Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Electronic Effect On Rhodium Diphosphine Catalyzed Hydroformylation: The Bite Angle Effect Reconsidered

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

11616 J. Am. Chem. Soc.

1998, 120, 11616-11626

Electronic Effect on Rhodium Diphosphine Catalyzed


Hydroformylation: The Bite Angle Effect Reconsidered
Lars A. van der Veen,† Maarten D. K. Boele,† Frank R. Bregman,† Paul C. J. Kamer,†
Piet W. N. M. van Leeuwen,*† Kees Goubitz,‡ Jan Fraanje,‡ Henk Schenk,‡ and Carles Bo§
Contribution from the Institute for Molecular Chemistry, UniVersity of Amsterdam,
Nieuwe Achtergracht 166, 1018 WV Amsterdam, The Netherlands, and
Department of Physical and Inorganic Chemistry, UniVersitat RoVira i Virgili,
Plaza Imperial Tarraco 1, 43005 Tarragona, Spain
ReceiVed June 5, 1998

Abstract: The electronic effect in the rhodium diphosphine catalyzed hydroformylation was investigated. A
series of electronically modified thixantphos ligands was synthesized, and their effects on coordination chemistry
and catalytic performance were studied. Phosphine basicity was varied by using p-(CH3)2N, p-CH3O, p-H,
p-F, p-Cl, or p-CF3 substituents on the diphenylphosphine moieties. X-ray crystal structure determinations of
the complexes (thixantphos)Rh(CO)H(PPh3) and (p-CH3O-thixantphos)Rh(CO)H(PPh3) were obtained. The
solutions structures of the (diphosphine)Rh(CO)H(PPh3) and (diphosphine)Rh(CO)2H complexes were studied
by IR and NMR spectroscopy. IR and 1H NMR spectroscopy showed that the (diphosphine)Rh(CO)2H
complexes consist of dynamic equilibria of diequatorial (ee) and equatorial-apical (ea) isomers. The equilibrium
compositions proved to be dependent on phosphine basicity; the ee:ea isomer ratio shifts gradually from almost
one for the p-(CH3)2N-substituted ligand to more than nine for the p-CF3-substituted ligand. Assignments of
bands to ee and ea isomers and the shifts in wavenumbers in the IR spectra were supported by calculations on
(PH3)2Rh(CO)2H, (PH3)2Rh(CO)2D, and (PF3)2Rh(CO)2H complexes using density functional theory. In the
hydroformylation of 1-octene and styrene an increase in l:b ratio and activity was observed with decreasing
phosphine basicity. Most remarkably for 1-octene the selectivity for linear aldehyde formation was between
92 and 93% for all ligands. These results indicate that the chelation mode in the (diphosphine)Rh(CO)2H
complexes per se is not the key parameter controlling the regioselectivity. Mechanistic explanations of the
effect of the natural bite angle on regioselectivity are reconsidered.

Introduction using rhodium diphosphite catalysts.14 A new highly active and


Rhodium-catalyzed hydroformylation is one of the most selective catalyst system based on binuclear rhodium complexes
important applications of homogeneous catalysis in industry.1-5 and tetraphosphine ligands has been developed by Stanley and
Since the discovery of the rhodium catalysts by Wilkinson,6,7 co-workers.15 Systematic studies of the correlation between
much effort has been made to enhance the regioselectivity of phosphine structure and catalytic performance are rare, however,
the reaction toward the formation of the more desirable linear and despite the development of a wide variety of phosphorus
aldehyde. High selectivities in the hydroformylation of 1-al- ligands, no detailed understanding concerning how phosphorus
kenes have been obtained for both diphosphite- and diphosphine- ligands control selectivity has emerged.
modified catalysts.8-13 Functionalized alkenes have been hy- The generally accepted dissociative mechanism for the
droformylated with high selectivity by Cuny and Buchwald rhodium-catalyzed hydroformylation as proposed by Wilkinson

is shown in Scheme 1.16 According to this mechanism for the
Department of Inorganic Chemistry, University of Amsterdam.
‡ Department of Crystallography, University of Amsterdam. hydroformylation of 1-alkenes, the selectivity is determined in
§ Universitat Rovira i Virgili. the step that converts a five-coordinate L2Rh(CO)H(alkene) into
(1) Parshall, G. W. Homogeneous Catalysis: the applications and
chemistry of catalysis by soluble transition metal complexes; Wiley: New (9) van Rooy, A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M.; Goubitz,
York, 1980. K.; Fraanje, J.; Veldman, N.; Spek, A. L. Organometallics 1996, 15, 835-
(2) Cornils, B. In New Syntheses with Carbon Monoxide; Falbe, J., Ed.; 847.
Springer-Verlag: Berlin, Heidelberg, 1980; pp 1-225. (10) Kranenburg, M.; van der Burgt, Y. E. M.; Kamer, P. C. J.; van
(3) Tolman, C. A.; Faller, J. W. In Homogeneous Catalysis with Metal Leeuwen, P. W. N. M. Organometallics 1995, 14, 3081-3089.
Phosphine Complexes; Pignolet, L. H., Ed.; Plenum: New York, 1983; pp (11) Billig, E.; Abatjoglou, A. G.; Bryant, D. R. (to Union Carbide Corp.)
81-109. U.S. Patent 4.769.498, 1988.
(4) Beller, M.; Cornils, B.; Frohning, C. D.; Kohlpaintner, C. W. J. Mol. (12) Devon, T. J.; Phillips, G. W.; Puckette, T. A.; Stavinoha, J. L.;
Catal. A: Chem. 1995, 104, 17-85. Vanderbilt, J. J. (to Eastman Kodak) U.S. Patent 4,694,109, 1987.
(5) Frohning, C. D.; Kohlpaintner, C. W. In Applied Homogeneous (13) Burke, P. M.; Garner, J. M.; Tam, W.; Kreutzer, K. A.; Teunissen,
Catalysis with Organometallic Compounds: a comprehensiVe handbook A. J. J. M.; Snijder, C. S.; Hansen, C. B. (to DSM/Du Pont de Nemours)
in two Volumes; Cornils, B., Herrmann, W. A., Eds.; VCH: Weinheim, WO 97/33854, 1997.
New York, Basel, Cambridge, Tokyo, 1996; Vol. 1, pp 27-104. (14) Cuny, G. D.; Buchwald, S. L. J. Am. Chem. Soc. 1993, 115, 2066-
(6) Evans, D.; Yagupsky, G.; Wilkinson, G. J. Chem. Soc. A 1968, 2660- 2068.
2665. (15) Broussard, M. E.; Juma, B.; Train, S. G.; Peng, W.-J.; Laneman, S.
(7) Brown, C. K.; Wilkinson, G. J. Chem. Soc. A 1970, 2753-2764. A.; Stanley, G. G. Science 1993, 260, 1784-1788.
(8) Buisman, G. J. H.; Vos, E. J.; Kamer, P. C. J.; van Leeuwen, P. W. (16) Evans, D.; Osborn, J. A.; Wilkinson, G. J. Chem. Soc. A 1968,
N. M. Tetrahedron: Asymmetry 1995, 6, 719-738. 3133-3142.

10.1021/ja981969e CCC: $15.00 © 1998 American Chemical Society


Published on Web 11/03/1998
Rhodium Diphosphine Catalyzed Hydroformylation J. Am. Chem. Soc., Vol. 120, No. 45, 1998 11617

Scheme 1 Table 1. Selected Data for Ligands 1-7


flexibility
ligand R βn,a deg range,a deg χi,b cm-1 σpc
1 N(CH3)2 109.1 92-124 1.8 -0.83
2 OCH3 106.9 91-123 3.4 -0.27
3 CH3 106.7 91-125 3.5 -0.17
4 H 106.4 91-127 4.3 0.00
5 F 106.6 92-128 5.0 0.06
6 Cl 107.8 91-126 5.6 0.23
7 CF3 109.3 92-128 6.6 0.54
a
The natural bite angle (βn) and the flexibility range were calculated
analogously to the method used by Casey and Whiteker.23 βn is defined
as the preferred chelation angle determined only by ligand backbone
constraints and not by metal valence angles. The flexibility range is
defined as the accessible range of bite angles within less than 3 kcal
mol-1 excess strain energy from the calculated natural bite angle.
b
Tolman χi values were abstracted from ref 28. c Hammett σp values
were abstracted from ref 29.

hydroformylation by using a series of p-substituted triphen-


ylphosphines.25 Diphosphines of different basicities were studied
by Unruh and Christenson.26 Both studies showed that less basic
either a linear or branched four-coordinate L2Rh(CO)(alkyl). phosphines afford higher reaction rates and higher linear over
For the linear rhodium alkyl species this step is irreversible at branched (l:b) ratios. In a recent study Casey and co-workers
moderate temperatures and sufficiently high pressures of CO. investigated the electronic effect of equatorial and apical
The structure of the alkene complex is therefore thought to play phosphines.27 They concluded that electron-withdrawing sub-
a crucial role in controlling the regioselectivity.17,18 However, stituents on phosphines in the equatorial position give high l:b
contrary to the L2Rh(CO)2H precursor complex, the L2Rh- ratios, while electron-withdrawing substituents on phosphines
(CO)H(alkene) intermediate has not been observed directly so in the apical position give low l:b ratios. The results of this
far. Brown and Kent have shown that the PPh3-modified pre- study, however, cannot be described in terms of electronic
cursor complex (PPh3)2Rh(CO)2H exists as a mixture of two effects only, since there are major steric differences between
rapidly equilibrating trigonal bipyramidal isomers in a diequa- the ee and ea coordinating diphosphines.
torial (ee) to equatorial-apical (ea) isomer ratio of 85:15.19,20 To study the exact nature of the electronic effect in the rho-
dium diphosphine catalyzed hydroformylation, we synthesized
a series of thixantphos10 ligands with varying basicity (1-7,

The stereoelectronic properties of phosphorus ligands have


a dramatic influence on the reactivity of hydroformylation cata-
lysts. Tolman introduced the concept of cone angle θ and the Table 1) and probed their catalytic performance and coordination
electronic parameter χ as measures of respectively the steric chemistry in rhodium complexes. In this series of ligands, steric
bulk and the electronic properties of phosphorus ligands.21,22 differences are minimal so purely electronic effects can be
Casey and co-workers developed the concept of the natural bite investigated. Here we report that the equilibrium between ee
angle as an additional characteristic of diphosphine ligands based and ea coordination in the (diphosphine)Rh(CO)2H complexes
on molecular mechanics calculations.23 The pronounced effect is dependent on phosphine basicity. Decreasing phosphine
of the natural bite angle on the selectivity in the hydroformy- basicity shifts the equilibrium to ee coordination. In the
lation was shown by Casey and co-workers as well as by our hydroformylation of 1-octene and styrene decreasing phosphine
group.10,17 basicity gives an increase both in l:b ratio and in activity, while
While a rational understanding of how ligand structure can increasing phosphine basicity gives the opposite effect. Most
steer the selectivity in the hydroformylation is now unfolding, remarkably, for 1-octene, selectivity for linear aldehyde remains
only limited attention has been paid to the electronic proper- virtually the same, 92-93%, due to the enhanced isomerization
ties of phosphine ligands and their influence on catalytic activity displayed by the less basic phosphines. These results
performance.24 In one of the few detailed studies, Moser and show that the electronic properties of the diphosphine ligands
co-workers investigated the effect of phosphine basicity in the and their coordination mode in the (diphosphine)Rh(CO)2H
(17) Casey, C. P.; Whiteker, G. T.; Melville, M. G.; Petrovich, L. M.; complexes do not influence the regioselectivity. Mechanistic
Gavney, J. A., Jr.; Powell, D. R. J. Am. Chem. Soc. 1992, 114, 5535-5543. implications and possible explanations for the bite angle effect
(18) Casey, C. P.; Petrovich, L. M. J. Am. Chem. Soc. 1995, 117, 6007- are discussed.
6014.
(19) Brown, J. M.; Kent, A. G. J. Chem. Soc., Perkin Trans. 2 1987, (24) Oswald, A. A.; Hendriksen, D. E.; Kastrup, R. V.; Mozeleski, E. J.
1597-1607. AdV. Chem. Ser. 1992, 230, 395-418.
(20) Throughout this work, notations ee and ea refer to respectively the (25) Moser, W. R.; Papile, C. J.; Brannon, D. A.; Duwell, R. A.;
diequatorial and equatorial-apical chelation modes of the diphosphine in Weininger, S. J. J. Mol. Catal. 1987, 41, 271-292.
trigonal bipyramidal rhodium complexes. (26) Unruh, J. D.; Christenson, J. R. J. Mol. Catal. 1982, 14, 19-34.
(21) Tolman, C. A. Chem. ReV. 1977, 77, 313-348. (27) Casey, C. P.; Paulsen, E. L.; Beuttenmueller, E. W.; Proft, B. R.;
(22) Tolman, C. A. J. Am. Chem. Soc. 1970, 92, 2953-2956. Petrovich, L. M.; Matter, B. A.; Powell, D. R. J. Am. Chem. Soc. 1997,
(23) Casey, C. P.; Whiteker, G. T. Isr. J. Chem. 1990, 30, 299-304. 119, 11817-11825.
11618 J. Am. Chem. Soc., Vol. 120, No. 45, 1998 Van der Veen et al.

Scheme 2a

a
(i) n-BuLi/TMEDA/Et2O/25 °C; (ii) ZnCl2/0 °C; (iii) PCl3/-110
°C; (iv) ArMgBr/THF/0 °C.

Results
Synthesis of Thixantphos Ligands Containing Electron-
Donating and Electron-Withdrawing p-Substituents. The
thixantphos ligands were synthesized by reacting the Grignard
reagent of the appropriate p-substituted aryl bromide with 4,5-
bis(dichlorophosphino)-2,7-dimethylphenoxathiin (8) (Scheme
2). Compound 8 was prepared by selective dilithiation of 2,7-
dimethylphenoxathiin followed by reaction with zinc chloride
and reaction of the resulting dizinc compound with an excess
of phosphorus trichloride at low temperature.30 Typical yields
of the ligands varied from 25% for p-CH3O-thixantphos (2) to
54% for p-(CH3)2N-thixantphos (1) based on 8. Table 1 shows
that variation of the substituent has only a minor effect on the
natural bite angle and flexibility range and, therefore, steric
effects can be neglected when evaluating the electronic influence
of these ligands in catalysis.31
Rhodium Complexes. To investigate the influence of
phosphine basicity on chelation behavior, the solution structures
of the (diphosphine)Rh(CO)H(PPh3) complexes and the (diphos-
phine)Rh(H)(CO)2 complexes were studied in great detail for
all ligands. Displacement of PPh3 in (PPh3)3Rh(CO)H by the
diphosphines 1-7 resulted in formation of the (diphosphine)-
Figure 1. X-ray structure and numbering scheme for (thixantphos)-
Rh(CO)H(PPh3) complexes 9a-15a. Crystals of the (diphos-
Rh(CO)H(PPh3) (12a) (hydrogen atoms have been omitted for clarity)
phine)Rh(CO)H(PPh3) complexes suitable for X-ray structure and optimized geometry of (PH3)3Rh(CO)H (bond lengths are given
determination were obtained for ligands 2 and 4. in angstroms).
The X-ray crystal structures of (p-CH3O-thixantphos)Rh-
(CO)H(PPh3) (10a) and (Thixantphos)Rh(CO)H(PPh3) (12a) are Table 2. Selected Bond Lengths and Angles for
(Thixantphos)Rh(CO)H(PPh3) (12a) and
very similar so only the latter is shown in Figure 1. Selected (p-CH3O-Thixantphos)Rh(CO)H(PPh3) (10a)
bond lengths and angles of both complexes are given in Table
2. The crystal structures reveal distorted trigonal bipyramidal (thixantphos)Rh(CO)H(PPh3) (p-CH3O-thixantphos)Rh(CO)H(PPh3)
(12a) (10a)
complex geometries with all phosphines occupying equatorial
sites. In the structures the Rh atom is located slightly above Bond Lengths (Å)
the equatorial plane defined by the three phosphorus atoms and Rh-P(1) 2.327(1) Rh-P(1) 2.313(2)
displaced toward the apical carbonyl ligand. The hydride ligand Rh-P(2) 2.318(1) Rh-P(2) 2.328(2)
was located in neither structures, but 1H NMR indicates that it Rh-P(3) 2.295(2) Rh-P(3) 2.292(2)
Rh-C(1)O 1.887(8) Rh-C(61)O 1.863(8)
occupies the apical site trans to the carbonyl ligand (vide infra).
The P-Rh-P bite angles are 111.73(5)° for the thixantphos Bond Angles (deg)
P(1)-Rh-P(2) 111.73(5) P(1)-Rh-P(2) 109.31(7)
ligand and 109.31(7)° for the p-CH3O-thixantphos ligand. The P(1)-Rh-P(3) 121.73(5) P(1)-Rh-P(3) 126.65(8)
small difference in observed bite angles between p-CH3O- P(2)-Rh-P(3) 124.53(5) P(2)-Rh-P(3) 121.65(7)
thixantphos and the unsubstituted thixantphos confirms our P(1)-Rh-C(1) 94.2(2) P(1)-Rh-C(61) 94.4(2)
assumption based on molecular mechanics that p-substitution P(2)-Rh-C(1) 94.9(2) P(2)-Rh-C(61) 96.9(3)
on the aryl rings has only minor steric effects. A noticeable P(3)-Rh-C(1) 95.0(2) P(3)-Rh-C(61) 94.2(3)
difference between comparable bond lengths in the complexes
is found in the Rh-CO distance. The decreased Rh-CO bond Calculated Structure for the (PH3)3Rh(CO)H Complex.
length for the p-CH3O-thixantphos complex (1.863(8) Å) To support the spectroscopic results of the (diphosphine)Rh-
compared to the thixantphos complex (1.887(8) Å) is a direct (CO)H(PPh3) complexes (vide infra), calculations were per-
result of increased back-bonding of the rhodium to the carbonyl formed on the (PH3)3Rh(CO)H complex using density functional
ligand (vide infra). theory (DFT). The geometry of the (PH3)3Rh(CO)H complex
was optimized under the constraints of the C3V symmetry group
(28) Dias, P. B.; de Piedade, M.; Simoes, J. A. M. Coord. Chem. ReV.
1994, 135/136, 737-807.
and characterized as a minimum from the vibrational analysis.
(29) Hansch, C.; Leo, A.; Taft, R. W. Chem. ReV. 1991, 91, 165-195. The structure is depicted in Figure 1. The calculated Rh-P
(30) Private communication of Dr. H. T. Teunissen and Prof. Dr. F. and Rh-C bond lengths and the H-Rh-C angle are in good
Bickelhaupt. agreement with the distorted trigonal bipyramidal symmetry
(31) It should be noted that in the MM2 calculations the same
approximate parameters are used for phosphorus and rhodium, despite observed in the X-ray structures of complexes 10a and 12a.
electronic changes. Recently Schmid and co-workers also reported on the optimized
Rhodium Diphosphine Catalyzed Hydroformylation J. Am. Chem. Soc., Vol. 120, No. 45, 1998 11619

Table 3. Selected IR Data for (Diphosphine)Rh(CO)H(PPh3)


Complexes 9a-15aa
ligand R ν1, cm-1 ν2, cm-1
9a N(CH3)2 1982 (s) 1909 (m)
10a OCH3 1989 (s) 1916 (m)
11a CH3 1999 (s) 1920 (m)
12a H 1999 (s) 1922 (w)
13a F 2001 (s) 1921 (m)
14a Cl 2006 (s) 1928 (m)
15a CF3 2009 (s) 1931 (m)
a
In KBr, carbonyl region.

structure of the (PH3)3Rh(CO)H complex obtained by DFT


calculations.32 The geometry parameters they obtained at the
LDA level of theory are very similar to the values presented
here.
NMR Data for (Diphosphine)Rh(CO)H(PPh3) Complexes
9a-15a. The presence of hydride ligands in all (diphosphine)-
Rh(CO)H(PPh3) complexes is confirmed by 1H NMR. The
rhodium hydride signals appear as broadened quartets or triplets
of doublets at δ ) -9.09 to -9.60 (td, 2J(P,H) ) 10-21 Hz).
The rhodium-proton coupling constants could not be resolved.
The phosphorus-proton coupling constants are consistent with
hydrides occupying apical positions cis to three equatorial
Figure 2. Optimized geometries of the (PH3)2Rh(CO)2H and (PF3)2-
phosphines in distorted trigonal bipyramidal complex geometries
Rh(CO)2H complexes together with the relative energy of the ee and
(vide supra). Evaluation of the 31P{1H} NMR spectra leads to ea isomers. Bond lengths are given in angstroms.
the conclusion that for all complexes the two phosphine moieties
of the diphosphine in the complex are equivalent. The diphos- Table 4. Calculated Structural Data for ee and ea Isomers of
phine P-PPh3 coupling constants (2J(P,P) ) 123-126 Hz) are Complexes (PH3)2Rh(CO)2H and (PF3)2Rh(CO)2Ha
also in agreement with structures in which all phosphines occupy (PH3)2Rh(CO)2H (PF3)2Rh(CO)2H
equatorial sites. ee ea ee ea
IR Data for (Diphosphine)Rh(CO)H(PPh3) Complexes
9a-15a. The IR frequencies of the (diphosphine)Rh(CO)- Rh-H (Å) 1.617 1.589 1.607 1.597
Rh-Pe/a (Å) 2.343 2.361/2.364 2.244 2.255/2.266
H(PPh3) complexes in the carbonyl region display a clear Rh-Ce/a (Å) 1.906/1.931 1.912 1.940/1.957 1.939
electronic effect (Table 3). The IR spectra of the complexes in C-Oe/a (Å) 1.162/1.159 1.163 1.152/1.148 1.153
KBr all have two absorption bands between 1900 and 2100 cm-1 H-Rh-Pe/a 84.2 85.5/179.7 82.4 84.9/177.5
due to combination bands of νRhH and νCO. Both bands show (deg)
a regular shift to higher wavenumbers with decreasing phosphine H-Rh-COe/a 82.7/177.4 84.6 82.2/179.5 82.3
(deg)
basicity. The lower electron density on the metal leads to a P-Rh-P 110.1 94.8 112.5 97.6
decrease in back-bonding from the rhodium to the carbonyl (deg)
ligand, and hence, higher CO stretching frequencies are a The subscripts e and a denominate equatorial and apical ligand
observed. The experimental results are in good correlation with positions.
the results of the DFT calculations on the (PH3)3Rh(CO)H
complex. The calculated IR spectrum also exhibits two decrease in Rh-P bond lengths (0.1 Å) is observed with
harmonic frequencies: one at 1911 cm-1 corresponding to the decreasing phosphine basicity, accompanied by a slight increase
symmetric combination of the Rh-H and the Rh-CO vibrations in P-Rh-P angle. The increase in P-Rh-P angle is, however,
and one at 2007 cm-1 corresponding to the antisymmetric most likely not an electronic but a steric effect. Unlike the other
combination of the same vibrations. parameters the effect of phosphine basicity on the Rh-H bond
Calculated Structures for (PH3)2Rh(CO)2H and (PF3)2Rh- length differs for each isomer; for the ee isomer the Rh-H
(CO)2H Complexes. To model the basicity-dependent features distance decreases with decreasing phosphine basicity, while
exhibited by the (diphosphine)Rh(CO)2H complexes (vide infra), for the ea isomer the opposite occurs. Comparison of the
DFT calculations were performed on the (PH3)2Rh(CO)2H and P-Rh-P angles of the calculated complexes with the flexibility
(PF3)2Rh(CO)2H complexes. Similar calculations of the equi- ranges in Table 1 shows that both ee and ea isomers of the
librium structures of four- and five-coordinate rhodium hy- (diphosphine)Rh(CO)2H complexes could be accessible for the
drido carbonyl complexes using PH3 as model ligand have thixantphos ligands.
been reported recently by Schmid and co-workers.32 The ee and HP NMR Data for (Diphosphine)Rh(CO)2H Complexes
ea isomers of both complexes were fully optimized imposing 9b-15b. For all diphosphine ligands, the formation of the
Cs symmetry (Figure 2). Bond lengths and angles are listed (diphosphine)Rh(CO)2H complexes was evidenced by the
in Table 4. A clear dependence on phosphine basicity is appearance of a doublet in the 31P NMR spectra and an apparent
observed for almost all parameters. Upon changing the basic triplet of doublets for the rhodium hydride in the 1H NMR
PH3 ligand for the less basic PF3 ligand, C-O distances de- spectra. Table 5 reveals that decreasing phosphine basicity gives
crease while Rh-C distances increase as a direct result of the decreasing rhodium-proton (1J(Rh,H)) and phosphorus-proton
reduced back-bonding to the carbonyl ligands. A large (2J(P,H)) coupling constants but increasing rhodium-phospho-
rus (1J(Rh,P)) coupling constants.33
(32) Schmid, R.; Herrmann, W. A.; Frenking, G. Organometallics 1997,
16, 701-708. (33) 103Rh NMR data will be reported elsewhere.
11620 J. Am. Chem. Soc., Vol. 120, No. 45, 1998 Van der Veen et al.

Table 5. Selected NMR Data for (Diphosphine)Rh(CO)2H


Complexes 9b-15ba
ligand R 1J(Rh,H), Hz 1J(Rh,P), Hz 2J % eeb
av(P,H), Hz
9b N(CH3)2 8.7 121 27.9 44-50
10b OCH3 7.5 124 21.6 56-62
11b CH3 7.2 126 17.6 63-69
12b H 6.6 128 14.7 69-75
13b F 6.3 130 11.0 76-83
14b Cl 6.0 132 8.4 81-88
15b CF3 4.5 134 3.6 89-96
a
In C6D6 at 298 K. b Calculated from the 2Jav(P,H) using trans
J(P,H) ) 106 Hz and cis 2J(P,H) e (2 Hz.
2

The rhodium-proton and phosphorus-proton coupling con-


stants and their dependence on phosphine basicity suggest that
these complexes exist as mixtures of ee and ea isomers that are
in dynamic equilibrium.19 Rapid interconversion of the two
isomers, presumably via pseudo-Berry rotations, interchanges
the phosphorus atoms, and therefore, averaged chemical shifts
and coupling constants are observed. As demonstrated by
Yagupsky and Wilkinson and recently by Casey and co-workers
for related iridium complexes,27,34 the ratio of the two isomers
can be calculated from the averaged observed coupling constant
if all separate values are known. Unfortunately the dynamic
equilibria of the complexes could not be frozen out even at 163
K, so the coupling constants for the individual isomers could
not be determined. Estimations of the complex mixture
compositions at room temperature were achieved using a cis
phosphorus-proton coupling of -2 to 2 Hz, and a trans
phosphorus-proton coupling of 106 Hz.35 From the calculated
complex mixture compositions it can be concluded that the
equilibrium between the ee and ea isomer is greatly influenced
by phosphine basicity. For the strongly electron-donating
N(CH3)2 substituent, the ea isomer is slightly favored, while
for the strongly electron-withdrawing CF3 substituent, the
equilibrium is shifted almost completely to the ee isomer.
The enhanced preference of the basic phosphines for apical
coordination and the preference of the nonbasic phosphines for
equatorial coordination is quantitatively demonstrated by our
DFT calculations on the ee and ea isomers of complexes (PH3)2-
Rh(CO)2H and (PF3)2Rh(CO)2H (Figure 2). The energy dif- Figure 3. HP-IR spectra for (diphosphine)Rh(CO)2H complexes 9b-
ference between the ee and ea isomer of the complex modified 15b in 2-Me-THF at 80 °C and 20 bar of CO/H2 (1:1) (carbonyl
region).
with PH3 as model ligand for a basic phosphine was calculated
to be 0.1 kcal mol-1 in favor of the ea isomer. This small
energy difference would correspond to an ee:ea equilibrium equilibrium between ee and ea complex isomers was supplied
composition of approximately 1:1 at room temperature.36 A by HP-IR spectroscopy. The spectra of the (diphosphine)Rh-
slight preference of the ea isomer over the ee isomer for the (CO)2H complexes all showed four absorption bands in the
PH3 ligand has also been reported by Schmid and co-workers carbonyl region (Figure 3).
and by Matsubara and co-workers.32,37 With PF3 as a model In an effort to assign the bands to ee and ea isomers the
ligand for a nonbasic phosphine the ee isomer was found to be (thixantphos)Rh(CO)2D complex was measured for comparison.
favored by 2.4 kcal mol-1. This corresponds to an ee:ea Upon H/D exchange, only ν1 and ν3 shift to lower wavenumbers
equilibrium ratio of approximately 50:1 at room temperature. (respectively 18 and 14 cm-1), and therefore, these two bands
The calculated shift of the equilibrium upon changing the basic are assigned to the carbonyls of the ee complex.38 The two
PH3 ligand for the less basic PF3 ligand is consistent with the remaining bands that do not shift, ν2 and ν4, belong consequently
observed shift in isomer composition with decreasing phosphine to the ea complex. From the disappearance of a low-frequency
basicity in the series of substituted thixantphos ligands. shoulder upon H/D exchange, it can be concluded that one of
the rhodium hydride vibrations is partly hidden under ν4.
HP-IR Data for (Diphosphine)Rh(CO)2H Complexes 9b-
To verify the empirical peak assignments, the theoretical IR
15b. Unambiguous evidence for the existence of a dynamic
spectra of the (PH3)2Rh(CO)2H and -D and (PF3)2Rh(CO)2H
(34) Yagupsky, G.; Wilkinson, G. J. Chem. Soc. A 1969, 725-733. complexes were calculated by DFT methods. Recently several
(35) A trans phosphorus-proton coupling of 106 Hz was observed in reports have been published concerning the calculation of
1H NMR spectrum of the ea isomer of (DPEphos)Rh(CO)2H at 173 K
(unpublished results). (38) H/D exchange will only effect the ee isomer because in this isomer
(36) By the calculation of the complex mixture composition entropy terms there is a trans relationship between the hydride and a carbonyl ligand which
are neglected. results in coupling of the vibrations. The disappearance of resonance
(37) Matsubara, T.; Koga, N.; Ding, Y.; Musaev, D. G.; Morokuma, K. interaction upon H/D exchange will lead to frequency shifts of the carbonyl
Organometallics 1997, 16, 1065-1078. bands of the ee isomer.
Rhodium Diphosphine Catalyzed Hydroformylation J. Am. Chem. Soc., Vol. 120, No. 45, 1998 11621

Table 6. Calculated Rh-H, Rh-D, and C-O Stretching Harmonic Frequencies, IR Intensities, and Assignments for the ee and ea Isomers
of Complexes (PH3)2Rh(CO)2D, (PH3)2Rh(CO)2H, and (PF3)2Rh(CO)2Ha
(PH3)2Rh(CO)2D (PH3)2Rh(CO)2H (PF3)2Rh(CO)2H
frequency, intensity, frequency,b intensity, frequency, intensity,
cm-1 km/mol assignment cm-1 km/mol assignment cm-1 km/mol
ee isomer
2006 638 (C-O)e+(C-O)a 2041 (2087) 532 (Rh-H)-(C-O)e-(C-O)a 2099 423
1963 924 (C-O)e-(C-O)a 1983 (2043) 705 (Rh-H)+(C-O)e+(C-O)a 2036 555
1414 14 Rh-D 1938 (1961) 371 (Rh-H)-(C-O)e+(C-O)a 2004 279
ea isomer
2069 (2049) 138 Rh-H 2062 52
1985 477 (C-O)+(C-O) 1985 (2061) 478 (C-O)+(C-O) 2042 374
1948 1223 (C-O)-(C-O) 1948 (2028) 1224 (C-O)-(C-O) 2010 1035
1470 61 Rh-D
a
Assignments based on the normal modes analysis, the subscripts e and a denominate equatorial and apical ligand positions, and the signs
indicate equal (+) or different (-) phase between the vibration of each bond. b The values in parentheses were reported by Schmid and co-
workers.32

Table 7. Results of the Hydroformylation of 1-Octene at 80 °Ca


harmonic frequencies of carbonyl ligands in metal-carbonyl
complexes and we have used the same methodology as described ligand R σp l:b ratiob % selectb % isomerb tofb,c
in some of them.39-41 The calculated harmonic vibrational 1 N(CH3)2 -0.83 44.6 93.1 4.8 28
frequencies, intensities, and the corresponding assignments based 2 OCH3 -0.27 36.9 92.1 5.3 45
on the normal modes analysis of the carbonyl vibrations are 3 CH3 -0.17 44.4 93.2 4.7 78
depicted in Table 6. Comparison of the frequencies of (PH3)2- 4 H 0.00 50.0 93.2 4.9 110
5 F 0.06 51.5 92.5 5.7 75
Rh(CO)2H and -D shows that H/D exchange only effects the 6 Cl 0.23 67.5 91.7 6.9 66
carbonyl frequencies of the ee isomer of the complex. The 7 CF3 0.54 86.5 92.1 6.8 158
calculated shift to lower frequency of the vibrations of the ee a Conditions: CO/H ) 1, P(CO/H ) ) 20 bar, ligand/Rh ) 5,
2 2
isomer is in perfect agreement with the observed spectra and substrate/Rh ) 637, [Rh] ) 1.00 mM, number of experiments ) 3. In
confirms the empirical peak assignment to ee and ea isomers. none of the experiments was hydrogenation observed. b Linear over
The fit of the absolute values of the calculated frequencies of branched ratio, percent selectivity for linear aldehyde, percent isomer-
(PH3)2Rh(CO)2H with the measured values of the series ization to 2-octene, and turnover frequency were determined at 20%
(thixantphos)Rh(CO)2H complexes (Figure 3) is striking. Only alkene conversion. c Turnover frequency ) (mol of aldehyde) (mol of
Rh)-1 h-1.
the calculated value for ν3 (1983 cm-1) is out of the range of
ν3 values obtained experimentally (1960-1982 cm-1).
respective (diphosphine)Rh(CO)2H complexes are identical. The
The effect of decreasing phosphine basicity on the theoretical
ee:ea equilibria are not influenced, and no other carbonyl
IR spectra is an overall increase in frequency and a decrease in
complexes are observed. The observation that the (diphosphine)-
intensity of the vibrations. The only vibration that does not
Rh(CO)2H complexes are the only rhodium carbonyl species
experience a positive shift of approximately 60 cm-1 upon
present during hydroformylation implies that the rate-determin-
exchanging PH3 for PF3, is the Rh-H frequency of the ea
ing step is located early in the catalytic cycle. Either CO
isomer. Instead the decreased Rh-H bond length (vide supra)
dissociation or, more likely, alkene coordination will be rate
results in a shift of 7 cm-1 to lower frequency. The low
determining (Scheme 1).26 Evidence for rate-limiting alkene
calculated intensities for the Rh-H vibrations of the ea isomers
coordination can be found in increasing reaction rates with
explain the absence of the corresponding Rh-H bands in the
decreasing CO pressure, increasing alkene concentration, and
spectra of the different (thixantphos)Rh(CO)2H complexes
decreasing phosphine basicity (vide infra).
(Figure 3). The calculated frequencies and intensities of the ee
Hydroformylation of 1-Octene. Hydroformylation of
isomers are in good agreement with the assignment of low-
1-octene was carried out at 80 °C and 20 bar of 1:1 CO/H2
frequency shoulders of ν4 to Rh-H signals in the observed
using a 1.0 mM solution of rhodium diphosphine catalyst pre-
spectra (Figure 3).
pared from Rh(CO)2(dipivaloylmethanoate) (dpm) and 5 equiv
The shift in isomer composition that was found in the NMR of ligand. The production of octene isomers, nonanal, and
spectra is confirmed and clearly visualized when comparing the 2-methyloctanal was monitored by gas chromatography. The
IR spectra of the series of complexes (Figure 3). Going from results of the experiments with ligands 1-7 are shown in Table
electron-donating substituents to electron-withdrawing substit- 7. Turnover frequencies were determined at 20% conversion.
uents, the intensities of the absorption bands of the ea isomer
Table 7 shows that, with the exception of ligands 5 and 6,
are diminished in favor of those of the ee isomer. Another trend
the rate of the reaction increases with decreasing phosphine
observed in the series is a regular shift to higher wavenumbers
basicity. An explanation for the deviant behavior of 5 and 6
of all four absorption bands with decreasing phosphine basicity.
can be incomplete catalyst formation or deactivation of the
These shifts are comparable to the ones found for the (diphos-
catalyst.
phine)Rh(CO)H(PPh3) complexes.
The inverse dependency of the reaction rate on the phosphine
The IR spectra of the rhodium species present during actual
basicity is often found in the rhodium-catalyzed hydroformy-
hydroformylation experiments with ligands 2, 4, and 7 and the
lation.25,26 Decreasing phosphine basicity facilitates CO dis-
(39) Ehlers, A. W.; Frenking, G.; Baerends, E. J. Organometallics 1997, sociation from the (diphosphine)Rh(CO)2H complex and en-
16, 4896-4902. hances alkene coordination to form the (diphosphine)Rh(CO)-
(40) Ehlers, A. W.; Ruiz-Morales, Y.; Baerends, E. J.; Ziegler, T. Inorg. H(alkene) complex, and therefore, the reaction rate increases.
Chem. 1997, 36, 5031-5036.
(41) Barckholtz, T. A.; Bursten, B. E. J. Am. Chem. Soc. 1998, 120, Weakening of the rhodium-carbonyl bond in the rhodium
1926-1927. carbonyl complexes by electron-withdrawing substituents on the
11622 J. Am. Chem. Soc., Vol. 120, No. 45, 1998 Van der Veen et al.

Scheme 3 Table 8. Results of the Hydroformylation of Styrene at 120 °Ca


ligand R σp l:b ratiob % selectb tofb,c
1 N(CH3)2 -0.83 0.75 43 1100
2 OCH3 -0.27 0.85 46 1800
3 CH3 -0.17 1.1 53 2800
4 H 0.00 1.3 56 3100
5 F 0.06 1.3 56 2900
6 Cl 0.23 1.4 59 3200
7 CF3 0.54 1.8 64 3400
a
Conditions: CO/H2 ) 1, P(CO/H2) ) 10 bar, ligand/Rh ) 10,
substrate/Rh ) 1746, [Rh] ) 0.50 mM, number of experiments ) 3.
In none of the experiments was hydrogenation observed. b Linear over
branched ratio, percent selectivity for linear aldehyde, and turnover
frequency were determined at 20% alkene conversion. c Turnover
frequency ) (mol of aldehyde) (mol of Rh)-1 h-1.

ligand is evidenced by an increase in carbonyl stretching selectivity for linear aldehyde. Although this may seem
frequencies (vide supra). paradoxical at first, this is not in disagreement with the results
Surprisingly, no electronic effect of the ligands on the obtained in the hydroformylation of 1-octene. Of course styrene
selectivity for linear aldehyde is observed.42 The selectivities behaves rather different in the hydroformylation compared to
for linear aldehyde are all between 92 and 93%. The l:b ratio 1-octene due to the formation of a stable η3 complex. The
and the isomerization to 2-octene, however, both increase with reactions were not performed under identical conditions, but
decreasing phosphine basicity. The increase in l:b ratio with the results obtained with both substrates can still be ex-
decreasing phosphine basicity can be attributed completely to plained similarly. As found for 1-octene the increase in l:b ratio
an increased tendency of the branched alkyl rhodium species with decreasing phosphine basicity is probably caused by
to form 2-octene instead of branched aldehyde (Scheme 3). The enhanced β-hydrogen elimination. However, for styrene, this
increasing electrophilicity of the rhodium center leads to a higher does not result in isomerization since β-hydrogen elimination
reactivity of the rhodium alkyl species toward CO dissociation will always result in reformation of styrene. The only effect
and β-hydrogen elimination.26 The total amount of all other of enhanced β-hydrogen elimination is an increase in l:b ratio
products (branched aldehyde and 2-octene) is 7-8% for all and selectivity for linear aldehyde, as was already explained
ligands. by Lazaroni and co-workers for the deuterioformylation of
The constant selectivity for linear aldehyde in the hydro- styrene.44,45
formylation of 1-octene implies that for the basic ligands the
l:b ratio reflects the regioselectivity of the formation of the Discussion
rhodium alkyl species. For the less basic ligands the increase
in l:b ratio results from the different behavior of branched and Our results are in contradiction with the earlier studies of
linear rhodium alkyls toward β-hydrogen elimination, as was Moser and co-workers and Unruh and Christenson.25,26 This can
already reported by Lazaroni and co-workers for the deuterio- be explained by the fact that in their ligand systems a strict
formylation of 1-hexene.43 The linear alkyl is mainly converted separation of steric and electronic effects may not have been
to linear aldehyde, while the branched alkyl partially generates obtained. In the series of substituted thixantphos ligands steric
2-octene. Since 2-octene is far less reactive in the hydroformy- differences are minimal and only electronic effects are displayed.
lation, its formation is essentially irreversible when 1-octene is In the present study we find that there is no influence of
still present. We conclude that for our catalytic system the ratio phosphine basicity on linear aldehyde selectivity.
of linear to branched rhodium alkyl species formed is determined Casey and co-workers correlated the regioselectivity of the
not by phosphine basicity but by steric constraints only. Since hydroformylation both with the natural bite angle and the
the natural bite angles for all ligands are virtually the same, chelation mode of diphosphines in the five-coordinate hydride
this steric parameter seems to explain best the observed constant complexes.17,18 No plausible explanation, however, has yet been
selectivity for linear aldehyde in this series of ligands. found for the observed trends.27 From our results it can be
Hydroformylation of Styrene. The hydroformylation of concluded that the coordination mode in the five-coordinated
styrene was carried out at 120 °C and 10 bar of CO/H2 (1:1) hydrido complexes by itself is not crucial in the determination
using a 0.50 mM solution of rhodium diphosphine catalyst of the regioselectivity of the reaction. Regioselectivity cannot
prepared from Rh(CO)2(dpm) and 10 equiv of ligand. Styrene be correlated with the (predominant) chelation mode of the
is a substrate having a distinct preference for the formation of ligands 1-7 in the (diphosphine)Rh(CO)2H complexes, since
the branched aldehyde due to the stability of the benzylic the ratio of ee and ea isomers varies from almost 1 for ligand
rhodium species, induced by the formation of a stable η3 1 to more than 9 for ligand 7, while the selectivity for linear
complex.3 The formation of the linear aldehyde can be enhanced aldehyde remains the same for all ligands. Instead, the natural
by using high temperature and low pressure.10,44 The production bite angle (obtained from MM2) seems to display a higher
of 1-phenylpropionaldehyde and 2-phenylpropionaldehyde was correlation with the regioselectivity! Naturally, the observed
monitored by gas chromatography. The results of the experi- structures of the (diphosphine)Rh(CO)2H complexes need not
ments with the ligands 1-7 are shown in Table 8. Turnover be identical to the structures of the (diphosphine)Rh(CO)H-
frequencies were determined at 20% conversion. (alkene) complexes, which escape direct observation.
In the hydroformylation of styrene a regular increase in rate (43) Lazzaroni, R.; Uccello-Barretta, G.; Benetti, M. Organometallics
with decreasing phosphine basicity is observed. The increase 1989, 8, 2323-2327.
in activity is accompanied with an increase in l:b ratio and (44) Lazzaroni, R.; Raffaelli, A.; Settambolo, R.; Bertozzi, S.; Vitulli,
G. J. Mol. Catal. 1989, 50, 1-9.
(42) Selectivity for linear aldehyde is the percentage of 1-nonanal of all (45) Lazzaroni, R.; Settambolo, R.; Raffaelli, A.; Pucci, S.; Vitulli, G.
products including 2-octene. J. Organomet. Chem. 1988, 339, 357-365.
Rhodium Diphosphine Catalyzed Hydroformylation J. Am. Chem. Soc., Vol. 120, No. 45, 1998 11623

Scheme 4 that formation of linear C is favored over that of branched C.


Modeling studies are currently in progress to substantiate this
further.
The existence of a transient four-coordinate L2Rh(CO)H
species in the hydroformylation was already postulated by
Wilkinson and co-workers.7,16,46 Yoshida and co-workers have
shown that coordinatively unsaturated trans L2Rh(CO)H com-
plexes can be stabilized and isolated by the use of sterically
bulky phosphine ligands (L ) PCy3 or P-i-Pr3).47 The X-ray
crystal structure of (PCy3)2Rh(CO)H was later reported by
Freeman and Young.48 The crystal structure has a square-planar
complex geometry with the two phosphines in a trans relation
and a P-Rh-P angle of 164.1°. Morokuma and co-workers
have calculated the potential and free energy profile of the full
cycle of ethylene hydroformylation catalyzed by the (PH3)2Rh-
(CO)H complex using ab initio molecular orbital theory.37,49
The optimized structure of the intermediate (PH3)Rh(CO)2H
complex was found to be square planar, and the structures of
the transition states between the (PH3)Rh(CO)2H and (PH3)-
Rh(CO)2H(C2H4) complex were found to be square pyramidal.
Herrmann and co-workers have recently calculated the equi-
librium structures of both the square-planar four-coordinate
(PH3)Rh(CO)2H and (PH3)2Rh(CO)H complex using DFT
theory.32 The ideal geometry of the (PH3)2Rh(CO)H complex
was found to have the phosphine groups positioned in a trans
relationship and bent toward the hydride.
Evidence for the assumption that even a thixantphos ligand
with calculated flexibility range for rhodium of 92-124° can
accommodate a trans square-planar geometry can be found in
the crystal structure of the (p-(CH3)2N-thixantphos)Ni(CN)2
complex.50,51 In this distorted square-planar d8 complex
p-(CH3)2N-thixantphos spans two trans positions with a P-Ni-P
angle of 151.5°. In the proposed transition states B (Scheme
The fact that in our system the ee:ea isomer ratio does not 4), the thixantphos or BISBI ligands also span trans positions
influence regioselectivity suggests that CO dissociation from with P-Rh-P angles around 150-160° and the phosphorus
both isomeric (diphosphine)Rh(CO)2H complexes gives the atoms are bent toward the hydride. This complex geometry
same reactive four-coordinate (diphosphine)Rh(CO)H interme- induced by the ligand backbone and the bite angle controls the
diate (A) in the hydroformylation cycle (Scheme 4). In this orientation of the substituents on the phosphorus atoms, which
context it is important to note that the isomerization reaction on their turn determine the relative stabilities of the transition
of ee and ea isomers is very fast. We were unable to freeze states B and/or of the isomers C.
the fluxional behavior in the NMR spectrum even at 163 K, The assumption that four-coordinate (diphosphine)Rh(CO)H
which indicates that the energy barrier for the interconversion complexes play a key role in the hydroformylation not only
of the two isomers is very low, probably less than 10 kcal mol-1. explains the effect of the bite angle on the regioselectivity but
Following CO dissociation from the (diphosphine)Rh(CO)2H also explains the effect on the activity of the reaction. The
complexes, the regioselectivity in the hydroformylation will be natural bite angle influences the relative stabilities of the
determined by two consecutive steps of alkene coordination and transient four-coordinate complexes and thus the activity of these
hydride migration. Here we suggest that the regioselectivity is intermediates. A diphosphine with a natural bite angle close
controlled by the alkene attack on the four-coordinate intermedi- to 90° cannot form a trans square-planar complex but will give
ate A via a square-pyramidal transition state B (Scheme 4). an energetically less favorable cis complex.32 A diphosphine
Increasing the bite angle of the diphosphine in intermediate A that can accommodate bite angles up to 160°, such as BISBI,17
results in increased embracing of the rhodium center by the can lead to the formation of a trans square-planar complex of
ligand and consequently in more steric hindrance for the alkene lower relative energy (two monodentate phosphines will of
entering the coordination sphere. The hindrance is different course give the most stable trans complex due to the lack of
for the pro-linear and pro-branched transition states B. The bite angle constraints). A higher stability of the transient four-
preference for the formation of pro-linear over pro-branched B coordinate (diphosphine)Rh(CO)H intermediate can result in a
is not only determined by the magnitude of the bite angle in
(46) Yagupsky, G.; Brown, C. K.; Wilkinson, G. J. Chem. Soc. A 1970,
intermediate A. Although two PPh3 ligands will induce a larger 1392-1401.
P-Rh-P angle in intermediate A, selectivities for linear (47) Yoshida, T.; Okano, T.; Ueda, Y.; Otsuka, S. J. Am. Chem. Soc.
aldehyde are much higher for diphosphines such as thixantphos 1981, 103, 3411-3422.
(48) Freeman, M. A.; Young, D. A. Inorg. Chem. 1986, 25, 1556-1560.
and BISBI, that will probably give bite angles of only (150° (49) Koga, N.; Jin, S. Q.; Morokuma, K. J. Am. Chem. Soc. 1988, 110,
(vide infra). Inspection of the molecular models shows that 3417-3425.
the backbone of the diphosphine plays a key role in the (50) Goertz, W. Ph.D. Thesis, Rheinisch-Westfälischen Technischen
determination of the relative stabilities of linear and branched Hochschule Aachen, 1998.
(51) Goertz, W.; Keim, W.; Vogt, D.; Englert, U.; Boele, M. D. K.; van
isomers of complex C. The backbone constrains the orientation der Veen, L. A.; Kamer, P. C. J.; van Leeuwen, P. W. N. M. J. Chem.
of the phenyl substituents of the diphosphines in such a way Soc., Dalton Trans. 1998, 2981-2988.
11624 J. Am. Chem. Soc., Vol. 120, No. 45, 1998 Van der Veen et al.

higher concentration of the intermediate and/or in a lower sistently in energy calculations as well as in the geometry optimiza-
activation energy for its formation. If the decrease in activation tions.58 Geometries were fully optimized under the constrains of the
energy for CO dissociation is accompanied by only a small Cs or C3V symmetry group. Second derivatives were evaluated
increase in activation energy for alkene coordination to the four- numerically with a two point formula.
coordinate complex, then both the higher reactant concentration b. MM2 Calculations. The molecular mechanics calculations were
performed using the CAChe WorkSystem version 3.9,59 on an Apple
and the lower activation energy can explain the observed
Power Macintosh 950, equipped with two CAChe CXP coprocessors.
increase in hydroformylation activity with increasing natural Calculations were carried out similarly to the method described by
bite angle.10,17 Casey and Whiteker,23 using a Rh-P bond length of 2.315 Å.
Minimization’s were done via the block-diagonal Newton-Raphson
Conclusion method, allowing the structures to converge fully with a termination
criterion of a rms factor of 0.0001 kcal mol-1Å-1 or less.
A pronounced effect of phosphine basicity on the chelation
General Procedure. All reactions were carried out using standard
mode of the ligands in the (diphosphine)Rh(CO)2H complexes Schlenk techniques under an atmosphere of purified argon or nitrogen.
was observed. The ee:ea isomer ratio, observed in the IR and Toluene and TMEDA were distilled from sodium, THF and diethyl
1H NMR spectra, showed a regular increase with decreasing
ether from sodium/benzophenone, and hexanes from sodium/benzophe-
phosphine basicity. The spectroscopic results were supported none/triglym. Methanol, ethanol, 1-propanol, and dichloromethane
by theoretical calculations using DFT. The electronic ligand were distilled from CaH2. Chemicals were purchased from Acros
effect on catalysis is reflected in both hydroformylation rate Chimica and Aldrich Chemical Co. (PPh3)3Rh(CO)H,60 Rh(CO)2-
and rate of β-hydrogen elimination. Most remarkably, in the (dpm),61 and 2,7-dimethylphenoxathiin62 were prepared according to
hydroformylation of 1-octene, the overall selectivity for linear literature procedures. Silica gel 60 (230-400 mesh) purchased from
aldehyde is not influenced by phosphine basicity. Apparently, Merck was used for column chromatography. Melting points were
determined on a Gallenkamp MFB-595 melting point apparatus in open
the chelation mode in the (diphosphine)Rh(CO)2H complexes
capillaries and are uncorrected. NMR spectra were obtained on a
per se does not determine the regiochemistry in the hydro- Bruker AMX 300 or a DRX 300 spectrometer. 31P and 13C spectra
formylation reaction and the selectivity for linear aldehyde were measured 1H decoupled. TMS was used as a standard for 1H
formation correlates better to the calculated natural bite angle. and 13C NMR and H3PO4 for 31P NMR. Mass spectroscopy was
Up to now the role of the four-coordinate L2Rh(CO)H measured on a JEOL JMS-SX/SX102A. Elemental analyses were
complexes in the hydroformylation has been underestimated. carried out on an Elementar Vario EL apparatus. Infrared spectra were
In finding explanations for ligand effects on catalytic perfor- recorded on a Nicolet 510 FT-IR spectrophotometer. HP-IR spectra
mance, attention is focused mainly on the step that converts were measured using a 20-mL homemade stainless steel autoclave
the more stable five-coordinate L2Rh(CO)H(alkene) complexes equipped with mechanical stirring and ZnS windows. Hydroformylation
into four-coordinate L2Rh(CO)(alkyl) complexes. However, as reactions were carried out in a 200-mL homemade stainless steel
autoclave. Syn gas (CO/H2, 1:1, 99.9%) and CO (99.9%) were
stated by Casey and co-workers, “The regioselectiVity of
purchased from Air Liquide. D2 was purchased from Hoekloos. Gas
hydroformylation is goVerned by a complex web of electronic chromatographic analyses were run on an Interscience HR GC Mega
and steric effects that haVe so far defied unraVeling.” 27 Our 2 apparatus (split/splitless injector, J&W Scientific, DB1 30-m column,
results clearly indicate that an explanation for the effect of the film thickness 3.0 mm, carrier gas 70 kPa He, FID detector) equipped
bite angle on the regioselectivity and activity in the rhodium with a Hewlett-Packard Data system (Chrom-Card).
diphosphine catalyzed hydroformylation is not found in the 4,5-Bis(dichlorophosphino)-2,7-dimethylphenoxathiin (8).30 At 0
structure of the five-coordinate (diphosphine)Rh(CO)2H com- °C, 55.0 mL of n-butyllithium (2.5 M in hexanes, 137 mmol) was added
plexes. The role of the four-coordinate (diphosphine)Rh(CO)H dropwise to a stirred mixture of 12.5 g of 2,7-dimethylphenoxathiin
complexes in the reaction can lead to an alternative plausible (54.8 mmol) and 21.0 mL of TMEDA (137 mmol) in 250 mL of diethyl
explanation of the bite angle effect. ether. The reaction mixture was slowly warmed to room temperature
and stirred for 16 h. Then the dark yellow solution was cooled to 0
°C, and 74.0 mL of zinc chloride (1.85 M in diethyl ether, 137 mmol)
Experimental Section
was added dropwise. The reaction mixture decolorized, and a white
Computational Details. a. DFT Calculations. All DFT calcula- precipitate was formed. After being stirred for 2 h at room temperature,
tions were performed with the Amsterdam Density Functional (ADF) the reaction mixture was cooled to -196 °C using liquid nitrogen. Then
program.52-54 The valence molecular orbitals were expanded in a large 275 mL of phosphorus chloride (3.2 mol) was distilled onto the frozen
basis set of Slater’s type orbitals (STOs), whereas the frozen core reaction mixture. The reaction mixture was warmed to -110 °C, and
approximation was used to treat the core electrons. The frozen core mechanical stirring was started as soon as possible. The temperature
orbitals have been taken from atomic calculations in a very large basis was allowed to rise slowly, and at -50 °C, the colorless reaction mixture
set, and quasirelativistic corrections using the Pauli formalism were became yellow. At room temperature, the reaction mixture was
applied to obtain the frozen core shells. For rhodium, the atomic evaporated in vacuo and the resulting residue was extracted with diethyl
orbitals up to 3d shell were kept frozen and a triple-ζ basis set with ether. The solvent was removed in vacuo, and the obtained product
one p polarization function was used. For phosphorus, carbon, oxygen, was crystallized from hexanes. Yield: 14.3 g of light yellow crystals
and fluor, the triple-ζ basis set with one d polarization function was (61%). Mp: 124-126 °C. 1H NMR (CDCl3): δ ) 7.60 (bs, 2H, H1,8),
selected, keeping the 2p for phosphorus and the 1s for C, O, and F 7.08 (bs, 2H, H3,6), 2.36 (s, 6H, CH3). 31P{1H} NMR (CDCl3): δ )
frozen. The basis set used for hydrogen was also a triple-ζ with one 155.4. 13C{1H} NMR (C6D6): δ ) 149.7 (t, J(P,C) ) 28.8 Hz, OC),
p function for polarization. For the energy calculations the local
exchange-correlation potential (LDA) parametrized by Vosko, Wilk, (56) Becke, A. D. Phys. ReV. A 1988, 38, 3098.
(57) Perdew, J. P. Phys. ReV. B 1986, 33, 8822-8824.
and Nusair55 was used, in combination with the gradient-corrected (58) Fan, L.; Versluis, L.; Ziegler, T.; Baerends, E. J.; Ravenek, W. Int.
(GGA) Becke’s56 functional for exchange and Perdew’s57 functional J. Quantum Chem., Quantum Chem. Symp. 1988, S22, 173.
for correlation (BP86). The GGA approach was applied self-con- (59) CAChe Scientific Inc., 18700 N.W. Walker Road, Building 92-01,
Beaverton, OR 97006.
(52) Guerra, C. F.; Visser, O.; Snijders, J. G.; te Velde, G.; Baerends, (60) Ahmad, N.; Levison, J. J.; Robinson, S. D.; Uttley, M. F. Inorg.
E. J. In Methods and Techniques in Computational Chemistry, METECC- Synth. 1974, 15, 59-60.
95; Clementi, E., Corongiu, G., Eds.; Cagliari: Stef, 1995; p 307. (61) Coolen, H. K. A. C.; van Leeuwen, P. W. N. M.; Nolte, R. J. M. J.
Baerends, E. J.; Ros, P. Chem. Phys. 1973, 2, 41. Org. Chem. 1996, 61, 4739-4747.
(54) te Velde, G.; Baerends, E. J. J. Comput. Phys. 1992, 99, 84. (62) Suter, C. M.; McKenzie, J. P.; Maxwell, C. E. J. Am. Chem. Soc.
(55) Vosko, S. H.; Wilk, L.; Nusair, M. Can. J. Phys. 1980, 58, 1200. 1936, 58, 717-720.
Rhodium Diphosphine Catalyzed Hydroformylation J. Am. Chem. Soc., Vol. 120, No. 45, 1998 11625

135.43, 130.9, 129.0, 128.6 (t, J(P,C) ) 64.9 Hz, C4,5), 119.5. IR (KBr, ) 21.7 (d, 1J(Rh,P) ) 124 Hz). HP-IR (2-MeTHF, cm-1): 2034
cm-1): 2923 (w), 2416 (w), 1436 (m), 1428 (m), 1141 (s), 1104 (s), (RhCO), 1990 (RhCO), 1966 (RhCO), 1942 (RhCO), 1595 (CO-dpm).
987 (s). (p-CH3-Thixantphos)Rh(CO)H(PPh3) (11a). This compound was
Thixantphos Ligands 1-7. In a typical experiment, a solution of prepared similarly to 9a. Orange crystals were obtained from toluene/
3.52 g of 4-bromo-N,N-dimethylaniline (17.6 mmol) in 20 mL of THF ethanol. 1H NMR (C6D6): δ ) 7.73 (m, 10H, PPh3), 7.46 (bm, 4H,
was added dropwise to a stirred mixture of 0.64 g of magnesium PPh3), 7.21 (s, 2H, H1,8), 7.08 (m, 1H, PPh3), 6.98 (m, 8H, PCCH),
turnings (26.4 mmol), activated with 0.05 mL of 1,2-dibromoethane 6.81 (d, 3J(H,H) ) 8.2 Hz, 4H, CH3CCH), 6.75 (d, 3J(H,H) ) 7.9 Hz,
(0.6 mmol), in 5 mL of THF at room temperature. The reaction mixture 4H, CH3CCH), 6.69 (s, 2H, H3,6), 2.12 (s, 12H, CH3 tolyl), 2.04 (s,
was stirred for three h. The reaction mixture was filtered and added 12H, CH3 tolyl), 1.78 (s, 6H, CH3), -9.23 (dt, 2J(P,H) ) 20 Hz, 2J(P,H)
to a stirred solution of 1.50 g of 8 (3.52 mmol) in 15 mL of THF at 0 ) 12 Hz, RhH). 31P{1H} NMR (C6D6): δ ) 43.6 (dt, 1J(Rh,P) ) 167
°C. The reaction mixture was slowly warmed to room temperature Hz, 2J(P,P) ) 123 Hz; PPh3), 26.8 (dd, 1J(Rh,P) ) 148 Hz, 2J(P,P) )
and stirred for another 3 h. The beige reaction mixture was hydrolyzed 124 Hz). IR (KBr, cm-1): 1999 (s, HRhCO), 1920 (m, HRhCO).
with 5 mL of water, and the solvent was removed in vacuo. The (p-CH3-Thixantphos)Rh(CO)2H (11b). This compound was pre-
resulting brown residue was dissolved in CH2Cl2 and washed with dilute pared similarly to 9b. 1H HP-NMR (C6D6): δ ) 7.51 (bs, 8H, PCCH),
HCl. The organic layer was separated, and the aqueous layer was 6.68 (d, 3J(H,H) ) 8.1 Hz, CH3CCH), 6.56 (bs, 2H, H1,8), 6.42 (d,
extracted CH2Cl2. The combined organic layers were dried with 2
J(P,H) ) 5.1 Hz, H3,6), 1.93 (s, 12H, CH3 tolyl), 1.60 (s, 6H, CH3),
MgSO4, and the solvent removed in vacuo. The resulting white powder -8.52 (dt, 1J(Rh,H) ) 7.2 Hz, 2J(P,H) ) 17.6 Hz, 1H, RhH). 31P-
was washed with hexanes and recrystallized from toluene. Yield: 1.43 {1H} HP-NMR (C6D6): δ ) 22.7 (d, 1J(Rh,P) ) 126 Hz). HP-IR
g of pure p-(CH3)2N-thixantphos (1) as a white microcrystalline powder (2-MeTHF, cm-1): 2035 (RhCO), 1992 (RhCO), 1969 (RhCO), 1943
(53%). All thixantphos ligands 1-7 were fully characterized. Detailed (RhCO), 1595 (CO-dpm).
descriptions of the syntheses and characterizations are included in the (Thixantphos)Rh(CO)H(PPh3) (12a). This compound was pre-
Supporting Information. pared similarly to 9a and was already described by Kranenburg and
(p-(CH3)2N-Thixantphos)Rh(CO)H(PPh3) (9a). A solution of co-workers.10 Orange yellow crystals suitable for X-ray structure
(PPh3)3Rh(CO)H (92 mg, 0.10 mmol) and 1 (85 mg, 0.11 mmol) in 3 determination were obtained from toluene/ethanol.
mL of toluene was stirred for 0.5 h at 70 °C. The solvent was removed X-ray Crystal Structure Determination of 12a. 12a crystallized
in vacuo ,and the resulting yellow solid was washed with ethanol. 1H in the triclinic space group P1, a ) 10.9776(6) Å, b ) 11.0845(9)
NMR (C6D6): δ ) 7.77 (m, 10H, PPh3), 7.50 (m, 4H, PPh3), 7.12 (m, Å, c ) 11.730(3) Å, R ) 113.77(1)°, β ) 94.81(1)°, γ ) 108.435(5)°,
1H, PPh3), 6.99 (m, 10H, PCCH, H1,8), 6.81 (bs, 2H, H3,6), 6.29 (d, V ) 1201.3(4) Å3, and Z ) 1. The data collection was carried out
3J(H,H) ) 8.6 Hz, 4H, (CH ) NCCH), 6.24 (d, 3J(H,H) ) 8.7 Hz, 4H,
3 2 at room temperature. The structure was solved by direct methods.
(CH3)2NCCH), 2.51 (s, 12H, (CH3)2N), 2.43 (s, 12H, (CH3)2N), 1.78 The hydrogen atoms were calculated. The structure was refined to
(s, 6H, CH3), -9.09 (dt, 2J(P,H) ) 21 Hz, 2J(P,H) ) 11 Hz, 1H, RhH). R ) 0.030 and Rw ) 0.038, for 4931 observed reflections. Crystal
31P{1H} NMR (C D ): δ ) 43.4 (dt, 1J(Rh,P) ) 167 Hz, 2J(P,P) )
6 6 data and collection parameters, atomic coordinates, bond lengths,
125 Hz; PPh3), 24.7 (dd, 1J(Rh,P) ) 147 Hz, 2J(P,P) ) 125 Hz). IR bond angles, and thermal parameters are included in the Supporting
(KBr, cm-1): 1982 (s, HRhCO), 1909 (m, HRhCO). Information.
(p-(CH3)2N-Thixantphos)Rh(CO)2H (9b). A solution of Rh(CO)2- (Thixantphos)Rh(CO)2H (12b). This compound was prepared
(acac) (13 mg, 50 µmol) and 1 (39 mg, 50 µmol) in 1.0 mL of C6D6 similarly to 9b and was already described by Kranenburg and
was pressurized with 20 bar of CO/H2 (1:1) and left overnight at 80 co-workers.10
°C. 1H HP-NMR (C6D6): δ ) 7.61 (m, 8H, PCCH), 6.75 (s, 2H, H1,8), (Thixantphos)Rh(CO)2D (12-Db). This compound was prepared
6.63 (d, 2J(P,H) ) 6.0 Hz, H3,6), 6.22 (d, 3J(H,H) ) 8.1 Hz, (CH3)2- in situ from 12b and D2. HP-IR (2-MeTHF, cm-1): 2019 (RhCO),
NCCH), 2.39 (s, 24H, (CH3)2N), 1.65 (s, 3H, CH3), -8.25 (dt, 1J(Rh,H) 1994 (RhCO), 1958 (RhCO), 1947 (RhCO), 1587 (CO-dpm).
) 8.7 Hz, 2J(P,H) ) 27.9 Hz, 1H, RhH). 31P{1H} HP-NMR (C6D6): (p-F-Thixantphos)Rh(CO)H(PPh3) (13a). This compound was
δ ) 19.8 (d, 1J(Rh,P) ) 121 Hz). HP-IR (2-MeTHF, cm-1): 2027 prepared similarly to 9a. 1H NMR (C6D6): δ ) 7.54 (m, 10H, PPh3),
(RhCO), 1983 (RhCO), 1960 (RhCO), 1935 (RhCO), 1596 (CO-dpm). 7.28 (m, 4H, PPh3), 7.09 (m, 1H, PPh3), 6.96 (m, 8H, PCCH), 6.82
(p-CH3O-Thixantphos)Rh(CO)H(PPh3) (10a). This compound (bs, 2H, H1,8), 6.61 (m, 8H, FCCH), 6.40 (bs, 2H, H3,6), -9.49 (dt,
was prepared similarly to 9a. Orange yellow crystals suitable for X-ray 2J(P,H) ) 18 Hz, 2J(P,H) ) 11 Hz, 1H, RhH). 31P{1H} NMR (C D ):
6 6
structure determination were obtained from toluene/ethanol. 1H NMR δ ) 42.7 (dt, 1J(Rh,P) ) 167 Hz, 2J(P,P) ) 124 Hz, PPh3), 26.5 (dd,
(C6D6): δ ) 6.68 (m, 10H, PPh3), 7.41 (m, 4H, PPh3), 7.04 (m, 1H, 1J(Rh,P) ) 149 Hz, 2J(P,P) ) 124 Hz). IR (KBr, cm-1): 2002 (s,

PPh3), 6.93 (m, 8H, PCCH), 6.80 (d, 4J(H,H) ) 1.6 Hz, 2H, H1,8), HRhCO), 1921 (m, HRhCO). Anal. Calcd for C57H42F4O2P3RhS: C,
6.61 (bs, 2H, H3,6), 6.51 (d, 3J(H,H) ) 8.6 Hz, 4H, CH3OCCH), 6.47 64.41; H, 3.99; S, 3.02. Found: C, 64.88; H, 4.28; S, 2.68.
(d, 3J(H,H) ) 8.7 Hz, 4H, CH3OCCH), 3.32 (s, 6H, CH3O), 3.25 (s, (p-F-Thixantphos)Rh(CO)2H (13b). This compound was prepared
6H, CH3O), 1.77 (s, 6H, CH3), -9.33 (dt, 2J(P,H) ) 20 Hz, 2J(P,H) ) similarly to 9b. 1H HP-NMR (C6D6): δ ) 7.19 (bs, 8H, PCCH), 6.70-
11 Hz, 1H, RhH). 31P{1H} NMR (C6D6): δ ) 43.4 (dt, 1J(Rh,P) ) 6.36 (ar), 1.58 (s, 6H, CH3), -8.97 (dt, 1J(Rh,H) ) 6.3 Hz, 2J(P,H) )
163 Hz, 2J(P,P) ) 124 Hz; PPh3), 25.7 (dd, 1J(Rh,P) ) 148 Hz, 2J(P,P) 11.0 Hz, 1H, RhH). 31P{1H} HP-NMR (C6D6): δ ) 23.3 (d, 1J(Rh,P)
) 124 Hz). IR (KBr, cm-1): 1989 (s, HRhCO), 1916 (m, HRhCO). ) 130 Hz). HP-IR (2-MeTHF, cm-1): 2041 (RhCO), 1997 (RhCO),
Anal. Calcd for C61H54O6P3RhS: C, 65.95; H, 4.90; S, 2.89. Found: 1975 (RhCO), 1950 (RhCO), 1588 (CO-dpm).
C, 66.14; H, 5.03; S, 3.23. (p-Cl-Thixantphos)Rh(CO)H(PPh3) (14a). This compound was
X-ray Crystal Structure Determination of 10a. 10a crystallized prepared similarly to 9a. 1H NMR (C6D6): δ ) 7.47 (dd, 3J(H,H) )
in the triclinic space group P1, a ) 10.7820(7) Å, b ) 11.707(3) 11 Hz, 3J(H,H) ) 8.3 Hz, 6H, PPh3), 7.36 (m, 4H, PPh3), 7.17 (bs,
Å, c ) 23.322(2) Å, R ) 86.737(9)°, β ) 78.718(6)°, γ ) 68.24(1)°, 4H, PPh3), 7.02 (m, 1H, PPh3), 6.87 (m, 16H), 6.74 (d, 4J(H,H) ) 1.2
V ) 2682.1(9) Å3, and Z ) 2. The data collection was carried out Hz, 2H, H1,8), 6.36 (bs, 2H, H3,6), 1.69 (s, 6H, CH3), -9.60 (q, 2J(P,H)
at room temperature. The structure was solved by direct methods. ) 14 Hz, 1H, RhH). 31P{1H} NMR (C6D6): δ ) 42.8 (dt, 1J(Rh,P) )
The hydrogen atoms were calculated. The structure was refined to R 165 Hz, 2J(P,P) ) 123 Hz; PPh3), 26.6 (dd, 1J(Rh,P) ) 149 Hz, 2J(P,P)
) 0.067 and Rw ) 0.085, for 9235 observed reflections. Crystal ) 123 Hz). IR (KBr, cm-1): 2006 (s, HRhCO), 1928 (m, HRhCO).
data and collection parameters, atomic coordinates, bond lengths, Anal. Calcd for C57H42Cl4O2P3RhS: C, 60.66; H, 3.75; S, 2.84.
bond angles, and thermal parameters are included in the Supporting Found: C, 61.14; H, 4.13; S, 2.78.
Information. (p-Cl-Thixantphos)Rh(CO)2H (14b). This compound was prepared
(p-CH3O-Thixantphos)Rh(CO)2H (10b). This compound was similarly to 9b. 1H HP-NMR (C6D6): δ ) 7.25-6.95 (ar), 6.83-6.68
prepared similarly to 9b. 1H HP-NMR (C6D6): δ ) 7.46 (bs, 8H, (ar), 1.59 (s, 6H, CH3), -9.04 (dt, 1J(Rh,H) ) 6.0 Hz, 2J(P,H) ) 8.4
PCCH), 6.73 (s, 2H, H1,8), 6.43 (bs, 8H, CH3OCCH), 6.30 (bs, 2H, Hz, 1H, RhH). 31P{1H} HP-NMR (C6D6): δ ) 23.7 (d, 1J(Rh,P) )
H3,6), 3.20 (s, 12H, CH3O), 1.64 (s, 6H, CH3), -8.25 (dt, 1J(Rh,H) ) 132 Hz). HP-IR (2-MeTHF, cm-1): 2042 (RhCO), 1999 (RhCO), 1977
7.5 Hz, 2J(P,H) ) 21.6 Hz, 1H, RhH). 31P{1H} HP-NMR (C6D6): δ (RhCO), 1952 (RhCO), 1574 (CO-dpm).
11626 J. Am. Chem. Soc., Vol. 120, No. 45, 1998 Van der Veen et al.

(p-CF3-Thixantphos)Rh(CO)H(PPh3) (15a). This compound was HP-FT-IR Experiments. In a typical experiment the HP-IR
prepared similarly to 9a. Orange crystals were obtained from toluene/ autoclave was filled with 2-5 equiv of ligand, 5 mg of Rh(CO)2(dpm),
ethanol. 1H NMR (C6D6): δ ) 7.43 (m, 10H, PPh3), 7.19 (m, 4H, and 15 mL of 2-MeTHF. The autoclave was purged three times with
PPh3), 7.02 (t, J ) 7.2 Hz, 8H, PCCH), 6.94 (m, 1H, PPh3), 6.85 (m, 15 bar of CO/H2 (1:1), pressurized to approximately 18 bar, and heated
8H, CF3CCH), 6.75 (d, 4J(H,H) ) 1.6 Hz, 2H, H1,8), 6.22 (bs, 2H, to 80 °C. Catalyst formation was followed in time by FT-IR and was
H3,6), 1.66 (s, 6H, CH3), -9.43 (q, 2J(P,H) ) 14 Hz, 1H, RhH). 31P- usually complete within 1 h.63
{1H} NMR (C6D6): δ ) 41.8 (dt, 1J(Rh,P) ) 164 Hz, 2J(P,P) ) 122 HP-NMR Experiments. In a typical experiment 0.4 mL of a
Hz, PPh3), 27.8 (dd, 1J(Rh,P) ) 149 Hz, 2J(P,P) ) 122 Hz). IR (KBr, solution of 50 µmol of Rh(CO)2(acac) and 50 µmol of ligand in 1.0
cm-1): 2009 (s, HRhCO), 1931 (m, HRhCO). Anal. Calcd for mL of C6D6 was transferred into a 0.5-cm sapphire NMR tube. The
C61H42F12O2P3RhS: C, 58.02; H, 3.35; S, 2.54. Found: C, 57.92; H, tube was pressurized with 20 bar of CO/H2 (1:1) and left overnight at
3.47; S, 2.78. 80 °C. For VT experiments half the amounts of rhodium and ligand
(p-CF3-Thixantphos)Rh(CO)2H (15b). This compound was pre- were used in 0.5 mL of THF-d8.
pared similarly to 9b. 1H HP-NMR (C6D6): δ ) 7.23 (bs, 8H, PCCH),
7.00 (d, 3J(H,H) ) 8.1 Hz, 8H, CF3CCH), 6.85 (d, 2J(P,H) ) 6.0 Hz, Acknowledgment. Financial support from SON/STW is
2H, H3,6), 6.71 (s, 2H, H1,8), 1.59 (s, 6H, CH3), -9.07 (dt, 1J(Rh,H) ) gratefully acknowledged, and we thank Dr. H. T. Teunissen
4.5 Hz, 2J(P,H) ) 3.6 Hz, 1H, RhH). 31P{1H} HP-NMR (C6D6): δ ) and Prof. Dr. F. Bickelhaupt for providing the experimental
24.7 (d, 1J(Rh,P) ) 134 Hz). HP-IR (2-MeTHF, cm-1): 2046 (RhCO), procedure for the synthesis of compound 8. DFT calculations
2004 (RhCO), 1982 (RhCO), 1957 (RhCO), 1607 (CO-dpm). were carried out on workstations purchased from funds provided
Hydroformylation Experiments. Hydroformylation reactions were by the Spanish Dirección General de Investigación Cientı́fica
carried out in an autoclave, equipped with glass inner beaker, a substrate y Técnica (DGICYT) under Project PB95-0639-C02-02 and
inlet vessel, a liquid sampling valve, and a magnetic stirring rod. The
by the CIRIT of the Generalitat de Catalunya under Project
temperature was controlled by an electronic heating mantle. In a typical
experiment, the desired amount of ligand was placed in the autoclave SGR97-17.
and the system was evacuated and heated to 50 °C. After 0.5 h, the
Supporting Information Available: Detailed description
autoclave was filled with CO/H2 (1:1) and a solution of Rh(CO)2(dpm)
(5 or 10 µmol) in 8.5 mL of toluene. The autoclave was pressurized
of the syntheses and characterizations of thixantphos ligands
to the appropriate pressure (6 or 16 bar), heated to reaction temperature, 1-7, HP-IR spectrum of 12-Db (carbonyl region), and tables
and stirred for 0.5 h (120 °C) or 1.5 h (80 °C) to form the active catalyst. of crystal data and collection parameters, atomic coordinates,
Then 1.0 mL of substrate (filtered over neutral activated alumina to bond lengths, bond angles, thermal parameters, and H-atom
remove peroxide impurities) and 0.5 mL of the internal standard coordinates for 10a and 12a (39 pages, print/PDF). See any
n-decane were placed in the substrate vessel, purged three times with current masthead page for ordering information and Web access
10 bar CO/H2 (1:1), and pressed into the autoclave with 10 or 20 bar instructions.
CO/H2 (1:1). At the 20% conversion level the reaction was stopped
by adding 0.25 mL of tri-n-butyl phosphite and cooling on ice. Samples JA981969E
of the reaction mixture were analyzed by temperature-controlled gas (63) Castellanos-Páez, A.; Castillón, S.; Claver, C.; van Leeuwen, P. W.
chromatography. N. M.; de Lange, W. G. J. Organometallics 1998, 17, 2543-2552.

You might also like