pmc logo imageJournal ListSearchpmc logo image
Logo of pnasPNAS Home page.Reference to the article.PNAS Info for AuthorsPNAS SubscriptionsPNAS About
Proc Natl Acad Sci U S A. 2007 April 24; 104(17): 6927–6932.
Published online 2007 April 13. doi: 10.1073/pnas.0610212104.
PMCID: PMC1855363
Research Articles, Chemistry
Coordination Chemistry of Saturated Molecules Special Feature
A delicate balance of complexation vs. activation of alkanes interacting with [Re(Cp)(CO)(PF3)] studied with NMR and time-resolved IR spectroscopy
Graham E. Ball, Christopher M. Brookes,§ Alexander J. Cowan,§ Tamim A. Darwish, Michael W. George,§ Hajime K. Kawanami,§ Peter Portius,§ and Jonathan P. Rourke
School of Chemistry, University of New South Wales, Sydney 2052, Australia;
§School of Chemistry, University of Nottingham, University Park, Nottingham NG7 2RD, United Kingdom; and
Department of Chemistry, University of Warwick, Coventry CV4 7AL, United Kingdom
To whom correspondence may be addressed. E-mail: g.ball/at/unsw.edu.au or Email: mike.george/at/nottingham.ac.uk
Edited by John E. Bercaw, California Institute of Technology, Pasadena, CA, and approved March 7, 2007
Author contributions: G.E.B., M.W.G., and J.P.R. designed research; G.E.B., C.M.B., A.J.C., T.A.D., M.W.G., H.K.K., P.P., and J.P.R. performed research; and G.E.B., A.J.C., T.A.D., and M.W.G. wrote the paper.
Received November 16, 2006.
Abstract
The organometallic alkane complexes Re(Cp)(CO)(PF3)(alkane) and Re(Cp)(CO)2(alkane) have been detected after the photolysis of Re(Cp)(CO)2(PF3) in alkane solvent. NMR and time-resolved IR experiments reveal that the species produced by the interaction of n-pentane with [Re(Cp)(CO)(PF3)] are an equilibrium mixture of Re(Cp)(CO)(PF3)(pentane) and Re(Cp)(CO)(PF3)(pentyl)H. The interaction of cyclopentane with [Re(Cp)(CO)(PF3)] most likely results in a similar equilibrium between cyclopentyl hydride and cyclopentane complexes. An increasing proportion of alkane complex is observed on going from n-pentane to cyclopentane to cyclohexane, where only a small amount, if any, of the cyclohexyl hydride form is present. In general, when [Re(Cp)(CO)(PF3)] reacts with alkanes, the products display a higher degree of oxidative cleavage in comparison with [Re(Cp)(CO)2], which favors alkane complexation without activation. Species with the formula Re(Cp)(CO)(PF3)(alkane) have higher thermal stability and lower reactivity toward CO than the analogous Re(Cp)(CO)2(alkane) complexes.
Keywords: alkane complexes, CH activation, multinuclear NMR, photochemistry
 
Intuitively, simple alkane molecules are among the most unlikely of ligands in coordination chemistry. There are no lone pairs or π electrons available for binding to the metal center. Only strong, single σ bonds are present, and it is through the coordination of C-H bonds to a metal center that complexation is most likely to occur (1). Complexes containing an intact σ bond of a covalent species X-H coordinated to a metal center are termed σ-complexes, and there has been widespread activity in this field recently (2, 3).

The prototypical σ complexes are the dihydrogen (H2) complexes first described by Kubas, which contain an H-H unit coordinated to a metal (2, 4). In the case of alkanes, the interaction of a C-H bond with a metal is particularly weak, largely because σ-complexation is often stabilized by a backbonding component that is highly inefficient for a C-H bond when compared with a H2 ligand. An understanding of this mode of bonding is interesting and important both from a theoretical point of view and to explain the role they play in the reactions of the C-H moiety.

The C-H activation reaction is a key step in potential functionalization of the normally inert alkane molecule, and its integration into catalytic processes is an important goal (5, 6). The C-H activation of alkanes has been shown to involve alkane complexes as intermediates in a number of elegant IR studies (57). After coordination, the C-H bond breaking step is usually fast and irreversible, precluding observation of the alkane complex with a relatively “slow” technique such as NMR spectroscopy. However, alkane complexes are not always unstable with respect to C-H activation. Two important examples that are formed photolytically are the long known M(CO)5(alkane) (M = Cr, Mo, W) systems (1, 8) and Re(Cp)(CO)2(n-heptane). This latter complex, characterized by time-resolved IR (TRIR) spectroscopy, had by far the longest lifetime measured (≈25 ms at 298 K) of any alkane complex in solution at the time (9). This crucial result paved the way for NMR studies of Re(Cp)(CO)2(alkane) (1012) and further IR studies (13), because at ≈183 K, the lifetime of these complexes (≈1 h) is sufficient for NMR experiments. The formation of long lived Re(Cp)(CO)2(alkane) complexes has recently prompted the study of the photolysis of Re(Cp′)(CO)3 (Cp′ = Cp, Cp*) in supercritical methane (scCH4) and liquid ethane. After photolysis in scCH4, a rapid equilibrium between Re(Cp′)(CO)2(CH4) and Re(Cp′)(CO)2(CH3)H is observed (14).

There are also many examples of what may be considered model complexes for the coordination of alkanes, namely agostic alkyl complexes (other than α-agostic species) (15, 16). In these compounds, the agostic η2-C,H interaction is stabilized by the chelate effect, because these ligands are tethered to the metal elsewhere.

Using combined IR, NMR, and theoretical studies, we have recently thoroughly characterized an organometallic xenon complex Re(iPrCp)(CO)(PF3)Xe (17, 18). It was noted that the stability of this complex was slightly higher than that of the closely related Re(iPrCp)(CO)2Xe species, formed in the same reaction mixture. Hence, the question was whether the introduction of a PF3 ligand would stabilize alkane complexation in species of the type Re(Cp)(CO)(PF3)(alkane) compared with Re(Cp)(CO)2(alkane). Alternatively, would the [Re(Cp)(CO)(PF3)] fragment display reactivity patterns more like those of [Re(Cp*)(CO)(PMe3)] (19) or [Re(Cp)(PMe3)2] (20) in which the presence of the phosphorus donor is known to promote activation of the C-H bond to form alkyl hydride species?

Our studies of the interaction of the [Re(Cp)(CO)(PF3)] fragment with three alkanes cyclopentane, cyclohexane, and pentane described here suggest behavior that is “in between” these two extremes of alkane complex vs. alkyl hydride, and this behavior is surprisingly dependent on the alkane.

Results and Discussion

NMR Studies.

Photolysis in cyclopentane. A solution of Re(Cp)(CO)2(PF3) (1) in 95% cyclopentane:5% pentane-d12 (added for locking purposes) was cooled to 185 K. Before photolysis, there was a single peak in the 1H NMR spectrum at δ 5.20 due to the Cp protons of 1, and one doublet in the 19F NMR spectrum at δ −2.56 (d, 1JFP = 1,232 Hz). The sample was irradiated with UV light for 1 min resulting in the depletion (≈15%) of the 19F resonance of 1 and the concomitant growth of two new resonances at δ −3.28 (dt, 1JPF = 1,211 Hz, 2 × JFH = 7.9 Hz) labeled species A and at δ −32.4 (d, 1JPF = 1,398 Hz), due to free PF3 (Fig. 1).

Fig. 1.Fig. 1.
282.4-MHz 19F NMR spectra of 1 in cyclopentane at 185 K. (Lower) Before photolysis. (Upper) After 1 min of photolysis. (Inset) Expansion of the highlighted half of resonance A.

This observation is consistent with previous experiments (17) where photolysis of 1 led to the loss of either a CO or a PF3 ligand to form [Re(Cp)(CO)(PF3)] and [Re(Cp)(CO)2]. These fragments can interact with the alkane solvent as outlined in Fig. 2. Resonance A can be assigned to either a cyclopentane (3a) or a cyclopentyl hydride (4a) complex or to a rapidly equilibrating mixture of both. The triplet splitting of resonance A, due to JFH, suggests a coupling to two effectively equivalent hydrogens.

Fig. 2.Fig. 2.
Possible photoproducts obtained from photolysis of 1 in alkanes.

1H NMR spectra showed a corresponding decrease in the signals of 1 upon irradiation. A quintet grows in at δ −2.29 in the 1H spectrum that is due to the bound CH2 unit of the cyclopentane ligand in the previously characterized Re(Cp)(CO)2(cyclopentane) (2a) (10), and a singlet at δ 4.92 is due to the Cp ligand of 2a. The alkane in complexes 2a–2c is known not to undergo oxidative cleavage to form alkyl hydrides under these conditions (1012).

In addition to the resonances due to 2a in the 1H spectrum, another peak due to a Cp ligand of a species with a higher concentration was also observed to grow in at δ 5.21. Intensities suggested that this peak was likely from the same species that produces the 19F resonance A, i.e., 3a/4a. Further photolysis (1 min) increased the concentration of all of the newly observed 19F and 1H signals. Significantly, there was no evidence for any shielded 1H resonance (δ < 0) expected for 3a/4a. As a result, the temperature of the NMR probe was raised to 215 K in 10-K steps in the absence of photolysis. As the temperature was increased, the 1H NMR spectra showed the emergence of a broad resonance A′ at δ −2.92, which sharpened and increased in intensity at higher temperature. A′ was assigned to the bound CH2 protons of cyclopentane in Re(Cp)(CO)(PF3)(cyclopentane) (3a), and resonance A is also assigned to 3a (Fig. 3). The intensity of signal A in the 19F NMR spectra was not temperature-dependent.

Fig. 3.Fig. 3.
Variable-temperature, 300-MHz 1H NMR spectra after photolysis of 1 in cyclopentane.

The apparent increase in concentration of 3a at higher temperature is an artifact of the selective excitation scheme used to suppress the resonances of the free alkane solvent. The selective excitation can cause broad peaks to be attenuated or even “lost” with the broadest peaks suffering the most severe attenuation, as was observed at 185 K. Temperature cycling between 215 and 185 K confirms that this is occurring as A′ disappears again at lower temperature. The cause of the broadening of the A′ resonance is the onset of a decoalescence phenomenon that is starting to occur, leading to increasing linewidths at lower temperatures. For the two hydrogens in a bound CH2 unit of 3a to be equivalent, it is necessary for them to average their environments through a combination of both (i) an exchange of complexed and uncomplexed hydrogens as shown in Fig. 4 and (ii) rotation of the cyclopentane moiety around the metal-ligand bond axis. A slowing of either the rate of exchange or rotation could lead to the observed decoalescence.

Fig. 4.Fig. 4.
Exchange of complexed C-H bonds in a bound CH2 unit.

The resonances associated with 2a started to disappear as the temperature increased above 185 K, due to the documented thermal decomposition of 2a (10). At 215 K, signals due to 3a (including A, A′) also decreased with time, due to the thermal decomposition of 3a. Clearly 3a is thermally more stable than 2a. Resonance A′ at 215 K shows a splitting attributed to a JPH coupling, confirmed by a 31P decoupling experiment. The 19F and 1H resonances of 3a were connected to the same species by using a pair of shift correlation experiments, 2D 1H-31P and 19F-31P. Both resonances A (19F) and A′ (1H) were shown to correlate to the same 31P frequency at δ 124.2, confirming that they are from the same species. The 19F-31P correlation experiment shows the 1H couplings in both the 19F and 31P dimensions seen earlier and reveals that JPH ≈ 26 Hz [see the supporting information (SI)].

Key indicators of the presence of an alkane complex from an NMR perspective are a bound CH2 or CH3 unit in which δ 1H is moderately shielded and, most importantly, a 1JCH value that is slightly reduced from that observed in a free alkane. For example, the protons of the bound CH2 group in 2a appear at δ −2.3 with 1JCH = 113 Hz (cf. cyclopentane δ 1.50; 1JCH 129.4 Hz) (10). In all of the alkane complexes observed in NMR studies to date, the exchange of the complexed and uncomplexed C-H bond has always been fast on the NMR time scale (Fig. 4). This means that the observed δ 1H and 1JCH values of the protons on the bound carbon are an average of the values for complexed and uncomplexed hydrogens. Density functional theory calculations suggest some deshielding of the 1H resonance and an increase in 1JCH of the uncomplexed hydrogen(s) in Re(Cp)(CO)2(cyclohexane) (12) compared with the free alkane. Therefore, the true shielding and reduction of 1JCH for the complexed hydrogen may be just over twice the observed changes in the averaged 1H NMR shift and 1JCH for a CH2 group compared with the free alkane. The trends observed in both agostic alkyl and dihydrogen complexes suggest that the more stretched the complexed C-H moiety becomes, the smaller 1JCH will become. Typically the 1JCH of a complexed C-H bond may be reduced by up to 50% relative to the uncomplexed unit (free, ≈125 Hz; complexed, 60–120 Hz) based on data from agostic species [although cases of 1JCH < 40 Hz are known (15)]. In an exchange averaged situation, one expects 1JCH in a bound CH2 of an alkane complex to be ≈90–128 Hz.

In the case where complete oxidative cleavage occurs, the complexed C-H bond is broken, and now the comparable coupling is a 2JCH interaction, which is expected to be small, typically <10 Hz, and likely of opposite sign to 1JCH. Examples of complexes that undergo a rapid exchange of alkyl and hydride protons, likely via an alkane complex intermediate, are known, e.g., [Cp*Os(dmpm)(CH3)H]+ (21).

It may not be possible to differentiate complexed alkanes from alkyl hydride species (4) that undergo rapid exchange of hydride and alkyl protons based on the δ 1H or δ 13C values alone, because they may be similar in these two scenarios. However, the averaging of a large 1JCH with a small (negative) 2JCH in the alkyl hydride case would be expected to produce an averaged JCH value of ≈55–70 Hz, significantly less than that observed for an alkane complex.

To gain insight into the nature of the interaction of cyclopentane in 3a, photolysis experiments in partially 13C-labeled cyclopentane (25% cyclopentane-13C1, 15% cyclopentane-d10, 60% normal cyclopentane, equivalent to ≈7% 13C total) were carried out. A 13C edited 1D 1H-{31P} experiment, which selects only the resonances of hydrogens directly attached to a 13C nucleus with simultaneous 31P decoupling, was used at 210 K, revealing JCH = 75 ± 2 Hz. A 2D 1H-13C correlated experiment provided δ 13C = −6.80 for the bound carbon of the cyclopentane of 3a compared with δ 13C = −31.2 for 2a. δ 13C of 3a is also higher than those reported for trans-Re(Cp)(PMe3)2(R)H (R = n-C6H13, c-C3H5, CH3), which vary from δ −14 to δ −39 (20). In lieu of more data, values of δ 13C of the carbon in the coordinated CH2 unit (Cα) are not a priori able to differentiate structures in this case.

The observed value of 1JCH ≈75 Hz for the interacting cyclopentane in 3a is significantly less than that reported for the complexed cyclopentane in 2a (112.9 Hz). It is also larger than the maximum value predicted for an averaged coupling in a cyclopentyl hydride complex 4a (≈70 Hz). This finding suggests at least three possibilities for the mode of coordination of the alkane in 3a. (i) The cyclopentane unit is coordinating to the metal center in a σ-alkane complex fashion with a highly stretched complexed C-H bond, significantly greater than the 1.15–1.17 Å calculated for the complexed C-H distance of Re(Cp)(CO)2(cyclohexane) (2c) (12). (ii) The complexed C-H bond of the coordinated cyclopentane has undergone oxidative cleavage to form the cyclopentyl hydride 4a, but the 1JCH value for the Cα-H is unusually large. (iii) There is rapid exchange between the σ-alkane and its corresponding alkyl hydride complex, and there is a sizeable amount of each component (3a and 4a) in equilibrium. In this latter case, the observed JCH would be the weighted average from all of the contributing structures in Fig. 5; i.e., 1JCH of an alkyl group, 1JCH of an uncomplexed C-H in a σ-alkane, 1JCH of a complexed (and possibly stretched) C-H in a σ-alkane, and 2JCH of a hydride. As an example, if the averaged 1JCH value in 3a were the same as 2a, 113 Hz, and if a best estimate averaged value of 1JCH for 4a, ≈61 Hz, were present, an equilibrium with a ratio of 3a:4a of 27:73 would be required to produce the observed value of 75 Hz. If more stretched C-H bonds in the alkane complex component are present, the relative amount of 3a would be increased. A similar approach may be applied to the use of JPH as well. The averaged value of JPH in the case of a cis-alkyl hydride 4a is expected to be much larger (estimate ≈33 Hz) than for the alkane complex 3a, which is estimated to be <10 Hz but increasing with C-H bond elongation. The observed value of 26 Hz suggests that the products have character that is closer to the alkyl hydride/highly stretched end of the range of possibilities.

Fig. 5.Fig. 5.
Exchange process for alkyl and hydride protons via alkane complexes.

Based on NMR studies alone, it is not possible to differentiate these three possibilities while there are exchange processes present at the lowest useable temperatures. Other methods must therefore be applied. In the case of IR spectroscopy, separate rather than averaged bands can often be observed for equilibrating mixtures allowing for the differentiation of single components that would be averaged in NMR experiments. This approach is described later and suggests the presence of an equilibrating mixture of 3a and 4a.

Photolysis in n-pentane. The photolysis of 1 was investigated in a mixture of 95% n-pentane and 5% n-pentane-d12 at 190 K, using a procedure similar to that described in the previous section. After photolysis, three 19F resonances were observed: a major product at δ −3.36 (dq, 1JFP = 1,212 Hz, 3 × JFH = 6.33 Hz), and two low intensity signals at δ −3.32 (1JFP = 1,213 Hz) and δ −3.20 (1JFP = 1,214 Hz) for which JFH could not be measured but is <5 Hz for δ −3.32. These three resonances were attributed to species of the formula Re(Cp)(CO)(PF3)(1-pentane) (3b-1), and tentatively Re(Cp)(CO)(PF3)(2-pentane) (3b-2) and Re(Cp)(CO)(PF3)(3-pentane) (3b-3), respectively (Fig. 6). The nomenclature implies that coordination to the metal center occurs through a C-H group of the C1, C2, or C3 of n-pentane.

Fig. 6.Fig. 6.
Resolution-enhanced, 470.6-MHz 19F NMR spectrum after photolysis of 1 in n-pentane at 190 K showing resonances from complexes of the formula 3b. [low asterisk], secondary photoproduct.

The three JFH splittings in the 19F resonance of the major product in n-pentane (3b-1) indicate that this product is formed by interaction of a CH3 group with the [Re(Cp)(CO)(PF3)] fragment, and because the three protons of the CH3 group couple equally to the 19F nuclei, they are likely exchanging positions rapidly, too. This assignment is supported by experiments in 2,2,4,4-pentane-d4 that produced exactly the same 19F resonance splittings (3 × JFH), whereas use of pentane-d12 results in no observable JFH splittings and a small isotope shift. 19F NMR spectra suggest a rapid exchange between the three hydrogens of the CH3 group even at 165 K. Exchange of complexed and uncomplexed hydrogens within the bound CH3 group in Re(Cp)(CO)2(1-pentane) (2b-1) is known to be fast at this temperature (11). 1H NMR spectra were also collected (at 190 K), and these showed three low-frequency 1H peaks due to the three different isomers of 2b (11). As in the case of cyclopentane, no 1H resonances in the δ < 0 region that could correspond to the largest signals in the 19F spectrum were observed. However, when the temperature was increased to 205 K, in the absence of UV light, a broad signal at δ −1.67 appeared due to 3b-1. Resonances attributed to isomers of 2b rapidly disappeared at this temperature. When the temperature was lowered to 160 K in the absence of UV light, the broad signal disappeared, whereas the 19F signals did not show any major changes. A decoalescence phenomenon is clearly operative in the case of 3b-1 as well.

The experiment was repeated by using n-pentane-13C1 at a higher temperature of 210 K. Two-dimensional 1H-13C-{31P} correlation experiments reveal that this broad signal is correlated with a shielded carbon signal at δ −15.1, and a 13C edited 1H NMR experiment shows JCH = 89 ± 2 Hz. Again, this is an averaged coupling that in this case involves three hydrogens rather than two in the case of cyclopentane. If this species were a 1-pentane complex, 3b-1, the exchange of one complexed C-H with two uncomplexed C-H bonds of a coordinated CH3 unit would be involved. Alternatively, if this species were an n-pentyl hydride, 4b, the coupling would be the average of one 2JCH value for the hydride and two alkyl 1JCH values. The observed value of JCH = 89 Hz is much lower than that of 117 Hz observed in 2b-1 and is at the upper limit of the range expected for an averaged coupling in an alkyl hydride system, predicted to be ≈71–89 Hz. Possible scenarios based on NMR data for 3b-1 are the following: (i) an alkane complex wherein the complexed C-H bond is stretched to near breaking; (ii) the system being pentyl hydride 4b with relatively high C-H coupling constants; or (iii) an equilibrium mixture of 4b and 3b where the equilibrium favors the alkyl hydride more than in the case of cyclopentane. TRIR experiments described below confirm that this third scenario is indeed the correct one.

In the case of 2b, there is a small thermodynamic preference for binding to the CH2 sites over the CH3 sites (11). Studies of species that induce C-H activation, such as [Tp′RhL] [Tp′ = Tris-(3,5-dimethylpyrazolyl)borate; L = CNCH2CMe3] have also shown a kinetic preference for binding at secondary sites (22). However, it is well known that there is a preference for C-H oxidative cleavage to occur at primary carbons (5, 6). In the case of [ReCp(CO)(PF3)] interacting with n-pentane, it is possible that the availability of an energetically more accessible primary alkyl hydride 4b leads to a preference for interaction with the primary carbon, the opposite of that which occurs with 2b. An alternative explanation is that incorporation of the sterically more demanding PF3 ligand leads to preferred binding of the less hindered primary carbon.

Photolysis in cyclohexane. The preferred conformation of a cyclohexane ring is a chair in which the axial and equatorial hydrogens within a CH2 group are chemically distinct. At temperatures around 190 K, the rate of inversion of the chair forms, the process that can make axial and equatorial protons equivalent from an NMR perspective, is slow. By removing the equivalency of the CH2 protons in the alkane molecule, the two dynamic processes described in the cyclopentane case, rotation of the alkane ligand and swapping of coordination of the metal center between the two CH2 hydrogens, will never make these two protons equivalent in the 1H NMR spectrum. Therefore, discrete resonances for axial and equatorial hydrogens of a bound cyclohexane moiety are expected.

By using a procedure similar to that used with cyclopentane, a solution of 1 in 50% cyclohexane:50% pentane-d12 was cooled to 190 K. Upon photolysis, four new 19F NMR resonances were observed to grow in simultaneously. One resonance at δ −3.13 (1JPF = 1,213 Hz), likely with two unresolved 3JFH couplings, both <5 Hz, was assigned to Re(Cp)(CO)(PF3)(cyclohexane) (3c). The other resonances were due to the three isomers of Re(Cp)(CO)(PF3)(pentane-d12) (3b-d12) previously assigned in the studies in n-pentane. A noticeable increase in the intensity of the 19F resonance of 3c and corresponding decrease in intensity of the signals due to 3b-d12 after the cessation of photolysis was observed, indicating a thermodynamic preference for 3c over 3b-d12.

1H NMR monitoring of the photolysis showed the appearance of two new broad signals at δ −2.73 (JPH ≈16 Hz) and δ −3.33 (JPH unresolved) that integrated one proton each with respect to five protons of a new peak in the Cp region at δ 4.93. The signals at δ −2.73 and δ −3.33 were assigned to the equatorial and axial protons, respectively, of a bound CH2 unit in the cyclohexane ligand of 3c. An unambiguous confirmation that these assignments are correct has not been obtained. Examination of the broad multiplet structures in the 1H NMR spectrum suggests that the resonance at δ −3.33 contains two larger axial–axial 3JHH splittings, consistent with it being an axial hydrogen and 2D exchange experiments favor this interpretation. A 2D 19F-31P correlation experiment revealed δ 31P = 118.3 for 3c, and the larger JPH ≈16 Hz was retained in this experiment.

Repeating the experiment by using uniformly labeled cyclohexane-[13C]6 reveals values of 1JCH of 93 ± 3 and 107 ± 3 Hz for the resonances at δ −2.73 and δ −3.33, respectively. A 1H-13C correlated experiment indicates that both of these 1H resonances are from the same bound CH2 moiety that has δ 13C = −17.6. This can be compared with a 1JCH = 125 ± 0.5 and 96.5 ± 0.5 Hz for the equatorial and axial protons, respectively, observed for the bound CH2 unit of the cyclohexane ligand in Re(Cp)(CO)2(cyclohexane) (2c), which has δ 13C = −22.4 (12). For further comparison, free cyclohexane at the same temperature has 1JCH = 128 Hz (both hydrogens) and δ 13C = 26.8. In the case of 2c, there was a clear preference for the complexation of axial over equatorial bonds, leading to a significantly more shielded 1H NMR shift and lower 1JCH value for the axial proton that complicates the comparison with 3c. It is possible that there is a steric factor playing a role here, the bulkier PF3 ligand tending to disfavor the coordination of the sterically more encumbered axial sites relative to the situation in 2c with the result that there appears to be almost no thermodynamic preference for axial or equatorial sites in 3c. Regardless of which are the axial and equatorial protons, the average of the two JCH couplings observed in 3c (99.5 Hz) is significantly less than the corresponding average in the alkane complex 2c (110.7 Hz). It is likewise clear that the average JCH coupling of the bound CH2 in 3c (99.5 Hz) is significantly larger than for the bound CH2 in cyclopentane complex 3a (75 Hz). This finding suggests that in 3c, the complexed C-H bonds may be more stretched in comparison with those found in 2c but less stretched than in 3a. Alternatively (or simultaneously), there may be a contribution from alkyl hydride isomers, Re(Cp)(CO)(PF3)(cyclohexyl)H 4c, that are in fast exchange with alkane complex isomers 3c. The relative amount of alkyl hydride isomers would be significantly lower for 3c than 3a. This proposed equilibrium between 3c and 4c is described in Fig. 7. The variety of possible orientations and geometries of the alkane, alkyl, or hydride ligands with respect to the PF3 and CO ligands means that the observed signals may be the average of numerous subtypes of different isomers.

Fig. 7.Fig. 7.
Possible equilibria between four different cyclohexane and cyclohexyl hydride complexes. Keq appears to be close to 1 in complex 3c.

The value of JCH for the axial hydrogen (107 Hz) is higher compared with the equatorial hydrogen (93 Hz) in bound cyclohexane and to a C-H in bound cyclopentane (75 Hz). This finding suggests that complexed axial C-H bonds retain a mostly, if not completely, alkane complex character. By comparison, complexed equatorial C-H bonds in cyclohexane and bound C-H in cyclopentane are becoming either progressively more stretched or have progressively greater contributions from the corresponding alkyl hydride isomer (i.e., K″ > K′ in this case; see Fig. 7).

TRIR Studies.

Pentane. The TRIR spectrum obtained after photolysis of 1 in n-pentane under CO (2 atm) is shown in Fig. 8. It is clear that the parent ν(CO) bands (1,941 and 2,004 cm−1) are bleached, and three transient bands can clearly be observed at ≈1,887, 1,923, and 1,953 cm−1. The transient bands are formed within the time resolution of the apparatus (20 ns). Previous TRIR studies on the photolysis of 1 in scXe (17) have shown both Re(Cp)(CO)2Xe and Re(Cp)(CO)(PF3)Xe are formed. The band at 1,923 cm−1 is assigned to Re(Cp)(CO)(PF3)(pentane) (3b). The band of 3b is not stable and decays monoexponentially [kobs = 9.1 (± 0.1) × 100 s−1, 292 ± 1 K] in the presence of CO to reform 1. Comparison with the ν(CO) band positions for known Re(Cp)(CO)2(alkane) complexes allows the transient bands at 1,887 and 1,953 cm−1 to be assigned to Re(Cp)(CO)2(pentane) (2b). However, the bands at 1,887 and 1,953 cm−1 show different kinetic behavior (Fig. 9a). The low-frequency band at 1,887 cm−1 decays monoexponentially [kobs = 1.01 (± 0.01) × 102 s−1, 292 ± 1 K]. The band at 1,953 cm−1 decays with biexponential kinetics. The faster component is in good agreement with the decay of 1,887 cm−1 [kobs = 9.6 (± 0.2) × 101 s−1, 291 ± 1 K], and this confirms the band's assignment to 2b. Further evidence for this assignment is that in the presence of CO, these bands decay to form Re(Cp)(CO)3 (5). The presence of two components in the decay of the 1,953 cm−1 band indicates that a second ν(CO) band is heavily overlapped with the 2b band at 1,953 cm−1, and the slower decay is very similar to the band at 1,923 cm−1 [kobs = 9.4 (± 0.9) × 100 s−1, 291 ± 1 K]. To obtain more information regarding this mystery peak, we monitored the photolysis of 5 in n-pentane by TRIR to determine precisely the ν(CO) bands of Re(Cp)(CO)2(pentane) (1,887 and 1,952 cm−1), and the ratio of bands is clearly different from those obtained in the fitting of the Re(Cp)(CO)2(PF3) experiment (Fig. 8).

Fig. 8.Fig. 8.
TRIR spectra showing 2b, 3b, and 4b. (a) FTIR of 1 in n-pentane under CO (2 atm). (b) TRIR difference spectra of the same solution 50 μs after photolysis (266 nm). (c) Expansion (1,965–1,875 cm−1) of TRIR spectrum of 1 in n-pentane. (more ...)
Fig. 9.Fig. 9.
Kinetic traces of ν(CO) bands at ≈1,883–1,887 and 1,948–1,952 cm−1 after photolysis of 1 in pentane at 293 ± 1 K (a), cyclopentane at 292 ± 1 K (b), and cyclohexane 293 ± 1 K (c), all under (more ...)

Multilorentzian fitting of the TRIR spectrum obtained after irradiation of 1 reveals the presence of an extra band at 1,955 cm−1. We have recently shown that photolysis of Re(Cp)(CO)3 in scCH4 generates Re(Cp)(CO)2(CH4) and Re(Cp)(CO)2(CH3)H that exist in a rapid equilibrium at room temperature in solution (14). We assign the overlapped band at 1,955 cm−1 to Re(Cp)(CO)(PF3)(pentyl)H (4b). The similar lifetimes of 3b and 4b also indicate that both may be present in a rapid equilibrium. The assignments of 3b and 4b have been further supported by density functional theory calculations in which we have computed the IR frequencies (see the SI).

Cyclopentane/cyclohexane. NMR experiments discussed above have indicated that the nature of the alkane solvent has a significant effect on the balance between coordination and activation of the alkane solvent. Experiments in cyclopentane have suggested the presence of a significant proportion of Re(Cp)(CO)(PF3)(cyclopentyl)H (4a), whereas in cyclohexane only a small fraction, if any, of Re(Cp)(CO)2(cyclohexyl)H (4c) was indicated to be present.

To investigate these results further, we carried out analogous experiments to those described above for pentane in cyclopentane and cyclohexane (see the SI). In both experiments, Re(Cp)(CO)(PF3)(alkane) and Re(Cp)(CO)2(alkane) (alkane = cyclopentane or cyclohexane) were observed. It was not possible to establish whether Re(Cp)(CO)(PF3)(alkyl)H was formed from inspection of the band ratios of Re(Cp)(CO)2(alkane) as described above for Re(Cp)(CO)(PF3)(pentyl)H (4b). Re(Cp)(CO)(PF3)(alkane) reacts with CO to reform the parent [cyclopentane: kobs = 3.3 (± 0.1) × 100 s−1, 292 ± 1 K; cyclohexane: kobs = 4.2 (± 0.2) × 100 s−1, 293 ± 1 K], slower than the corresponding reaction of Re(Cp)(CO)2(alkane) to form Re(Cp)(CO)3. We found no evidence from the kinetics of the ν(CO) band at ≈1,950 cm−1 of Re(Cp)(CO)2(cyclohexane) (2c) for the presence of another species. NMR studies suggest that that the amount of cyclohexyl hydride species 4c present is much less than in the case of pentane or cyclopentane. Any possible concentration of 4c would be small and therefore difficult to detect due to the overlapping of the IR spectral features. In the TRIR experiment in cyclopentane, we examined the decay of the two ν(CO) bands of Re(Cp)(CO)2(cyclopentane) (2a) and both bands decay at a similar rate [1,883 cm−1: kobs = 4.4 (± 0.2) × 101 s−1; 1,951 cm−1: kobs = 3.8 (± 0.1) × 101 s−1, both at 292 ± 1 K]. The slightly increased lifetime of the band at 1,951 cm−1 may be due to a small concentration of 4a, having ν(CO) bands overlapped with those of 2a again consistent with NMR studies. However, this assignment cannot be made solely on the TRIR kinetics.

Conclusions

Combined NMR and TRIR data reveal that the species produced by the interaction of n-pentane with [Re(Cp)(CO)(PF3)] are an equilibrium mixture of Re(Cp)(CO)(PF3)(pentane) (3b) and Re(Cp)(CO)(PF3)(pentyl)H (4b). This contrasts with the [Re(Cp)(CO)2] fragment, which only forms alkane complexes with n-pentane and has a slight preference for complexation of secondary C-H bonds. The relative concentrations of 3b and 4b may be temperature-dependent, with lower temperatures leading to an increase of the relative concentration of alkyl hydride, but accurate equilibrium constants are difficult to extract. Data for the products from the interaction of cyclopentane with [Re(Cp)(CO)(PF3)] again suggest that an equilibrium between the cyclopentyl hydride 4a and the isomeric cyclopentane complex 3a exists. What is clear is that there is a trend in both the NMR and TRIR data on going from n-pentane to cyclopentane to cyclohexane with a progressive favoring of the proportion of alkane complex present. Hence, in the case of cyclohexane, even at 180 K there is only a small amount, if any, of the cyclohexyl hydride form 4c present. The NMR data suggest that there is little preference for the interaction of [Re(Cp)(CO)(PF3)] with either axial or equatorial hydrogens in cyclohexane but that interaction with the equatorial site is more likely to result in the formation of 4c, based on the lower JCH and higher JPH values in the equatorial case. Alternatively, the JCH and JPH values may be indicative of a more stretched C-H interaction in the case of binding of the equatorial hydrogen. All of the alkane complex-alkyl hydride equilibria in Figs. 4, 5, and 7 are fast on the NMR time scale at 215 K, but decoalescence phenomena are observed at lower temperatures for 3a/4a and 3b/4b, possibly due to a slowing of rotation and/or swapping of complexed and uncomplexed hydrogens in an alkane ligand.

Although CO and PF3 ligands are considered similar in terms of their π-acceptor properties, there is a general shift toward products that display a higher degree of oxidative cleavage when [Re(Cp)(CO)(PF3)] reacts with alkanes in comparison with [Re(Cp)(CO)2], which favors alkane complexation without activation. The complexes incorporating a PF3 ligand have greater thermal stability and react more slowly with CO.

Materials and Methods

Re(Cp)(CO)2(PF3) (1) was synthesized by using the literature procedure (17).

NMR/Photolysis Experiments. NMR samples of 1 (≈1 mg) in a mixture of appropriate alkane and pentane-d12 (530 μl) were prepared. These samples were precooled (160–223 K) in the probe of a Bruker (Rheinstetten, Germany) DPX 300, DMX 500, or DMX 600 NMR spectrometer. Light from a 100-W Hg arc lamp was delivered to the top of the solution by using a single-core fiber-optic as described in refs. 10 and 11. Progress of reactions was monitored by using 19F and 1H NMR spectroscopy either on separate, similar samples or by alternating experiment type. Conventional 19F spectra were obtained and were 1H-coupled. 1H NMR spectra were obtained by using an excitation sculpting scheme to suppress the resonances of the free alkane (11). Collected NMR spectra and further experimental details are given in the SI.

IR Experiments. Re(Cp)(CO)3 (Strem Chemicals, Newburyport, MA; 99%) was used as received. Cyclopentane (Fluka, Buchs, Switzerland; HPLC grade), pentane (Lancaster, Ward Hill, MA; >99%), and cyclohexane (Aldrich, St. Louis, MO; HPLC grade, >99.9%) were distilled from CaH2 and degassed before use. All experiments were carried out by using a CaF2 cell (Harrick, Ossining, NY; pathlength of 0.5–1.5 mm) under CO (2 atm) at room temperature. Fresh solution was flowed into the cell after every UV laser shot. Details of the diode laser-based TRIR apparatus are described in ref. 23. Briefly, the IR source is a continuous-wave IR diode laser (MDS 1100; Mütek, Herrsching, Germany). In these experiments, the change in IR transmission at one IR frequency was measured by a fast MCT detector, after UV excitation of the sample by a pulsed Nd:YAG laser (Quanta-Ray GCR-12; Spectra Physics, Mountain View, CA; 266 nm), which initiates the photochemical reactions. A spectrum was built up on a point-by-point basis in custom software, by repeating this measurement at different IR frequencies.

Table 1.Table 1.
ν(CO) band positions of complexes recorded at 292–293 K
Supplementary Material
Supporting Information
Acknowledgments

This work was supported by FUJIFILM Imaging Colorants (A.J.C.), the European Union (P.P.), the Engineering and Physical Sciences Research Council (M.W.G.), SASOL (C.M.B.), the Australian Research Council (G.E.B.), the University of Nottingham (A.J.C.), the University of New South Wales (G.E.B.), and a J. W. T. Jones Fellowship from the Royal Society of Chemistry (to J.P.R.).

Abbreviation

TRIRtime-resolved IR.

Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/cgi/content/full/0610212104/DC1.
References
1.
Hall, C; Perutz, RN. Chem Rev. 1996;96:3125–3146. [PubMed]
2.
Kubas, GJ. J Organomet Chem. 2001;635:37–68.
3.
Crabtree, RH. Angew Chem Int Ed Engl. 1993;32:789–805.
4.
Heinekey, DM; Oldham, WJ., Jr Chem Rev. 1993;93:913–926.
5.
Labinger, JA; Bercaw, JE. Nature. 2002;417:507–514. [PubMed]
6.
Crabtree, RH. J Organomet Chem. 2004;689:4083–4091.
7.
Bengali, AA; Schultz, RH; Moore, CB; Bergman, RG. J Am Chem Soc. 1994;116:9585–9589.
8.
Brown, CE; Ishikawa, Y; Hackett, PA; Rayner, DM. J Am Chem Soc. 1990;112:2530–2536.
9.
Sun, X-Z; Grills, DC; Nikiforov, SM; Poliakoff, M; George, MW. J Am Chem Soc. 1997;119:7521–7525.
10.
Geftakis, S; Ball, GE. J Am Chem Soc. 1998;120:9953–9954.
11.
Lawes, DJ; Geftakis, S; Ball, GE. J Am Chem Soc. 2005;127:4134–4135. [PubMed]
12.
Lawes, DJ; Darwish, TA; Clark, T; Harper, JB; Ball, GE. Angew Chem Int Ed. 2006;45:4486–4490.
13.
Childs, GI; Colley, CS; Dyer, J; Grills, DC; Sun, X-Z; Yang, J; George, MW. J Chem Soc Dalton Trans. 2000:1901–1906.
14.
Cowan, AJ; Portius, P; Jina, OS; Grills, DC; McMaster, J; George, MW. Proc Natl Acad Sci USA. 2006;104:6933–6938.
15.
Brookhart, M; Green, MLH; Wong, LL. Prog Inorg Chem. 1988;36:1–124.
16.
Brookhart, M; Green, MLH. J Organomet Chem. 1983;250:395–408.
17.
Ball, GE; Darwish, TA; Geftakis, S; George, MW; Lawes, DJ; Portius, P; Rourke, JP. Proc Natl Acad Sci USA. 2005;102:1853–1858. [PubMed]
18.
McMaster, J; Portius, P; Ball, GE; Rourke, JP; George, MW. Organometallics. 2006;25:5242–5248.
19.
Klahn-Oliva, AH; Singer, RD; Sutton, D. J Am Chem Soc. 1986;108:3107–3108.
20.
Wenzel, TT; Bergman, RG. J Am Chem Soc. 1986;108:4856–4867.
21.
Gross, CL; Girolami, GS. J Am Chem Soc. 1998;120:6605–6606.
22.
Vetter, AJ; Flaschenriem, C; Jones, WD. J Am Chem Soc. 2005;127:12315–12322. [PubMed]
23.
George, MW; Poliakoff, M; Turner, JJ. Analyst. 1994;119:551–560.