Next Article in Journal
Methyl 2-(1-methyl-3-oxoisoindolin-1-yl)acetate
Previous Article in Journal
2-(Fluoromethyl)-4,7-dimethoxy-1-methyl-1H-benzimidazole
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Planar-Chiral P and N-bridged Diferrocenes

by
Philipp Honegger
1,*,
Lukas Scheibelberger
2 and
Michael Widhalm
3
1
Institute of Computational Biological Chemistry, University of Vienna, Währinger Straße 17, 1090 Vienna, Austria
2
Institute of Organic Chemistry, University of Vienna, Währinger Straße 38, 1090 Vienna, Austria
3
Institute of Chemical Catalysis, University of Vienna, Währinger Straße 38, 1090 Vienna, Austria
*
Author to whom correspondence should be addressed.
Molbank 2020, 2020(2), M1130; https://doi.org/10.3390/M1130
Submission received: 1 April 2020 / Revised: 25 April 2020 / Accepted: 29 April 2020 / Published: 4 May 2020
(This article belongs to the Section Organic Synthesis)

Abstract

:
We describe the synthesis of planar-chiral diferrocene compounds 4, 7 and 12, intended as key precursors for a new family of asymmetric catalyst ligands with less complex structures than their popular equivalents. In contrast to conventional 1,2-disubstituted ferrocenes containing center-chiral and planar-chiral elements, these compounds are purely planar-chiral due to the absence of α-substiuted ethyl groups. The title compounds 4, 7 and 12 were obtained from known precursors in 87%, 59% and 60% yields, respectively.

Graphical Abstract

1. Introduction

Disubstituted ferrocenyl compounds have been used as powerful asymmetry-inducing units in organic synthesis; if one cyclopentadienyl ring is substituted with two different functional groups, the resulting compound displays planar chirality [1,2]. Ferrocene-based ligands have been used successfully in asymmetric catalysis [3,4,5,6,7]. Their range of applications is impressive, and beyond academic research, ligands like Xyliphos (Scheme 1) have been employed in industrial-scale asymmetric hydrogenations [8].
The majority of their syntheses are based on 1-N,N-dimethylamino-ethylferrocene, well known as Ugi’s amine. Its popularity is based on simple large-scale preparation of both enantiomerically pure forms and perfect stereocontrol of ortho-metallation/substitution, enabling the diastereoselective synthesis of planar-chiral disubstituted ferrocenyl compounds [9]. In a subsequent reaction, the dimethylamino group at the α-carbon of the ethyl side chain can be displaced by other nucleophiles stereo-retentively [10]. This is shown in Scheme 2.
In many cases, the asymmetric induction may be even improved further by combining two of these asymmetric ferrocene units in one molecule. For instance, the P,P,P-diferrocene ligand Pigiphos [11] (Scheme 1) has been employed in asymmetric acetalization [12], hydroamination [13,14,15,16], hydrophosphination [17], carbonyl reduction [18,19], 1,3-dipolar cycloaddition [20], olefin cyclopropanation [21] and Nazarov cyclization [22]. Stanphos (Scheme 1) is the cyclized version of Pigiphos [23] and has been used as a P,P-ligand in asymmetric hydroalkoxylations [24] and the α-hydroxylation of 1,3-ketoesters [25]. Zirakzadeh et al. developed ferrocene-based PNP-pincer ligands for Fe(II)-catalyzed asymmetric hydrogenation [26]. Their diferrocene compounds are linked by a secondary amine, imine [27] or two secondary imines [28] (Scheme 1).
We attempted to access similar compounds containing oxygen, nitrogen or sulfur at the ethyl side chain. However, we noticed that when replacing phosphorous (R2PPh) with sulfur (R2S) or nitrogen (R2NPh), eliminations yielding vinylferrocenes as decay products become a serious side reaction [29], thereby limiting the practical value of such compounds. In this article, we present two possible synthetic workarounds:
Designing diferrocenyl compounds with methyl instead of ethyl chains, effectively removing the possibility of elimination. As we cannot use the convenient diastereoselective ortho-lithiation nor the stereo-retentive SN1 reaction of Ugi’s amine, Xiao et al. have developed a diastereoselective ortho-lithiation protocol of the chiral methylferrocene equivalent with a homochiral O-methylephedrine auxiliary attached to the amino group [30]. This approach will yield compounds 4 and 7.
Instead of the α-substituted ethyl group, we changed to a β-substituted one in order to disfavor elimination processes. To date, there are barely any examples of said type of ferrocene compounds in the literature. With such structures, we get rid of stereocenters at the ethylene side chains, simplifying the stereochemistry to pure planar chirality. This approach would yield compounds like 12.

2. Results

2.1. N-bridged Diferrocenyl Compound 4

Our synthetic route towards precursors for compounds resembling Pigiphos [11] is illustrated in Scheme 3. First, the ferrocenyl-O-methylephedrine derivative 1 was diastereoselectively lithiated and converted to bromide 2 using a modified procedure with C2Br2Cl4 instead of C2Br2F4 [30]. For removal of the auxiliary ammonium iodide, salt 3 was prepared under standard conditions [30].
The activated leaving group enabled the coupling of two homochiral ferrocene units of configuration Rp with bidentate nucleophiles to yield purely planar-chiral diferrocenyl compounds. We used 0.5 equiv of benzyl amine as a nucleophile to obtain diferrocene amine 4 from 3 following a modified procedure [30], but were not able to obtain the desired compound, instead diferrocenyl amine 4 was obtained in 65% yield in DMF under microwave irradiation (supporting information document, Figures S1–S3).
Encouraged by the outcome, we optimized the yield by performing this transformation in two steps. First, ammonium salt 3 was treated with a large excess of benzyl amine, yielding 90% of monoferrocenyl amine 5 (supporting information document, Figures S4–S7). Subsequently, 5 was coupled with one equiv. of 3 to give 4 in 87% yield over two steps.

2.2. P-bridged Diferrocenyl Compound 7

Similar to diferrocene 8, as published by Hayashi et al. [24,31,32], we planned to couple two auxiliary-substituted ferrocene molecules 1 via Cl2PPh to obtain a phosphorous-linked diferrocene. However, our preliminary experiments failed, possibly due to the steric repulsion of the large ephedrine auxiliary. Instead, we present an efficient workaround to obtain a similar diferrocene precursor.
To overcome the problem of steric hindrance in 1 we replaced the bulky auxiliary prior to coupling with a smaller substituent. The sterically less demanding and purely planar-chiral amine 6 was obtained from ammonium salt 3 and dimethylamine in an autoclave according to reference [30]. We then followed an optimized ortho-lithiation protocol [30] to couple 6 with 0.5 equiv of Cl2PPh and obtained a compound identified as the desired homochiral phosphorous-linked diferrocene (Rp,Rp)-7 in 59% yield (supporting information document, Figures S8–S11) and some dehalogenated product of 6.
Compound 7 strongly resembles Hayashi’s diamine diferrocene compound 8 [24,31,32], but the stereochemical complexity is reduced from configuration (S,S,Rp,Rp) to (Rp,Rp), since it is entirely planar-chiral. We thus propose compound 7 as a new key precursor for less complicated equivalents of ferrocene-based ligands, such as Gipiphos [33] and Stanphos [23,24,25], with the additional advantage of hindered elimination at the alkyl chains.

2.3. 1,2-Substituted Diferrocenyl Phosphinsulfide 12

In order to synthesize further purely planar-chiral diferrocenyl phosphines (Scheme 4), divinyl substituted diferrocenylphosphine 9 was chosen as a key intermediate, easily accessible from diamine (R,R,Sp,Sp)-8 [34] (Scheme 4). To introduce substituents at terminal carbon atoms, hydroboration of the double bond is well established, leading predominantly to anti-Markownikow products.
Since we noticed in a pre-experiment that treatment of 9 with BH3*THF followed by oxidative work-up yielded an inseparable mixture of P-borane complexes with primary and secondary OH groups, we changed to 9-borabicyclo(3.3.1)nonane (9-BBN) to improve the regioselectivity. Treatment with H2O2/OH afforded the desired diol isolated as phosphine oxide(Sp,Sp)-10 (supporting information document, Figures S12–S15), but in moderate yield (25%) and difficult to purify chromatographically due to its high polarity. In contrast, with P-sulfide 11 (supportung information document, Figures S16 and S17) the hydroboration step proceeded more efficiently giving diol (Sp,Sp)-12 in 60% yield (supporting information document, Figures S18–S21). Both new 1,2-substituted diferrocenyl phosphines contained some residual eluent moisture which could not be removed in high vacuum. Drying by heating was not possible since we observed product decomposition.
We have thus produced new homochiral compounds 10 and 12 with the unusual 1,2-substitution pattern rarely published for ethylferrocenes. In contrast to Hayashi’s diferrocenyl precursor 8 with (R,R,Sp,Sp), these only include two planar chiral moieties with (Sp,Sp) configuration, thereby reducing the stereochemical complexity.

3. Materials and Methods

3.1. General

Routine NMR spectra were recorded on a 400 MHz Bruker AVIII 400 spectrometer operating at 400.27 MHz (1H), 100.66 MHz (13C) and 162.04 MHz (31P) with an autosampler. 1H-NMR spectra and 13C{1H}-NMR spectra used for substance characterization were recorded on a Bruker AVIII 700 spectrometer (Bruker Biospin, Billerica, MA, USA) at 700.40 MHz (1H) and 176.13 MHz (13C). 13C{1H}-NMR spectra were recorded in J-modulated mode. NMR chemical shifts are referenced to non-deuterated CHCl3 residual shifts: at 7.26 ppm for 1H-NMR, at 77.00 ppm for CDCl3 for 13C-NMR and at 0.00 ppm for 85% H3PO4 for 31P-NMR. HR-MS were recorded by a Bruker Maxis ESI oa-RTOF mass spectrometer equipped with a quadrupole analyzer ion guide. Melting points were measured on a Reichelt Thermovar Kofler apparatus, uncorrected.
Absolute THF was dried by distillation from sodium benzophenonketyl under Ar. Et2O and n-pentane were dried by distillation over LiAlH4. Reaction progress was monitored by TLC, SiO2 sheets with F254 fluorescent indicator. Heating by microwave irradiation was performed in a START-1500 oven (EMLS microwave laboratory systems, Leutkirch, Germany). Preparative column chromatography was carried out using a Biotage Isolera One automated flash chromatography instrument with self-packed columns (Macherey-Nagel silica gel 60M, particle size 40–63 μm). All the other chemicals were analytical grade and used without further purification.
Published procedures were followed to obtain 3 [30], 6 [30], 8 [24] and 9 [34].

3.2. Synthesis

3.2.1. N-(2Rp-Bromoferrocenylmethyl)-N-methyl-1-methoxy-1-phenylprop-2-ylamine 2

Ferrocenyl compound 1 (synthesized according to reference [30]) (3.396 g, 9.00 mmol) was dissolved in 90 mL of dry n-pentane under Ar. The solution was cooled to −78 °C. To the suspension 6.4 mL of tert-BuLi (1.7 M, 10.9 mmol, 1.2 eq) was added dropwise and the reaction mixture was stirred for 1.5 h at −78 °C, then for 2.5 h at −30 °C. The reaction mixture was cooled to −78 °C, and 4.40 g of 1,2-dibromotetrachloroethane (13.50 mmol, 1.5 eq) dissolved in 13.5 mL of dry THF was added dropwise. The mixture was stirred at −78 °C for 20 min and allowed to warm up to RT overnight. To the reaction mixture, 90 mL of saturated aqueous NaHCO3 solution was added and the organic layer was separated. The aqueous layer was repeatedly extracted with Et2O (2 × 100 mL, 1 × 50 mL). The combined organic fractions were washed with water and brine and dried over MgSO4. The solvent was removed under reduced pressure and the residue was purified by column chromatography (SiO2; 17% (Et2O + 1.5% Et3N) in heptane) yielding 3.513 g of 2 (86%) as a dark orange oil. NMR spectra are in agreement with reference [30].

3.2.2. N,N-Bis(2Rp-bromoferrocenylmethyl)benzylamine 4

Route 1: Ammonium iodide salt 3 [30] (100 mg, 0.17 mmol) was suspended in 8 mL of benzene and 2 mL of DMF in a glass tube. To the suspension, 9 μL BnNH2 (0.08 mmol, 0.5 eq) was added and the suspension was stirred under microwave irradiation at 110 °C for 10 min. The reaction mixture was cooled to RT, volatiles were removed under reduced pressure and 25 mL of Et2O was added. The organic mixture was washed with water (2 × 30 mL) and brine (1 × 30 mL), and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified via column chromatography (SiO2; 5–15% (Et2O + 1.5% Et3N) in heptane) yielding 35.7 mg of amine 4 (65%) as an orange glassy solid.
Route 2: Amine 5 (250 mg, 0.65 mmol) was dissolved in 8 mL of benzene and 2 mL of DMF in a glass tube. To the suspension, 389 mg of ammonium iodide salt 3 (0.65 mmol, 1.00 eq) was added and the mixture was stirred under microwave irradiation at 110 °C for 15 min. The reaction mixture was cooled to RT, volatiles were removed under reduced pressure and 35 mL of Et2O was added. The organic layer was washed with water (2 × 40 mL) and brine (1 × 40 mL), and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified by column chromatography (SiO2, 5–15% (Et2O + 1.5% Et3N) in heptane) yielding 417 mg of amine 4 (97%).
1H-NMR (700 MHz, CDCl3) δ = 7.35 (m, 2H); 7.30 (m, 2H); 7.22 (m, 1H); 4.40 (m, 2H); 4.22 (m, 2H); 4.06 (m, 2H); 4.04 (s, 10H); 3.60 (d, J = 14.2 Hz, 1H); 3.57 (s, 4H); 3.47 (d, J = 14.1 Hz, 1H) ppm. 13C-NMR (700 MHz, CDCl3) δ = 140.16 (Cq); 128.64 (CH); 128.06 (CH); 126.68 (CH); 83.66 (Cq); 80.45 (Cq); 71.12 (CH); 69.95 (CH); 68.35 (CH); 66.03 (CH); 56.71 (CH2); 51.74 (CH2) ppm. HR-MS: m/z calculated for C29H27Br2Fe2N; [M]+ 658.9209, found: 658.9226.

3.2.3. N-(2Rp-Bromoferrocenylmethyl)benzylamine 5

Ammonium iodide salt 3 (100 mg, 0.17 mmol) was suspended in 8 mL of benzene and 1 mL of DMF in a glass tube, and 182 μL of BnNH2 (1.67 mmol, 10 eq) was added. The suspension was stirred under microwave irradiation at 110 °C for 15 min. The reaction mixture was cooled to RT, volatiles were removed under reduced pressure and 25 mL of Et2O was added. The organic layer was washed with water (2 × 30 mL) and brine (1 × 30 mL), and dried over MgSO4. The solvent was removed under reduced pressure and the crude product was purified by column chromatography (SiO2; 10–30% (Et2O + 1.5% Et3N) in heptane) yielding 58 mg of amine 5 (90%) as a yellow oil. 1H-NMR (700 MHz, CDCl3) δ = 7.36–7.32 (m, 4H); 7.28–7.24 (m, 1H); 4.42 (m, 1H); 4.21 (m, 1H); 4.14 (s, 5H); 4.09 (pt, J = 2.5 Hz, 1H); 3.84 (d, J = 13.3 Hz, 1H); 3.79 (d, J = 13.4 Hz, 1H); 3.78 (d, J = 13.3 Hz, 1H); 3.55 (d, J = 13.4 Hz, 1H) ppm. 13C-NMR (700 MHz, CDCl3) δ = 140.24 (Cq); 128.38 (CH); 128.09 (CH); 126.91 (CH); 84.90 (Cq); 79.80 (Cq); 70.98 (Cq); 70.11 (CH); 67.32 (CH); 65.98 (CH); 53.06 (CH2); 46.94 (CH2) ppm. HR-MS: m/z calculated for C18H18BrFeN; [M + H]+ 384.0050, found: 384.0042.

3.2.4. 1,1ʺ-(Phenylphosphinidene)bis[(2Rp)-2-[(dimethylamino)-methyl]]ferrocene 7

Amine 6 [30] (1.042 g, 3.236 mmol) was dissolved in 3 mL of dry Et2O under Ar. The yellow-to-orange solution was cooled to −40 °C and 2 mL of n-BuLi (1.6 M; 3.2 mmol, 1.0 eq) was added dropwise to the reaction mixture. The solution was stirred at this temperature for 2 h, and then cooled to −78 °C. To the solution, 220 μL of Cl2PPh (1.62 mmol, 0.50 eq) was added. The suspension was stirred for another 20 min at this temperature and allowed to warm to RT overnight. The reaction mixture was quenched with 5 mL of aqueous saturated NaHCO3 solution. Residues were dissolved by adding some DCM and the aqueous layer was extracted with DCM (2 × 10 mL). The combined organic fractions were washed with water (3 × 35 mL) and brine (3 × 35 mL), and dried over MgSO4. The solvent was removed under reduced pressure and the residue was purified by column chromatography (SiO2, 65–100% (Et2O + 1.5% Et3N) in heptane) yielding 561 mg diaminophosphine 7 (59%) as a brown oil. 1H-NMR (600 MHz, CDCl3) δ = 7.65–7.61 (m, 2H); 7.31–7.26 (m, 3H); 4.47 (m, 1H); 4.41 (m, 1H); 4.33 (pt, J = 2.4 Hz, 1H); 4.29 (m, 1H); 4.19 (pt, J = 2.4 Hz, 1H); 4.12 (m, 1H); 4.06 (s, 5H); 3.97 (dd, J = 13.2, 2.2 Hz, 1H); 3.67 (s, 5H); 3.47 (d, J = 13.2 Hz, 1H); 3.30 (d, J = 13.3 Hz, 1H); 3.19 (dd, J = 13.2, 1.7 Hz, 1H); 2.36 (s, 6H); 1.76 (s, 6H) ppm. 13C-NMR (600 MHz, CDCl3) δ = 140.57 (d, JCP = 6.5 Hz, Cq); 134.38 (d, JCP = 23.0 Hz, CH); 128.48 (d, JCP = 0.9 Hz, CH); 127.51 (d, JCP = 8.6 Hz, CH); 91.18 (d, JCP = 31.5 Hz, Cq); 88.46 (d, JCP = 23.0 Hz, Cq); 80.17 (d, JCP = 5.9 Hz, Cq); 75.92 (d, JCP = 12.9 Hz, Cq); 72.50 (d, JCP = 5.4 Hz, CH); 71.71 (d, JCP = 5.3 Hz, CH); 71.64 (d, JCP = 3.4 Hz, CH); 70.31 (d, JCP = 3.5 Hz, CH); 69.92 (CH); 69.60 (CH); 69.41 (CH); 67.99 (CH); 58.63 (d, JCP = 11.2 Hz, CH2); 57.73 (d, JCP = 8.4 Hz, CH2); 46.02 (CH3); 44.74 (CH3) ppm. 31P-NMR δ = −44.15 (s) ppm. HRMS: m/z calculated for C32H37Fe2N2P; [M + H]+ 593.1471, found: 593.1457.

3.2.5. 1,1ʺ-(Phenylphosphinideneoxide)di[(2Sp)-2-(-1-hydroxy-2-ethyl)]ferrocene 10

1,1ʺ-(Phenylphosphinideneoxide)di[(2Sp)-2-(-1-hydroxy-2-ethyl)]ferrocene 10 [35]: Divinylphosphine 9 (56 mg, 0.11 mmol) was dissolved in 1 mL of absolute THF under Ar. The orange solution was cooled in an ice bath and 4 mL of 9-BBN solution (0.5 M in THF, 2 mmol, 19 eq) was added dropwise. The solution was stirred under Ar at RT for 24 h; then 0.7 mL of aqueous 3 M KOAc and 6 mL of aqueous H2O2 (30%, 59 mmol, 554 eq) were added in that order, and the solution was stirred for another 24 h. The solution was extracted with DCM (2 × 5 mL). The combined organic fractions were washed with water (15 mL) and brine (15 mL), and dried over MgSO4. The solvent was removed under reduced pressure and the product was purified by column chromatography (Al2O3, 50–100% EtOAc in heptane, 0–20% MeOH in EtOAc) yielding 16 mg of dihydroxyphosphineoxide 10 (25%). 1H-NMR (600 MHz, CDCl3) δ = 7.77-7.73 (m, 2H); 7.53-7.44 (m, 3H); 4.55 (m, 1H); 4.45 (m, 1H); 4.36 (m, 1H); 4.31 (q, J = 2.3 Hz, 1H); 4.30–4.26 (m, 2H); 4.28 (s, 5H); 4.11 (q, J = 7.2 Hz, 1H); 4.09 (m, 1H); 4.00 (m, 1H); 3.94 (s, 5H); 3.39–3.24 (m, 3H); 3.05–2.92 (m, 2H); 2.57–2.43 (m, 2H) ppm. 13C-NMR δ = 135.89 (d, JCP = 108.9 Hz, Cq); 131.28 (d, JCP = 2.7 Hz, CH); 130.82 (d, JCP = 9.6 Hz, CH); 128.05 (d, JCP = 12.0 Hz, CH); 91.69 (d, JCP = 10.8 Hz, Cq); 88.18 (d, JCP = 11.8 Hz, Cq); 76.09 (d, JCP = 115.6 Hz, Cq); 73.29 (d, JCP = 15.0 Hz, CH); 72.36 (d, JCP = 10.1 Hz, CH); 72.09 (d, JCP = 10.2 Hz, CH); 71.25 (d, JCP = 118.1 Hz, Cq); 71.24 (d, JCP = 15.7 Hz, CH); 70.20 (CH); 70.11 (CH); 69.91 (d, JCP = 11.0 Hz, CH); 68.98 (d, JCP = 11.2 Hz, CH); 64.01 (CH2); 62.54 (CH2); 31.73 (CH2); 31.10 (CH2) ppm. 31P-NMR δ = 32.70 (s) ppm. HR-MS: m/z calculated for C30H31Fe2O3P; [M + Na]+ 605.0602, found: 605.0592; [M + K]+ 621.0341, found: 621.0327.

3.2.6. 1,1‘‘-(Phenylphosphinidenesulfide)di[(2Sp)-2-vinyl]ferrocene 11

1,1‘‘-(Phenylphosphinidenesulfide)di[(2Sp)-2-vinyl]ferrocene 11 [36]: Divinylphosphine 9 (240 mg, 0.45 mmol) and sulfur (92 mg, 2.87 mmol, 6.34 eq) were dissolved in 3.6 mL of toluene under Ar. The solution was stirred and heated under reflux for 4 h, and then cooled to RT. The solvent was removed under reduced pressure and the residue was purified by column chromatography (SiO2; 0–10% EtOAc in heptane), yielding 255 mg of divinylphosphinesulfide 11 (quantitative) as orange crystals. M.p.: 191–192 °C (decomposition). 1H-NMR (600 MHz, CDCl3) δ = 8.11–8.04 (m, 1H); 7.73–7.67 (m, 2H); 7.47–7.37 (m, 3H); 6.95–6.88 (m, 1H); 5.52–5.46 (m, 1H); 5.24–5.17 (m, 1H); 5.21 (m, 1H); 4.86–4.82 (m, 2H); 4.71 (m, 1H); 4.36 (s, 5H); 4.28 (m, 1H); 4.26 (m, 1H); 4.23 (s, 5H); 3.72 (m, 1H); 3.51 (m, 1H) ppm. 13C-NMR δ = 134.63 (d, JCP = 87.8 Hz, Cq); 134.37 (CH); 132.97 (CH); 132.18 (d, JCP = 10.6 Hz, CH); 130.92 (d, JCP = 2.9 Hz, CH); 127.58 (d, JCP = 12.3 Hz, CH); 112.13 (CH2); 111.61 (CH2); 88.19 (d, JCP = 11.6 Hz, Cq); 86.45 (d, JCP = 11.5 Hz, Cq); 79.27 (Cq); 78.63 (Cq); 75.62 (d, JCP = 11.8 Hz, CH); 74.48 (d, JCP = 12.7 Hz, CH); 71.40 (CH); 71.10 (CH); 69.72 (d, JCP = 10.4 Hz, CH); 68.66 (d, JCP = 10.5 Hz, CH); 68.02 (d, JCP = 9.0 Hz, CH); 67.70 (d, JCP = 9.1 Hz, CH) ppm. 31P-NMR (400 MHz, CDCl3) δ = 40.54 (s) ppm. HR-MS: m/z calculated for C30H27Fe2PS; [M]+ 562.0270, found: 562.0270.

3.2.7. 1,1‘‘-(Phenylphosphinidenesulfide)di[(2Sp)-2-(-1-hydroxy-2-ethyl)]ferrocene 12

Divinylphosphinesulfide 11 (530 mg, 0.94 mmol) was dissolved in 10 mL of absolute THF under Ar. The orange solution was cooled in an ice bath and 16 mL of 9-BBN solution (0.5 M in THF, 8 mmol, 8.5 eq) was added dropwise. The solution was stirred under Ar at RT for 24 h, and then cooled in an ice bath; 7 mL of aqueous 3 M KOAc and 5 mL of aqueous H2O2 (30%, 49 mmol, 52 eq) were added in that order and the solution was stirred at RT for another 24 h. The solution was extracted with DCM (2 × 12 mL). The combined organic fractions were washed with water (50 mL) and brine (50 mL) and dried over MgSO4. The solvent was removed under reduced pressure and the product was purified by column chromatography (SiO2, 0–50% EtOAc in heptane) yielding 337 mg of dihydroxide 12 (60%) as an orange oil. Residual eluent moisture (heptane) could not be removed in high vacuum. Solvent removal by heating was not possible due to decomposition of the product (visible by formation of a black solid). 1H-NMR (700 MHz, CDCl3) δ = 7.79 (dd, J = 13.0, 7.6 Hz, 2H); 7.49 (pt, J = 7.5 Hz, 1H); 7.44 (pt, J = 7.9 Hz, 2H); 4.53 (s, 1H); 4.42 (s, 1H); 4.35 (s, 5H); 4.22 (s, 1H); 4.20 (s, 1H); 4.12 (s, 5H); 4.09 (m, 2H); 3.93 (s, 1H); 3.70 (s, 1H); 3.66 (dt, J = 15.3, 6.5 Hz, 1H); 3.28 (m, 1H); 3.14 (m, 1H); 3.13 (m, 1H); 2.70–2.64 (m, 1H); 2.53–2.48 (m, 1H) ppm. 13C-NMR δ = 134.90 (d, JCP = 87.1 Hz, Cq); 132.13 (d, JCP = 10.4 Hz, CH); 131.11 (d, JCP = 2.8 Hz, CH); 127.69 (d, JCP = 12.2 Hz, CH); 90.48 (d, JCP = 13.0 Hz, Cq); 87.45 (d, JCP = 12.7 Hz, Cq); 80.25 (d, JCP = 95.7 Hz, Cq); 74.13 (d, JCP = 12.8 Hz, CH); 74.04 (d, JCP = 95.6 Hz, Cq); 72.99 (d, JCP = 12.9 Hz, CH); 72.26 (d, JCP = 9.4 Hz, CH); 71.97 (d, JCP = 9.8 Hz, CH); 70.70 (CH); 70.44 (CH); 68.82 (d, JCP = 10.2 Hz, CH); 67.67 (d, JCP = 10.6 Hz, CH); 63.26 (CH2); 62.07 (CH2); 31.19 (CH2); 30.78 (CH2) ppm. 31P-NMR δ = 40.32 (s) ppm. HR-MS: m/z calculated for C30H31Fe2O2PS; [M]+ 598.0481, found: 598.0479.

Supplementary Materials

Figure S1: 1H-NMR spectrum of compound 4. Figure S2: 13C-NMR spectrum of compound 4. Figure S3: HR-MS spectrum of compound 4. Figure S4: 1H-NMR spectrum of compound 5. Figure S5: 13C-NMR spectrum of compound 5. Figure S6: HR-MS of compound 5. Figure S7: HR-MS of compound 5 (detailed). Figure S8: 1H-NMR spectrum of compound 7. Figure S9: 13C-NMR spectrum of compound 7. Figure S10: 31P-NMR spectrum of compound 7. Figure S11: HR-MS of compound 7. Figure S12: 1H-NMR spectrum of compound 10. Figure S13: 13C-NMR spectrum of compound 10. Figure S14: 31P-NMR spectrum of compound 10. Figure S15: HR-MS of compound 10. Figure S16: 1H-NMR spectrum of compound 11. Figure S17: 13C-NMR spectrum of compound 11. Figure S18: 1H-NMR spectrum of compound 12. Figure S19: 13C-NMR spectrum of compound 12. Figure S20 31P-NMR spectrum of compound 12. Figure S21: HR-MS of compound 12.

Author Contributions

P.H.: Synthesis planning, literature search, experimental synthesis, writing of the manuscript; L.S.: Experimental synthesis; M.W.: Resources, spectral interpretation, proofreading of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Open Access Funding by the University of Vienna is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Schaarschmidt, D.; Lang, H. Selective Syntheses of Planar-Chiral Ferrocenes. Organometallics 2013, 32, 5668–5704. [Google Scholar] [CrossRef]
  2. Toma, Š.; Csizmadiova, J.; Meciarova, M.; Šebesta, R. Ferrocene phosphane-heteroatom/carbon bidentate ligands in asymmetric catalysis. Dalton Trans. 2014, 43, 16557–16579. [Google Scholar] [CrossRef] [PubMed]
  3. Dai, L.-X.; Hou, X.-L. Chiral Ferrocenes. In Asymmetric Catalysis; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  4. Štěpnička, P. Ferrocenes: Ligands, Materials and Biomolecules; John Wiley & Son: Chichester, UK, 2008. [Google Scholar]
  5. Manoury, E.; Poli, R. Phosphine-Containing Planar Chiral Ferrocenes: Synthesis, Coordination Chemistry and Applications to Asymmetric Catalysis. In Phosphorus Compounds. Catalysis by Metal Complexes; Peruzzini, M., Gonsalvi, L., Eds.; Springer: Dordrecht, the Netherlands, 2011. [Google Scholar]
  6. Dwadnia, N.; Roger, J.; Pirio, N.; Cattey, H.; Hierso, J.-C. Input of P,N-(phosphanyl, amino)-ferrocene hybrid derivatives in late transition metals catalysis. Coord. Chem. Rev. 2018, 355, 74–100. [Google Scholar] [CrossRef]
  7. Butt, N.A.; Liu, D.; Zhang, W. The Design and Synthesis of Planar Chiral Ligands and Their Application to Asymmetric Catalysis. Synlett 2014, 25, 615–630. [Google Scholar] [CrossRef]
  8. Blaser, H.-U.; Pugin, B.; Spindler, F.; Thommen, M. From a chiral switch to a ligand portfolio for asymmetric catalysis. Acc. Chem. Res. 2007, 40, 1240–1250. [Google Scholar] [CrossRef]
  9. Marquarding, D.; Klusacek, H.; Gokel, G.; Hoffmann, P.; Ugi, I. Stereoselective syntheses. VI. Correlation of central and planar chirality in ferrocene derivatives. J. Am. Chem. Soc. 1970, 92, 5389–5393. [Google Scholar] [CrossRef]
  10. Gokel, G.; Marquarding, D.; Ugi, I. Stereoselective syntheses. VIII. Retentive nucleophilic displacements of.alpha.-substituted alkylferrocenes. J. Org. Chem. 1972, 20, 3052–3058. [Google Scholar] [CrossRef]
  11. Barbaro, P.; Togni, A. A New Chiral Tridentate Ferrocenyl Ligand. Synthesis and Characterization of Its Palladium(II) and Nickel(II) Complexes. Organometallics 1995, 14, 3570–3573. [Google Scholar] [CrossRef]
  12. Barbaro, P.; Bianchini, C.; Oberhauser, W.; Togni, A. Synthesis and characterization of chiral bis-ferrocenyl triphosphine Ni (II) and Rh (III) complexes and their use as catalyst precursors for acetalization reactions. J. Mol. Catal. 1999, 145, 139–146. [Google Scholar] [CrossRef]
  13. Fadini, L.; Togni, A. Asymmetric hydroamination of acrylonitrile derivatives catalyzed by Ni(II)-complexes. Tetrahedron: Asymmetry 2008, 19, 2555–2562. [Google Scholar] [CrossRef]
  14. Fadini, L.; Togni, A. Asymmetric Catalytic Hydroamination of Activated Olefins in Ionic Liquids. Helv. Chim. Acta 2007, 90, 411–424. [Google Scholar] [CrossRef]
  15. Gischig, S.; Togni, A. Synthesis and Coordination Chemistry of a New Chiral Tridentate PCP N-Heterocyclic Carbene Ligand Based on a Ferrocene Backbone. Organometallics 2004, 23, 2479–2487. [Google Scholar] [CrossRef]
  16. Gischig, S.; Togni, A. PdII Complexes of Tridentate PCP N-Heterocyclic Carbene Ligands: Structural Aspects and Application in Asymmetric Hydroamination of Cyano Olefins. Eur. J. Inorg. Chem. 2005, 2005, 4745–4754. [Google Scholar] [CrossRef]
  17. Sadow, A.D.; Togni, A. Enantioselective Addition of Secondary Phosphines to Methacrylonitrile: Catalysis and Mechanism. J. Am. Chem. Soc. 2005, 127, 17012–17024. [Google Scholar] [CrossRef] [PubMed]
  18. Barbaro, P.; Bianchini, C.; Togni, A. Synthesis and Characterization of Ruthenium (II) Complexes Containing Chiral Bis (ferrocenyl)−P3 or−P2S Ligands. Asymmetric Transfer Hydrogenation of Acetophenone. Organometallics 1997, 16, 3004–3014. [Google Scholar] [CrossRef]
  19. Barbaro, P.; Bianchini, C.; Giambastiani, G.; Togni, A. Ruthenium(II) Complexes with Triphosphane Ligands Combining Planar, Phosphorus, and Carbon Chirality: Application to Asymmetric Reduction of Trifluoroacetophenone. Eur. J. Inorg. Chem. 2003, 2003, 4166–4172. [Google Scholar] [CrossRef]
  20. Milosevic, S.; Togni, A. Enantioselective 1,3-Dipolar Cycloaddition of C,N-Cyclic Azomethine Imines to Unsaturated Nitriles Catalyzed by NiIIPigiphos. J. Org. Chem. 2013, 78, 9638–9646. [Google Scholar] [CrossRef]
  21. Lee, H.M.; Bianchini, C.; Jia, G.; Barbaro, P. Styrene Cyclopropanation and Ethyl Diazoacetate Dimerization Catalyzed by Ruthenium Complexes Containing Chiral Tridentate Phosphine Ligands. Organometallics 1999, 18, 1961–1966. [Google Scholar] [CrossRef]
  22. Walz, I.; Togni, A. Ni(II)-catalyzed enantioselective Nazarov cyclizations. Chem. Comm. 2008, 4315–4317. [Google Scholar] [CrossRef]
  23. Broggini, D. Phosphine and Carbene Ligands for Asymmetric Catalysis Containing Two Planar-Chiral Ferrocenyl Units. Ph.D. Thesis, Swiss Federal Institute Of Technology Zurich, Zurich, Switzerland, 2003. [Google Scholar]
  24. Barreiro, E.; Broggini, D.; Adrio, L.; White, A.; Schwenk, R.; Togni, A.; Hii, K. Gold (I) complexes of conformationally constricted chiral ferrocenyl phosphines. Organometallics 2012, 31, 3745–3754. [Google Scholar] [CrossRef]
  25. Smith, A.M.R.; Rzepa, H.S.; White, A.J.P.; Billen, D.; Hii, K.K. Delineating Origins of Stereocontrol in Asymmetric Pd-Catalyzed α-Hydroxylation of 1,3-Ketoesters. J. Org. Chem. 2010, 75, 3085–3096. [Google Scholar] [CrossRef] [PubMed]
  26. Zirakzadeh, A.; Kirchner, K.; Roller, A.; Stöger, B.; Widhalm, M.; Morris, R.H. Iron(II) Complexes Containing Chiral Unsymmetrical PNP′ Pincer Ligands: Synthesis and Application in Asymmetric Hydrogenations. Organometallics 2016, 35, 3781–3787. [Google Scholar] [CrossRef]
  27. Zirakzadeh, A.; Kirchner, K.; Roller, A.; Stöger, B.; Carvalho, M.; Ferreira, L. Synthesis, coordination behavior and structural features of chiral iron(ii) PNP diferrocene complexes. RSC Adv. 2016, 6, 11840–11847. [Google Scholar] [CrossRef]
  28. Zirakzadeh, A.; de Aguilar, S.R.M.M.; Widhalm, M.; Mereiter, K.; Kirchner, K. Iron(II) complexes featuring chiral PNNP diferrocene: Synthesis and characterization of potential hydrogenation catalysts. J. Organomet. Chem. 2016, 819, 260–265. [Google Scholar] [CrossRef]
  29. Honegger, P.; Widhalm, M. (2S,4R,6R,8S)-4,6-Dimethyl-1-phenyl-diferroceno-1-phosphines. Molbank 2019, 4, 1098. [Google Scholar] [CrossRef] [Green Version]
  30. Xiao, L.; Kitzler, R.; Weissensteiner, W. O-Methylephedrine: A Versatile and Highly Efficient ortho-Directing Group. Synthesis of Enantiopure 1,2-Disubstituted Ferrocene Derivatives. J. Org. Chem. 2001, 66, 8912–8919. [Google Scholar] [CrossRef] [PubMed]
  31. Hayashi, T.; Mise, T.; Fukushima, M.; Kagotani, M.; Nagashima, N.; Hamada, Y.; Matsumoto, A.; Kawakami, S.; Konishi, M.; Yamamoto, K.; et al. Asymmetric Synthesis Catalyzed by Chiral Ferrocenylphosphine–Transition Metal Complexes. I. Preparation of Chiral Ferrocenylphosphines. Bull. Chem. Soc. Jpn. 1980, 53, 1138–1151. [Google Scholar] [CrossRef] [Green Version]
  32. Butler, I.; Cullen, W.; Rettig, S. Synthesis of derivatives of [.alpha.(dimethylamino)ethyl]ferrocene via lithiation reactions and the structure of 2-[.alpha.-(dimethylamino)ethyl]-1,1’,3-tris(trimethylsilyl)ferrocene. Organometallics 1986, 5, 1320–1328. [Google Scholar] [CrossRef]
  33. Fadini, L.; Togni, A. Ni (II)-PPP complexes: From the hydroamination of activated olefins to the synthesis of β-amino acids. Chimia 2004, 58, 208–211. [Google Scholar] [CrossRef]
  34. Honegger, P.; Widhalm, M. (2S,4R,8S)-4-Methyl-1-phenyl-diferroceno-5-Z-ethylene-1-phosphinoxide. Molbank 2020, 2020, 1105. [Google Scholar] [CrossRef] [Green Version]
  35. Meng, X.; York, E.A.; Liu, S.; Edgar, K.J. Hydroboration–oxidation: A chemoselective route to cellulose ω-hydroxyalkanoate esters. J. Carb. Pol. 2015, 133, 262–269. [Google Scholar] [CrossRef] [PubMed]
  36. Honegger, P.; Widhalm, M. Formaldehyde-1,1″-(phenylphosphinidenesulfide)bis[(2S)-2-[(1R)-1-(methylamino)-ethyl]]ferrocene-aminal. Molbank 2019, 4, 1090. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Ferrocene-based catalysts Xyliphos (i), Pigiphos (ii), NHC-Pigiphos (iii) and Stanphos (iv), and pincer ligands recently developed by Zirakzadeh et al. (v–viii).
Scheme 1. Ferrocene-based catalysts Xyliphos (i), Pigiphos (ii), NHC-Pigiphos (iii) and Stanphos (iv), and pincer ligands recently developed by Zirakzadeh et al. (v–viii).
Molbank 2020 m1130 sch001
Scheme 2. Stereochemical properties of Ugi’s amine: diastereoselective lithiation followed by reaction with electrophile E [9] and stereo-retentive SN1 reaction with nucleophile Nu at the α-carbon of the alkyl sidechain [10].
Scheme 2. Stereochemical properties of Ugi’s amine: diastereoselective lithiation followed by reaction with electrophile E [9] and stereo-retentive SN1 reaction with nucleophile Nu at the α-carbon of the alkyl sidechain [10].
Molbank 2020 m1130 sch002
Scheme 3. Synthesis of diferrocenes 4 and 7.
Scheme 3. Synthesis of diferrocenes 4 and 7.
Molbank 2020 m1130 sch003
Scheme 4. Synthesis of diferrocene 12.
Scheme 4. Synthesis of diferrocene 12.
Molbank 2020 m1130 sch004

Share and Cite

MDPI and ACS Style

Honegger, P.; Scheibelberger, L.; Widhalm, M. Planar-Chiral P and N-bridged Diferrocenes. Molbank 2020, 2020, M1130. https://doi.org/10.3390/M1130

AMA Style

Honegger P, Scheibelberger L, Widhalm M. Planar-Chiral P and N-bridged Diferrocenes. Molbank. 2020; 2020(2):M1130. https://doi.org/10.3390/M1130

Chicago/Turabian Style

Honegger, Philipp, Lukas Scheibelberger, and Michael Widhalm. 2020. "Planar-Chiral P and N-bridged Diferrocenes" Molbank 2020, no. 2: M1130. https://doi.org/10.3390/M1130

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