1993, 455, 219.
98 Eilbracht, P.; Kranemann, C. L.; Brfacker, L. Eur. J. Org. Chem 1999, 1907.
99 Rische, T.; Eilbracht, P. Tetrahedron 1999, 55, 3917.
100 Rische, T.; Barfacker, L.; Eilbracht, P. Eur. J. Org. Chem. 1999, 653.
101 Rische, T.; Eilbracht, P. Tetrahedron 1998, 54, 8441.
102 Barfacker, L.; Rische, T.; Eilbracht, P. Tetrahedron 1999, 55, 7177.
103 Kranemann, C. L.; Kitsos-Rzychon, B.; Eilbracht, P. Tetrahedron 1999, 55, 4721.
104 Rische, T.; Eilbracht, P. Tetrahedron 1999, 55, 1915.
105 Rische, T.; Muller, K-S.; Eilbracht. P. Tetrahedron 1999, 55, 9801.
106 Breit, B. Tetrahedron Lett. 1998, 39, 5163.
107 Barfacker, L.; El Tom, D.; Eilbracht, P. Tetrahedron Lett. 1999, 40, 4031.
108 Hollman, C.; Eilbracht, P. Tetrahedron Lett. 1999, 40, 4313.
109 Breit, B.; Zahn, S. K. Angew. Chem. Int. Ed. 1999, 38, 969.
110 Pino, P.; Barca, G. in Organic Synthesis via Metal Carbonyls, vol 2, (Eds. Wender, I.;
Pino, P,), Wiley-Interscience, New York, 1977, p. 419.
111 Tkatehenko, I. in Comprehensive Organometallic Chemistry. vol. 8, (Eds. Wilkinson,
Sir G. Stone, F. G. A.; Abel, E.), Pergamon Press, Oxford, 1982.
112 Botteghi, C.; Salomon, Ch. Tetrahedron Lett. 1974, 4285.
113 Johnson, J. R.; Cuny, G. D.; Buschwald, S. L. Angew. Chem. Int. Ed. Engl. 1995, 34,
1760.
114 Doyama, K.; Joh, T.; Takahashi, S. Tetrahedron Lett. 1986, 27, 4497.
115 Doyama, K.; Joh, T.; Shiohara, T.; Takahashi, S. Bull Chem. Soc. Jpn. 1988, 61, 4353.
116 Campi, E. M.; Jackson, W. R. Aust. J. Chem. 1989, 42, 471.
117 Campi, E. M.; Jackson, W. R.; Nilsson, Y. Tetrahedron Lett. 1991, 32, 1093.
6. Hydroformylation in organic synthesis 187
1 I8 Joh, T.; Doyama, K.; Onitsuka, T.; Shiohara, S.; Takahashi, S. Organometallics 1991,
10, 2493.
119 Laine, R. M.; Crawford, E. J. J. Mol. Catal. 1988, 44, 357.
120 Cavinato, G.; Toniolo, L. J. Mol. Catal. 1996, 105, 9.
121 Joh, T.; Fujiwara, K.; Takahashi, S. Bull. Chem. Soc. Jpn. 1993, 66, 978.
122 Joh, T.; Doyama, K.; Fujiwara, K.; Maeshima, K.; Takahashi, S. Organometallics 1991,
10, 508.
123 Hirao, K.; Morii, N.; Joh, T.; Takhashi, S. Tetrahedron Lett. 1995, 36, 6243.
124 Sugioka, T.; Zhang, S.-W.; Mori, N.; Joh, T.; Takahashi, S. Chem. Lett. 1996, 249.
125 Yoneda, E.; Sugioka, T.; Hirao, K.; Zhang, S.-W.; Takahashi, S. J. Chem. Soc., Perkin
Trans. I 1998, 477.
126 Yoneda, E.; Kaneko, T.; Zhang, S-W.; Takahashi, S. Tetrahedron Lett. 1998, 39, 5061.
127 Sugioka, T.; Yoneda, E.; Onitsuka, K.; Zhang, S-W.; Takahashi, S. Tetrahedron Lett.
1997, 38 , 4989.
128 a) Murai, S.; Sonoda, N. Angew. Chem. Int. Ed. Engl. 1979, 18, 837. b) Chatani, N.;
Kajiwata, Y.; Nishimura, H.; Murai, S. Organometallics 1991, 10, 21. c) Chatani, N.;
Murai, S. Synlett 1996, 414.
129 Matsuda, I.; Ogiso, A.; Sato, S.; Isumi, Y. J. Am. Chem. Soc. 1989, 111, 2332.
130 a) Monteil, F.; Matsuda, I.; Alper, H. J. Am. Chem. SOC. 1995, 117, 4419. b) Ojima, I.,
Vidal, E.; Tzamarioudaki, M.; Matsuda, I. J. Am. Chem. SOC. 1995, 117, 6797.
131 Ojima, I.; Fracchiolla, D. A.; Donovan, R. J.; Banerji, P. J. Org. Chem. 1994, 59, 7594.
132 a) Doyle, M. P.; Shanklin, M. S. Organometallics 1993, 12, 11. b) Doyle, M. P.;
Shanklin, S. Organometallics 1994, 13, 1081.
133 Zhou, J.-Q.; Alper, H. Organometallics 1994, 13, 1586.
134 a) Jung, M. E.; Gaede, B. Tetrahedron 1979, 35, 621. b) Carter, M. J.; Fleming, I.;
Percival, A. J. Chem. Soc., Perkin Trans. I 1981, 2415.
135 Pillot, J.; Dunogues. J.; Calas, R. J. J. Chem. Res. Synop. 1977, 268.
136 Fleming, I.; Perry, D. A. Tetrahedron 1981, 37, 4027.
137 Matsuda, I.; Fukuta, Y.; Tsuchihashi, T.; Nagashima, H.; Itoh, K. Organometallics
1997, 16, 4327.
138 Ojima, I.; Vidal, E.; Tzamarioudaki, M.; Matsuda, J. J. Am. Chem. Soc. 1995, 117,
6797.
139 Brfacker, L.; Hollmann, Ch.; Eilbracht, P. Tetrahedron 1998, 54, 4493.
Chapter 7
189
P. W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 189-202.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
190 Chapter 7
reactant and product phase. See Chapter 8 for an overview of the other
separation techniques in use for hydroformylation.
In contrast to heterogeneous catalysts, their homogeneous counterparts
don't resist high temperatures or harsh work-up conditions mostly due to
damage of weak carbon-metal bonds or the loss of ligand stabilization of the
central metal atom of the complexes. The avoidance of (traces) of water or
air at such conditions is also often a parameter that is critical to the success
of the systems. Thus the conditions of the separation of the reaction product
from the solution, in which the catalyst has been dissolved, must be as mild
as possible. A textbook example of such a modern technology is the
application of two- or multi-phase reaction systems. One of the most
important developments in 1980-2000 in the area of homogeneous catalysis
is the successful development of water-stable as well as highly water-soluble
catalyst systems and the consecutive introduction of the aqueous two-phase
technology.
What limitations have to be accounted for if multi-phase systems are
applied is shown in the following listing of possible disadvantages:
- low reaction rates and the resulting low space-time yields due to
insufficient solubility of the reactants or complex interface processes such
as mass-transport limitations,
- surfactant effects that might favor the enrichment of the reaction product
or side products in the reaction media,
- slow or incomplete phase separation,
- the need of high reaction volumes,
- corrosion effects,
- decrease of the chemoselectivity (e.g. pH-controlled side-reactions, etc.).
as the solvent for the Ni based catalyst. The linear higher alkenes are not
soluble in the catalyst phase and can thus easily be separated. This concept is
not applied in hydro formylation.
The third concept is used in an IFP process for the dimerisation of
alkenes. Hydroformylation has been successfully conducted in ionic liquids
in the laboratory [ 1].
Fluorinated organic liquids, concept four, have not been applied in
industry, because of technical and economic disadvantages (expensive
ligands and solvents). Hydroformylation has been investigated in fluorinated
solvents [2] (see Chapter 10).
BINAS
Kinetics.
When two-phase technology is applied, phase transfer phenomena can
play a dominant role in determining reaction rates and kinetics. The
hydroformylation of alkenes in a two-phase water/organic solvent system is
defined by the reaction of two gases, carbon monoxide and hydrogen, with
gaseous or liquid alkenes like ethene, propene or higher homologues.
Dependent on mass transport and reaction conditions (temperature, partial
pressures of the gases, stirring) and reactions partners (type of alkene), the
reaction of the alkene with CO and H2 takes place at the organic-aqueous
interface or in the aqueous bulk phase. No published data are available for
propene or butene.
One study has been carried out using octene-1, which is easy to handle,
thus yielding reliable results for the derivation of kinetic data, although
solubilities change with the composition of the system. Using the Rh/TPPTS
catalyst in an aqueous solution a rate law was determined for the two-phase
194 Chapter 7
k[alkene]o[Rh][H2][CO]
Ro=
(1+KH2[H2])(1+Kco[CO]2)
The mixing of the reaction phases is also important for the reaction rate
so as to provide maximum liquid-liquid interfacial area in cases where the
reaction occurs mostly at the liquid-liquid interface and not in the bulk
aqueous phase. This can be achieved by high agitation speeds, optimized
stirrer type and geometry. In aqueous two-phase systems also the hold-up of
the aqueous phase plays an important role on the reaction rate. With too high
hold-up volumes the reaction rate starts to decrease (phase inversion yielding
the dispersed phase as reaction place which would be, in this case, the
organic phase). Thus it is recommended to keep the aqueous catalyst
solution at an optimized level to provide high productivity with regard to the
aldehyde formation. When using co-solvents like lower alcohols,
polyethylene glycol or acetonitrile the biphasic character of the system is
retained. The rate of the hydroformylation can be enhanced by several times,
albeit at the cost of a lower selectivity as a result of e.g. formation of acetals
or ethers.
7. Aqueous biphasic hydroformyla tion 195
the simple Rh/TPP system. Due to the sufficient strong coordination of the
TPPTS-ligand to the Rh metal-center almost no Rh leaching into the organic
phase is observed - a performance prerequisite with respect to the desired
economy of the process. Even with the low solubility of propene or butene in
water, the high reactivity of the used catalyst system is high enough to
provide a very efficient hydroformylation rate. To increase the solubility of
propylene in the water phase, plant trials were performed with polyethylene
glycol with various chain lengths, acting as an auxiliary-agent for better
propylene solubility. Good reaction rates (rate increase of more than 10 %)
were achieved, demonstrating a simple tool to increase productivity or to
switch to milder conditions. In summary, the major advantages of the
TPPTS-system are the simple, automatically performed mild catalyst
recycle, the comparable low Rh-inventory (rhodium savings), the high 1:b
ratio of ~20/1 and the beneficial low energy consumption of the whole
process as well as the improved utilization of propylene and synthesis gas
(CO/H2).
The water-based process is normally operated at a pH-value between 5
and 6 to provide reaction conditions, which depress unwanted side reactions
of the formed aldehydes. It turned out that a careful pH-control is beneficial
for various operation parameters of the protic reaction media: catalyst
reactivity and selectivity are directly influenced by and thus dependent on
the H3O+ concentration present in the aqueous phase.
providing a simple method for the single conversion of the terminal alkene.
The catalyst isomerizes a small fraction of butene-1 to butene-2. The latter is
not converted under standard reaction conditions. Thus the butanes and the
unconverted butene-2 are separated form the reaction products in a stripping
column. Advantageously the RCH/RP technology can be applied to the
hydroformylation of C3- as well as C4-alkenes in the same production unit;
only minor changes (e.g. reaction temperature, partial pressures) are
necessary to switch between the different feedstock streams [2 13. This is not
the case for alkenes with five or more carbon atoms. Nevertheless the latter
feedstocks correspond to approximately 25 % of the worldwide oxo-
business.
Today the Ruhrchetme/Rhne-Poulene Process (RCH/RP) is successfully
operated at two locations in the world. In the early 1980s a 100,000 ton scale
process was developed and commercialized at the Oberhausen site of
Ruhrchemie AG (now Celanese AG). The first unit was started up in 1984.
In 1988 and in 1999 two additional lanes went on stream. The actual
combined capacity of all three units at Oberhausen (combined with the
conversion of not reacted propylene in a high-pressure hydroformylation set
up) amounts to more than 500,000 tons of C4-products. The 1:b-C4-
aldehydes produced are converted further to the corresponding C4-alcohols
or after aldol condensation and hydrogenation to 2-ethyl- 1 -hexanol (2-EH).
With an annual volume of 360,000 tons of 2-EH the Oberhausen facility
became the worlds largest 2-EH operating unit.
In 1996 the RCH/RP-process was licensed to Hanwha Corp. in South
Korea. With a total capacity of 120,000 tons the unit started up successfully
in 1997. The aldehydes are dedicated mainly for the manufacturing of
butanols.
All units are running well without any major problems being observed.
The process can be operated for long periods of time at steady and stable
conditions. In Table 1 typical process conditions and the composition of the
oxo crude product are shown as fifteen years averages.
7.2.6 Economics
Table 1. Reaction conditions and composition of crude product of the RCH/RP process
(15 years average)
Both factors together with the reduced fixed costs and the usage of an
own technology (no license fees in case of Celanese plants) makes the
RCH/RP process ca. 10 % cheaper in manufacturing costs (costs for ligand
synthesis already included) compared to classical rhodium process applying
a homogeneous phosphine-modified catalyst.
References
1 Chauvin, Y.; Mumann, L.; Olivier, H. Angew. Chem. 1995, 107, 2941.
2 Horvth, T.; Rbai, J. Science, 1994, 266, 72.
3 a)Kuntz, E. D. CHEMTECH, 1987, 17 (9), 570. b) Kuntz, E. G., Fr. Pat. 2,314,910,
1977 (to Rhone-Poulene); Chem. Abstr. 1977, 87, 101944.
4 Bhanage, B.M.; Studies in hydroformylation of olefins using transition metal cataysts,
Ph. D. Thesis, University of Pune, 1995: b) Deshpande, R. M.; Chaudhari, R. V. Ind.
Eng. Chem. Res. 1988. 27, 1996.
5 Frohning, C. D.; Kohlpaintner C. W. in Aqueous-Phase Organometallic Catalysis (Eds.:
B. Cornils, W. A. Herrmnann), VCH, Weinheim, 1998, p. 295.
6 Chaudhari, R. V.; Bhanage B. M. in Aqueous-Phase Organometallic Catalysis (Eds.:
Cornils, B.; Herrmann, W. A.; VCH, Weinheim, 1998, p. 290; b) (Cornils, B.; Wiebus,
E., EP 0,158,246, 1985 (to Ruhrchemie AG).
7 Horvth, I.; Kastrup, R. V.; Oswald, A. A.; Mozeleski, E. J. Catal. Letters 1989, 2, 85.
8 Cornils, B.; Herrmann, W. A.; Aqueous Phase Organometallic Catalysis, Wiley-VCH
Weinheim, 1998, p. 306 ff.
9 Abatjoglou, A. G.; Bryant, D. R. US Patent, 4,731,486, 1988 (to Union Carbide
Cooperation). Chem. Abstr. 1988, 109, 57042.
10 Wasserscheid, P.; Salzer, A. XXXIII. Jahrestreffen Deutscher Katalytiker, Tagungsband
p. 66ff.
11 Bahrmann, H.; Schulte, M. DE 19756945 (to Celanese). Chem. Abstr. 1999, 131, 75261.
12 Cornils, B.; Herrmann, W. A. Aqueous Phase Organometallic Catalysis, Wiley-VCH
Weinheim, 1998, p. 306 ff.
13 Greiner, R.; Burger, B. XXXIII. Jahrestreffen Deutscher Katalytiker, Tagungsband p.
119 ff.
14 Sandee, A. J.; van der Veen, L. A.; Reek J. N. H.; Kamer, P. C. J.; Lutz, M.; Spek, A.
L.; van Leeuwen, P. W. N. M. Angew. Chem. Int. Ed. Engl. 1999, 38, 3231.
15 Bayer, E.; Schurig, V. Chemtech 1976, 212.
16 Jin, Z.; Zheng, X.; Fell, B. J. Mol. Catal. A.; Chem. 1997, 116, 55.
17 Grtner, R.; Cornils, B.; Springer, H.; Lappe, P., DE 3.235.030, to Ruhrchemie AG
(1982).
18 Kulpe, J. A. Dissertation, 1989, Technische Universitt Munchen; b) Kohlpaintner, C.
W. Dissertation, 1992, Techniche Universitt Munchen.
19 Frohning, C. D.; Kohlpaintner, C. W. in Aqueous-Phase Organometallic Catalysis (Eds.:
Cormils, B.; Herrmann, W. A. VCH, Weinheim, 1998, p. 304.
20 Herrmann, W. A.; Kohlpaintner, C. W.; Bahrmann, H.; Konkol, W. J. Mol. Catal. 1992,
73, 191; b) Herrmann, W. A.; Kohlpaintner, C. W.; Manetsberger, R. B.; Bahrmann, H. ;
Kottmann, H. J. Mol. Catal. 1995, 97, 65; c) Bahrmann, H.; Bach, H.; Frohning, C. D.;
Kleiner, H. J.; Lappe, P.; Peters, D.; Regnat, D.; Herrmann, W. A. J. Mol. Catal. 1997,
116,49; d) Herrmann, W. A.; Kohlpaintner, C. W. Angew. Chem. Int. Ed. Engl. 1993,
32, 1524.
21 a) Frohning, C. D.; Kohlpaintner, C. W. in Aqueous-Phase Organometallic Catalysis
(Eds.: Cornils, B.; Herrmann, W. A.; VCH, Weinheim, 1998, p. 302; b) Bahrmann, H.;
Frohning, C. D.; Heymanns, P.; Kalbfell, H.; Lappe, P.; Peters, D.; Wiebus, E. J. Mol.
Catal. 1997, 116, 35.
Chapter 8
Peter Arnoldy
Shell International Chemicals, Shell Research and Technology Centre Amsterdam, Badhuisweg
3, 1031 CM Amsterdam, The Netherlands.
8.1 Introduction
203
P.W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 203-231.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
204 Chapter 8
8.2 Economics
The following calculation shows how product value determines the need
for effective Rh recycle:
Using a Rh price of $30,000/kg and a Rh Concentration in the reactors of 300 ppmw, the
Rh value in the reactors is 9 $/kg reactor content. Assuming 50%w product in the reactors, a
typical petrochemical product value of $1/kg and the need to limit Rh cost to 1% of
manufacturing costs, leads to the need to recycle Rh not less than 1800 times before it can be
wasted. Using the same assumptions but taking a fine chemical product value of %100/kg
results in need for only 18 recycles. Going buck to the petrochemical example, I % Rh cost is
equivalent to a physical Rh loss with crude product of only 0.33 ppmw. assuming that there
are no other Rh loss pathways! Or in chemical terms, at a product mol weight of 72 (butanal),
a turnover is required of 4.3 million mol product/mol Rh!
206 Chapter 8
While catalyst stability affects directly the catalyst cost, other catalyst
performance elements (activity and selectivity) also have their impact.
Catalyst selectivity (alkene loss to aldehyde byproduct, paraffin, alcohol,
heavy ends) determines alkene variable costs, but also capital costs, via
process simplifications and increased production of desired product in the
same facilities. Chemoselectivity to aldehydes is high for all Rh catalysts.
By-products can include aldehyde isomers, low-reactive alkene isomers,
alcohols, alkanes, and heavy ends. Some aldehyde isomers (generally
branched) have a significant value (e.g., isobutanal, some branched detergent
alcohol constituents).
Of all by-products, the formation of heavy ends constitutes the biggest
problem, generally not so much from a product point of view, but rather
from a process point of view. Heavy-ends accumulation can pose serious
process problems, since it can lead to a forced Rh-containing bleed. Figure 2
gives a survey of heavy ends formation reactions, as they can take place in
reactors and other hot places like distillation bottoms.
As has become clear from the above, catalyst cost (Rh and ligand) should
be controlled very carefully. Especially Rh containment is the dominant
theme in development of Rh-based processes. Besides direct catalyst losses,
also catalyst deactivation can take place, which leads to a forced bleed of
deactivated catalyst. Rh can be recovered from such bleeds as well as from
crude product. Below, a survey is given of potential degradation/
deactivation, loss and recovery routes [7]. The chemistry of the degradation
reactions of the ligands will be discussed in Chapter 9, while in this chapter
we will concentrate on the process aspects.
the reactor system (when sufficient separation of catalyst and heavy ends
is not possible) or (ii) because of catalyst deactivation by interaction of
Rh with poisons, like ligand degradation products (see 8.4.2) or external
poisons such as sulfur species and dienes.
Figure 3. Process type I: stripping reactor (gas recycle process Union Carbide)
There is no need for staging via use of more reactors in series, because of
the presence of propene recycle. Butanal linearity is ca. 92%, circa 2%
paraffin is produced, but no butanols. A syn gas bleed is required to get rid
of inert gases (methane, nitrogen, propane), and it is inevitable that some
propene is lost via this bleed. Due to syn gas recycling, H2/CO ratios above
10 are achieved. A large reactor freeboard and a de-mister pad are required
to ensure that no liquid (containing Rh!) is entrained with the gas. Crude
product is condensed at high pressure and subsequently degassed and
stabilized by removal of residual propene. Level control is maintained via
gas flow and crude product recycling. BASF has developed a similar
process, using a slightly higher temperature (110 oC) and lower TPP
concentration, resulting in a smaller gas recycle, but a somewhat lower
linearity (86%) [32].
Figure 4. Process type 11: liquid recycle, distillative separation (Union Carbide process)
Figure 4 gives a typical process scheme for the Union Carbide process.
Reactor conditions are similar to those in the stripping reactor case,
temperature (90 oC) and Rh concentration (250 ppmw) are slightly lower.
The reactor is stirred for good gas/liquid mass transfer. Liquid product is
depressurized and transferred hot to an evaporation section, where Rh
catalyst, dissolved in mainly heavy ends, is separated and recycled. Probably
short-contact time distillation equipment (wiped- or falling-film evaporators)
is applied to protect the Rh catalyst. There is in principle a choice in
distillation conditions: atmospheric distillation leads to relatively high
temperature (thermal strain) with potential for Rh plating and aldehyde
oligomerization to heavy ends. Vacuum distillation would lead to lower
temperature, but oxygen ingress results in some oxidation of phosphorus
ligands and aldehyde; therefore vacuum distillation seems to be not
preferred. For lower aldehydes like butanal, atmospheric distillation is
applied. With increasing molecular weight of products (and heavy ends),
vacuum distillation becomes a requirement. Process for higher aldehydes are
less attractive, because of such harsh vacuum distillation conditions (with
limited stability of aldehydes and catalyst), but also due to the unavoidable
accumulation of heavy ends and the related need for a (catalyst-containing)
heavy ends bleed plus Rh recovery section.
8. Process aspects of rhodium-catalyzed hydroformylation 215
reaction leads to (i) too low reaction rates because of low substrate solubility
or mass transfer limitations, or (ii) inefficient phase separation, because of
emulsification. The two separate phases form after reaction just by cooling
or, more typically, by addition of a solvent. In order to have efficient
extraction after the reaction, the distribution coefficients over two liquid
phases of catalyst and products should be sufficiently different. Generally
one phase is aqueous, and the other one is apolar. Given the product
(a)polarity, ligand design will result in the opposing ligand (a)polarity. Two
type IV concepts will be described: use of apolar catalyst for polar product
(type IVA), and use of polar catalyst for apolar product (type IVB).
Figure 6. Process type IV: one-phase reaction, extractive separation; type IVA: apolar
catalyst, polar product (Kurarays 1 ,4-butanediol process)
(ii) ligand oxidation [50-53]. Very mild process conditions (ca. 60-65 C, 2-
2.5 bar) can be used, because of the high reactivity of AA compared with
unfunctionalized alkenes. These mild conditions will help to suppress heavy
ends (acetals) formation. Reaction takes place in one phase using an apolar
catalyst in an aromatic solvent like toluene. Multi-stage extraction with
water (ca. 30 C, water/organic phase volume ratio of ca. 1:1) leads to
recovery of the hydrophilic product in the aqueous phase, while essentially
all catalyst is left in the organic phase and can be recycled. Rh losses via the
aqueous phase are negligible (10 ppbw). Note that heavy ends as well as
TPPO (formed by oxidative addition of AAs C-O bond to Rh [52]) are less
polar and will accumulate in the apolar catalyst recycle, leading to need for
some bleed from the catalyst recycle. The reaction is very sensitive to CO
pressure. A too low a CO pressure results in low selectivities (production of
more propanal and propanol via AA isomerization and hydrogenation,
respectively) and catalyst deactivation. A too high a CO pressure leads to
lower linearity (< 70% rather than ca. 87%) as well as Rh catalyst loss with
the aqueous product phase. In order to achieve the optimum CO pressure,
good control of reaction rate and avoidance of mass transfer limitation (good
mixing) are important. The H2/CO ratio of syngas feed is ca. 4; CO
starvation is probably prevented by high syngas recycling rates, but addition
of CO in a second stage is an alternative option. AA concentrations should
be kept low: high AA conversions are preferred and back-mixing in the
reactors will help. Typical AA conversions are ca. 98%, propanal and
propanal selectivities are ca. 7 and 3 mol%, respectively. Recently Arco has
been able to achieve a linearity improvement (up to ca. 95%) by use of a
group 8 metal (Fe or Ru) as an additive [54]. There still seems to be scope in
Rh-catalyzed hydroformylation for BDO production: Lyondell has
announced to build 126 kt/a new capacity in the Netherlands in 2001, in
addition to their current 55 kt/a capacity in the US, which would be the
largest single stream BDO unit in the world [55].
Figure 7. Process type IV: one-phase reaction, extractive separation; type IVB: polar catalyst,
apolar product (Union Carbide's higher alkene hydroformylation process)
220 Chapter 8
a
C6-C14 higheralcohols M 1970 II none 23 Ch. 8.6.3
1 -alkenes
g
ethene propanol C,U 1974 I TPP 400 Ch. 8.5.1
The use of a process of the extractive type (type III or IVB) seems most
logical for these higher alkenes, since this enables efficient separation under
mild conditions of aldehydes and heavy ends from polar catalyst. For type III
processes, the major problem is the low solubility of higher alkenes in water
(the presence of salts in the aqueous layer making it even worse) [73,80],
thus limiting the reaction essentially to the interface. Combinations of the
following basic approaches can be used: (i) increase the alkene solubility in
the polar phase, (ii) increase the interfacial area, (iii) use a phase-transfer
catalyst, and (iv) enhance the reaction rate at the interface. To achieve these
objectives, either additives are used (generally selecting the Rh/TPPTS
catalyst) or the phosphine structure is modified. Most approaches, however,
seem to fail in the two-phase system. Once the problems around catalyst
activity are solved, significant problems are still present around loss of
catalyst and additives. Type IVB processes seem to be most convenient. In
8.5.4.2, the process developed by Union Carbide has been described. Many
groups are exploring similar approaches [73,81].
References
58 Abatjoglou, A. G.; Bryant, D. R.; Peterson, R. R., Eur. Pat. Appl. 350922, 1989; U.S.
Pat. 5,180,854, 1993 (to Union Carbide); Chem. Abstr. 1990, 113, 80944.
59 Paetzold, E.; Kinting, A.; Oehme, G., J. Prakt. Chem. 1987, 329, 725.
60 Breslow, D. S.; Skolnik, H., Multi-Sulfur and Sulfur and Oxygen Five- and Six-
Membered Heterocycles, Interscience, New York, 1966; Part 1, pp. 76-95; Part 2, pp.
775-789.
61 Abatjoglou, A. G.; Peterson, R. R.; Bryant, D. R., in Catalysis of Organic Reactions
(Malz, R. E., Ed.), Chem. Ind. 68, Dekker, New York, 1996, pp 133-139.
62 Billig, E.; Abatjoglou, A. G.; Bryant, D. R., U.S. Pat. 4,748,261, 1988; U.S. Pat.
4,885,401, 1989 (to Union Carbide); Chem. Abstr. 1987, 107, 7392.
63 Bryant, D. R.; Nicholson, J. C.; Briggs, J. R.; Packett, D. L.; Maher, J. M., U.S. Pat.
5,886,235, 1999 (to Union Carbide); Chem. Abstr. 1999, 130, 254055.
64 van Rooy, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Goubitz, K.; Fraanje, J.;
Veldman, N.; Spek, A. L., Organometallics 1996, 15, 835.
65 Cuny, G. D.; Buchwald, S. L., J. Am. Chem. Soc. 1993, 115, 2066.
66 Nicholson, J. C.; Bryant, D. R.; Nelson, J. R, U.S. Pat. 5,763,679, 1998 (to Union
Carbide); Chem. Abstr. 1998, 129, 55721,
67 Nicholson, J. C.; Bryant, D. R.; Nelson, J. R.; Briggs, J. R.; Packett, D. L.; Maher, J. M.,
U.S. Pat. 5,874,639, 1999 (to Union Carbide); Chem. Abstr. 1999, 130, 184071.
68 Billig, E.; Bryant, D. R., U.S. Pat. 5.763,670, 1998 (to Union Carbide); Chem. Abstr.
1998, 129, 55719; U.S. Pat. 5,767,321, 1998 (to Union Carbide); Chem. Abstr. 1998,
129, 82975.
69 Bryant, D. R.; Nicholson, J.C., U.S. Pat. 5,763,671, 1998 (to Union Carbide); Chem.
Abstr. 1998, 129, 69096; U.S. Pat. 5,789,625, 1998 (to Union Carbide); Chem. Abstr.
1998, 129, 166637.
70 Bryant, D. R.; Nicholson, J. C.; Briggs, J. R.; Packett, D. L.; Maher, J. M., US. Pat.
5,917,095, 1999 (to Union Carbide); Chem. Ahstr. 1999, 131, 60316.
71 Bryant, D. R.; Leung, T. W.; Billig, E.; Eisenschmid, T. C.; Nicholson, J. C.; Briggs, J.
R.; Packett, D. L.; Maher, J. M., U.S. Pat. 5,874,640, 1999 (to Union Carbide); Chem.
Abstr. 1999, 130, 184072.
72 Leung, T. W.; Bryant, D. R.; Shaw, B. L., U.S. Pat. 5,731,472, 1998 (to Union Carbide);
Chem. Abstr. 1998, 128, 245454.
73 Bahrmann, H., in Aqueous-Phase Organometallic Catalysis (Cornils, B; Herrmann,
W. A., Eds.), Wiley-VCH, Weinheim 1998, pp. 306-321.
74 Tano, K.; Sato, K.; Okoshi, T, Ger. Pat. 3338340, 1984 (to Mitsubishi Chemical); Chem.
Abstr. 1984, 101, 170705.
75 Miyazawa, C.; Hiroshi, M.; Tsuboi, A.; Hamano, K., Eur. Pat. Appl. 272608, 1988 (to
Mitsubishi Chemical); Chem. Abstr. 1988, 109, 15 1931.
76 Wilne, C., in Alpha Olefins Applications Handbook (Lappin, G. R.; Sauer, J.D., Eds.),
Marcel Dekker, New York, 1989, pp. 139-199.
77 Wagner, J. D.; Lappin, G. R.; Zietz, J. R., in Kirk-Othmer Encyclopedia of Chemical
Technology, 4th ed., Vol. 1, 1991, pp. 893-913.
78 Celanese, Chem. Eng. 1981, 88 (22), 68.
79 Unruh, J. D.; Strong, J. R.; Koski, C. L., in Alpha Olefins Applications Handbook
(Lappin, G. R.; Sauer, J. D., Eds.), Marcel Dekker, New York, 1989, pp. 311-327.
80 Hanson, B. E., in Aqueous-Phase Organometallic Catalysis (Cornils, B; Herrmann,
W. A., Eds.), Wiley-VCH, Weinheim 1998, pp. 181-188.
81 Fell, B., Tenside Surf Det. 1998, 35, 326.
82 Hort, E. V.; Taylor, P., in Kirk-Othmer Encyclopedia of Chemical Technology, 4th ed.,
Vol.l, 1991, pp. 209-216.
8. Process aspects of rhodium-catalyzed hydroformylation 231
83 Chen, S. C.; Chu, C. C., Lin, F. S.; Chou, J. Y., U.S. Pat. 5,426,250, 1995 (to Darien
Chemical).
84 Ichikawa, S.; Ohgomori, Y.; Sumitani, N.; Hayashi, H.; Imanari, M., Ind. Eng. Chem.
Res. 1995, 34, 971.
85 Weissermel, K., Arpe, H.-J., Industrial Organic Chemistry, 2nd ed., VCH, Weinheim,
1993, pp. 235-262.
86 Ohgomori, Y.; Suzuki, N.; Sumitani, N., J. Mol. Catal. 1998, 133, 289.
87 Fell, B.; Hermanns, P.; Bahrmann, H., J. prakt. Chem. 1998, 340, 459.
88 Hansen, C.B.; Teunissen, A.J.J.M., Eur. Pat. Appl. 712828 1995 (to DSM, du Pont);
Chem. Abstr. 1996, 125, 114182.
89 Wissing, E.; Teunissen, A. J.; Hansen, C. B.; van Leeuwen, P. W. N. M.; van Rooy, A.;
Burgers, D., Int.PCT Pat. Appl. WO 96/16923 1996 (to DSM, Du Pont); Chem. Abstr.
1996, 125, 114185.
90 Burke, P. M.; Gamer, J. M.; Tam, W.; Kreutzer, K. A., Teunissen, A. J. J. M.; Snijder,
C. S.; Hansen, C.B., Int. PCT Pat. Appl. WO 97/33854 1997 (to DSM, Du Pont); Chem.
Abstr. 1997, 127, 294939; U.S. Pat. 5,874,641 1999 (to DSM, Du Pont); Chem. Abstr.
1999, 130, 198123.
91 Borman, P. C.; Gelling, O. J., Eur. Pat. Appl. 839787 1998 (to DSM); Chem. Abstr.
1998, 128, 323141.
92 Bunel, E. E., Int. PCT Pat. Appl. WO 99106345 1999 (to du Pont, DSM); Chem. Abstr.
1999, 130, 167872.
93 Argyropoulos, J. N.; Bryant, D. R.,; Morrison, D. L.; Stockman, K.E., Int. PCT Pat.
Appl. WO 99138832, 1999 (to Union Carbide); Chem. Abstr. 1999, 131, 131509.
94 Tokitoh, Y.; Yoshimura, N., U.S. Pat. 4,663,468, 1987 (to Kuraray); Chem. Abstr. 1986,
104, 148739.
95 Tokitoh, Y.; Yoshimura, N., U.S. Pat. 4,808,737, 1989; Eur. Pat. Appl. 223103, 1989 (to
Kuraray); Chem. Abstr. 1987, 107, 154896.
96 Matsumoto, M.; Yoshimura, N.; Tamura, M., U.S. Pat. 4,510,332, 1985 (to Kuraray);
Chem. Abstr. 1984, 100, 22329.
97 Tokitoh, Y.; Yoshimura, N., U.S. Pat. 4,808,756, 1989 (to Kuraray); Chem. Abstr. 1987,
107, 96332.
98 Omatsu, T.; Tokitoh, Y.; Yoshimura, N., Eur. Pat. Appl. 303060, 1989 (to Kuraray);
Chem. Abstr. 1989, 111, 38870.
99 Tokuda, Y.; Ruda, K.; Suzuki, S., Jap. Pat. 11071318 1999 (to Kuraray); Chem. Abstr.
1999, 130, 222989.
100 Chem. Week 1994, Jul. 27, p. 26; Eur. Chem. News 1994, Oct. 3, p. 28.
Chapter 9
9.1 Introduction
L3RhA + H2 + CO L&h(CO)H + HA
A = acetate, 2,4-pentanedionate
High temperatures and low pressures accelerate the reaction. The catalyst
HRh(CO)4 must lose one molecule of CO before coordination of the alkene
substrate can take place (this is neither the catalysts resting state nor the
rate-determining step; for the sake of simplicity it is represented this way).
Thus, loss of CO is an intricate part of the catalytic cycle, which includes the
danger of a complete loss of ligands giving precipitation of the metal or one
of the carbonyl clusters. Precipitation of cobalt metal is quite common in
cobalt catalyzed hydroformylation, but because of the high cost of rhodium
(a thousand times more expensive than cobalt) precipitation of rhodium has
to be avoided under all circumstances. Addition of a phosphine or phosphite
ligand stabilizes the rhodium carbonyl species forming HRh(CO)4-n(PR3)n
complexes as discussed in Chapters 3-8. In most instances when ligands are
present, other complexes are formed upon decomposition rather than
rhodium carbonyl clusters or metal (vide infra).
Phosphorus free catalysts are used successfully both in industry and the
laboratory [3-5] and care should be taken to maintain conditions at which no
rhodium cluster or metal is formed. The precipitate observed first is that of
Rh4(CO)12.
The most commonly used phosphines such as tpp, tppms, and tppts have
been well tested for their toxicity owing to their use on industrial scale. No
particular dangers have been reported. The same holds for many phosphites,
which are also marketed as anti-oxidants. Apart from these compounds the
phosphorus compounds and their intermediates should be handled with the
same care as the toxic analogs that are known. Toxicity levels may be
comparable to common agrochemicals. Below we show one example of such
a group of phosphite ligands that turned out to be toxic (Table 1, Figure 2).
Most of the ligands we are using in the laboratory have not been tested.
The latter mechanism has only been observed for palladium compounds
as yet, although it would also be feasible for rhodium(I) compounds giving
an anionic rhodium species and a phosphonium fragment. The actual
phosphorus carbon bond breaking occurs upon the reverse reaction, when the
phosphonium ion adds oxidatively to a low-valent metal.
Oxidative addition of P-C bond to a low-valent metal. Oxidative addition
of C-Br or C-CI bonds is an important reaction in cross-coupling type
catalysis, and while the reaction of a P-C bond is very similar, the breaking
of carbon-to-phosphorus bonds is not a useful reaction in homogeneous
catalysis. The reverse reaction, making of a carbon-to-phosphorus bond via
palladium or nickel catalysis, is becoming increasingly more important for
the synthesis of new phosphines [9]. P-C bond breaking is an undesirable
side-reaction that occurs in systems containing transition metals and
phosphine ligands and it leads to deactivation of the catalysts. The oxidative
addition of a phosphine to a low-valent transition-metal can be most easily
understood by comparing the Ph2P- fragment with a Cl- or Br- substituent at
the phenyl ring; electronically they are very akin, c.f. Hammett parameters
and the like. The phosphido anion formed during this reaction will usually
lead to bridged structures, which are extremely stable. Decomposition of
ligands during hydroformylation has been reported both for rhodium and
cobalt catalysts [10-12].
Thermal decomposition of RhH(CO)(PPh3)3, in the absence of H2 and
CO, leads to a stable cluster shown in Figure 4 containing 2-PPh2 fragments
[ 13] as was studied by Bryants group at Union Carbide. Presumably clusters
of this type also form in hydroformylation plants on the long run. Recovery
of rhodium from these inert clusters is a tedious operation. Reaction of the
cluster mixture with reactive organic halides such as allyl chloride has been
described to give allyldiphenylphosphine and rhodium chloride, which can
be easily extracted into a water layer. [ 14].
9. Catalyst preparation and decomposition 239
Figure 6. Orthometallation
while the aryl phosphites are not. Acids, carbenium ions, and metals catalyze
the Arbusov rearrangement. Many examples of metal catalyzed
decomposition reactions have been reported [38, 39] (see Figure 14).
Figure 16. Stability of phosphites at 160 C after 23 hours (see text for explanation; numbers
give percentage of decomposition)
9. Catalyst preparation and decomposition 247
Dienes and alkynes are poisons for many alkene processes. In polyolefin
manufacture they must be carefully removed as they deactivate the catalyst
completely. Insertion of conjugated dienes is even much more rapid than
insertion of ethene and propene. The resulting -allyl species are inactive as
catalysts.
In rhodium catalyzed hydroformylation the effect is less drastic and often
remains unobserved, but surely diene impurities obscure the kinetics of
alkene hydroformylation [42]. Because the effect is often only temporary we
summarize it here under dormant sites. Hydroformylation of conjugated
alkadienes is much slower than that of alkenes, but also here alkadienes are
more reactive than alkenes toward rhodium hydrides [43, 44]. Stable -allyl
complexes are formed that undergo very slowly insertion of carbon
monoxide (Figure 17). The resting state of the catalyst will be a -allyl
species and less rhodium hydride is available for alkene hydroformylation.
Thus, alkadienes must be thoroughly removed as described by Garland [45],
especially in kinetic studies. It seems likely that 1,3- and 1,2-diene
impurities in 1 -alkenes will slow down, if not inhibit, the hydroformylation
of alkenes.
References
1 (a) Mieezynska, E.; Trzeciak, A. M.; Zilkowski, J. J. J. Mol. Catal. 1993, 80, 189. (b)
Buhling, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Mol. Catal. A: Chemical,
1995, 98, 69.
2 Coolen, H. K. A. C.; van Leeuwen, P. W. N. M.; Nolte, R. J. M. J. Org. Chem. 1996, 61,
4739.
3 Onada, T. ChemTech, September 1993 , 34.
4 Lazzaroni, R.; Pertici, P.; Bertozzi, S.; Fabrizi, G. J. Mol. Cat. 1990, 58, 75.
5 Vidal, J. L.; Walker, W. E. Inorg. Chem. 1981, 20, 249.
6 Bellet, E. M.; Casida, J. E. Science, December 1973, 182, 1135.
7 Garrou, P. E. Chem. Rev. 1985, 85, 171.
8 Garrou, P. E.; Dubois, R. A.; Jung, C. W. ChemTech, February 1985, 123.
9 Cai, D.; Payack, J. F.; Bender, D. R.; Hughes, D. 1.; Verhoeven, T. R.; Reider, P. J. J.
Org. Chem. 1994, 51 ,629.
10 Chini, P.; Martinengo, S.; Garlaschelli, G. J. Chem. Soc. Chem. Commun. 1972, 709.
11 Dubois, R. A.; Garrou, P. E. Organometallics, 1986, 5, 466.
12 Harley, A. D.; Guskey, G. J.; Geoffroy, G. L. Organometallics, 1983, 2, 53.
13 Billig, E.; Jamerson, J. D.; Pruett, R. L. J. Organomet. Chem. 1980 192, C49.
14 Miller, D. J.; Bryant, D. R.; Billig, E.; Shaw, B. L. U.S. Pat 4,929,767 (to Union Carbide
Chemicals and Plastics Co.) 1990; Chem. Abstr. 1991, 113, 85496.
15 Herrmann W. A.; Kohlpaintner, C. W. Angew. Chem. Int. Ed. Engl. 1993, 32, 1524.
16 Goel, A. B. Inorg. Chim. Acta, 1984, 84, L25.
17 Sakakura, T. J. Organometal. Chem. 1984, 267, 17 1.
18 Abatjoglou A. G.; Bryant, D. R. Organometallics 1984, 3, 932.
19 Kong, K-C.; Cheng, C-H. J. Amer. Chem. SOC. 1991, 113, 6313.
20 Sisak, A.; Ungvry, F.; Kiss, G. J. Mol. Catal. 1983, 18, 223.
250 Chapter 9
49 Bonds, W. D.; Brubaker, C. H.; Chandrasekaran, E. S.; C. Gibsons, Grubbs, R. H.; Kroll,
L. C. J. Am. Chem. Soc. 1975, 97, 2128.
50 van Rooy, A.; Buisman, G. J. H.; Roobeek, C. F.; Sablong, R.; van Leeuwen, P. W. N.
M. unpublished results.
51 Parshall, G. W.; Knoth, W. H.; Schunn, R. A. J. Am. Chem. SOC. 1969, 91, 4990.
52 Coolen, H. K. A. C.; van Leeuwen, P. W. N. M.; Nolte, R. J. M. J. Organomet. Chem.
1995, 496, 159.
53 Fernandez, E.; Ruiz, A.; Claver, C.; Castillon, S.; Chaloner, P. A.; Hitchcock, P. B.
Inorg. Chem. Comm. 1999 ,2, 21.
Chapter 10
10.1 Introduction
It has been shown that a number of active sites of enzymes contain two
metal ions that give high activity via a cooperative bimetallic mechanism
253
P.W.N.M. van Leeuwen and C. Claver (eds.), Rhodium Catalyzed Hydroformylation, 253-279.
2000 Kluwer Academic Publishers. Printed in the Netherlands.
254 Chapter 10
Figure I. The proposed mechanism for the bimetallic hydroformylation using ligand 1
The bridge between the two metal centers is too large in ligands 6 and 7,
which prevents the cooperation between the metal centers. Indeed activities
and selectivities were found that were similar to the monometallic species.
Another important indication supporting this bimetallic cooperativity was an
256 Chapter 10
Figure 2. Schematic presentation of the aggregation of amphiphiles into micelles, bilayers and
vesicles
Figure 3. Some of the amphilic phosphine ligans that have been reported
Figure 4. Schematic representation of catalysis using amphiphilic ligands that form vesicles
very large interfacial area, which results in very efficient catalysis for
organic substrates. Furthermore, the catalyst stays completely on the support.
The important issues are the generality of the concept, the robustness of the
catalyst system and the influence of the thickness of the water layer [16].
This water layer has an enormous impact on the catalytic activity (figure 6).
If this layer is too thin the activity of the catalyst is much lower due to a
decrease of the catalyst mobility. The catalyst is bound to the silica resulting
in a heterogeneous system.
This was verified by 31P NMR spin relaxation measurements, which
show that the spin relaxation declines with increasing water content [ 17]. It
was observed in the hydroformylation of 1-heptene using a
HRh(CO)(TPPTS)3-SAPC that the TOF increases with a factor as high as
100 when going from 2.9 to 9 wt. % water (table 3).
10. Novel developments in hydroformylation 261
At higher water contents of the support the water layer becomes too thick
and the substrate has to diffuse into the water layer, or the catalyst has to
diffuse to the interface. The result is a decrease in catalyst-product contact
time leading to lower activities. This sensitivity towards water is a drawback
of this otherwise attractive concept. Horvth performed experiments using
substrates with different solubilities in water and showed that, under optimal
conditions, this solubility did not influence the activity [ 18]. Furthermore, he
performed a hydroformylation reaction in a continuous system and even
under reaction conditions no leaching of rhodium complex was detected. The
water obviously leaches if the SAPC is used in a continuous flow system,
which in a practical application should be compensated for by using water-
saturated organic solvents.
Figure 6. The influence of the water contents of the hydrophilic support on the relative
catalytic activity
as SAPC; they were much more selective than the SAPC known so far [ 19b].
Recycling experiments showed that these catalysts retained their activity and
selectivity in at least ten consecutive runs, whereas under similar conditions
the TPPTS based catalyst showed a reduced performance in the fourth run.
Mortreux compared the activity of the SAPC catalysts with that of the
homogeneous analogue in the hydroformylation of methyl acrylate [20].
They observed an activity for the SAPC that was strongly dependent on the
amount of water present in the system. More remarkably, the optimized
activity of the SAPC was higher than that of the homogeneous systems. This
effect was ascribed to the polar interactions between the substrate and the
silica support. This effect was not observed for nonpolar substrates such as
propene, which supported the hypothesis.
In order to show the versatility of the method Davis extended the concept
to other hydrophilic liquids such as ethylene glycol and glycerol [21]. The
reactions then take place at the hydrophilic-hydrophobic liquid interface. In
this specific example the supported-phase concept was used for the
asymmetric reduction using a ruthenium catalyst. The obtained
enantioselectivity was higher than that of the SAPC, which was ascribed to
the decrease in hydrolysis of chloro-ligand from the ruthenium complex.
Naughton et al. used this supported homogeneous film catalysis concept for
the hydroformylation reaction using TPPTS as the ligand [22]. The low
molecular weight polyethylene glycol resulted in the formation of an
effective film.
13 14 15
Below the critical temperature of CO2, thus in the liquid state, almost no
activity was observed, presumably due to the low solubility of the precursor
in liquid.
On using phosphine ligands the selectivity of the reaction increased, up
to 99% aldehyde was formed. The selectivity for the linear product increased
upon changing the P/Rh ratio from 4 to 10; the linear to branched (1:b) ratio
increased from 3.2 to 5.6. A further increase in P/Rh ratio did not improve
the selectivity further, but a dramatic drop in activity was observed. Similar
effects are observed for reactions performed in organic solvents, indicating
that the reactive species is indeed the phosphine-rhodium complex.
Interestingly, the phosphite ligand 15 also gives a high chemo- and regio-
selectivity, whereas in organic solvents under similar conditions this type of
catalysts gives a lot of isomerized products. -H elimination of the rhodium
alkyl complex results in the internal alkenes, whereas the CO insertion leads
to the aldehyde (see figure 6, chapter 4). The high CO concentration in
scCO2 favors the CO insertion, which results in higher selectivities
compared to the analogous reactions performed in organic media.
The catalyst-product separation was also studied. By changing the
pressure and temperature after the reaction they created a phase separation
from which the aldehydes could be collected as a colorless liquid (containing
less than 1 ppm rhodium). The catalysts of ligands 13-15 were used in 5
consecutive runs without significant changes in the catalytic performance.
The concept can be extended to asymmetric hydroformylation of styrene
in scCO2 using BINAPHOS as the chiral ligand. The initial experiments
using this ligand clearly show that the active catalyst in scCO2 does not
involve the chiral ligand [27]. The low solubility of the ligand was the major
problem, since the catalyst precursor in absence of ligand also gave a high
activity, but of course no enantioselectivity. Therefore a BINAPHOS
similar using temperatures above and below the Tc of CO2. A slightly higher
regioselectivity was also observed on using 16 compared to BINAPHOS in
the homogeneous phase, which was ascribed partly to the ligand and partly
to the solvent.
Erkey studied the hydroformylation of 1 -octene in scCO2 using several
different fluoralkyl and fluoralkoxy functionalized aryl phosphines [29].
They found an increasing activity of the rhodium catalyst with decreasing
basicity of the phosphine ([3,5-(CF3)2C 6H3] 3P > [4-CF3C6H4]3P and [3-CF3
C6H4]3P > [4-CF3OC6H4]3P). The basicity was measured by the carbonyl
stretching frequencies of the complexes.
The low solubility of the catalyst in scCO2 is not necessarily a problem.
Cole-Hamilton used rhodium trialkylphosphine complexes for the
hydroformylation of 1-hexene in scCO2 [30]. The rhodium concentration
was about 6.5 mM with a P:Rh ratio of 6. The catalyst performed similarly
in scCO2 compared to experiments performed in toluene. The reaction rate
decreased with increasing PEt3, whereas higher Pco or PH2 resulted in an
enhanced reaction rate. The selectivity for the linear product slightly
increased on using scCO2 compared to toluene (1:b = 2.4 and 2.1
respectively).
Although the scCO2 approach is very elegant, one should keep in mind
that it is a relatively expensive method, especially for the bulk chemical
industry. For each recycling step the system has be depressurized and
pressurized again with the newly added substrate. This process requires more
powerful pumps and more capital expensive reactors than for example
biphasic aqueous phase process. Furthermore, the low solubility of potential
interesting substrates might hamper the commercialization of scCO2 in the
fine chemical industry.
Table 5. Effect of the phosphine concentration on the selectivity and activity of the
hydroformylation of 1 -decene a
b
[L] mM P/Rh TOF 1:b
-1 -1
(mol.mol .h )
22.1 27 2160 3.3
22.5 39 1764 3.3
41.3 76 1440 4.4
82.1 79 756 5.0
152.2 103 288 6.3
304.0 102 144 7.8
23 Not given 1908 3.2
c
23 Not given 22572 3.2
d
23 Not given 792 2.3
a
Conditions: 50/50 vol. % toluene/C6F11CF3, T=100 C, pressure=11 bar CO/H2 (1:1),
b c d
initial [I-decene]= 1 M. P[CH2CH2(CF2)5CF3]3 PPh3 P[(CH2)7CH3]3.
At higher concentrations of the ligand a lower activity and higher 1:b ratio
was observed, which is in line with results previously obtained in organic
solvents. The activity of the fluorous pony-tail phosphine ligand was similar
to the alkyl phosphine, but remarkably the selectivity was identical to the
PPh3 system. No explanation was found for this effect. A minor effect of the
fluorous phase on the catalytic performance of the Rh/PPh3 was observed.
The activity was approximately 30 % lower in 50/50 vol % toluene/C6F11CF3
compared to toluene, and the 1:b ratio slightly increased. This effect was
attributed to differences in gas solubilities.
The fluorous biphase catalyst recovery concept was tested by performing
nine consecutive reactions in a batch reactor. A total loss of 4.2% of rhodium
was detected, and the decreasing 1:b ratio suggested some ligand leaching.
The total turnover number reached with the system was 35,000 mole of
aIdehyde/mole rhodium, with a rhodium loss of 1.18 ppm per mole of
aldehyde. Further optimization of this system, i.e. the use of heavier fluorous
solvents and longer pony tails, should decrease the amount of catalyst and
fluorous solvent leaching.
it was suggested that this is due to steric congestion of the more compact
dendrimers. Additional Soxhlet extraction experiments showed that rhodium
leached under catalytic conditions; both CO and substrate promote metal
leaching from the support. The phosphite-modified polymers grafted on
silica did not show any rhodium leaching when the hydroformylation was
performed under the proper conditions 139]. A constant conversion was
observed in a continuous flow experiment that was performed for 10 days in
benzene (residence time 3.3h).
Functionalized dendrimers offer new solutions to the catalyst/product
separation problem of homogeneous catalysts and might give catalysts with
unique novel properties. However, if these systems will ever be attractive
alternatives for commercial processes remains to be seen. They have to
compete with other systems consisting of cheaper polymer materials. Metal
leaching and catalyst decomposition are serious problems that should be
addressed. Especially the influence of the local very high concentration of
metal sites in the case of periphery functionalized dendrimers can promote
both these effects. The general problem of metal leaching can be solved by
the proper choice of ligands, as will be clear from an example given in the
next section.
Figure 10. The phosphine functionalized PAMAM dendrimers anchored on a silica support
17d (cf. run 7 and 10). This ligand is frozen in the polymer matrix in a
different conformation, since it is co-polymerized at three positions of the
ligand. Some of these conformations are not able to form the rhodium
complex that gives the enantioselective hydroformylation catalyst. In
contrast, the other ligands have more freedom and can therefore form the
proper catalyst after the polymerization. The degree of cross-linking has no
significant influence on the selectivity of the reaction (runs 2, 4, 5, and 11).
The reaction can be performed in hexane, but the ee was slightly lower. The
recycling experiments clearly show the drawback of these types of polymers.
The polymers were partly crushed by the stirring bar and approximately 50%
of the initially charged rhodium was removed after the first run.
Figure 12. Two different routes to prepare sol-gel immobilized transition metal complexes
These silica immobilized systems prepared via the sol-gel method are
very promising organic-inorganic hybrid materials. The chemistry takes
place at the interface of the materials suppressing the detrimental influence
of the support [46c]. A series of network modifiers for the sol-gel process
are available, which can be used to optimize the support. This tool to further
improve the catalyst performance already proved to be valuable for several
reactions [46c]. For the hydroformylation reaction this still needs to be
explored.
18
a strong preference for the heptane layer. So far, none of the smart polymer
concepts has been tested for the hydroformylation reaction. However, the
current rapid developments in polymer technology and the already described
interesting systems makes the usage of smart functionalized polymers a
promising development.
Figure 13. The functionalized cyclodextrin building block as supramolecular catalyst for two-
phase catalysis. The actual catalysis is proposed to take place at the phase boundary
Interestingly, the selectivity for the linear product increased on using the
supramolecular system; the 1:b ratio was 3.2 compared to 1.5 for the parent
catalyst. A better test for the selectivity of this supramolecular catalyst is
allylbenzene, a substrate that easily isomerizes to methyl styrene. Again the
hydroformylation reaction was very selective and a higher 1:b ration was
observed on using the supramolecular system. The presence of an excess of
toluene, a guest molecule that competes for the cavity with the substrate,
resulted in a lower regioselectivity. This indicates that the formation of the
supramolecular complex is responsible for the higher selectivity.
19
Initial experiments aiming at recycling of the catalysts showed that the
supramolecular system was mainly in the water layer after separation from
the organic phase. Reuse of the aqueous phase revealed that 50% of the
catalytic activity was retained. Prior the work of Reetz several groups
studied the effect of cyclodextrins as inverse phase transfer materials on the
276 Chapter 10
and available in large quantities, which makes them suitable for commercial
applications.
10.5 Conclusions
References
35 (a) Reetz, M. T.; Lohmer, G.; Schwickardi, R. Angew. Chem. Int. Ed. Engl. 1997, 36,
1526. (b) Brinkman, N.; Giebel, D.; Lohmer, G.; Reetz, M. T.; Kragl, U. J. Catal. 1999,
183, 163.
36 de Groot, D.; Eggeling, E. B.; de Wilde, J. C.; Kooijman, H.; van Haaren, R. J.; van der
Made, A. W.; Spek, A. L.; Vogt, D.; Reek, J. N. H.; Kamer, P. C. J.; van Leeuwen, P. W.
N. M. Chem. Commun. 1999, 1623.
37 Hovestad, N. J.; Eggeling, E. B.; Heidbchel, H. J.; Jastrzebski, J. T. B. H.; Kragl, U.;
Keim, W.; Vogt, D.; van Koten, G. Angew. Chem. Int. Ed. Engl. 1999, 38, 1655.
38 (a) Bourque, S. C.; Maltais, F.; Xiao, W. -J.; Tardif, O.; Alper, H.; Arya, P.; Manzer, L.
E. J .Am. Chem. SOC. 1999, 121, 3035. (b) Bourque, S. C.; Alper, H.; J. Am. Chem. Soc.
2000, 122, 956. (c) Arya, P.; Rao, N. V.; Singkhonrat, J.; Alper, H.; Bourque, S. C.;
Manzer, L. E. J. Org. Chem. 2000, 65, 1881.
39 Jongsma, T.; van Aert, H.; Fossen, M.; Challa, G.; van Leeuwen, P. W. N. M. J. Mol.
Catal. 1993, 83, 37.
40 Hartley, F. R.. Supported metal complexes, a new generation catalysts, Reidel,
Dordrecht, 1985.
41 (a) Pomogailo, A. D.; Whrle, D. in Macromolecule-Metal Complexes, Ciardelli, F.;
Tsuchida, E.; Whrle, D. Eds., Springer, Berlin, 1996. (b) Keim, W.; Driessen-Hlscher,
B. in Handbook of Heterogenous Catalysis, vol 1, Ertl, G.; Knzinger, H.; Weitkamp, J.
Eds. WILEY-VCH, Weinheim, 1997.
42 (a) Nozaki, K.; Itoi, Y.; Shibahara, F.; Shirakawa, E.; Ohta, T.; Takaya, H.; Hiyama, T.
J. Am. Chem. Soc. 1998, 120, 4051. (b) Nozaki, K.; Shibahara, F.; Itoi, Y.; Shirakawa,
E.; Ohta, T.; Takaya, H.; Hiyama, T. Bull. Chem. Soc. Jpn. 1999, 72, 191 1.
43 Farrell, M. O.; van Dyke, C. H.; Boucher, L. J.; Metlin, S. J. J. Organomet. Chem. 1979,
172, 367.
44 Deschler, U.; Kleinschmit, P.; Panster, P.; Angew. Chem. Int. Ed. Engl. 1986, 25, 236.
45 (a) Monaco, S. J.; Ko, E. I. CHEMTECH, 1998, 23. (b) Ingersoll, C. M.: Bright, F. V.
CHEMTECH, 1997, 26.
46 (a) Lindner, E.; Kemmler, M.; Mayer, H. A.; Wegner, P. J. Am. Chem. Soc. 1994, 116,
348. (b) Lindner, E.; Schneller, T.; Auer, F.; Wegner, P.; Mayer, H. A. Chem. Eur. J.
1997, 3, 1833. (c) Lindner, E.; Schneller, Auer, F.; Mayer, H. A. Angew. Chem. Int. Ed.
Engl. 1999, 38, 2155.
47 Wieland, S.; Panster, P. Catal. Org. React. 1994, 62, 383.
48 Sandee, A. J.; van der Veen, L. A.; Reek. J. N. H.; Kamer, P. C. J.; Lutz, M.; Spek, A.
L.; van Leeuwen, P. W. N. M. Angew. Chem. lnt. Ed. Engl. 1999, 38, 3231.
49 (a) Bergbreiter, D. E.; Catal. Today, 1998, 42, 389. (b) Bergbreiter, D. E.; Zhang, L.;
Mariagnanam, V. M. J. Am. Chem. SOC. 1993, 115, 9295. (c) Bergbreiter, D. E.; Liu, Y.-
S.; Osburn, P. L. J. Am. Chem. Soc. 1998, 120, 4250.
50 (a) Coolen, H. K. A. C.; van Leeuwen, P. W. N. M.; Nolte, R. J. M. Angew. Chem. Int.
Ed. Engl. 1992 ,31, 906. (b) Coolen, H. K. A. C.; Meeuwis, J. A. M.; van Leeuwen, P.
W.N. M.;Nolte,R. J. M. J. Am. Chem. Soc. 1995, 117, 11906.
51 Reetz, M. T.; Waldvogel, S. R. Angew. Chem. Int. Ed. Engl. 1997, 36, 865. (b) Reetz, M.
T. J. Heterocyclic Chem. 1998, 35, 1056. (c) Reetz, M. T. Catal. Today, 1998, 42, 399.
52 For hydroformylation reactions with cyclodextrins as additives: (a) Anderson, J. R.;
Campi, E. A.; Jackson, W. R. Catal. Lett. 1991, 9, 55. (b) Monflier, E.; Fremy, G.;
Castanet, Y.; Mortreux, A. Angew. Chem. Int. Ed. Engl. 1995, 34, 2269. (c) Monflier, E.;
Tilloy, S.; Fremy, G.; Castanet, Y.; Mortreux, A. Tetrahedron Lett. 1995, 36, 9481. (d)
Tilloy, S.; Bertoux, F.; Mortreux, A.; Monflier, E. Catal. Today, 1999, 48, 245.
281
Index
BDPP 134
Berry mechanism 114
bimetallic catalysts 253
BINAP 3 4 134
BINAPHOS 4 7 108 124 264 270
binaphthyl 125
BINAS 193
biphasic hydroformylation 135 136
BIPHEPHOS 147 170 178
BIPHOLOPHOS 136
BISBI ligands 7 82 83 84 85 86
87 136 147 258
BISBIS 192
bisphenol bridges 45
bite angle 5 8 10 82-99 135
BPPM 132
bridge length 45
catalyst decomposition 235 236 237 238 239 240
241 242 243 244 245 246
247 248 249
bulky diphosphite catalysts 45 108
bulky phosphite 6 7 37 38 39 40
41 42 43 44 151 158
1,4-butanediol 217 225
2-butene 46
DEGUPHOS 134
dendrimer supported catalysts 268
detergent alcohols 223
detergents 201
deuterioformylation 16 24 25 26 27 30
31 33 65 84
dihydrofurans 150
dihydropyrans 151
dimer formation rhodium complexes 54 64 72 247
dimethoxyacrolein 153
dimethylbut-1-ene 18 25
dinuclear species 137
DIOCOL 133
DIOP 2 78 82 84 133
dioxaphosphepine 120
DIPAMP 2 4
diphosphines as ligands 76-102 131 132 133 134 135
136 137 138 139 140
diphosphites as ligands 44-59 108-120
dirhodium species 54 65 69 137
dissociation of CO 74 93 100 101 102
distillative separation 214
1,1-disubstituted alkenes 149 164
dormant sites 247
double-bond isomerization 207
DPBS 219
DPPB, dppb 217
DuPhos 3 4
electronic parameter 8
electron-withdrawing phosphines 79
enantiofacial selection 128
energetics 100 101 102
ethyl acrylate 139 154
evaporation 211 212 213 214
extraction 211 215 216 217 218 219
220 225
extraction after one-phase reaction 216
Josiphos 4
Kinetic studies 43 47 69 70 71 72
97 98 99 100 101 102
112
kinetics two-phase catalysis 193 194 195
kinetics, fluorous phase 266
Ligand effects 1 2 3 4 5 6
7 8 9 10 11 12
13 14 38 66 67 68
76-99
ligand loss 209 235 236 237 238 239
240 241 242 243 244 245
246
liquid recycle process 213
long chain alkenes 200
Nanofiltration 268
NAPHOS 137 193
natural bite angle 82 88 96 97 98 99
NMP 220
NMR spectroscopy 49 50 51 52 68 69
80 91
nylon monomers 226
Terpenes 167
Tischenko reaction 207
tpp 6 63 64 65 66 67
68 69 70 71 72 73
74 75 212 213 221
tppms 5 7 221
tppts 5 6 7 191 215 221
TPPTS, see tppts
transition state geometry 98
trifluoropropene 18 153
triphenyl phosphite 38
triphenylphosphine, see tpp
tris(ortho tert-butylphenyl)phosphite 37 38 39 40 41 42
43 44 221
tris(2,2,2-trifluoroethyl) phosphite 38
tum-stile mechanism 53 114 138
two-phase catalysis 190
two-phase reaction 215
vinylpyrrole 153
Wilkinson 64 72
Wittig reaction 177
Xantphos ligands 87 88 89 90 91 92
93 94 95 147 201 259
262 273
xylofuranose 122