• Keine Ergebnisse gefunden

Limits of Activity: Weakly Coordinating Ligands in Arylphosphinesulfonato Palladium(II) Polymerization

III. Index of Complexes

3. Limits of Activity: Weakly Coordinating Ligands in Arylphosphinesulfonato Palladium(II) Polymerization

Catalysts

A

3.1 Introduction

The coordination strength of the monodentate ligand L introduced with the catalyst precursors [(P^O)PdMe(L)] (MeO1-L; P^O = κ2-P,O-Ar2PC6H4SO2O with Ar = 2-MeOC6H4) has a major impact on the catalytic activity in homo- and copolymerizations due to the equilibrium (P^O)PdR(L) + monomer (P^O)PdR(monomer) + L, which accompanies chain-growth. Thus, stronger coordinating ligands shift the equilibrium towards the dormant species MeO1-L. So far monodentate ligands, e.g. PPh3, tmeda, pyridine, 2,6-lutidine, DMSO and derivatives thereof have been used (Figure 3-1).24,25,38,43,45,86

Alternatively, carbon based ligands occupying two coordination sites, e.g. η3-allyl or η12-2-methoxycyclooct-5-enyl are suitable precursors to initiate chain growth.20,87

Figure 3-1. Reported complexes MeO1-L with neutral, monodentate ligands L.

By comparison to the aforementioned N- and P-based ligands dimethylsulfoxide (DMSO) binds less strongly to the metal center and is more readily displaced by olefinic substrates. This enabled homooligomerization of methyl acrylate (MA) and the isolation of ethylene-methyl acrylate copolymers with more than 50 mol% MA incorporation.43 Here, the weak coordination strength of DMSO permitted polymerization at low ethylene pressures and thus high MA/ethylene ratios. For completeness it should be mentioned that entirely 'base-free' species of the molecular composition [(P^O)PdMe] have been isolated25,44,49 or synthesized in situ by abstraction of L from (P^O)PdMe(L).21,45-47,87

However, so far no improved polymerization activities in comparison to ‘base-coordinated’ compounds have

been reported. For in situ activated catalysts this might be due to incomplete activation or side reactions by activation reagents or catalyst precursors.21,45,87 For isolated material the reported low solubility likely renders part of the catalyst inactive.44

The significantly higher activity observed with DMSO- vs. pyridine-coordinated catalyst precursors suggests studies of further weaker coordinating ligands. Phosphine oxides (O=PR3) as a ligand class lend themselves for this purpose, as they are easily accessible from the corresponding phosphines and exhibit a defined coordination site at the oxygen atom.

Furthermore, a great variety of phosphines is commercially available and allows for electronic and steric fine-tuning. While chelating, hemilabile ligands (X^O; X = N,P,O; O = phosphine oxide), and especially the phosphine-phosphine oxide ligands have attracted much attention in homogeneous catalysis,88-93 the application of monodentate tertiary phosphine oxides in homogenous catalysis is rare,94-98 even though coordination towards metal centers is well studied.99 With respect to ethylene polymerization, Starzewski et al. reported a remarkable effect with a catalyst system based on a hemilabile κ2-P,O ligand, Ni(cod)2 and an additional monodendate ligand. By changing the monodentate ligand from triphenylphosphine to the corresponding phosphine oxide molecular weight of the produced polymer could be significantly increased from low-molecular weight waxes to high molecular weight polyethylene (Mη > 106 g/mol).97,98

3.2 Results and Discussion

3.2.1 Coordination Strength of Phosphine Oxides

Since the coordination strength is influenced by steric as well as electronic properties, both parameters should be varied independently. Here, the cone angle θ and the electronic parameter χ of the corresponding phosphines enable an educated selection of phosphine oxides (χ is defined as the difference between the ν(CO) (A1) absorption in L-Ni(CO)3 and the ν(CO) (A1) absorption in (t-Bu)3P-Ni(CO)3).100 For this study OPBu3, OPOct3, OPPh3, OP(o-Tol)3, and OP(p-CF3C6H4)3,100,101 for which θ– and χ–parameters have been reported, as well as the even more electron deficient OP(3,5-(CF3)2C6H3)3 were selected (Figure 3-2). For OP(3,5-CF3C6H3)3 no parameter data was available, but coordination strength can be expected to be rather low due to increased electron deficiency in comparison to OP(p-CF3C6H4).

Phosphine oxides not available commercially were synthesized by phosphine oxidation with H2O2 (c.f. Chapter 7.2.1).

Figure 3-2. Coordination strength of phosphine oxides based on parameters θ and χ of the corresponding phosphines.33,100

Regarding electronic properties, the comparison of the literature derived electronic parameters with the observed 31P NMR shifts of the phosphine oxides show a good correlation. With increasing electron deficiency χ increases, while δ decreases. It is important to note that the 31P NMR shift of phosphine oxides is a reliable measure of the basicity as opposed to the 31P NMR shift of phosphines which is also influenced by sterics.102 As expected, the aryl phosphine oxides exhibit a weaker basicity than OPBu3. In comparison to OPPh , steric bulk is increased by introduction of a methyl group in ortho-position in

OP(o-is enlarged in comparOP(o-ison to the phosphine. In contrast, the introduction of a CF3-group in para-position in OP(p-CF3C6H4)3 does not change steric properties but increases electron deficiency. In OP(3,5-(CF3)2C6H3)3 electron deficiency is further increased while the steric influence is believed to be similar to OP(p-CF3C6H4)3 (θ = 145° vs. ~151° for P(3,5-Me2C6H3)3).103

The relative coordination strength of these phosphine oxides compared to DMSO (KOPR3) was determined by 1H NMR spectroscopy. The 1H resonance of DMSO in CD2Cl2 is downfield shifted from 2.54 ppm (free DMSO) to 2.95 ppm by complexation to the Pd center in MeO1-dmso. Here, the exact coordination mode of DMSO remains unclear, but for a related complex an S-coordination was observed by X-Ray analysis.43 Partial replacement of DMSO by OPR3 leads to a high field shift due to a fast equilibrium between Pd-bound and uncoordinated DMSO. From the shift difference the ratio between MeO1-dmso and MeO1-L and consequently KL at 25 °C was calculated (cf. Chapter 7.1.10).

Table 3-1. KL for MeO1-dmso + L 1-L + DMSO.

a∆δ is low due to limited solubility of OPR3 consequently inaccuracy of KL is enhanced.

The results are summarized in Table 3-1. Whereas both alkyl phosphine oxides coordinate slightly stronger than DMSO (KOPBu3 = 3.5, KOPOct3 = 3.3), the more bulky and electron-deficient OPPh3 exhibits KOPPh3 = 0.2 for the equilibrium MeO1-dmso + OPR3

MeO1-OPR3 + DMSO. An even weaker coordination is evident for the comparison with OP(o-Tol)3 (KOPTol3 = 0.03), and OP(p-CF3C6H4)3 (KOP(pCF3Ar)3 = 0.04, Table 3-1, compare KL for MeOH and 2,6-lutidine). Hence, coordination strength can be controlled by either steric bulk or electron deficiency over a large range. The introduction of a second electron withdrawing group in OP(3,5-(CF3)2C6H3)3 further reduces coordination strength significantly (KOP(3,5CF3Ar)3 ~ 0.001). Note, that in a related study KL was determined vs. DMSO for ethyl acetate (KEA < 10-2), methyl ethyl sulfone (KMES < 10-2), propionic acid (KPrA < 0.1),

N,N-dimethylacetamide (KDMAcA ≈ 0.4), N-methylacetamide (KMAcA ≈ 0.6) and acrylonitrile (KACN ≈ 1).56

In order to investigate the influence of the temperature on the coordination equilibrium, KOPPh3 was determined in the temperature range between -25 °C and 25 °C. Only minor changes of KOPPh3 in this temperature range have been observed (Table 3-2).

Table 3-2. KL at different Temperatures ([MeO1-dmso] = 0.016 mol L-1; equiv. L: 10.3).

Plotting the natural logarithm of the equilibrium constants against the temperature in a Van’t Hoff analysis allowed for the determination of ∆H°DMSO/OPPh3 = 8 kJ mol-1,

Figure 3-3. Van’t Hoff plot for the equilibrium MeO1-dmso + OPPh3 MeO1-OPPh3 + DMSO determined by variable temperature 1H NMR spectroscopy from T = -25 °C to 25 °C.

It is assumed that a similar small temperature dependence applies to all KOPR3 and that

according to the equilibrium [(P^O)PdR(L)] + ethylene [(P^O)PdR(ethylene)] + L, complexes MeO1-L derived from phosphine oxides with a weaker coordination strength than DMSO, i.e. from OPPh3, OP(o-Tol)3, OP(p-CF3C6H4)3, and OP(3,5-(CF3)2C6H3)3 are expected to exhibit higher turnover frequencies than MeO1-dmso as long as saturation kinetics are not reached. Consequently, such complexes MeO1-L represent valuable synthetic targets for highly active single component catalysts.

3.2.2 Complex Synthesis and Characterization

For the synthesis of phosphine oxide complexes MeO1-OPR3 standard procedures were not applicable since they are either based on introduction of the ligand with the Pd-precursor as with L = TMEDA from [(tmeda)PdMe2],20,24,38,87

or subsequent ligand substitution by a stronger coordinating ligand.25,38,44,45,62,86

Notably, the weaker coordinating DMSO could be introduced by substitution of TMEDA. This substitution occurs since TMEDA is removed from the equilibrium ½ (MeO1)2-tmeda + DMSO MeO1-dmso + ½ TMEDA under vacuum due to the considerably higher volatility of TMEDA vs. DMSO.43 However, an analogous procedure, e.g. solvent evaporation from a mixture of (MeO1)2-tmeda and phosphine oxide in high boiling solvents, did not result in the isolation of clean products. In an alternative approach, multinuclear 'base-free' palladium alkyl complexes which are accessible e.g. by pyridine or lutidine abstraction with B(C6F5)325,44 may be suitable precursors for the preparation of phosphine oxide complexes MeO1-OPR3. However, a more convenient synthesis starts from [{(MeO1-Cl)-µ-Na}2)]61 and chloride abstraction in the presence of phosphine oxides is expected to generate MeO1-OPR3 if the presence of stronger coordinating ligands is avoided (Scheme 3-1).

Scheme 3-1. Synthesis of MeO1-L.

The viability of this general route was demonstrated by the synthesis and isolation of

MeO1-dmso. As expected, also the complexes MeO1-OPBu3 and MeO1-OPOct3 with the