Next Article in Journal
Synthesis, 18F-Radiolabelling and Biological Characterization of Novel Fluoroalkylated Triazine Derivatives for in Vivo Imaging of Phosphodiesterase 2A in Brain via Positron Emission Tomography
Next Article in Special Issue
Ambiphilic Frustrated Lewis Pair Exhibiting High Robustness and Reversible Water Activation: Towards the Metal-Free Hydrogenation of Carbon Dioxide
Previous Article in Journal
Comparative Studies on Polyphenolic Composition, Antioxidant and Diuretic Effects of Nigella sativa L. (Black Cumin) and Nigella damascena L. (Lady-in-a-Mist) Seeds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactivity of Amine/E(C6F5)3 (E = B, Al) Lewis Pairs toward Linear and Cyclic Acrylic Monomers: Hydrogenation vs. Polymerization

Department of Chemistry, Colorado State University, Fort Collins, CO 80523-1872, USA
*
Author to whom correspondence should be addressed.
Molecules 2015, 20(6), 9575-9590; https://doi.org/10.3390/molecules20069575
Submission received: 29 April 2015 / Revised: 20 May 2015 / Accepted: 21 May 2015 / Published: 26 May 2015
(This article belongs to the Special Issue The Reactivity of Frustrated Lewis Pairs)

Abstract

:
This work reveals the contrasting reactivity of amine/E(C6F5)3 (E = B, Al) Lewis pairs toward linear and cyclic acrylic monomers, methyl methacrylate (MMA) and biorenewable γ-methyl-α-methylene-γ-butyrolactone (γMMBL). While mixing of 2,2,6,6-tetramethylpiperidine (TMP) and B(C6F5)3 leads to a frustrated Lewis pair (FLP), Et3N reacts with B(C6F5)3 to form disproportionation products, ammonium hydridoborate ionic pair and iminium zwitterion. On the other hand, the stoichiometric reaction of either TMP or Et3N with Al(C6F5)3 leads to clean formation of a classic Lewis adduct (CLA). Neither TMP nor Et3N, when paired with E(C6F5)3, polymerizes MMA, but the Et3N/2B(C6F5)3 pair promotes transfer hydrogenation of MMA to form methyl isobutyrate. In contrast, the amine/E(C6F5)3 pairs promote rapid polymerization of γMMBL carrying the more reactive exocyclic methylene moiety, achieving full conversion in less than 3 min even at a low catalyst loading of 0.0625 mol %. TMP is more effective than Et3N for the polymerization when paired with either the borane or the alane, while the alane exhibits higher polymerization activity than the borane when paired with Et3N. Overall, the TMP/Al(C6F5)3 system exhibits the highest polymerization activity, achieving a maximum turn-over frequency of 96,000 h−1 at 0.125 mol % of catalyst loading, producing high molecular weight PγMMBL with Mn = 1.29 × 105 g∙mol−1.

Graphical Abstract

1. Introduction

The combination of various highly acidic Lewis acids (LAs) and sterically hindered Lewis bases (LBs) provides opportunities to construct different types of “frustrated Lewis pairs” (FLPs) [1,2,3,4,5,6,7,8,9], in which the direct LA-LB interaction is either precluded or significantly weakened due to the steric crowdedness. More significantly, this unquenched, orthogonal reactivity allows the cooperative activation of small molecules [10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25], catalytic hydrogenation [26,27,28,29,30,31,32] and new reactivity/reaction development [33,34,35,36,37,38,39,40,41,42,43,44,45]. While the original FLP systems pioneered by Stephan and Erker focused mostly on boron-based LAs and phosphine-based LBs, the scopes of the suitable candidates for the FLP chemistry have recently been extended to other group 13 and 14 elements as LAs (e.g., organoaluminum species, silylium cations, and borenium cations), and group 14 and 15 elements as LBs (e.g., amines, pyridines, and carbenes). Furthermore, a considerable amount of efforts has been directed to the design of unimolecular FLPs and chiral versions of FLPs for asymmetric transformations [46,47,48,49,50,51,52,53,54,55].
We have been interested in developing new polymerization methods enabled by Lewis pairs, namely Lewis Pair Polymerization (LPP) [56], to synthesize different classes of polymers. In this type of polymerization, Lewis pairs are involved in monomer substrate activation, chain initiation, as well as chain propagation and termination/transfer steps (Scheme 1). We have demonstrated, for example, FLPs or classic Lewis adducts (CLAs) consisting of bulky phosphine or N-heterocyclic carbene (NHC) LBs and the highly acidic Al(C6F5)3 LA are capable of promoting rapid polymerization of conjugated polar alkenes such as linear acrylic monomer methyl methacrylate (MMA) and cyclic biorenewable γ-methyl-α-methylene-γ-butyrolactone (γMMBL) into high molecular weight polymers [57,58,59]. Such LPP system is also effective for the polymerization of monomers featuring the C=C-C=N functionality, including 2-vinyl pyridine and 2-isopropenyl-2-oxazoline [60,61]. Controlled Ring-opening (co)polymerization of other monomers such as lactide and lactones into cyclic polymers has been achieved by employing Zn(C6F5)2-based Lewis pairs [62], while radical polymerization of styrene is successfully mediated by a persistent FLP-NO aminoxyl radical generated by N,N-cycloaddition of a cyclohexylene-linked intramolecular P/B system to nitric oxide [63]. More recently, we found that the boron-based LA, in combination with various NHC or phosphine LBs, can be highly effective for polymerization of MMA and γMMBL: while the B/P FLP systems exhibit no or negligible polymerization reactivity, the P→B and NHC→B CLAs have been found to exhibit unexpectedly high activity for the polymerization of γMMBL [64]. This unique polymerization method has now been employed to the regio-controlled polymerization of dissymmetric divinyl polar monomers [57,65].
A search of the current array of LPP systems reveals that the combination of LAs and amines is scarce and not well-established [58]. In contrast, the B/N system has been studied extensively in the area of the FLP chemistry with respect to the LA-LB interaction, small molecule activation [25], catalysis [28,44,66,67] as well as mechanistic investigations [68]. In this context, we set out to examine the polymerization reactivity of two widely used amines, triethylamine (Et3N) and 2,2,6,6-tetramethylpiperidine (TMP), coupled with two strong organo-LAs, E(C6F5)3 (E = B, Al). We envisioned that the introduction of such sterically encumbered amines with different structural and electronic properties will present an opportunity to probe the polymerization activity of the amine/E(C6F5)3 (E = B or Al) system toward two representative acrylic monomers, linear MMA and cyclic γMMBL (Scheme 1).
Scheme 1. Generalized chain initiation and propagation mechanism for polymerization of conjugated polar alkenes carrying a functional group (FG) by Lewis pairs through zwitterionic active species or intermediates [56], and the structure of LAs, LBs and monomers examined in this study.
Scheme 1. Generalized chain initiation and propagation mechanism for polymerization of conjugated polar alkenes carrying a functional group (FG) by Lewis pairs through zwitterionic active species or intermediates [56], and the structure of LAs, LBs and monomers examined in this study.
Molecules 20 09575 g004

2. Results and Discussion

2.1. Reaction of Amines and E(C6F5)3 in the Absence of Monomer

A survey of the recent literatures indicates that the coordination chemistry and structural characterizations of several amine/E(C6F5)3 pairs have been reported. Preceding the discovery of the FLP system, earlier studies focused on systems featuring direct B-N and Al-N bond formation with less sterically hindered amines. For instance, Lancaster and co-workers [69] reported that both the borane and alane E(C6F5)3 form stable adducts with a variety of primary and secondary amines. In addition, in all of the cases, the two Lewis acids show invariable reactivity towards these amines. On the other hand, we observed markedly different polymerization activity between the borane and the alane, when used as (co)catalysts and monomer activators for both transition-metal-catalyzed and LPP processes [59,70,71,72,73]. In general, the alane is more effective than borane due to its higher Lewis acidity, oxophilicity and larger radius. Our recent work showed that the alane displays clear advantages over the borane in several different types of FLP-type catalysis [74]. In this LPP work, we chose two sterically congested amines, Et3N and TMP, as the LB to pair with E(C6F5)3 (Scheme 1).
These two amines exhibit FLP-type reactivity when coupled with the borane [25,75], but their reaction with the alane has not yet been examined. An earlier study showed that, mixing of B(C6F5)3 and Et3N results in a disproportionation reaction of the Lewis pair to form ammonium hydridoborate (C6F5)3B-H···H-N(C2H5)3, concomitant with equimolar iminium zwitterion (C6F5)3B-CH2-CH=N(C2H5)2 [75] (Scheme 2). B(C6F5)3 is known to abstract a hydride from the α-position of a bulky amine with α-hydrogens and form an FLP with the one without α-hydrogens such as TMP. The latter system was utilized for dihydrogen activation and CO2 reduction [25,44]. We speculated that switching from B(C6F5)3 to its congener Al(C6F5)3 could achieve different coordination reactivity. As expected, the alane, considered possessing higher Lewis acidity and less steric hindrance, forms a dative bond with both Et3N and TMP. In the stoichiometric reaction between Al(C6F5)3 and Et3N, the crystalline adduct Et3N·Al(C6F5)3 was isolated in a yield of 81%. Only one set of ethyl and one set of C6F5 resonances were presented in the 1H- and 19F-NMR spectra, respectively (Figure 1). Apparently, the direct Al-N bonding prevents other reactions such as hydride abstraction, which stands in contrast to the borane. Similarly, the TMP·Al(C6F5)3 adduct was isolated as colorless crystals in a yield of 85%, in which the complex formation was confirmed by both 1H- and 19F-NMR spectroscopy (Figure 2). In the 1H-NMR spectrum, the N-H signal is significantly downfield shifted to 3.71 ppm (cf. 0.64 ppm for the free TMP), as a result of the enhanced acidity after coordination to the electron-deficient alane. This adduct is stable in solution up to 2 days, which precludes the possibility of catalyst decomposition (e.g., protonolytic cleavage of the Al-C bond by the N-H proton) during the course of polymerization (vide infra). In addition, an intramolecular FLP (5) [67] was also included for the LPP study.
Scheme 2. Different amine/E(C6F5)3 (E = B or Al) systems employed for this LPP study and their corresponding reactivity in the absence of monomer.
Scheme 2. Different amine/E(C6F5)3 (E = B or Al) systems employed for this LPP study and their corresponding reactivity in the absence of monomer.
Molecules 20 09575 g005
Figure 1. 1H- and 19F- (inset) spectra (C6D6) of adduct Et3N·Al(C6F5)3.
Figure 1. 1H- and 19F- (inset) spectra (C6D6) of adduct Et3N·Al(C6F5)3.
Molecules 20 09575 g001
Figure 2. 1H- and 19F- (inset) spectra (C6D6) of adduct TMP·Al(C6F5)3.
Figure 2. 1H- and 19F- (inset) spectra (C6D6) of adduct TMP·Al(C6F5)3.
Molecules 20 09575 g002

2.2. Stoichiometric Hydrogenation of MMA to Methyl Isobutyrate by Et3N/B(C6F5)3

We first examined the activity of the current amine/LA pairs toward polymerization of the linear methacrylate, MMA, but found no polymerization activity under our standard polymerization conditions for all five amine/LA systems. To gain further insight into this observation, we carried out NMR scale reactions for the amine/LA pairs with equimolar MMA (Figure 3). Accordingly, upon mixing stoichiometric amounts of Et3N/B(C6F5)3/MMA, only the signals of MMA and the products of the disproportionation reaction of Et3N/B(C6F5)3 were observed, regardless of the mixing sequence. Interestingly, addition of more B(C6F5)3 into the above mixture resulted in the consumption of MMA and formation of a new species, which was identified as methyl isobutyrate, Me2CHC(=O)OMe, a product of C=C double bond hydrogenation of MMA (Scheme 3). This transfer hydrogenation was fast and clean, and all of the MMA was consumed in the presence of an additional equivalent of B(C6F5)3. Noteworthy is the inability to generate the ammonium enolborate initiation species, which was further supported by the presence of the geminal protons (=CH2) NMR signals derived from a 1:1:1 mixture of TMP/B(C6F5)3/MMA (see Supporting Information). Similarly, mixing of a preformed Et3N·Al(C6F5)3 adduct with MMA only resulted in the replacement of Et3N with MMA, as indicated by the generation of free Et3N and the remaining geminal protons (=CH2) signals. Overall, these NMR experiments showed that the combination of Et3N or TMP with E(C6F5)3 provides insufficient activation for the polymerization of MMA, either due to the reduction of MMA into the inactive methyl isobutyrate, or the steric bulkiness around the amine, which impairs the effective formation of the active species for polymerization.
Figure 3. 1H-NMR spectra (CD2Cl2) of a mixture of Et3N/B(C6F5)3/MMA in the ratio of 1:1:1 (top), 1:1.5:1 (middle) and 1:2:1 (bottom, some amount of Et3N-H···H-B(C6F5)3 remains due to a slight deficiency of MMA).
Figure 3. 1H-NMR spectra (CD2Cl2) of a mixture of Et3N/B(C6F5)3/MMA in the ratio of 1:1:1 (top), 1:1.5:1 (middle) and 1:2:1 (bottom, some amount of Et3N-H···H-B(C6F5)3 remains due to a slight deficiency of MMA).
Molecules 20 09575 g003
Scheme 3. Stoichiometric hydrogenation of MMA to methyl isobutyrate by Et3N/B(C6F5)3.
Scheme 3. Stoichiometric hydrogenation of MMA to methyl isobutyrate by Et3N/B(C6F5)3.
Molecules 20 09575 g006

2.3. Characteristic of γMMBL Polymerization by Amine/E(C6F5)3

Next, we set out to investigate activity of the amine/E(C6F5)3 systems toward cyclic γMMBL, the results of which were summarized in Table 1. The intramolecular FLP system 5 yielded no polymer formation up to 24 h, under the current standard conditions and with or without another equivalent of B(C6F5)3 (entries 17 and 18). Mixing of equimolar γMMBL and 5 in CD2Cl2 led to formation of adduct γMMBL·5 via the Lewis acidic borane site in 5, but no active species or polymer formation. This result implies that the inability of 5 to polymerize γMMBL is attributed to the lack of initiation by the bulky amine site in 5. On the other hand, the intermolecular Lewis pair systems 1 to 4 initiated rapid and quantitative polymerization of γMMBL into PγMMBL. Under our current standard polymerization conditions ([γMMBL]:[LA]:[LB] = 200:2:1, 0.5 mL γMMBL, 2.0 mL CH2Cl2, RT), Et3N/LA system 1 (B) or 3 (Al) polymerized γMMBL with full conversion in 1 or 0.5 min, giving PγMMBL with Mn = 3.12 or 1.92 × 104 g∙mol−1 (Đ = 2.31 or 2.02), thus achieving an averaged turn-over frequency (TOF) of 12,000 or 24,000 h−1 for the borane or alane, respectively (entries 1 and 3). Increasing the ratio to 800:2:1 resulted in an increase of Mn to 5.13 or 3.30 × 104 g∙mol−1 for the B and Al LA, respectively, although a significantly prolonged polymerization period of 60 or 30 min was needed to achieve full monomer conversion (entries 2 and 4). Overall, the alane showed two-fold higher activity over the borane when coupled with Et3N. Moreover, the reverse addition sequence, in which equimolar Et3N and LA were premixed, followed by addition to a CH2Cl2 solution of γMMBL, was also effective, albeit a slightly slower rate compared to the standard procedure, with or without an additional equivalent of LA. A stoichiometric reaction of Et3N/B(C6F5)3/γMMBL revealed that all of the monomer was rapidly consumed, with generation of PγMMBL (see Supporting Information). We reasoned that ammonium hydridoborate, formed by premixing of Et3N and B(C6F5)3, can initiate the polymerization via the nucleophilic hydride, whereas the Et3N·Al(C6F5)3 adduct formed via premixing can dissociate in the presence of γMMBL and thus promote the subsequent initiation and bimolecular, activated monomer propagation.
As mentioned above, the combination of the alane or the borane with TMP offered more well-defined coordination chemistry (i.e., either CLA or FLP). Hence, the TMP/LA pairs were investigated for LPP with 6 different [monomer]:[LA]:[TMP] ratios, ranging from 50:2:1 to 1600:2:1, as well as different addition sequences. Impressively, the polymerizations were finished within 1 min for a ratio up to of 800:2:1 (entries 5–9 and 11–15), and even under lower catalyst loading (0.0625 mol % relative to TMP) conditions (1600:2:1), the polymerizations were completed in 3 min (entries 10 and 16). Thus, the polymerization by the B/TMP and Al/TMP Lewis pairs was very rapid with TOF up to 96,000 h−1 (for Al), providing polymers with Mn ranging from medium 6.75 × 104 to high 2.55 × 105 g∙mol−1 (Đ = 1.54 to 1.92) and medium 6.96 × 104 to high 1.38 × 105 g∙mol−1 (Đ = 2.16 to 2.42) for the borane and alane, respectively. Interestingly, under the current standard conditions, if the LAs were premixed with TMP and added to a CH2Cl2 solution of the monomer, the polymerization proceeded rapidly as well. This result confirmed that the TMP·Al(C6F5)3 adduct formed via premixing can rapidly dissociate in the presence of γMMBL monomer and promote the subsequent initiation and bimolecular, activated monomer propagation [58,59], while the B/TMP FLP should be active regardless of the addition sequence.
Table 1. Selected results of polymerization of γMMBL by LA/amine LPs a.
Molecules 20 09575 i001
Table 1. Selected results of polymerization of γMMBL by LA/amine LPs a.
Molecules 20 09575 i001
EntryLP[M]/[Base]Time (min)Conv. b (%)TOF (h-1)Mnc (kg∙mol−1)Đ d (Mw/Mn)
1B/Et3N (1)200110012,00031.22.31
2 8006010080053.12.47
3Al/Et3N (3)2000.510024,00019.22.02
4 80030100160033.02.26
5B/TMP (2)500.5100600067.5 d1.54
6 1000.510012,000100 d1.64
7 2000.510024,000145 d1.78
8 4000.510048,000156 d1.92
9 800110048,000253 d1.64
10 1600310032,000255 d1.66
11Al/TMP (4)500.5100600069.62.41
12 1000.510012,00077.22.29
13 2000.510024,0001132.16
14 4000.510048,0001192.38
15 8000.510096,0001292.21
16 1600310032,0001382.16
17B-TMP (5)20014400---
18 e5 + B20014400---
a Conditions: carried out at RT (ca. 25 °C) in CH2Cl2 with 2.5 mL of the total solution; b Monomer conversion measured by 1H-NMR spectroscopy; c Number-average molecular weight (Mn) and molecular weight distribution (Đ) determined by GPC in DMF relative to PMMA standards; d Bimodal distribution, with an additional small peak (5.0% to 18.3%) on the higher molecular weight region; e Carried out with an additional equivalent of B(C6F5)3.

3. Experimental Section

3.1. Materials, Reagents, and Methods

All syntheses and manipulations of air- and moisture-sensitive materials were carried out in flamed Schlenk-type glassware on a dual-manifold Schlenk line, on a high-vacuum line, or in an inert gas-filled glovebox. NMR-scale reactions were conducted in Teflon-valve-sealed J. Young-type NMR tubes. HPLC-grade organic solvents were first sparged extensively with nitrogen during filling 20 L solvent reservoirs and then dried by passage through activated alumina (for Et2O, THF, and CH2Cl2) followed by passage through Q-5 supported copper catalyst (for toluene and hexanes) stainless steel columns. Benzene-d6 and toluene-d8 were dried over sodium/potassium alloy and vacuum-distilled or filtered, whereas CD2Cl2 and CDCl3 were dried over activated Davison 4 Å molecular sieves. HPLC-grade dimethyl formamide (DMF) was degassed and dried over CaH2 overnight, followed by vacuum distillation (CaH2 was removed before distillation). NMR spectra were recorded on Varian Inova 300 (300 MHz, 1H; 75 MHz, 13C; 282 MHz, 19F) or a Varion 400 MHz spectrometer. Chemical shifts for 1H-, 19F- and 13C- spectra were referenced to internal solvent resonances and are reported as parts per million relative to SiMe4.
Methyl methacrylate (MMA) was purchased from Sigma-Aldrich Co. (St. Louis, MO, USA), while γ-methyl-α-methylene-γ-butyrolactone (γMMBL) was purchased from TCI America (Portland, OR, USA). These monomers were first degassed and dried over CaH2 overnight, followed by vacuum distillation. Further purification of MMA involved titration with neat tri(n-octyl)aluminum to a yellow end point [76], followed by distillation under reduced pressure. All purified monomers were stored in brown bottles and stored inside a glovebox freezer at −30 °C. Amines including triethylamine and 2,2,6,6-tetramethylpiperidine (TMP) were purchased from Sigma-Aldrich Co (St. Louis, MO, USA), distilled over CaH2, and store brown bottles in a glovebox prior to use. BHT-H was recrystallized from hexanes prior to use. Tris(pentafluorophenyl)borane, obtained as a research gifts from Boulder Scientific Company (Mead, CO, USA), was further purified by sublimation under vacuum. Tris(pentafluorophenyl)alane·(toluene)0.5 [77,78,79] and 1-(2-[bis(pentafluorophenyl)boryl]benzyl)-2,2,6,6-tetramethylpiperidine [67] were synthesized according to literature procedures.

3.2. Isolation of Adduct Et3N·Al(C6F5)3

Al(C6F5)3·(toluene)0.5 (135 mg, 0.235 mmol) and Et3N (32.6 μL, 0.235 mmol, 1.0 equiv) were dissolved in toluene (1 mL) in a glovebox. The mixture was allowed to react for 10 min, then layered with hexanes (4 mL), and placed in a freezer and recrystallized at −35 °C. The titled product was isolated as colorless crystals. Yield: 120 mg (81%). 1H-NMR (C6D6, 23 °C): δ 2.60 (q, J = 7.2 Hz, 6H, CH2CH3), 0.55 (t, 9H, CH2CH3). 13C-NMR (C6D6, 23 °C): δ 150.1, 141.9, 137.4 (C6F5, ipso-C6F5-Al not observed), 47.4 (CH2), 8.8 (CH3). 19F-NMR (C6D6, 23 °C): δ −119.0 (d, 3J = 19.4 Hz, 6F, o-F), −152.0 (t, 3J = 19.9 Hz, 3F, p-F), −160.9 (m, 6F, m-F) ppm.

3.3. Isolation of Adduct TMP·Al(C6F5)3

Al(C6F5)3·(toluene)0.5 (141 mg, 0.246 mmol) and TMP (42 μL, 0.246 mmol, 1.0 equiv) were dissolved in toluene (1 mL) in a glovebox. The mixture was allowed to react for 10 min, then layered with hexanes (3 mL), and placed in a freezer and recrystallized at −35 °C. The titled product was isolated as colorless crystals and kept in the freezer to avoid decomposition. Yield: 140 mg (85%). 1H-NMR (C6D6, 23 °C): δ 3.72 (br, 1H, NH), 1.09 (br, 6H, CH3), 0.93 (br, 6H, CH3), 0.90 (br, 4H, CH2), 0.75 (m, 2H, CH2). 13C-NMR (C6D6, 23 °C): δ 150.2, 142.1, 137.5 (C6F5, ipso-C6F5-Al not observed), 60.4, 40.6, 32.7, 24.7, 16.2 (TMP). 19F-NMR (C6D6, 23 °C): δ −119.1 (d, 3J = 16.1 Hz, 6F, o-F), −151.7 (t, 3J = 20.3 Hz, 3F, p-F), −161.2 (pst, 3J = 18.3 Hz, 6F, m-F) ppm.

3.4. Stoichiometric Hydrogenation of MMA by Et3N/2B(C6F5)3 to Methyl Isobutyrate

A CD2Cl2 (0.7 mL) solution of MMA (5.0 L, 0.0469 mmol) in a J. Young NMR tube was added B(C6F5)3 (24 mg, 0.0469 mmol), followed by addition of Et3N (6.54 L, 0.0469 mmol). The mixture was allowed to react for 10 min before 1H-NMR measurement, which indicated formation of ammonium hydridoborate (C6F5)3B-H···H-N(C2H5)3 and iminium zwitterion (C6F5)3B-CH2-CH=N(C2H5)2 but without consuming any MMA (Scheme 3, Figure 3). In the next step, a second equivalent of B(C6F5)3 was transferred to the same NMR tube and the reaction progress was monitored by 1H-NMR. The transfer hydrogenation went to completion yielding methyl isobutyrate.

3.5. Typical Procedure for the NMR Scale Reactions (TMP/B(C6F5)3/MMA as an Example)

A CD2Cl2 (0.7 mL) solution of MMA (5.0 L, 0.0469 mmol) in a J. Young NMR tube was added B(C6F5)3 (24 mg, 0.0469 mmol), followed by addition of TMP (7.98 L, 0.0469 mmol). The mixture was allowed to react for 10 min before 1H-NMR measurement, which indicated the retention of the geminal protons of MMA.
In the case of the alane, a different addition sequence was adopted, in which the preformed amine/Al adduct was transferred to the corresponding monomer in specific deuterated solvents as the solvent will decompose the alane upon mixing directly.

3.6. General Polymerization Procedures

Polymerizations were performed either in 25 mL flame-dried Schlenk flasks interfaced to the dual-manifold Schlenk line for runs using external temperature bath, or in 20 mL glass reactors inside the glovebox for ambient temperature (ca. 25 °C) runs. In a typical polymerization procedure, a predetermined amount of B(C6F5)3 or Al(C6F5)3 was first dissolved in a monomer (0.5 mL of MMA or γMMBL, 200 equiv relative to the LB) and 2.0 mL of solvent (CH2Cl2) inside a glovebox. The polymerization was started by rapid addition of a solution of a LB (1 equiv of an amine) in 0.5 mL of CH2Cl2 via a gastight syringe to the above mixture containing the LA and monomer under vigorous stirring. The amount of the monomer was fixed for the varied [M]/[LB] ratio runs. In a second procedure of polymerization, the premixed LA/LB system (or the intramolecular Lewis pair) was dissolved in 0.5 mL of toluene (for the alane system since the alane is known to decompose in CH2Cl2 in the absence of a monomer) or CH2Cl2 (for the borane system), followed by addition to a CH2Cl2 (2.0 mL) solution of monomer (0.5 mL) to initiate the polymerization. After the measured time interval, a 0.2 mL aliquot was taken from the reaction mixture via syringe and quickly quenched into a 4-mL vial containing 0.6 mL of undried “wet” CDCl3 stabilized by 250 ppm of BHT-H; the quenched aliquots were later analyzed by 1H-NMR to obtain the percent monomer conversion data. After the polymerization was stirred for the stated reaction time and then the polymer was immediately precipitated into 200 mL of methanol, stirred for 1 h, filtered, washed with methanol, and dried in a vacuum oven at 50 °C overnight to a constant weight.

3.7. Polymer Characterizations

Polymer number-average molecular weights (Mn) and molecular weight distributions (Đ = Mw/Mn) were measured by gel permeation chromatography (GPC) analyses carried out at 40 °C and a flow rate of 1.0 mL∙min−1, with DMF as the eluent, on a Waters University 1500 GPC instrument equipped with one PLgel 5 μm guard and three PLgel 5 μm mixed-C columns (Polymer Laboratories; linear range of molecular weight = 200–2,000,000). The instrument was calibrated with 10 PMMA standards, and chromatograms were processed with Waters Empower software (version 2002).

4. Conclusions

In summary, Lewis pairs consisting of bulky amine Et3N and TMP as LBs and bulky E(C6F5)3 as strong organo-LAs were employed for investigation into their reactivity toward two acrylic monomers, including linear MMA and cyclic γMMBL. While TMP and B(C6F5)3 form an FLP, Et3N reacts with B(C6F5)3 to form disproportionation products ammonium hydridoborate ionic pair and iminium zwitterion. On the other hand, the stoichiometric reaction of Et3N and TMP with Al(C6F5)3 leads to clean formation CLAs.
In the case of linear MMA, the Et3N/2B(C6F5)3 pair promotes transfer hydrogenation of MMA to form methyl isobutyrate. The TMP/B(C6F5)3 FLP is also incapable of polymerizing MMA, attributable to the steric hindrance and low nucleophilicity at N in TMP, which resulted in no formation of the active propagating species. In contrast, the amine/LA pairs promote rapid polymerization of γMMBL carrying the more reactive exocyclic methylene moiety, achieving full conversion in less than 3 min even at a low catalyst loading of 0.0625 mol %. The Al/Et3N pair displays higher activity when compared with the B/Et3N pair. Meanwhile, TMP is more efficient than Et3N when paired with either the borane or the alane. Of the four LA/LB pairs investigated in this study, the TMP/Al(C6F5)3 pair exhibits the highest activity, achieving a maximum TOF of 96,000 h−1 at 0.125 mol % of catalyst loading or 32,000 h−1 at 0.0625 mol %, producing high molecular weight PγMMBL with Mn = 1.29 × 105 or 1.38 × 105 g∙mol−1, respectively. As compared with the previous LBs employed for LPP, such as phosphines and NHCs, the current LPP system utilizing the inexpensive, readily available and more stable and environmentally friendly amine LBs will provide a greener alternative approach toward bio-renewable polymers based on γMMBL and other monomers.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/20/06/9575/s1.

Acknowledgments

This work was supported by the National Science Foundation (NSF-1150792 and NSF-1507702). We thank Boulder Scientific Co. for the research gift of B(C6F5)3.

Author Contributions

J.C. and E.Y.-X.C. formulated the research idea and participated in the design of the experiments and preparation of the manuscript. J.C. carried out the actual experiments and interpreted the data. All authors have read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stephan, D.W. Frustrated Lewis pairs: From concept to catalysis. Acc. Chem. Res. 2015, 48, 306–316. [Google Scholar] [CrossRef] [PubMed]
  2. Stephan, D.W.; Erker, G. Frustrated Lewis pair chemistry of carbon, nitrogen and sulfur oxides. Chem. Sci. 2014, 5, 2625–2641. [Google Scholar] [CrossRef]
  3. Erker, G.; Stephan, D.W. (Eds.) Frustrated Lewis Pairs I & II; Springer-Verlag: Berlin, Germany, 2013.
  4. Stephan, D.W. “Frustrated Lewis pair” hydrogenations. Org. Biomol. Chem. 2012, 10, 5740–5746. [Google Scholar] [CrossRef] [PubMed]
  5. Erker, G. Frustrated Lewis pairs: Some recent developments. Pure Appl. Chem. 2012, 84, 2203–2217. [Google Scholar] [CrossRef]
  6. Erker, G. Organometallic frustrated Lewis pair chemistry. Dalton Trans. 2011, 40, 7475–7483. [Google Scholar] [CrossRef] [PubMed]
  7. Stephan, D.W.; Erker, G. Frustrated Lewis pairs: Metal-free hydrogen activation and more. Angew. Chem. Int. Ed. 2010, 49, 46–76. [Google Scholar] [CrossRef] [PubMed]
  8. Stephan, D.W. Frustrated Lewis pairs: A new strategy to small molecule activation and hydro-genation catalysis. Dalton Trans. 2009, 17, 3129–3136. [Google Scholar] [CrossRef] [PubMed]
  9. Stephan, D.W. “Frustrated Lewis pairs”: A concept for new reactivity and catalysis. Org. Biomol. Chem. 2008, 6, 1535–1539. [Google Scholar] [CrossRef] [PubMed]
  10. Lawrence, E.J.; Oganesyan, V.S.; Hughes, D.L.; Ashley, A.E.; Wildgoose, G.G. An electrochemical study of frustrated Lewis pairs: A metal-free route to hydrogen oxidation. J. Am. Chem. Soc. 2014, 136, 6031–6036. [Google Scholar] [CrossRef] [PubMed]
  11. Sajid, M.; Elmer, L.M.; Rosorius, C.; Daniliuc, C.G.; Grimme, S.; Kehr, G.; Erker, G. Facile carbon monoxide reduction at intramolecular frustrated phosphane/borane Lewis pair templates. Angew. Chem. Int. Ed. 2013, 52, 2243–2246. [Google Scholar] [CrossRef] [PubMed]
  12. Dobrovetsky, R.; Stephan, D.W. Stoichiometric metal-free reduction of CO in syn-gas. J. Am. Chem. Soc. 2013, 135, 4974–4977. [Google Scholar] [CrossRef] [PubMed]
  13. Appelt, C.; Slootweg, J.C.; Lammertsma, K.; Uhl, W. Reaction of a P/Al-based frustrated Lewis pair with ammonia, borane, and amine-boranes: Adduct formation and catalytic dehydrogenation. Angew. Chem. Int. Ed. 2013, 52, 4256–4259. [Google Scholar] [CrossRef] [PubMed]
  14. Bertini, F.; Lyaskoyskyy, V.; Timmer, B.J.J.; de Kanter, F.J.J.; Lutz, M.; Ehlers, A.W.; Slootweg, J.C.; Lammertsma, K. Preorganized frustrated Lewis pairs. J. Am. Chem. Soc. 2012, 134, 201–204. [Google Scholar] [CrossRef] [PubMed]
  15. Schafer, A.; Reissmann, M.; Schafer, A.; Saak, W.; Haase, D.; Muller, T. A new synthesis of triarylsilylium ions and their application in dihydrogen activation. Angew. Chem. Int. Ed. 2011, 50, 12636–12638. [Google Scholar] [CrossRef] [PubMed]
  16. Marwitz, A.J.V.; Dutton, J.L.; Mercier, L.G.; Piers, W.E. Dihydrogen activation with tBu3P/B(C6F5)3: A chemically competent indirect mechanism via in situ-generated p-tBu2P-C6F4-B(C6F5)2. J. Am. Chem. Soc. 2011, 133, 10026–10029. [Google Scholar] [CrossRef] [PubMed]
  17. Lu, Z.P.; Cheng, Z.H.; Chen, Z.X.; Weng, L.H.; Li, Z.H.; Wang, H.D. Heterolytic cleavage of dihydrogen by “frustrated Lewis pairs” comprising bis(2,4,6-tris(trifluoromethyl)phenyl)borane and amines: Stepwise versus concerted mechanism. Angew. Chem. Int. Ed. 2011, 50, 12227–12231. [Google Scholar] [CrossRef] [PubMed]
  18. Ekkert, O.; Kehr, G.; Frohlich, R.; Erker, G. P-C bond activation chemistry: Evidence for 1,1-carboboration reactions proceeding with phosphorus-carbon bond cleavage. J. Am. Chem. Soc. 2011, 133, 4610–4616. [Google Scholar] [CrossRef] [PubMed]
  19. Appelt, C.; Westenberg, H.; Bertini, F.; Ehlers, A.W.; Slootweg, J.C.; Lammertsma, K.; Uhl, W. Geminal phosphorus/aluminum-based frustrated Lewis pairs: C-H versus C equivalent to C activation and CO2 fixation. Angew. Chem. Int. Ed. 2011, 50, 3925–3928. [Google Scholar]
  20. Ines, B.; Holle, S.; Goddard, R.; Alcarazo, M. Heterolytic S-S bond cleavage by a purely carbogenic frustrated Lewis pair. Angew. Chem. Int. Ed. 2010, 49, 8389–8391. [Google Scholar] [CrossRef] [PubMed]
  21. Grimme, S.; Kruse, H.; Goerigk, L.; Erker, G. The mechanism of dihydrogen activation by frustrated Lewis pairs revisited. Angew. Chem. Int. Ed. 2010, 49, 1402–1405. [Google Scholar] [CrossRef] [PubMed]
  22. Momming, C.M.; Otten, E.; Kehr, G.; Frohlich, R.; Grimme, S.; Stephan, D.W.; Erker, G. Reversible metal-free carbon dioxide binding by frustrated Lewis pairs. Angew. Chem. Int. Ed. 2009, 48, 6643–6646. [Google Scholar] [CrossRef] [PubMed]
  23. Holschumacher, D.; Bannenberg, T.; Hrib, C.G.; Jones, P.G.; Tamm, M. Heterolytic dihydrogen activation by a frustrated carbene-borane Lewis pair. Angew. Chem. Int. Ed. 2008, 47, 7428–7432. [Google Scholar]
  24. Chase, P.A.; Stephan, D.W. Hydrogen and amine activation by a frustrated Lewis pair of a bulky N-heterocyclic carbene and B(C6F5)3. Angew. Chem. Int. Ed. 2008, 47, 7433–7437. [Google Scholar] [CrossRef] [PubMed]
  25. Sumerin, V.; Schulz, F.; Nieger, M.; Leskelä, M.; Repo, T.; Rieger, B. Facile heterolytic H2 activation by amines and B(C6F5)3. Angew. Chem. Int. Ed. 2008, 47, 6001–6003. [Google Scholar] [CrossRef] [PubMed]
  26. Hounjet, L.J.; Bannwarth, C.; Garon, C.N.; Caputo, C.B.; Grimme, S.; Stephan, D.W. Combinations of ethers and B(C6F5)3 function as hydrogenation catalysts. Angew. Chem. Int. Ed. 2013, 52, 7492–7495. [Google Scholar] [CrossRef] [PubMed]
  27. Greb, L.; Daniliuc, C.G.; Bergander, K.; Paradies, J. Functional-group tolerance in frustrated Lewis pairs: Hydrogenation of nitroolefins and acrylates. Angew. Chem. Int. Ed. 2013, 52, 5876–5879. [Google Scholar] [CrossRef] [PubMed]
  28. Chernichenko, K.; Madarasz, A.; Papai, I.; Nieger, M.; Leskela, M.; Repo, T. A frustrated-Lewis-pair approach to catalytic reduction of alkynes to cis-alkenes. Nat. Chem. 2013, 5, 718–723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Miller, A.J.M.; Labinger, J.A.; Bercaw, J.E. Homogeneous CO hydrogenation: Dihydrogen activation involves a frustrated Lewis pair instead of a platinum complex. J. Am. Chem. Soc. 2010, 132, 3301–3303. [Google Scholar] [CrossRef] [PubMed]
  30. Eros, G.; Mehdi, H.; Papai, I.; Rokob, T.A.; Kiraly, P.; Tarkanyi, G.; Soos, T. Expanding the scope of metal-free catalytic hydrogenation through frustrated Lewis pair design. Angew. Chem. Int. Ed. 2010, 49, 6559–6563. [Google Scholar] [CrossRef] [PubMed]
  31. Axenov, K.V.; Kehr, G.; Frohlich, R.; Erker, G. Catalytic hydrogenation of sensitive organometallic compounds by antagonistic N/B Lewis pair catalyst systems. J. Am. Chem. Soc. 2009, 131, 3454–3455. [Google Scholar] [CrossRef] [PubMed]
  32. Ashley, A.E.; Thompson, A.L.; O’Hare, D. Non-metal-mediated homogeneous hydrogenation of CO2 to CH3OH. Angew. Chem. Int. Ed. 2009, 48, 9839–9843. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, X.W.; Kehr, G.; Daniliuc, C.G.; Erker, G. Internal adduct formation of active intramolecular C-4-bridged frustrated phosphane/borane Lewis pairs. J. Am. Chem. Soc. 2014, 136, 3293–3303. [Google Scholar] [CrossRef] [PubMed]
  34. Rocchigiani, L.; Ciancaleoni, G.; Zuccaccia, C.; Macchioni, A. Probing the association of frustrated phosphine-borane Lewis pairs in solution by NMR spectroscopy. J. Am. Chem. Soc. 2014, 136, 112–115. [Google Scholar] [CrossRef] [PubMed]
  35. Henthorn, J.T.; Agapie, T. Dioxygen reactivity with a ferrocene-Lewis acid pairing: Reduction to a boron peroxide in the presence of tris(pentafluorophenyl)borane. Angew. Chem. Int. Ed. 2014, 53, 12893–12896. [Google Scholar] [CrossRef] [PubMed]
  36. Sajid, M.; Kehr, G.; Wiegand, T.; Eckert, H.; Schwickert, C.; Pottgen, R.; Cardenas, A.J.P.; Warren, T.H.; Frohlich, R.; Daniliuc, C.G.; et al. Noninteracting, vicinal frustrated P/B-Lewis pair at the norbornane framework: Synthesis, characterization, and reactions. J. Am. Chem. Soc. 2013, 135, 8882–8895. [Google Scholar] [CrossRef] [PubMed]
  37. Menard, G.; Hatnean, J.A.; Cowley, H.J.; Lough, A.J.; Rawson, J.M.; Stephan, D.W. C-H bond activation by radical ion pairs derived from R3P/Al(C6F5)3 frustrated Lewis pairs and N2O. J. Am. Chem. Soc. 2013, 135, 6446–6449. [Google Scholar] [CrossRef] [PubMed]
  38. Zhao, X.X.; Stephan, D.W. Olefin-borane “van der Waals complexes”: Intermediates in frustrated Lewis pair addition reactions. J. Am. Chem. Soc. 2011, 133, 12448–12450. [Google Scholar] [CrossRef] [PubMed]
  39. Menard, G.; Stephan, D.W. Stoichiometric reduction of CO2 to CO by aluminum-based frustrated Lewis pairs. Angew. Chem. Int. Ed. 2011, 50, 8396–8399. [Google Scholar] [CrossRef] [PubMed]
  40. Chapman, A.M.; Haddow, M.F.; Wass, D.F. Frustrated Lewis pairs beyond the main group: Cationic zirconocene-phosphinoaryloxide complexes and their application in catalytic dehydrogenation of amine boranes. J. Am. Chem. Soc. 2011, 133, 8826–8829. [Google Scholar] [CrossRef] [PubMed]
  41. Cardenas, A.J.P.; Culotta, B.J.; Warren, T.H.; Grimme, S.; Stute, A.; Frohlich, R.; Kehr, G.; Erker, G. Capture of NO by a frustrated Lewis pair: A new type of persistent N-oxyl radical. Angew. Chem. Int. Ed. 2011, 50, 7567–7571. [Google Scholar] [CrossRef] [PubMed]
  42. Momming, C.M.; Kehr, G.; Wibbeling, B.; Frohlich, R.; Schirmer, B.; Grimme, S.; Erker, G. Formation of cyclic allenes and cumulenes by cooperative addition of frustrated Lewis pairs to conjugated enynes and diynes. Angew. Chem. Int. Ed. 2010, 49, 2414–2417. [Google Scholar] [CrossRef] [PubMed]
  43. Menard, G.; Stephan, D.W. Room temperature reduction of CO2 to methanol by Al-based frustrated Lewis pairs and ammonia borane. J. Am. Chem. Soc. 2010, 132, 1796–1797. [Google Scholar] [CrossRef] [PubMed]
  44. Berkefeld, A.; Piers, W.E.; Parvez, M. Tandem frustrated Lewis pair/tris(pentafluorophenyl) borane-catalyzed deoxygenative hydrosilylation of carbon dioxide. J. Am. Chem. Soc. 2010, 132, 10660–10661. [Google Scholar] [CrossRef] [PubMed]
  45. Alcarazo, M.; Gomez, C.; Holle, S.; Goddard, R. Exploring the reactivity of carbon(0)/borane-based frustrated Lewis pairs. Angew. Chem. Int. Ed. 2010, 49, 5788–5791. [Google Scholar] [CrossRef] [PubMed]
  46. Lindqvist, M.; Borre, K.; Axenov, K.; Kótai, B.; Nieger, M.; Leskelä, M.; Pápai, I.; Repo, T. Chiral molecular tweezers: Synthesis and reactivity in asymmetric hydrogenation. J. Am. Chem. Soc. 2015, 137, 4038–4041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wei, S.M.; Du, H.F. A highly enantioselective hydrogenation of silyl enol ethers catalyzed by chiral frustrated Lewis pairs. J. Am. Chem. Soc. 2014, 136, 12261–12264. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, J.; Lalancette, R.A.; Jäkle, F. Chiral organoborane Lewis pairs derived from pyridylferrocene. Chem. Eur. J. 2014, 20, 9120–9129. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, Y.B.; Du, H.F. Chiral dienes as “ligands” for borane-catalyzed metal-free asymmetric hydrogenation of imines. J. Am. Chem. Soc. 2013, 135, 6810–6813. [Google Scholar] [CrossRef] [PubMed]
  50. Chen, J.; Lalancette, R.A.; Jäkle, F. Synthesis and Lewis acid properties of a ferrocene-based planar-chiral borenium cation. Chem. Commun. 2013, 49, 4893–4895. [Google Scholar] [CrossRef] [PubMed]
  51. Ghattas, G.; Chen, D.J.; Pan, F.F.; Klankermayer, J. Asymmetric hydrogenation of imines with a recyclable chiral frustrated Lewis pair catalyst. Dalton Trans. 2012, 41, 9026–9028. [Google Scholar] [CrossRef] [PubMed]
  52. Sumerin, V.; Chernichenko, K.; Nieger, M.; Leskela, M.; Rieger, B.; Repo, T. Highly active metal-free catalysts for hydrogenation of unsaturated nitrogen-containing compounds. Adv. Synth. Catal. 2011, 353, 2093–2110. [Google Scholar] [CrossRef]
  53. Chen, J.; Venkatasubbaiah, K.; Pakkirisamy, T.; Doshi, A.; Yusupov, A.; Patel, Y.; Lalancette, R.A.; Jäkle, F. Planar chiral organoborane Lewis acids derived from naphthylferrocene. Chem. Eur. J. 2010, 16, 8861–8867. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, D.J.; Wang, Y.T.; Klankermayer, J. Enantioselective hydrogenation with chiral frustrated Lewis pairs. Angew. Chem. Int. Ed. 2010, 49, 9475–9478. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, D.J.; Klankermayer, J. Metal-free catalytic hydrogenation of imines with tris(perfluorophenyl)borane. Chem. Commun. 2008, 18, 2130–2131. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, E.Y.X. Polymerization by classical and frustrated Lewis pairs. Top. Curr. Chem. 2013, 334, 239–260. [Google Scholar] [PubMed]
  57. Chen, J.; Chen, E.Y.X. Lewis pair polymerization of acrylic monomers by N-heterocyclic carbenes and B(C6F5)3. Isr. J. Chem. 2015, 55, 216–225. [Google Scholar] [CrossRef]
  58. Zhang, Y.T.; Miyake, G.M.; John, M.G.; Falivene, L.; Caporaso, L.; Cavallo, L.; Chen, E.Y.X. Lewis pair polymerization by classical and frustrated Lewis pairs: Acid, base and monomer scope and polymerization mechanism. Dalton Trans. 2012, 41, 9119–9134. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, Y.T.; Miyake, G.M.; Chen, E.Y.X. Alane-based classical and frustrated Lewis pairs in polymer synthesis: Rapid polymerization of MMA and naturally renewable methylene butyrolactones into high-molecular-weight polymers. Angew. Chem. Int. Ed. 2010, 49, 10158–10162. [Google Scholar] [CrossRef] [PubMed]
  60. He, J.H.; Zhang, Y.T.; Falivene, L.; Caporaso, L.; Cavallo, L.; Chen, E.Y.X. Chain propagation and termination mechanisms for polymerization of conjugated polar alkenes by [Al]-Based frustrated Lewis pairs. Macromolecules 2014, 47, 7765–7774. [Google Scholar] [CrossRef]
  61. He, J.H.; Zhang, Y.T.; Chen, E.Y.X. Synthesis of pyridine- and 2-oxazoline-functionalized vinyl polymers by alane-based frustrated Lewis pairs. Synlett 2014, 25, 1534–1538. [Google Scholar]
  62. Piedra-Arroni, E.; Ladaviere, C.; Amgoune, A.; Bourissou, D. Ring-opening polymerization with Zn(C6F5)2-based Lewis pairs: Original and efficient approach to cyclic polyesters. J. Am. Chem. Soc. 2013, 135, 13306–13309. [Google Scholar] [CrossRef] [PubMed]
  63. Sajid, M.; Stute, A.; Cardenas, A.J.P.; Culotta, B.J.; Hepperle, J.A.M.; Warren, T.H.; Schirmer, B.; Grimme, S.; Studer, A.; Daniliuc, C.G.; et al. N,N-Addition of frustrated Lewis pairs to nitric oxide: An easy entry to a unique family of aminoxyl radicals. J. Am. Chem. Soc. 2012, 134, 10156–10168. [Google Scholar] [CrossRef] [PubMed]
  64. Xu, T.Q.; Chen, E.Y.X. Probing site cooperativity of frustrated phosphine/borane Lewis pairs by a polymerization study. J. Am. Chem. Soc. 2014, 136, 1774–1777. [Google Scholar] [CrossRef] [PubMed]
  65. Jia, Y.B.; Ren, W.M.; Liu, S.J.; Xu, T.Q.; Wang, Y.B.; Lu, X.B. Controlled divinyl monomer polymerization mediated by Lewis pairs: A powerful synthetic strategy for functional polymers. ACS Macro Lett. 2014, 3, 896–899. [Google Scholar] [CrossRef]
  66. Eisenberger, P.; Bailey, A.M.; Crudden, C.M. Taking the F out of FLP: Simple Lewis acid-base pairs for mild reductions with neutral boranes via borenium ion catalysis. J. Am. Chem. Soc. 2012, 134, 17384–17387. [Google Scholar] [CrossRef] [PubMed]
  67. Sumerin, V.; Schulz, F.; Atsumi, M.; Wang, C.; Nieger, M.; Leskela, M.; Repo, T.; Pyykko, P.; Rieger, B. Molecular tweezers for hydrogen: Synthesis, characterization, and reactivity. J. Am. Chem. Soc. 2008, 130, 14117–14119. [Google Scholar] [CrossRef] [PubMed]
  68. Schulz, F.; Sumerin, V.; Heikkinen, S.; Pedersen, B.; Wang, C.; Atsumi, M.; Leskela, M.; Repo, T.; Pyykko, P.; Petry, W.; et al. Molecular hydrogen tweezers: Structure and mechanisms by neutron diffraction, NMR, and deuterium labeling studies in solid and solution. J. Am. Chem. Soc. 2011, 133, 20245–20257. [Google Scholar] [CrossRef] [PubMed]
  69. Mountford, A.J.; Lancaster, S.J.; Coles, S.J.; Horton, P.N.; Hughes, D.L.; Hursthouse, M.B.; Light, M.E. Intra- and intermolecular N-H···F-C hydrogen-bonding interactions in amine adducts of tris(pentafluorophenyl)borane and -alane. Inorg. Chem. 2005, 44, 5921–5933. [Google Scholar] [CrossRef] [PubMed]
  70. Ning, Y.; Zhu, H.P.; Chen, E.Y.X. Remarkable Lewis acid effects on polymerization of functionalized alkenes by metallocene and lithium ester enolates. J. Organomet. Chem. 2007, 692, 4535–4544. [Google Scholar] [CrossRef]
  71. Rodriguez-Delgado, A.; Chen, E.Y.X. Single-site anionic polymerization. Monomeric ester enolaluminate propagator synthesis, molecular structure, and polymerization mechanism. J. Am. Chem. Soc. 2005, 127, 961–974. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, E.Y.X.; Kruper, W.J.; Roof, G.; Wilson, D.R. “Double activation” of constrained geometry and ansa-metallocene group 4 metal dialkyls: Synthesis, structure, and olefin polymerization study of mono-and dicationic aluminate complexes. J. Am. Chem. Soc. 2001, 123, 745–746. [Google Scholar] [CrossRef] [PubMed]
  73. Bolig, A.D.; Chen, E.Y.X. Reversal of polymerization stereoregulation in anionic polymerization of MMA by chiral metallocene and non-metallocene initiators: A new reaction pathway for metallocene-initiated MMA polymerization. J. Am. Chem. Soc. 2001, 123, 7943–7944. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, J.; Chen, E.Y.X. Elusive silane alane complex [Si-H···Al]: Isolation, characterization, and multifaceted frustrated-Lewis-pair-type catalysis. Angew. Chem. Int. Ed. 2015. [Google Scholar] [CrossRef]
  75. Di Saverio, A.; Focante, F.; Camurati, I.; Resconi, L.; Beringhelli, T.; D’Alfonso, G.; Donghi, D.; Maggioni, D.; Mercandelli, P.; Sironi, A. Oxygen-bridged borate anions from tris(pentafluorophenyl)borane: Synthesis, NMR characterization, and reactivity. Inorg. Chem. 2005, 44, 5030–5041. [Google Scholar] [CrossRef] [PubMed]
  76. Allen, R.D.; Long, T.E.; Mcgrath, J.E. Preparation of high-purity, anionic-polymerization grade alkyl methacrylate monomers. Polym. Bull. 1986, 15, 127–134. [Google Scholar] [CrossRef]
  77. Feng, S.G.; Roof, G.R.; Chen, E.Y.X. Tantalum (V)-based metallocene, half-metallocene, and non-metallocene complexes as ethylene-1-octene copolymerization and methyl methacrylate polymerization catalysts. Organometallics 2002, 21, 832–839. [Google Scholar] [CrossRef]
  78. Lee, C.H.; Lee, S.J.; Park, J.W.; Kim, K.H.; Lee, B.Y.; Oh, J.S. Preparation of Al(C6F5)3 and its use for the modification of methylalumoxane. J. Mol. Catal. A Chem. 1998, 132, 231–239. [Google Scholar] [CrossRef]
  79. Biagini, P.; Lugli, G.; Abis, L.; Andreussi, P. Organometallic derivatives of group IIIA and process for their preparation. U.S. Patent 5602269, 11 February 1997. [Google Scholar]
  • Sample Availability: Not available.

Share and Cite

MDPI and ACS Style

Chen, J.; Chen, E.X.-Y. Reactivity of Amine/E(C6F5)3 (E = B, Al) Lewis Pairs toward Linear and Cyclic Acrylic Monomers: Hydrogenation vs. Polymerization. Molecules 2015, 20, 9575-9590. https://doi.org/10.3390/molecules20069575

AMA Style

Chen J, Chen EX-Y. Reactivity of Amine/E(C6F5)3 (E = B, Al) Lewis Pairs toward Linear and Cyclic Acrylic Monomers: Hydrogenation vs. Polymerization. Molecules. 2015; 20(6):9575-9590. https://doi.org/10.3390/molecules20069575

Chicago/Turabian Style

Chen, Jiawei, and Eugene X.-Y. Chen. 2015. "Reactivity of Amine/E(C6F5)3 (E = B, Al) Lewis Pairs toward Linear and Cyclic Acrylic Monomers: Hydrogenation vs. Polymerization" Molecules 20, no. 6: 9575-9590. https://doi.org/10.3390/molecules20069575

Article Metrics

Back to TopTop