Next Article in Journal
Tetrahydroborates: Development and Potential as Hydrogen Storage Medium
Next Article in Special Issue
Mechanistic Implications for the Ni(I)-Catalyzed Kumada Cross-Coupling Reaction
Previous Article in Journal
Unique Hydrogen Desorption Properties of LiAlH4/h-BN Composites
Previous Article in Special Issue
Redox-Induced Aromatic C–H Bond Functionalization in Metal Complex Catalysis from the Electrochemical Point of View
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

3-Methylindole-Based Tripodal Tetraphosphine Ruthenium Complexes in N2 Coordination and Reduction and Formic Acid Dehydrogenation

by
Fenna F. Van de Watering
,
Nicol Heijtbrink
,
Jarl Ivar Van der Vlugt
,
Wojciech I. Dzik
,
Bas De Bruin
* and
Joost N. H. Reek
*
Homogeneous, Supramolecular and Bio-Inspired Catalysis Group, Van’t Hoff Institute for Molecular Sciences, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
*
Authors to whom correspondence should be addressed.
Submission received: 27 September 2017 / Revised: 18 October 2017 / Accepted: 25 October 2017 / Published: 30 October 2017

Abstract

:
The ruthenium(II) complexes RuCl2L1H, RuCl2L1CF3, RuCl2L1OMe and RuCl2L2H were synthesized from [Ru(η6-benzene)Cl(μ-Cl)]2 and the corresponding tripodal tris-3-methylindolephosphine-based ligands L1H, L1CF3, L1OMe, and L2H. Stoichiometric reduction of these complexes with KC8 yielded the corresponding ruthenium(0) dinitrogen complexes. The latter complexes were studied in the N2 reduction with chlorosilanes and KC8, yielding stoichiometric amounts of the silylamines. The synthesized ruthenium(II) complexes are also active catalysts for the formic acid dehydrogenation reaction.

Graphical Abstract

1. Introduction

Electrocatalytic reduction reactions that store energy by the conversion of H+, CO2, or N2, are important in the context of the transformation to a society based on sustainable energy. Therefore, the development of efficient catalysts that enable these reactions is needed. Homogeneous tripodal ligand transition metal complexes have shown beneficial reactivity over mono- or bidentate ligands in many organic transformations [1,2,3]. In recent years, tripodal ligands coordinated to molybdenum, iron, and cobalt were shown to lead to complexes that are capable of the reduction of N2 to ammonia or silylamines [4,5,6,7,8,9,10]. Also, reversible hydrogen storage in CO2 was shown to be possible with iron and cobalt based on tripodal ligands [11,12,13]. Previously, we reported the coordination of ligand L1H (Figure 1) to rhodium [14,15] and ruthenium [16]. This ligand, which features three 1-(3-methylindolyl)diphenylphosphine groups tethered to a central phosphorus atom, stabilizes ruthenium in the oxidation states +II, +I, and 0. The isolated low oxidation state RuIClL1H and Ru0(N2)L1H complexes revealed interesting one-electron reactivity, as they were able to abstract chlorine atoms from either chloroform or dichloromethane, leading to formation of the corresponding RuII complex. In this paper, we expand the coordination chemistry of ruthenium and 3-methylindole-based ligands [17,18,19,20,21,22,23] and investigate how modification of the ligand scaffold and its electronic properties influence the stability and activity of these complexes in dinitrogen coordination and subsequent reduction as well as in formic acid dehydrogenation. Ligand L1H was modified using functionalized P(Ar)2 units, wherein the arene carries either electron-withdrawing –CF3 groups (L1CF3) or electron-donating –OMe groups (L1OMe) in the para-position (Figure 1). Additionally, the known structural isomer, L2H (Figure 2) [14], in which the central P-donor is bound to the three nitrogen atoms of the 2-(3-methylindolyl)diphenylphosphine moiety, was also used in these studies.

2. Results

Ligands L1H and L2H were prepared in two steps from 3-methylindole as previously reported by our group [14,15]. The new ligands L1CF3 and L1OMe were prepared in a similar way to L1H, by reacting tris-2-(3-methylindolyl)phosphine [17] with the corresponding chlorodiarylphosphine (Figure 1).
The reaction of L1H with [Ru(η6-benzene)Cl(μ-Cl)]2 in refluxing THF for 16 h led to the formation of the octahedral complex RuCl2L1H in quantitative yield (Figure 2). Surprisingly, the other ligands did not react with this ruthenium precursor in refluxing THF. Even after 36 h, only trace amounts of the target complexes were observed using ex-situ 31P NMR spectroscopy, and mainly signals of the unreacted ligand were present. Reactions in refluxing THF/toluene (1:3) mixture at 120 °C for a prolonged time of 64 h did lead to the formation of complexes RuCl2L1CF3, RuCl2L1OMe, and RuCl2L2H in >50% isolated yields.
All three L1-based RuCl2L1 complexes show similar 31P NMR spectra: three signals in a ratio 1(td):2(t):1(td) (Figure 3). This splitting pattern indicates the formation of a symmetric complex. Ruthenium(II) complexes tend to form octahedral 18-valence electron complexes, and X-ray analysis and NMR spectroscopy confirmed that complex RuCl2L1H indeed features an octahedral geometry with the two chlorides in mutual cis position and all four phosphines bound to the metal center [15]. The above factors and the similarity of the 31P NMR spectra of the other L1-based complexes point to the same octahedral geometry around the metal center.
Layering of a dichloromethane solution of RuCl2L1CF3 with pentane resulted in the formation of crystals suitable for single-crystal X-ray diffraction. The crystal structure (Figure 4) of RuCl2L1CF3 indeed shows an overall octahedral geometry with the chlorido ligands in cis position, in accordance with the observed 31P NMR splitting pattern. The P1–Ru–P2 angle in RuCl2L1CF3 (157.78(6)°) is comparable to the one previously reported for RuCl2L1H (160.04(3)°) (Table 1). The P1–Ru, P2–Ru and P3–Ru bond lengths are all slightly smaller in RuCl2L1CF3 (2.3097(19) Å, 2.3445(19) Å and 2.2383(17) Å, respectively) compared with RuCl2L1H (2.3189(9) Å, 2.3727(9) Å and 2.2671(9) Å). The –CF3 groups clearly point away from the metal center, so steric effects do not play a significant role in the coordination chemistry. The Cs symmetry observed in solution by 31P NMR spectroscopy for both RuCl2L1H and RuCl2L1CF3 is not present in the solid state, as the unit cell contains two rotamers. This becomes evident when the symmetry-equivalent molecules in the crystal structure of RuCl2L1CF3 are viewed from the bottom: the indolyl moieties either all point clockwise, or counter-clockwise (Figure 5, showing the bottom view of RuCl2L1CF3 and a schematic representation). In solution, these two rotamers seem to rapidly interchange on the NMR timescale, resulting in one signal for the two mutually trans phosphines. Likely, the methyl groups of the indolyl moiety do not create enough steric repulsion to suppress this process. Despite several attempts, no crystals suitable for X-ray diffraction of the RuCl2L1OMe complex were obtained.
Interestingly, complex RuCl2L2H, which is based on the linkage isomer of L1H, ligand L2H, displays a different 31P NMR spectrum (Figure 6). The four signals, two triplets of doublets and two doublets of doublets of doublets with an AX pattern, each integrating one phosphine, indicate the formation of a complex with a different symmetry. Crystals obtained by layering of a DCM solution of RuCl2L2H with pentane at 5 °C were of sufficient quality for X-ray crystallography. The corresponding structure of complex RuCl2L2H (Figure 7) helps to understand the origin of the AX pattern, as the bottom view displays much more steric hindrance around the central phosphorus atom P4 (vide infra).
However, different than the L1-based complexes, where the methyl groups are pointing towards the central phosphine (Figure 5), in complex RuCl2L2H, the bulkier phenyl rings of the indolyl moiety point towards the central phosphine (Figure 8). From the crystal structure, it appears that these phenyl rings should encounter more hindrance toward inversion and it therefore seems likely that this process is associated with a high energy barrier in solution. As a result, the two phosphine atoms (P1 and P2, Figure 7) in trans position become inequivalent, leading to the observed AX pattern in the 31P NMR spectrum as a result of the P1–P2 coupling.
The crystal structure of RuCl2L2H further reveals minor differences in the coordination around the ruthenium center compared with RuCl2L1H (Table 1). The P1–Ru, P2–Ru, and P3–Ru bond lengths become slightly larger (2.3295(8) Å, 2.4079(8) and 2.2913(8) Å), but the P4–Ru bond (2.1363(8) Å) shortens, which is in line with a more π-acidic character of the tris(amido)phosphine P4 atom.
Two-electron reduction of RuCl2L1H under an N2 atmosphere led to the formation of the RuN2L1H complex, which features a characteristic infrared band corresponding to the coordinated N2 at 2125 cm−1 (Table 2) [15]. To explore if the electronic properties of the ligands L1CF3, L1OMe, and L2H have an influence on the binding of N2 to Ru0, the corresponding N2 complexes were generated in situ by reacting the ruthenium(+II) complexes with two equivalents of KC8 (Figure 9). The extent of activation of the N2 bond was quantified by measuring the N≡N stretch frequency with infrared spectroscopy, as this is a direct measure for the ligand-dependent π-back-donation. Table 2 shows the N2 stretch frequencies of the Ru0N2L complexes.
As expected, the complex coordinated with the less electron-rich, CF3 substituted ligand shows a band with a higher wavenumber in the infrared spectrum (2136 cm−1), as a result of weaker metal-to-N2 π-back-donation. The complex coordinated to the more electron-rich OMe substituted ligand shows a band at 2113 cm−1, thus revealing stronger π-back-donation to N2. Clearly, the peripheral phosphines induce measurable electronic effects on the d-orbitals of the ruthenium center; the phosphines cis to the N2 ligand become stronger σ-donors and weaker π-acceptors in L1OMe and vice versa for L1CF3, which changes the binding properties of the N2 ligand. Going from RuN2L1H (2125 cm−1) to the structural isomer RuN2L2H (2136 cm−1), the N2 stretch frequency increases by 11 cm−1. This is likely a result of the more π-acidic pivotal P-atom (atom P4, Figure 10, top) of the tris(amido)phosphine donor ligand, which competes with the N2 ligand for π-back-donation from the same metal d-orbitals. Thus, substitution of the phosphine trans to the N2 ligand for a more π-acidic group leads to weaker π-back-donation of Ru to N2. These results show that the electronic properties of the dinitrogen complexes can be tuned by ligand design, either by manipulating the substitutions on the donor atoms coordinated cis to N2 or by changing the π-acidity of the atom bound trans to N2.
Layering of a THF solution of RuN2L1CF3 with pentane resulted in the formation of single crystals suitable for X-ray diffraction (Figure 10). Similar to the RuN2L1H complex reported previously [15], RuN2L1CF3 also shows a trigonal bipyramidal geometry. The N≡N bond is slightly shorter in RuN2L1CF3 (1.064(5)°) than in RuN2L1H (1.085(5)°) (Table 1), which is in accordance with the observed trend in the activation of the N2 stretch frequency (Table 2). In solution, these complexes remain trigonal bipyramidal, as judged from the occurrence of a doublet and a quartet in all 31P NMR spectra.
The dinitrogen ligand in all of these ruthenium complexes is weakly coordinating as judged by the relatively small IR-shifts compared to that of free N2 (ν = 2359 cm−1) (Table 2). Preliminary experiments show no dinitrogen reduction to ammonia when complex RuCl2L1H was subjected to excess reductant (KC8) in the presence of acid ({H(OEt2)2}{BArF24}). Therefore, we set out to explore the reactivity of the tripodal tetraphosphine ruthenium(II) complexes for dinitrogen reduction with chlorosilanes (Figure 11), which follows a different mechanism [24].
A mixture of the RuCl2L complexes with 100 equiv of KC8 and 100 equiv of chlorotrimethylsilane in 10 mL THF at room temperature under an atmosphere of dinitrogen was stirred for 1 day, whereafter an aliquot of the reaction mixture was analyzed by GC. Modest yields of ~1.5 equivalents of (Me3Si)3N relative to Ru were detected, regardless of the complex used (Table 3). The use of metallic sodium as reductant or longer reaction times did not increase the overall yield. In the absence of ruthenium only traces of (Me3Si)3N could be detected.
To gain more insight on the N2 activation using these novel ruthenium complexes, the reduction was followed in time for complexes RuCl2L1H and RuCl2L2H, analyzing aliquots of the reaction mixture every hour. The reaction profile shows an incubation period of 1 h before product formation is observed (Figure 12). Such an incubation period was also reported for various other iron and cobalt systems that were investigated for N2 reduction under comparable conditions [24,25,26], and which all showed similar activities for the formation of tris(trimethylsilyl)amine, regardless of the initial catalyst structure. It was speculated that the catalytically active species are generated after ligand dissociation from these iron and cobalt pre-catalysts. The exact nature of the reported catalysts could not be elucidated, as no dinitrogen-containing species were detected in these previous studies; however, test experiments pointed to a molecular nature of the catalyst [24,25].
Given that our complexes also show an incubation period and that electronic effects induced by ligand modification do not affect the catalytic activity, we speculate that our Ru-complexes are all pre-catalysts. To further test this hypothesis, we performed a catalytic reaction with [Ru(η6-benzene)Cl(μ-Cl)]2 in the absence of any tripodal indolyl-phosphine ligand under the same reaction conditions. Comparable amounts (1.29 equiv relative to Ru) of silylamine were formed in this control experiment, which indeed suggests ligand dissociation to activate the catalyst. Thus, these ligands cannot prevent metal leaching under the strongly reducing conditions used. Ex-situ 31P NMR spectroscopy of the RuCl2L1H catalytic system showed the absence of the signals of the ruthenium complex pointing to decomposition of the pre-catalyst under the applied conditions.
Next, we investigated the potential of these Ru-species in the catalytic dehydrogenation of formic acid (Figure 13), given our interest in the reversible storage of H2 into liquid fuels [27,28,29,30,31,32,33]. For Ru-catalysts, formic acid dehydrogenation is accelerated when more electron-rich ligands are employed, as was shown by Himeda et al., who used a series of bipyridine ligands with various substituents at the para position (–OH, –OMe, –Me, –CO2H, and –H) [34]. With the various substituted tripodal tetraphosphine ligands L1 and L2 in hand, we anticipated to observe similar effects for the corresponding complexes in the formic acid dehydrogenation. The RuCl2L complexes are coordinatively saturated, which necessitates chloride abstraction using AgBF4 to allow for substrate activation. Addition of 1.2 mmol of formic acid under nitrogen to a refluxing solution of a mixture of Ru-precatalyst (8.5 µmol) and AgBF4 (17 µmol) in 3 mL THF pre-activated for 2 h (initial concentration of formic acid: 0.4 M, substrate to catalyst ratio: 140) led to gas production within 1–5 min. The reactions were monitored volumetrically with a gas burette for one hour (see experimental section for a schematic representation). A typical time profile of the dehydrogenation reaction with catalyst RuCl2L1OMe is displayed in Figure 14. The overall results of the different catalysts used in the dehydrogenation of formic acid are presented in Table 4. The gas outlet was analyzed using gas chromatography, which confirmed presence of both CO2 and H2, while no traces of CO were detected (detection limit of the GC for carbon monoxide was δ = 10 ppm). Under the described conditions, all the substrate is used within less than 2 h. Subsequent addition of aliquots of formic acid allowed for continuation of the reaction without an observed decrease in activity for at least 6 h.
Complexes RuCl2L1OMe, RuCl2L1H, and RuCl2L2H follow the same trend as noticed by Himeda [34], with the most electron-rich complex RuCl2L1OMe giving the highest activity (Turnover frequency (TOF) = 121 h−1) followed by RuCl2L1H (TOF = 76 h−1) and the least electron-rich RuCl2L2H (TOF = 33 h−1). Interestingly, the RuCl2L1CF3 complex (TOF = 124 h−1) shows a comparable, even slightly higher activity to the electron-rich RuCl2L1OMe complex. The reason for this unusual observation is currently not clear. The TOFs found for the RuCl2L complexes in the formic acid decomposition reaction are relatively low compared to reported systems based on Ir and Ru. For instance TOF of 487,500 h−1 was obtained with [Cp*Ir(L)Cl]Cl (L = 2,2’-bis-2-imidazoline) in neat HCOOH at 90 °C [36], TOF of 24,000 h−1 in boiling HCOOH/dioxane using [K(dme)2][RuH(trop2dad)] (trop2dad = 1,4-bis(5H-dibenzo[a,d]cyclohepten-5-yl)-1,4-diazabuta-1,3-diene [37], TOF of 4556 h1 with the more related Ru(H)2(P4) complexes (P4 = meso-1,1,4,7,10,10-hexaphenyl-1,4,7,10-tetraphosphadecane) in HCOOH/N,N-dimethyloctylamine/propylene carbonate mixtures at 80 °C [38]. However, these initial results show that the RuCl2L complexes are active in the dehydrogenation of formic acid and that the reactivity can be steered by ligand modification. Further reactivity studies with these complexes should be performed to elucidate the mechanism of the reaction and to find the maximum capacities (stability, temperature, solvent) of these catalysts.

3. Materials and Methods

All reactions were carried out under an atmosphere of nitrogen or argon using standard Schlenk techniques or in the glovebox. Reagents were purchased from commercial suppliers and used without further purification. THF, pentane, hexane, and Et2O were distilled from sodium benzophenone ketyl. These solvents were degassed using the freeze–pump–thaw method (three cycles) and stored under dinitrogen atmosphere. CH2Cl2 was distilled from CaH2 under dinitrogen. NMR spectra (1H, 31P, and 13C{1H, 31P}) were measured on a Bruker DRX 500, Bruker AV 400, Bruker DRX 300, or on a Bruker AV 300 spectrometer (Bruker, Billerica, MA, USA). IR spectra (ATR mode) were recorded with a Bruker Alpha-p FT-IR spectrometer (Bruker, Billerica, MA, USA). High resolution mass spectra were recorded on a JEOL AccuTOF LC, JMS-T100LP mass spectrometer (JEOL, Tokyo, Japan) using cold electron-spray ionization (CSI) at −40 °C. GC measurements were performed on a Shimadzu GC-17A Gas Chromatograph (Shimadzu Corporation, Kyoto, Japan) with a Supelco SPB-1 fused silica capillary column. Tris-2-(3-methylindolyl)phosphine [14], L1H [14], L2H [14] and potassium graphite (KC8) [39] were prepared according to literature procedures. Chlorobis[4-(trifluoromethyl)phenyl]phosphine (Alfa Aesar, Haverhill, MA, USA) and chlorobis[4-methoxyphenyl]phosphine (Sigma Aldrich, St. Louis, MO, USA) are commercially available chemicals and were used as received.
Tris-2-(3-methyl-N-di[4-(trifluoromethyl)phenyl]phosphinoindolyl)phosphine (L1CF3) was prepared in the analogous way to L1H using chlorobis[4-(trifluoromethyl)phenyl]phosphine: To a solution of tris-2-(3-methylindolyl)phosphine (1.9 g, 4.53 mmol) in THF (50 mL) was added n-BuLi (2.5 M in hexanes, 5.7 mL, 14.3 mmol) at −78 °C. The resulting solution was stirred for 1 h and chlorobis[4-(trifluoromethyl)phenyl]phosphine (3.5 mL, 14.0 mmol) was added. The reaction mixture was stirred for 16 h and allowed to warm slowly to room temperature. The resulting suspension was concentrated in vacuo and dissolved in CH2Cl2 (total amount of 50 mL including washing of the pads). The suspension was filtered through a pad of basic alumina and subsequently through a pad of SiO2. The solvent was removed under reduced pressure and resulted in a yellow foam. This yellow foam was purified by SiO2 chromatography using a gradient from pure hexane to 2% Et2O in hexane. Yield: 2.66 g (42%). 1H NMR (300 MHz CDCl3) δ: 7.44 (d, J = 8.6 Hz, 15H), 7.33–7.22 (m, 6H), 7.21–7.14 (m, 6H), 7.11 (t, J = 7.6 Hz, 3H), 6.93 (t, J = 7.8 Hz, 3H), 6.59 (d, J = 8.4 Hz, 3H), 2.14 (s, 9H). 19F NMR (282 MHz, CDCl3) δ = −60.78 ppm. 31P NMR (121 MHz, CDCl3) δ: 32.63 (d, J = 151.2 Hz, 3P), −76.09 (q, J = 151.6 Hz, 1P). 13C NMR (75 MHz, CDCl3) δ: 1140.28, 140.13, 139.77, 139.53, 139.36, 133.67, 131.91, 131.71, 131.62, 131.43, 129.27, 125.60, 125.53, 125.42, 125.33, 123.72, 122.00, 121.21, 119.50, 113.97, 9.94 ppm.
Tris-2-(3-methyl-N-di[4-methoxyphenyl]phosphinoindolyl)phosphine (L1OMe) was prepared in the analogous way to L1H using chlorobis[4-methoxyphenyl]phosphine: To a solution of tris-2-(3-methylindolyl)phosphine (2.20 g, 5.22 mmol) in THF (50 mL) was added n-BuLi (2.5 M in hexanes, 6.6mL, 16.44 mmol) at −78 °C. The resulting solution was stirred for 1 h and chlorobis[4-methoxyphenyl]phosphine (4.83 g, 17.23 mmol) dissolved in 10 mL THF was added. The reaction mixture was stirred for 3 days and allowed to warm to room temperature. The resulting suspension was concentrated in vacuo, dissolved in CH2Cl2 (total amount of 50 mL including washing of the pads) and filtered through a pad of basic alumina. Evaporation of the solvent and trituration with Et2O yielded a white solid which was recrystallized from vapor diffusion evaporation of Et2O to a concentrated THF solution. Washing with Et2O and drying under vacuum yielded the ligand in pure form. Yield: 4.03 g (67%). 1H NMR (300 MHz, (CD3)2CO): δ 7.46 (d, J = 7.9, 1.0 Hz, 3H), 7.39–7.27 (m, 6H), 7.20–7.07 (m, 6H), 7.05–6.95 (t, 3H), 6.94–6.81 (m, 12H), 6.77 (d, 6H), 3.74 (s, 18H), 2.01 (s, 9H) ppm. 31P NMR (122 MHz, (CD3)2CO): δ 36.83 (d, J = 166.0 Hz, 3P), −76.21 (q, J = 166.7, 1P) ppm. 13C NMR (75 MHz, (CD3)2CO) δ: 161.42, 141.33, 141.16, 134.63, 134.33, 134.06, 134.03, 127.95, 127.78, 127.58, 127.38, 124.22, 123.07, 120.83, 119.56, 115.06, 114.91, 114.86, 114.81, 114.77, 55.52, 9.87 ppm.
RuCl2L1CF3: L1CF3 (1.3732 g, 0.99 mmol) and [Ru(η6-benzene)Cl(μCl)]2 (249.3 mg, 0.50 mmol) were suspended in THF (4 mL) and toluene (8 mL) and stirred at 60 °C for 64 h. After cooling, the yellow precipitated complex was filtered, washed with toluene (1 × 2 mL) and pentane (3 × 3 mL), and dried overnight in the vacuum oven at 40 °C. Yield: 0.9914 g (64%) of a yellow solid. Recrystallization of the complex by layering a DCM solution with pentane at 5 °C gave crystals suitable for X-ray diffraction analysis. 1H NMR (300 MHz, CDCl3): δ 7.72 (dd, J = 10.8, 8.0 Hz, 3H), 7.59 (dt, J = 10.5, 5.3 Hz, 4H), 7.53–7.31 (m, 8H), 7.24 (t, J = 7.6 Hz, 4H), 7.18–7.01 (m, 7H), 6.99–6.81 (m, 7H), 6.11 (d, J = 8.5 Hz, 2H), 5.86 (d, J = 8.5 Hz, 1H), 2.93 (s, 6H), 2.69 (s, 3H) ppm. 31P NMR (122 MHz, CDCl3): δ 99.83 (dt, J = 31.5, 26.4 Hz, 1P), 78.90 (t, J = 26.3 Hz, 2P), 49.00 (dt, J = 30.3, 26.7 Hz, 1P) ppm. 13C NMR (75 MHz, CDCl3): δ 134.83, 134.73, 133.95, 133.82, 132.74, 132.63, 132.31, 132.18, 131.43, 131.31, 129.54, 128.96, 128.73, 128.49, 128.18, 128.05, 127.88, 127.74, 127.68, 127.58, 127.44, 126.91, 126.76, 126.44, 126.30, 125.15, 124.71, 124.64, 122.90, 122.45, 122.32, 120.76, 120.22, 117.80, 115.44, 115.00, 13.64, 12.27, 11.99. ppm. Mass Analysis (CSI) [C69H45Cl1F18N3P4Ru]+: calc: 1518.1024 found 1518.1067.
RuCl2L1OMe: L1OMe (1.02 g, 0.88 mmol) and [Ru(η6-benzene)Cl(μCl)]2 (227.9 mg, 0.45 mmol) were suspended in THF (4 mL) and toluene (8 mL) and stirred at 120 °C for 3 days leaving a pale brown solution. Upon cooling of the reaction mixture, an off-white solid precipitated. This solid was filtered and washed with hexane (3 × 3 mL) and dried overnight in the vacuum oven at 40 °C, resulting in the product as a brown-white solid. Yield: 1.15 g of an off-white solid (97.8%). 1H NMR (300 MHz, CDCl3): δ 7.66 (d, J = 8.0 Hz, 2H), 7.58 (d, J = 7.9 Hz, 1H), 7.44–7.32 (m, 4H), 7.32–7.24 (m, 1H), 7.24–7.02 (m, 10H), 6.89 (q, J = 7.8, 6.9 Hz, 4H), 6.77 (t, J = 7.8 Hz, 1H), 6.68 (d, J = 8.5 Hz, 4H), 6.28 (d, J = 8.5 Hz, 2H), 6.24 (d, J = 7.9 Hz, 2H), 6.12 (d, J = 8.5 Hz, 4H), 5.90 (d, J = 8.5 Hz, 1H), 3.70 (d, J = 1.8 Hz, 12H), 3.62 (s, 6H), 2.84 (s, 6H), 2.60 (s, 3H) ppm. 31P NMR (121 MHz, CDCl3): δ 107.24 (dt, J = 32.0, 26.0 Hz, 1P), 83.78 (t, J = 26.6 Hz, 2P), 54.51 (dt, J = 30.9, 26.9 Hz, 1P) ppm. 13C{1H,31P} NMR (75 MHz, CDCl3): δ = 159.95, 159.46, 139.85, 139.70, 135.39, 135.35, 135.27, 133.02, 132.98, 132.86, 123.72, 123.69, 121.47, 121.27, 120.20, 117.09, 113.38, 113.29, 112.88, 112.72, 112.03, 111.96, 111.89, 77.43, 77.00, 76.58, 54.90, 54.82, 54.79, 11.92, 11.68 ppm. Mass Analysis (CSI) [C69H63ClN3O6P4Ru]+: calc: 1290.2399 found: 1290.2400.
RuCl2L2H: L2H (587.3 mg, 0.60 mmol) and [Ru(η6-benzene)Cl(μ-Cl)]2 (150.1 mg, 0.30 mmol) were suspended in THF (3 mL) and toluene (6 mL) and stirred at 120 °C for 64 h. After cooling, 20 mL of hexane was added, which resulted in the precipitation of a yellow solid. The solid was filtered, washed with hexane (3 × 3 mL), and dried overnight in the vacuum oven at 60 °C. Yield: 477.5 mg (69.1%) of a yellow solid. Recrystallization of the complex by layering a DCM solution with pentane at 5 °C gave crystals suitable for X-ray diffraction analysis. 1H NMR (300 MHz, CDCl3): δ 7.90 (dd, J = 11.6, 7.7 Hz, 2H), 7.84–7.65 (m, 6H), 7.60 (d, J = 7.8 Hz, 1H), 7.53 (t, J = 9.3 Hz, 3H), 7.45 (d, J = 7.8 Hz, 1H), 7.37 (t, J = 7.6 Hz, 1H), 7.32–7.12 (m, 9H), 7.11–6.92 (m, 6H), 6.90–6.75 (m, 7H), 6.70 (d, J = 8.4 Hz, 1H), 6.36 (t, J = 8.4 Hz, 3H), 6.23 (dd, J = 7.5 Hz, 2H), 1.97 (s, 3H), 1.72 (s, 3H), 1.40 (s, 3H) ppm. 31P NMR (121 MHz, CDCl3): δ 144.43 (dt, J = 38.9, 37.8 Hz, 1P), 39.22 (dt, J = 45.0, 23.6 Hz, 1P), 24.70 (ddd, J = 328.6, 31.0, 21.1 Hz, 1P), 16.62 (ddd, J = 331.4, 39.7, 25.6 Hz, 1P) ppm. 13C NMR (75 MHz, CDCl3): δ 134.83, 134.73, 133.95, 133.82, 132.74, 132.63, 132.31, 132.18, 131.43, 131.31, 129.54, 128.73, 128.49, 128.18, 128.05, 127.88, 127.74, 127.68, 127.58, 127.44, 126.91, 126.76, 126.44, 126.31, 125.15, 124.71, 124.64, 122.90, 122.45, 122.32, 120.76, 120.22, 117.80, 115.44, 115.00, 13.64, 12.27, 11.99 ppm. Mass Analysis (CSI) [C63H51ClN3P4Ru]+: calc: 1110.1789; found: 1110.1779.

3.1. Reduction of RuCl2L to RuN2L

RuCl2L (±17 µmol) and KC8 (2–3 equiv) were transferred to a flame-dried Schlenk flask in the glovebox. THF (2 mL) was added and the suspension was stirred for two to three hours; after filtration, a red solution was obtained. Part of this solution was used for in-situ infrared spectrometry, part of the solution was used for in-situ 31P NMR analysis, and the rest of the solution was set for crystallization via slow diffusion evaporation of pentane. Applying this procedure to RuCl2L1OMe led to formation of a purple solution after filtration, which did not show an infrared band after two hours. However, after one day, the solution turned more red and the corresponding infrared band could be detected. The infrared analysis of the thus obtained solutions yielded the corresponding infrared bands as given in Table 2. The 31P-NMR analysis showed the presence of one doublet and one quartet, indicative of a symmetric species, which corresponds with the proposed trigonal bipyramidal geometry.

3.2. Catalysis: Dinitrogen Reduction

The catalyst (33 µmol) and KC8 (100 equiv) were transferred to a 100 mL flame-dried Schlenk flask equipped with a glass stirring bar in the glovebox. 10 mL of dry THF was added, followed by the addition of the chlorotrimethylsilane (100 equiv). The suspension was stirred overnight. A quantity of 10 µL of n-decane was added as internal standard, whereafter a filtered aliquot of the reaction mixture was measured on the Gas-GC.

3.3. Catalysis in Time: Dinitrogen Reduction

The same procedure as for the normal nitrogen reduction was used, adding the 10 µL of n-decane right away and taking a filtered aliquot of the reaction mixture every hour, which was subsequently measured on the Gas-GC.

3.4. Catalysis: Formic Acid Dehydrogenation

Hydrogen evolution was initiated by the addition of formic acid (45 μL, 1.2 mmol) to a solution of catalyst (±8.5 μmol) in refluxing THF (3 mL) in a 10 mL reaction flask equipped with a condenser. The gas was cooled via a condenser, with the outlet connected to a burette filled with water (Figure 15). The displacement of the water level in the burette was measured in time. The second burette was also filled with water and used to compensate for pressure buildup, keeping the water levels at equal height. The results are shown in Table 5.
TOFH2 was calculated using Equation (1) [40].
T O F H 2 = V o b s · 0.5 V m · n c a t
Vobs: measured gas volume displacement cylinder [mL]; Vm: molar gas volume: 24.49 [mL/mmol]; ncat: amount of catalyst [mmol].

3.5. X-ray Diffraction Analysis

Crystallographic data was obtained using a Bruker D8 Quest Eco diffractometer equipped with a Triumph monochromator. The intensities were integrated with the SAINT software package [41]. Multiscan absorption correction and scaling was performed with SADABS [42]. The structure was solved with Intrinsic Phasing Methods using SHELXT [43]. Least-squares refinement was performed with SHELXL 2013 [44] against F2 of all reflections. Non-hydrogen atoms were refined freely with anisotropic displacement parameters. All hydrogen atoms were located in difference Fourier maps and refined with a riding model. All structures have solvent-accessible voids filled with disordered solvent. Their contribution to the structure factors in the refinement was taken into account with the PLATON/SQUEEZE approach [45] (Table 6).
RuCl2L1CF3: The structure contains voids (2544 Å3 per unit cell) filled with disordered solvent molecules. Their contribution to the structure factors was secured by back-Fourier transformation using the SQUEEZE routine of the PLATON package [45], resulting in 671 electrons per unit cell which corresponds to 16 disordered molecules of CH2Cl2 per unit cell. Alerts A and most Alerts B generated by the IUCr checkCIF program resulted from large displacement parameters of F atoms implying intense rotation of the CF3 groups of the L1CF3 ligand. Thirteen FCF reflections below θ-min were obscured by the beam-stop which generated Alert B.
RuCl2L2H: The structure contains voids (1518 Å3 per unit cell) filled with disordered solvent molecules. Their contribution to the structure factors was secured by back-Fourier transformation using the SQUEEZE routine of the PLATON package [45], resulting in 503 electrons per unit cell which corresponds to 12 disordered molecules of CH2Cl2 per unit cell. Eleven FCF reflections below θ-min were obscured by the beam-stop which generated Alert B by the IUCr checkCIF program.
RuN2L1CF3: The structure contains voids (2224 Å3 per unit cell) filled with disordered solvent molecules. Their contribution to the structure factors was secured by back-Fourier transformation using the SQUEEZE routine of the PLATON package [45], resulting in 514 electrons per unit cell which corresponds to 12 disordered molecules of pentane per unit cell. Alerts A generated by the IUCr checkCIF program resulted from large displacement parameters of F atoms implying intense rotation of the CF3 groups of the L1CF3 ligand. Thirteen FCF reflections below θ-min were obscured by the beam-stop which generated Alert B.
CCDC 1574140, 1574141, 1574142 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge CB2 1EZ, UK; Fax: +44-1223-336033; E-mail: [email protected]).

4. Conclusions

In conclusion, we have shown the formation of new ruthenium complexes based on a tripodal 3-methylindolephosphine scaffold. The ligand has two structural (“linkage”) isomers, and, for one of those isomers, several analogues with electron-withdrawing and -donating groups were prepared. These ligands were successfully coordinated to [Ru(η6-benzene)Cl(μ-Cl)]2, resulting in the formation of octahedral complexes with the two chlorides in cis position. The stoichiometric reduction of these complexes with KC8 yielded the corresponding dinitrogen complexes. Electronic effects in the ligands translated to a change in the N≡N stretch frequency of the coordinated dinitrogen. The complexes were studied in the N2 reduction with chlorosilanes and KC8, yielding stoichiometric amounts of the silylamines. When the reaction was followed in time, an incubation period was observed. This incubation period, and the near-identical activities found for all complexes tested, suggests that these species are mere pre-catalysts. The complexes were also employed for formic acid dehydrogenation. Turnover frequencies between 33 and 124 h−1 were reached, depending on the ligand used, clearly showing that the electronic properties of the ligand do have an effect in this reaction. Further studies with these complexes have to be performed to elucidate the reaction mechanism and stability of the catalysts.

Supplementary Materials

The following are available online at www.mdpi.com/2304-6740/5/4/73/s1, Figures S1–S4, cif and cif-checked files.

Acknowledgments

We thank the National Research School Combination Catalysis (NRSC-C) and the Netherlands Organization for Scientific Research (NWO-CW, VENI grant 722.013.002 for W.I.D.) and the University of Amsterdam (RPA Sustainable Chemistry) for funding. We thank Jan-Meine Ernsting for assistance with NMR spectroscopy and Ed Zuidinga for mass spectrometry measurements.

Author Contributions

Fenna F. van de Watering, Wojciech I. Dzik, Bas de Bruin and Joost N. H. Reek conceived and designed the experiments; Fenna F. van de Watering performed the experiments; Nicol Heijtbrink performed control experiments; Fenna F. van de Watering, Bas de Bruin, Jarl Ivar van der Vlugt and Wojciech I. Dzik analyzed the data; Fenna F. van de Watering, Jarl Ivar van der Vlugt, Wojciech I. Dzik, Bas de Bruin and Joost N. H. Reek wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Field, L.D.; Messerle, B.A.; Smernik, R.J. Synthesis and Properties of Iron(II) Hydride Complexes Containing the Tripodal Tetraphosphine Ligand P(CH2CH2PMe2)3. Inorg. Chem. 1997, 36, 5984–5990. [Google Scholar] [CrossRef] [PubMed]
  2. Bruneau, C.; Dixneuf, P.H. Ruthenium in Catalysis; Bruneau, C., Dixneuf, P.H., Eds.; Springer: Rennes, France, 2014; Volume 48. [Google Scholar]
  3. Bianchini, C. Stoichiometric and catalytic functionalization reactions of alkynes at transition metal complexes stabilized by tripodal polyphosphine ligands. Pure Appl. Chem. 1991, 63, 829–834. [Google Scholar] [CrossRef]
  4. Yandulov, D.V.; Schrock, R.R.; Rheingold, A.L.; Ceccarelli, C.; Davis, W.M. Synthesis and Reactions of Molybdenum Triamidoamine Complexes Containing Hexaisopropylterphenyl Substituents. Inorg. Chem. 2003, 42, 796–813. [Google Scholar] [CrossRef] [PubMed]
  5. Mankad, N.P.; Whited, M.T.; Peters, J.C. Terminal FeI–N2 and FeII···H–C Interactions Supported by Tris(phosphino)silyl Ligands. Angew. Chem. Int. Ed. 2007, 46, 5768–5771. [Google Scholar] [CrossRef] [PubMed]
  6. Moret, M.E.; Peters, J.C. Terminal Iron Dinitrogen and Iron Imide Complexes Supported by a Tris(phosphino)borane Ligand. Angew. Chem. Int. Ed. 2011, 50, 2063–2067. [Google Scholar] [CrossRef] [PubMed]
  7. Creutz, S.E.; Peters, J.C. Catalytic Reduction of N2 to NH3 by an Fe–N2 Complex Featuring a C–Atom Anchor. J. Am. Chem. Soc. 2014, 136, 1105–1115. [Google Scholar] [CrossRef] [PubMed]
  8. Siedschlag, R.B.; Bernales, V.; Vogiatzis, K.D.; Planas, N.; Clouston, L.J.; Bill, E.; Gagliardi, L.; Lu, C.C. Catalytic Silylation of Dinitrogen with a Dicobalt Complex. J. Am. Chem. Soc. 2015, 137, 4638–4641. [Google Scholar] [CrossRef] [PubMed]
  9. Del Castillo, T.J.; Thompson, N.B.; Peters, J.C. A Synthetic Single-Site Fe Nitrogenase: High Turnover, Freeze-Quench 57Fe Mössbauer Data, and a Hydride Resting State. J. Am. Chem. Soc. 2016, 138, 5341–5350. [Google Scholar] [CrossRef] [PubMed]
  10. Yandulov, D.V.; Schrock, R.R. Catalytic reduction of dinitrogen to ammonia at a single molybdenum center. Science 2003, 301, 76–78. [Google Scholar] [CrossRef] [PubMed]
  11. Li, Y.-N.; Ma, R.; He, L.-N.; Diao, Z.-F. Homogeneous hydrogenation of carbon dioxide to methanol. Catal. Sci. Technol. 2014, 4, 1498–1512. [Google Scholar] [CrossRef]
  12. Wang, W.H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. CO2 Hydrogenation to Formate and Methanol as an Alternative to Photo- and Electrochemical CO2 Reduction. Chem. Rev. 2015, 115, 12936–12973. [Google Scholar] [CrossRef] [PubMed]
  13. Federsel, C.; Jackstell, R.; Boddien, A.; Laurenczy, G.; Beller, M. Ruthenium-Catalyzed Hydrogenation of Bicarbonate in Water. ChemSusChem 2010, 3, 1048–1050. [Google Scholar] [CrossRef] [PubMed]
  14. Wassenaar, J.; de Bruin, B.; Siegler, M.A.; Spek, A.L.; Reek, J.N.H.; van der Vlugt, J.I. Activation of H2 by a highly distorted RhII complex with a new C3-symmetric tripodal tetraphosphine ligand. Chem. Commun. 2010, 46, 1232–1234. [Google Scholar] [CrossRef] [PubMed]
  15. Wassenaar, J.; Siegler, M.A.; Spek, A.L.; De Bruin, B.; Reek, J.N.H.; van der Vlugt, J.I. Versatile New C3-Symmetric Tripodal Tetraphosphine Ligands; Structural Flexibility to Stabilize CuI and RhI Species and Tune Their Reactivity. Inorg. Chem. 2010, 49, 6495–6508. [Google Scholar] [CrossRef] [PubMed]
  16. van de Watering, F.F.; van der Vlugt, J.I.; Dzik, W.I.; de Bruin, B.; Reek, J.N.H. Metalloradical Reactivity of RuI and Ru0 Stabilized by an Indole-Based Tripodal Tetraphosphine Ligand. Chem. Eur. J. 2017, 23, 12709–12713. [Google Scholar] [CrossRef] [PubMed]
  17. Yu, J.O.; Browning, C.S.; Farrar, D.H. Tris-2-(3-methylindolyl)phosphine as an anion receptor. Chem. Commun. 2008, 1020–1022. [Google Scholar] [CrossRef] [PubMed]
  18. Yu, J.O.; Lam, E.; Sereda, J.L.; Rampersad, N.C.; Lough, A.J.; Browning, C.S.; Farrar, D.H. 2-Indolylphosphines, a New Class of Tunable Ligands: Their Synthesis, Facile Derivatization, and Coordination to Palladium(II). Organometallics 2005, 24, 37–47. [Google Scholar] [CrossRef]
  19. Yu, J.O. An Exploration of the Structural, Electronic, and Anion Binding Properties of 2-Indolylphosphines. Ph.D. Thesis, University of Toronto, Toronto, ON, Canada, 2008. [Google Scholar]
  20. Ciclosi, M.; Lloret, J.; Estevan, F.; Lahuerta, P.; Sanaú, M.; Pérez-Prieto, J. A C3-Symmetric Palladium Catalyst with a Phosphorus-Based Tripodal Ligand. Angew. Chem. Int. Ed. 2006, 45, 6741–6744. [Google Scholar] [CrossRef] [PubMed]
  21. Ciclosi, M.; Estevan, F.; Lahuerta, P.; Passarelli, V.; Perez-Prieto, J.; Sanaú, M. Synthesis and reactivity of the novel hydride derivative RhHCl(TIMP3) (HTIMP3 = tris[1-(diphenylphosphino)-3-methyl-1H-indol-2-yl]methane). Dalton Trans. 2009, 2290–2297. [Google Scholar] [CrossRef] [PubMed]
  22. Penno, D.; Koshevoy, I.O.; Estevan, F.; Sanaú, M.; Ubeda, M.A.; Pérez-Prieto, J. Synthesis of a New C3-Symmetric Tripodal P4-Tetradentate Ligand and Its Application to the Formation of Chiral Metal Complexes. Organometallics 2010, 29, 703–706. [Google Scholar] [CrossRef]
  23. Kuo, YY.; Haddow, M.F.; Jamieson, A.L.; Owen, G.R. Synthesis, structural characterisation and catalytic application of dichloro(η6-p-cymene){diphenyl(3-methyl-2-indolyl)phosphine}-ruthenium(II) in the transfer hydrogenation of ketones. Transit. Met. Chem. 2013, 38, 641–648. [Google Scholar] [CrossRef]
  24. Imayoshi, R.; Tanaka, H.; Matsuo, Y.; Yuki, M.; Nakajima, K.; Yoshizawa, K.; Nishibayashi, Y. Cobalt-Catalyzed Transformation of Molecular Dinitrogen into Silylamine under Ambient Reaction Conditions. Chem. Eur. J. 2015, 21, 8905–8909. [Google Scholar] [CrossRef] [PubMed]
  25. Yuki, M.; Tanaka, H.; Sasaki, K.; Miyake, Y.; Yoshizawa, K.; Nishibayashi, Y. Iron-catalysed transformation of molecular dinitrogen into silylamine under ambient conditions. Nat. Commun. 2012, 3, 1254. [Google Scholar] [CrossRef] [PubMed]
  26. Dzik, W.I. Silylation of Dinitrogen Catalyzed by Hydridodinitrogentris(Triphenylphosphine)Cobalt(I). Inorganics 2016, 4, 21. [Google Scholar] [CrossRef]
  27. Oldenhof, S.; de Bruin, B.; Lutz, M.; Siegler, M.A.; Patureau, F.W.; van der Vlugt, J.I.; Reek, J.N.H. Base-Free Production of H2 by Dehydrogenation of Formic Acid Using an Iridium–bisMETAMORPhos Complex. Chem. Eur. J. 2013, 19, 11507–11511. [Google Scholar] [CrossRef] [PubMed]
  28. Oldenhof, S.; Lutz, M.; de Bruin, B.; van der Vlugt, J.I.; Reek, J.N.H. Dehydrogenation of formic acid by Ir–bisMETAMORPhos complexes: Experimental and computational insight into the role of a cooperative ligand. Chem. Sci. 2015, 6, 1027–1034. [Google Scholar] [CrossRef]
  29. Oldenhof, S.; van der Vlugt, J.I.; Reek, J.N.H. Hydrogenation of CO2 to formic acid with iridiumIII(bisMETAMORPhos)(hydride): The role of a dormant fac-IrIII(trihydride) and an active trans-IrIII(dihydride) species. Catal. Sci. Technol. 2016, 6, 404–408. [Google Scholar] [CrossRef]
  30. Jongbloed, L.S.; de Bruin, B.; Reek, J.N.H.; Lutz, M.; van der Vlugt, J.I. Reversible cyclometalation at RhI as a motif for metal–ligand bifunctional bond activation and base-free formic acid dehydrogenation. Catal. Sci. Technol. 2016, 6, 1320–1327. [Google Scholar] [CrossRef]
  31. Van de Watering, F.F.; Lutz, M.; Dzik, W.I.; de Bruin, B.; Reek, J.N.H. Reactivity of a Ruthenium–Carbonyl Complex in the Methanol Dehydrogenation Reaction. ChemCatChem 2016, 8, 2752–2756. [Google Scholar] [CrossRef] [PubMed]
  32. De Boer, S.Y.; Korstanje, T.J.; La Rooij, S.R.; Kox, R.; Reek, J.N.H.; van der Vlugt, J.I. Ruthenium PNN(O) Complexes: Cooperative Reactivity and Application as Catalysts for Acceptorless Dehydrogenative Coupling Reactions. Organometallics 2017, 36, 1541–1549. [Google Scholar] [CrossRef]
  33. Trincado, M.; Sinha, V.; Rodriguez-Lugo, R.E.; Pribanic, B.; de Bruin, B.; Grützmacher, H. Homogeneously catalysed conversion of aqueous formaldehyde to H2 and carbonate. Nat. Commun. 2017, 8, 14990. [Google Scholar] [CrossRef] [PubMed]
  34. Himeda, Y.; Miyazawa, S.; Hirose, T. Interconversion between formic acid and H2/CO2 using rhodium and ruthenium catalysts for CO2 fixation and H2 storage. ChemSusChem 2011, 4, 487–493. [Google Scholar] [CrossRef] [PubMed]
  35. Scholten, J.D.; Prechtl, M.H.G.; Dupont, J. Decomposition of Formic Acid Catalyzed by a Phosphine-Free Ruthenium Complex in a Task-Specific Ionic Liquid. ChemCatChem 2010, 2, 1265–1270. [Google Scholar] [CrossRef]
  36. Wang, Z.; Lu, S.-M.; Li, J.; Wang, J.; Li, C. Unprecedentedly High Formic Acid Dehydrogenation Activity on an Iridium Complex with an N,N-Diimine Ligand in Water. Chem. Eur. J. 2015, 21, 12592–12595. [Google Scholar] [CrossRef] [PubMed]
  37. Rodríguez-Lugo, R.E.; Trincado, M.; Vogt, M.; Tewes, F.; Santiso-Quinones, G.; Grützmacher, H. A homogeneous transition metal complex for clean hydrogen production from methanol-water mixtures. Nat. Chem. 2013, 5, 342–347. [Google Scholar] [CrossRef] [PubMed]
  38. Mellone, I.; Bertini, F.; Peruzzini, M.; Gonsalvi, L. An active, stable and recyclable Ru(II) tetraphosphine-based catalytic system for hydrogen production by selective formic acid dehydrogenation. Catal. Sci. Technol. 2016, 6, 6504–6512. [Google Scholar] [CrossRef]
  39. Weitz, I.S.; Rabinovitz, M. The application of C8K for organic synthesis: Reduction of substituted naphthalenes. J. Chem. Soc. Perkin Trans. 1993, 1, 117–120. [Google Scholar] [CrossRef]
  40. Nielsen, M.; Alberico, E.; Baumann, W.; Drexler, H.-J.; Junge, H.; Gladiali, S.; Beller, M. Catalysis: A step closer to a methanol economy. Nature 2013, 495, 85–89. [Google Scholar] [CrossRef] [PubMed]
  41. Bruker. SAINT-Plus; Bruker AXS Inc.: Madison, WI, USA, 2001. [Google Scholar]
  42. Sheldrick, G.M. SADABS and TWINABS; Universität Göttingen: Göttingen, Germany, 2008. [Google Scholar]
  43. Sheldrick, G.M. SHELXT; Universität Göttingen: Göttingen, Germany, 2012. [Google Scholar]
  44. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  45. Spek, A.L. Structure validation in chemical crystallography. Acta Cryst. 2009, D65, 148–155. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Synthesis of ligands L1R by reacting tris-2-(3-methylindolyl)phosphine with the corresponding chlorodiarylphosphine.
Figure 1. Synthesis of ligands L1R by reacting tris-2-(3-methylindolyl)phosphine with the corresponding chlorodiarylphosphine.
Inorganics 05 00073 g001
Figure 2. Ruthenium complexes of L1R and L2H.
Figure 2. Ruthenium complexes of L1R and L2H.
Inorganics 05 00073 g002
Figure 3. 31P-NMR spectra of L1R-based ruthenium(II) dichlorido complexes.
Figure 3. 31P-NMR spectra of L1R-based ruthenium(II) dichlorido complexes.
Inorganics 05 00073 g003
Figure 4. ORTEP drawing of RuCl2L1CF3 (CCDC 1574141) showing one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Figure 4. ORTEP drawing of RuCl2L1CF3 (CCDC 1574141) showing one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Inorganics 05 00073 g004
Figure 5. Drawing of bottom view of RuCl2L1CF3 of both isomers, showing only the (3-methylindolyl)phosphine part. Insert: schematic representation of the two rotamers with the 3-methylindolyl moiety twisted clockwise (left) or counter-clockwise (right).
Figure 5. Drawing of bottom view of RuCl2L1CF3 of both isomers, showing only the (3-methylindolyl)phosphine part. Insert: schematic representation of the two rotamers with the 3-methylindolyl moiety twisted clockwise (left) or counter-clockwise (right).
Inorganics 05 00073 g005
Figure 6. 31P NMR spectrum of RuCl2L2H.
Figure 6. 31P NMR spectrum of RuCl2L2H.
Inorganics 05 00073 g006
Figure 7. ORTEP drawing of RuCl2L2H (CCDC 1574140) showing a side view of one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Figure 7. ORTEP drawing of RuCl2L2H (CCDC 1574140) showing a side view of one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Inorganics 05 00073 g007
Figure 8. Bottom view of X-ray structure of RuCl2L2H of both rotamers, showing only the (3-methylindolyl)phosphine part. Insert: schematic representation of the two rotamers where the 3-methylindolyl moiety is twisted clockwise (left) or counter-clockwise (right).
Figure 8. Bottom view of X-ray structure of RuCl2L2H of both rotamers, showing only the (3-methylindolyl)phosphine part. Insert: schematic representation of the two rotamers where the 3-methylindolyl moiety is twisted clockwise (left) or counter-clockwise (right).
Inorganics 05 00073 g008
Figure 9. Reaction of RuCl2L complexes with two equivalents of KC8 in presence of N2 forming the corresponding Ru0(N2)L.
Figure 9. Reaction of RuCl2L complexes with two equivalents of KC8 in presence of N2 forming the corresponding Ru0(N2)L.
Inorganics 05 00073 g009
Figure 10. ORTEP drawing of RuN2L1CF3 (top, CCDC 1574142) showing one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Figure 10. ORTEP drawing of RuN2L1CF3 (top, CCDC 1574142) showing one of the two rotamers. Thermal ellipsoids are set at 50% probability. Solvent molecules and hydrogen atoms have been omitted for clarity.
Inorganics 05 00073 g010
Figure 11. Reduction of dinitrogen with KC8, Me3SiCl and RuCl2L forming (Me3Si)3N.
Figure 11. Reduction of dinitrogen with KC8, Me3SiCl and RuCl2L forming (Me3Si)3N.
Inorganics 05 00073 g011
Figure 12. Dinitrogen reduction over a period of 4 h with catalysts RuCl2L1H and RuCl2L2H. Turnover number (TON) is defined as equivalents of (Me3Si)3N formed per ruthenium atom.
Figure 12. Dinitrogen reduction over a period of 4 h with catalysts RuCl2L1H and RuCl2L2H. Turnover number (TON) is defined as equivalents of (Me3Si)3N formed per ruthenium atom.
Inorganics 05 00073 g012
Figure 13. Formic acid decomposition by the RuL(BF4)2 complexes.
Figure 13. Formic acid decomposition by the RuL(BF4)2 complexes.
Inorganics 05 00073 g013
Figure 14. Hydrogen evolution from formic acid with catalyst RuCl2L1OMe as detected volumetrically with a gas burette. After 1 h, the burette contained 20 mL of H2.
Figure 14. Hydrogen evolution from formic acid with catalyst RuCl2L1OMe as detected volumetrically with a gas burette. After 1 h, the burette contained 20 mL of H2.
Inorganics 05 00073 g014
Figure 15. Hydrogen evolution set up.
Figure 15. Hydrogen evolution set up.
Inorganics 05 00073 g015
Table 1. Selected bond lengths (Å) and angles (°) for RuL complexes.
Table 1. Selected bond lengths (Å) and angles (°) for RuL complexes.
RuCl2L1H [a]RuCl2L1CF3RuCl2L2HRuN2L1HRuN2L1CF3
Ru1–P12.3189(9)2.3097(19)2.3295(8)2.2747(12)2.2613(10)
Ru1–P22.3727(9)2.3445(19)2.4079(8)2.2752(11)2.2702(10)
Ru1–P32.2671(9)2.2383(17)2.2913(8)2.2774(11)2.2554(10)
Ru1–P42.1932(9)2.2023(17)2.1363(8)2.2133(11)2.2193(9)
Ru1–Cl12.4869(9)2.4869(16)2.4829(7)
Ru1–Cl22.4471(9)2.4487(16)2.4705(7)
Ru1–N1 2.011(4)2.066(3)
N1–N2 1.085(5)1.064(5)
P1–Ru1–P2160.04(3)157.78(6)159.24(3)122.85(4)119.79(4)
P1–Ru1–P3102.23(3)101.39(7)101.30(3)115.80(4)119.28(4)
P2–Ru1–P394.27(3)97.42(7)93.60(3)118.33(4)118.16(4)
P1–Ru1–P487.06(3)86.13(6)86.13(3)84.14(4)84.54(4)
P2–Ru1–P482.67(3)83.12(6)80.63(3)83.91(4)84.13(4)
P3–Ru1–P486.85(3)87.02(6)85.62(3)84.63(4)84.73(3)
[a] Taken from reference [16].
Table 2. IR frequency of the selected Ru0(N2)L complexes.
Table 2. IR frequency of the selected Ru0(N2)L complexes.
ComplexνN2 (cm−1)
RuN2L1CF32136
RuN2L1H2125
RuN2L1OMe2113
RuN2L2H2136
Table 3. Formation of (Me3Si)3N with RuCl2L complexes.
Table 3. Formation of (Me3Si)3N with RuCl2L complexes.
ComplexEquiv (Me3Si)3N [a]
RuCl2L1CF30.92 [b]
RuCl2L1H1.83 [b]
RuCl2L1OMe1.74 [b]
RuCl2L2H1.40 [b]
[Ru(η6-benzene)Cl(μ-Cl)]21.29 [c]
-trace
[a] Conditions: ~33 μmol of the complex was suspended together with KC8 (100 equiv) in 10 mL THF whereafter the chlorotrimethylsilane (100 equiv) was added. The suspension was stirred for 23 h at room temperature; [b] Average of two runs; [c] Equivalents per ruthenium atom.
Table 4. Activities of RuCl2L catalysts in the formic acid decomposition reaction.
Table 4. Activities of RuCl2L catalysts in the formic acid decomposition reaction.
ComplexTOF (h−1) [a]
RuCl2L1CF3124
RuCl2L1H76
RuCl2L1OMe121
RuCl2L2H33
[b]
[{RuCl2(p-cymene)}2]1540 [c]
[a] Conditions: ~8.5 µmol preactivated catalyst (2 equiv AgBF4 RuCl2L. 2 h), 3 mL THF, 45 μL (1.2 mmol) HCOOH (catalyst to substrate ratio: 1:140), reflux temperature, 1 h reaction time, Turnover frequency (TOF, defined as number of equivalents of H2 produced by the catalyst in one hour) is the average of two catalytic runs. The gas was collected volumetrically with a burette; [b] Conditions: 3 mL THF, 45 μL (1.2 mmol) HCOOH, reflux temperature, 30 min reaction time. [c] Conditions: ~6.5 µmol catalyst, 7 mmol ionic liquid, 17.6 mmol HCOOH (catalyst to substrate ratio: 1:2700), 80 °C, 7 h reaction time. Data taken from ref. [35].
Table 5. Duplo measurements for formic acid dehydrogenation by the depicted catalysts.
Table 5. Duplo measurements for formic acid dehydrogenation by the depicted catalysts.
ComplexTOF (h−1) #1TOF (h−1) #2TOF (h−1) Average
RuCl2L1CF3120.8127.2124
RuCl2L1H79.371.975.6
RuCl2L1OMe123.5117.4120.5
RuCl2L2H22.442.732.6
Table 6. Crystallographic data for RuCl2L1CF3, RuCl2L2H and RuN2L1H.
Table 6. Crystallographic data for RuCl2L1CF3, RuCl2L2H and RuN2L1H.
ComplexRuCl2L1CF3RuCl2L2HRuN2L1CF3
Empirical formulaC69H45Cl2F18N3P4Ru + solventC63H51Cl2N3P4Ru, 2(CH2Cl2) + solventC69H45F18N5P4Ru + solvent
FW1553.93 a1315.77 a1511.05 a
Temperature [K]150150150
RadiationMo KαMo KαMo Kα
Wavelength [Å]0.710730.710730.71073
Cryst syst.monoclinicmonoclinicmonoclinic
Space groupC 2/cP 21/nP 21/c
a [Å]21.594(3)19.1879(9)14.2851(6)
b [Å]27.753(4)18.1604(8)18.9090(8)
c [Å]23.938(3)19.7879(9)29.8721(12)
α [deg.]909090
β [deg.]93.631(3)91.012(2)97.011(2)
γ [deg.]909090
Volume [Å3]14317(3)6894.2(5)8008.6(6)
Z844
Colorpale yellowyellowdark red
θ-max25.13525.03026.450
Density [kg·m−3]1.442 a1.268 a1.253 a
Absorp. Coeff. [mm−1]0.472 a0.591 a0.356 a
F(000)6240 a2688.03040 a
R1/wR2/S0.0715/0.1975/1.0530.0412/0.1435/1.1340.0666/0.2002/1.436
a Excluding the disordered solvent contribution.

Share and Cite

MDPI and ACS Style

Van de Watering, F.F.; Heijtbrink, N.; Van der Vlugt, J.I.; Dzik, W.I.; De Bruin, B.; Reek, J.N.H. 3-Methylindole-Based Tripodal Tetraphosphine Ruthenium Complexes in N2 Coordination and Reduction and Formic Acid Dehydrogenation. Inorganics 2017, 5, 73. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics5040073

AMA Style

Van de Watering FF, Heijtbrink N, Van der Vlugt JI, Dzik WI, De Bruin B, Reek JNH. 3-Methylindole-Based Tripodal Tetraphosphine Ruthenium Complexes in N2 Coordination and Reduction and Formic Acid Dehydrogenation. Inorganics. 2017; 5(4):73. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics5040073

Chicago/Turabian Style

Van de Watering, Fenna F., Nicol Heijtbrink, Jarl Ivar Van der Vlugt, Wojciech I. Dzik, Bas De Bruin, and Joost N. H. Reek. 2017. "3-Methylindole-Based Tripodal Tetraphosphine Ruthenium Complexes in N2 Coordination and Reduction and Formic Acid Dehydrogenation" Inorganics 5, no. 4: 73. https://0-doi-org.brum.beds.ac.uk/10.3390/inorganics5040073

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop