A Dinuclear Tricyclic Transition State Model for Carbonyl Addition of

May 25, 2018 - Abstract. In the presence of a catalytic amount of chiral BINOL derivatives (or BINOLs) a mixture of various organometallic compounds w...
0 downloads 0 Views 2MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

A Dinuclear Tricyclic Transition State Model for Carbonyl Addition of Organotitanium Reagents: DFT Study on the Activity and Enantioselectivity of BINOLate Titanium Catalysts Toshiro Harada J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.8b00712 • Publication Date (Web): 25 May 2018 Downloaded from http://pubs.acs.org on May 26, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

A Dinuclear Tricyclic Transition State Model for Carbonyl Addition of Organotitanium Reagents: DFT Study on the Activity and Enantioselectivity of BINOLate Titanium Catalysts Toshiro Harada* Faculty of Molecular Chemistry and Engineering, Kyoto Institute of Technology, Matsugasaki, Sakyo-ku, Kyoto, 606-8585, Japan [email protected]

Abstract: In the presence of a catalytic amount of chiral BINOL derivatives (or BINOLs) a mixture of various organometallic compounds with Ti(OiPr)4 undergoes enantioselective addition to aldehydes and ketones. Although the catalyst and reacting nucleophile of the reaction have been elucidated to be (BINOLate)Ti2(OiPr)6 and RTi(OiPr)3, respectively, little is known about the properties of short-lived intermediates and transition structures. In this work, the mechanism of this reaction is investigated with the aid of DFT (M06) calculations. The study provides support for the following mechanistic understandings: (i) The direct racemic reaction proceeds through a pathway involving initial aggregation of RTi(OiPr)3 with Ti(OiPr)4 followed by carbonyl addition of the resulting dinuclear aggregate. (ii) The enantioselective reaction takes place through a pathway involving initial ligand exchange of RTi(OiPr)3 with (BINOLate)Ti2(OiPr)6 followed by the addition of the resulting chiral dinuclear titanium species via a chiral BINOLate-chelated, tricyclic transition structure. (iii) The enantioselective pathway is favorable not because BINOLate ligands accelerate the carbonyl addition but because the ligands stabilize the chiral dinuclear species against deaggregation through a chelating bridge. (iv) The chiral transition structure serves as a model accounting for the re-face addition generally observed in the reaction of aldehydes with (R)-BINOLs. Introduction Catalytic enantioselective carbonyl addition of organometallic reagents has attracted great interests for its synthetic utility as a fundamental method for constructing chiral molecular frameworks.1–9 Of various approaches developed so far, chiral titanium catalysis derived from BINOL and H8-BINOL derivatives (BINOLs),10,11 such as 1a-c and 2a-c, have contributed significantly in expanding the scope of the reaction with respect to carbonyl substrates and organometallic reagents (Scheme 1). Not only aldehydes but also ketones undergo enantioselective addition with the BINOL-derived titanium catalyst system as illustrated in representative reactions in Table 1. With the BINOL catalyst system, enantioselective alkylation, arylation (including heteroarylation), and vinylation have been realized through the use of a variety of readily available organometallic reagents, such as organozinc, 12–17 -boron,18–20 aluminum,21–27 -magnesium,28–33 and -titanium reagents34–39

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme 1 Enantioselective Carbonyl Addition of Organometallic Reagents Catalyzed by Chiral Titanium Complexes Derived from BINOLs

Despite such diversity, several common features are found in these reactions. First, enantiofacial selectivity is governed solely by the absolute configuration of the axially chiral ligands irrespective of organometallic nucleophiles employed. Aldehydes always undergo selective re- and si-face addition with (R)- and (S)-BINOLs, respectively.40 Interestingly, the opposite selectivity is observed for ketones. Ketones undergo si- and re-face addition with the (R)- and (S)-ligands, respectively. Secondly, reactions are generally carried out with more than stoichiometric amount of a titanium alkoxide, specifically Ti(OiPr)4, except those with organotitanium reagents (Table 1). It has been revealed that Ti(OiPr)4 plays a dual role in generation of a dinuclear titanium complex 3, as a chiral catalyst, through ligand exchange (eq 1)41–46 and in formation of organotitanium reagent RTi(OiPr)3 (4), as a reacting species, through transmetalation (eq 2).[47,42,22] Thirdly, the reactions exhibit high enantioselectivity (typically >90% ee) although the organotitanium reagent 4 is known to undergo facile addition to aldehydes without the aid of catalysis.48,49,38 It should be noted that such feature is uncommon in other catalytic enantioselective carbonyl addition reactions, in which non- or less-reactive organometallic reagents are employed as the nucleophiles50–52 with the aid of ligand accelerated catalysis.53

ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Table 1 Representative Examples of Enantioselective Carbonyl Addition Catalyzed by BINOLDerived Chiral Titanium Complexes ref. entry RM (equiv) Ti(OiPr)4 BINOL mol% selec(equiv) s tivitya addition to aldehydes (Scheme 1-a) 1a 1 Et2Zn 3.0 1.4 20 re 12,13 2a 2 3.0 1.4 20 re 14 1b 3 3.0 1.4 0.5 re 15 2b 4 3.0 1.4 2 re 16 2b 5b RZnBr (R = alkyl) 2.7 7.2 5 re 18 2b 6 RBR’2, (R = alkyl, vinyl) 1.5 3.0 5 re 19,20 2a 7 Et3Al 3.0 1.2 20 re 21 2a 8 Ar3Al 1.2 1.3 10 re 22 9 RAlMe2 (R = vinyl) 3.0 3.0 2b 5 re 26,27 1b 10 RMgX (R = alkyl) 2.2 5.8 2 re 28,31 1a 11b 2.0 0.9–1.4 15–20 re 30,32 2b 12 ArMgBr 1.2 3.0 2 re 29,32 1a 13 RTi(OiPr)3 (R = alkyl) 1.2–2.0 0–2.0 10 re 35 2a 14 ArTi(OiPr)3 1.2 0 3–10 re 34 2c 15 1.5 0–0.2 0.25 re 38 addition to ketones (Scheme 1-b) 1a 16 Ar3Al 2.5 5.0 10 si 23 17 RAliBu2 (R = vinyl) 1.6 3.0 1a 10 si 25 1a 18 ArTi(OiPr)3 1.5 0.3 10 si 36 1c 19 1.5 0 2 si 39 a Enantiofacial selectivity of the carbonyl compounds when (R)-BINOLs are used. bMgBr2 was used as an additive. bBis[2-(N,Ń-dimethylamino)ethyl]ether was used as an additive. Recent reports from this laboratory revealed that the catalytic activity of the chiral titanium complexes is remarkably enhanced by the use of BINOL 1b,c and 2b,c, bearing a sterically demanding aryl group at the 3-position of parent BINOL (1a) and H8-BINOL (2a), respectively. 54 The enhanced activity of the resulting catalysts enabled us to expand the scope of organometallic nucleophiles (entries 5, 6, 9, 10, and 12 in Table 1) and reduce the amount the chiral ligands to a practical level (0.25–2 mol%) keeping high enantioselectivity and short reaction time (typically 2–3 h) (entries 4, 10, 12, 15, and 19). Although catalysts derived from parent ligands 1a and 2a exhibit high enantioselective at higher catalyst loadings (10–20 mol%), reduction in the amount of the catalysts significantly degrades the enantioselectivity owing to the concurrent direct addition of the nucleophiles leading to racemic products. The 3-substituted BINOL derived catalysts, on the other hand, is of high turnover frequency, overwhelming the direct racemic reaction,55 thus exhibiting high activity at the low catalyst loadings.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Because of the synthetic versatility and high activity, studies have been carried out toward the elucidation of the mechanism of the enantioselective carbonyl addition catalyzed by BINOL derived chiral titanium complexes.36,41,42,43,44 Although the previous studies have provided valuable information on the solution structures and fluxional equilibrium of catalytically relevant species (vide infra), little is known about the properties of experimentally inaccessible short-lived intermediates and transition structures.56 In this work, we have performed density functional theory (DFT) calculations to shed light on the nature of catalysis by BINOL derived chiral titanium complexes, from which we expected to answer (1) what the detailed mechanism of the direct racemic reaction is; (2) what the detailed mechanism of the catalytic enantioselective reaction is; (3) why the catalytic pathway is favorable relative to the direct pathway; and (4) what the transition structure model rationalizing observed enantioselectivity of the BINOL derived catalysts is. The study have revealed a novel dinuclear tricyclic transition structure, which provides a critical basis in considering the above questions (Scheme 2). Scheme 2 Dinuclear Transition Structures for Carbonyl Addition of Organometallic Nucleophiles

Computational Details The quantum chemical calculations were carried out with the Gaussian 0957 program package. Geometry optimizations and frequency calculations were performed at the M0658,59/BS1 level in CH2Cl2 using the SMD solvation model.60 BS1 designates a mixed basis set of LANL2DZ61,62 for the titanium atom and 6-31G(d)63 for other atoms. The results of frequency calculations were examined to confirm that each structure is a local minimum (no imaginary frequency) or a transition state (only one imaginary frequency). Thermodynamic corrections at 273.15 K (0 °C, the temperature frequently employed in the catalytic enantioselective addition) and 1 atm in CH2Cl2 were obtained at the M06/BS1 level. Considering from the size of molecules (up to 99 atoms), Grimme’s quasi-harmonic approximations (QHAs)64–66 were applied with the frequency cut-off value of 100 cm-1 to correct the effect of small frequency vibrational modes. The translational entropy in dichloromethane was evaluated by applying the Whitesides method,67 where the volume of a dichloromethane molecule and the density of dichloromethane solvent were taken as 56.6 Å3 and 1.3266 g/cm3, respectively.68 Natural bond order (NBO) calculations were performed by the NBO program69 using the wave function obtained from the M06/BS1 level in CH2Cl2 with the SMD model. The 3D structures were prepared using CYLview.70 Results and Discussion Model Reaction of Formaldehyde with MeTi(OMe)3 (5a) and MeTi2(OMe)7 (6a). In our computational modeling, methylation of formaldehyde with mononuclear MeTi(OMe)3 (5a) was examined first. The calculations show that the reaction starts with initial coordination of the carbonyl oxygen atom to the titanium center of 5a to form association complex AC-5a with a pseudo-trigonal-bipyramidal titanium center, followed by methyl addition to give product alkoxide P-5a via four-membered transition structure TS-5a (Figure 1). The Gibbs free energy change (ΔG) of the association step is +4.6 kcal/mol, in accord with lower Lewis acidity of titanium alkoxides.49d A free energy barrier for this pathway is +16.7 kcal/mol and overall reaction is significantly exergonic (–38.0 kcal/mol).

ACS Paragon Plus Environment

Page 4 of 19

Page 5 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Figure 1. Calculated energy profile for the addition of MeTi(OMe)3 (5a) and MeTi2(OMe)7 (6a) to formaldehyde. The relative Gibbs and zero-point corrected potential energy (in parentheses) obtained from the M06/BS1 calculations in CH2Cl2 (SMD) are given in kcal/mol. The participation of dimeric species in the carbonyl addition of Grignard reagents and organolithium reagents has been demonstrated experimentally.71,72 Recent theoretical computational studies have revealed that these reactions, as well as the reaction of bimetallic organozincates (R3ZnLi), proceeds via a ladder-like transition state (Scheme 2).73–75 Considering the tendency of titanium alkoxides to form aggregate complexes, 76–80 the reaction of formaldehyde with dinuclear MeTi2(OMe)7 (6a) was also chosen as a model reaction. Dinuclear aggregate complex 6a is calculated to be thermodynamically stable, requiring +11.8 kcal/mol free energy for deaggregation to give mononuclear 5a and Ti(OMe)4 (Figure 1). The two titanium centers of 6a are both penta-coordinated pseudo-trigonal-bipyramidal, in which each μ-oxo bridge takes an apical and an equatorial position of one and the other titanium atom, respectively (Figure 2). This structure is generally found in the X-ray structural analysis of relevant dinuclear titanium alkoxide complexes.35,42,79–84 The reaction of formaldehyde with 6a first forms association complex AC-6a with the free energy change of +2.1 kcal/mol. The value is smaller than that of AC-5a (+4.6 kcal/mol), suggesting higher Lewis acidity of 6a. Ti2 of AC6a is hexa-coordinated pseudo-octahedral, coordinated to the oxygen lone pair of the aldehyde in the carbonyl plane. The C–C bond formation occurs as the carbonyl carbon attached to Ti2 approaches the methyl group on Ti1 through a characteristic transition structure TS-6a, in which the carbonyl oxygen interacts simultaneously with Ti2 and Ti1 in the direction coplanar and perpendicular to the carbonyl plane, respectively.85 As Ti1 and Ti2 are also bridged by the two μoxo oxygen atoms, O3 and O4, TS-6a has a tricyclic structure. Both Ti1 and Ti2 centers are hexacoordinated pseudo-octahedral, where Ti1–O3–O4 plane and Ti2–O3–O4 plane are not coplanar, bending about 38°. All attempts to locate a ladder-like transition state in which carbonyl oxygen atom interacts with a single titanium center resulted in a change of the geometry to generate TS6a with a simultaneous interaction with the two titanium centers.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Calculated structures of titanium complexes and key intermediates in the methylation of formaldehydes with 5a and 6a with distance in angstrom and NBO charges in parentheses. TS-6a is located +9.0 kcal/mol in free energy above the reactants (6a and formaldehyde). The value is significantly smaller than that of monocyclic transition structure TS-5a, indicating that the reaction pathway involving dinuclear aggregate 6a is preferable at least for the model reaction of the methoxy derivatives. The pathway via TS-6a leads to the formation of dinuclear product complex P-6a with high free energy gain (–37.8 kcal/mol). In the transition structure TS-5a and TS-6a, the CH3–Ti bond length is elongated 6.0% and 5.3%, respectively, as compared with the corresponding association complexes (Figure 2). The extent of bond elongation is considerably smaller than a value (21%) reported for the addition of Me3ZnLi to formaldehyde via a ladder-like transition state.75 In addition, the forming C–C bond lengths are relatively long in the transition structures (2.355 and 2.377 Å for TS-5a and TS-6a, respectively).86 Furthermore, the natural bond orbital (NBO) charge change of the methyl carbon atom is very small from AC-5a to TS-5a (–1.035  –1.017) and almost zero from AC-6a to TS6a (–0.993  –0.994), showing that the additional charge separation is not required in the transition states for the C–C bond formation. All these results, together with high exergonicity of the reactions, indicate that TS-5a and TS-6a lay early on the C–C bond forming step. The positive charge on the carbonyl carbon is decreased from the associate complexes to the transition structures. In accord with the lower free energy barrier, the charge change of the process for 6a (+0.321  +0.250) is smaller than that for of 5a (+0.317  +0.219), implying that TS-6a is even earlier than TS-5a. Model Reaction of Formaldehyde with MeTi(OiPr)3 (5b), MeTi2(OiPr)7 (6b), and Me((R)BINOLate)Ti2(OiPr)5 (7b). Previous studies on the structure of titanium alkoxides Ti(OR)4 have shown the formation of polynuclear aggregates [Ti(OR)4]n in solution, where n, or an association number, depends on the size of the alkoxy group.76–78 The titanium alkoxide aggregates undergo rapid ligand exchange to form fluxional mixture of several species.87,88 The mean value of the association number n of the mixture in benzene is about 3 for Ti(OR)4 (R = Et, Pr, Bu) and that for sterically demanding Ti(OiPr)4 is close to 1.89 On the other hand, a dinuclear aggregate structure is highly favorable for titanium complexes derived from BINOLs and Ti(OiPr)4.36,38,41– 46,79 Thus, BINOLs react with two equivalent of Ti(OiPr)4 to form dinuclear complex 3 (eq 1) and the reaction of BINOLs and Ti(OiPr)4 in a 1:1 molar ratio results in the formation of dimeric aggregate complex 8 (eq 3). In solution, complex 3 is in equilibrium with complex 8 and Ti(OiPr)4.41–44 Under the reaction conditions of the catalytic enantioselective carbonyl additions,

ACS Paragon Plus Environment

Page 6 of 19

Page 7 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

where Ti(OiPr)4 is employed in large excess with respect to the ligands, the equilibrium of eq 3 is considerably shifted to the right-hand side and dimeric complex 8 is negligible.38

The DFT-calculation of the model reaction showed that the pathway involving of MeTi2(OMe)7 (6a), via tricyclic transition structure TS-6a, is energetically more favorable than that involving MeTi(OMe)3 (5a), via four-membered transition structure TS-5a (Figure 1). However, the lower aggregation tendency of sterically demanding Ti(OiPr)4 suggests that the formation of isopropoxy derivative MeTi2(OiPr)7 (6b) from MeTi(OiPr)3 (5b) and Ti(OiPr)4 is unfavorable at equilibrium (eq 4). Accordingly, in considering the real reaction, a lower concentration of 6b at the equilibrium should be also taken into account. On the other hand, dinuclear BINOLate complex Me((R)BINOLate)Ti2(OiPr)5 (7b) might be stabilized by virtue of the chelating bridge of the bidentate ligand and not prone to deaggregation (eq 5). Indeed, Ph((S)-BINOLate)Ti2(OiPr)5, a phenyl derivative of 7b, has been prepared by Gau and coworkers by treatment of [((S)BINOLate)Ti(OiPr)2]2 with PhTi(OiPr)3, and its structure has been determined by X-ray crystallography.36 It has been proposed that the chiral dinuclear complex undergoes enantioselective carbonyl addition under the catalytic conditions. The dinuclear BINOLate complex 7b, as a chiral methylation reagent, would undergo enantioselective carbonyl addition via a tricyclic transition structure. Because of the anticipated stability of 7b against deaggregation, a pathway involving 7b, even with a catalytic amount, could well overwhelm the direct pathway involving 6b, a minor component of the aggregation equilibrium.

To test the above hypothesis, the model reaction of formaldehyde with isopropoxy derivative 5b, 6b, and 7b was examined. As there are many conformers (3n; n = 3, 7, and 5 for 5b, 6b, and 7b, respectively) with respect to the isopropyl groups, we did not take a comprehensive approach in calculations but focused on a representative conformer. Although being less quantitative, the approach provided a basis for examining the above hypothesis. Mononuclear reagent 5b reacts with formaldehyde via association complex AC-5b and fourmembered transition structure TS-5b with a free energy profile almost similar to that of methoxy derivative 5a (Figure 3, 4, and S-1). The pathway has a free energy barrier of +18.7 kcal/mol. On

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the other hand, a free energy barrier for the addition of dinuclear reagent 6b via AC-6b and tricyclic transition structure TS-6b is increased by +3.7 kcal/mol in comparison with that of the methoxy derivative 6a, presumably owing to the steric congestion around the isopropyl groups in the transition structure of the two hexa-coordinated pseudo-octahedral titanium centers. However, the free energy barrier of 12.7 kcal/mol is sufficiently lower than that of 5b.90

Figure 3. Calculated energy profile for the addition of MeTi(OiPr)3 (5b), MeTi2(OiPr)7 (6b), and Me((R)-BINOLate)Ti2(OiPr)5 (7b) to formaldehyde. The relative Gibbs and zero-point corrected potential energy (in parentheses) obtained from the M06/BS1 calculations in CH2Cl2 (SMD) are given in kcal/mol.

Figure 4 Calculated structures of titanium complexes and key intermediates in the methylation of formaldehydes with 7b. The addition of dinuclear BINOLate species 7b94 proceeds with initial formation of association complex AC-7b with a free energy loss of +4.7 kcal/mol, followed by tricyclic transition structure TS-7b(a), which is the most probable structure of several other isomeric transition structures rationalizing the experimentally observed enantioselectivity as will be discussed in the next section. The free energy barrier of this pathway (+12.2 kcal/mol) is almost same as that involving 6b, implying that the ligand acceleration is not prominent.

ACS Paragon Plus Environment

Page 8 of 19

Page 9 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Previously, the association number n of MeTi(OiPr)3 (5b) was determined by cryoscopy as 1.12 in benzene (0.026 M).89 From this value, it is evaluated that the molar fraction of dimer [MeTi(OiPr)3]2 is 12% and ΔG for the deaggregation is +1.0 kcal/mol at 0 °C. To check the validity of the current DFT method, ΔG for the deaggregation of [MeTi(OiPr)3]2 to give 5b in dichloromethane was calculated using M06/BS1. A free energy change of +2.3 kcal/mol is obtained for this process. Accordingly, the calculated energy is in agreement with the experimentally estimated value while being with some overestimation. A free energy change of +2.7 kcal/mol is obtained for deaggregation of MeTi2(OiPr)7 (6b) to give MeTi(OiPr)3 (5b) and Ti(OiPr)4 (Figure 3). As noted before, the dinuclear aggregate structures are highly stabilized against deaggregation by the μ-oxo bridge of BINOLates (eq 5). Indeed, the calculations showed that deaggregation of Me((R)-BINOLate)Ti2(OiPr)5 (7b) to give mononuclear ((R)-BINOLate)Ti(OiPr)2 (9b) and 5b is highly unfavorable with a free energy of +13.4 kcal/mol. The thermodynamic stability of dinuclear aggregate 7b against deaggregation as well as the free energy profile of the more realistic model reaction of isopropoxy derivatives 5b, 6b, and 7b provides strong support for the hypothesis that rationalizes the preferential catalytic enantioselective reaction overriding the direct racemic reaction. Although the present results indicate that direct racemic reaction proceeds with dinuclear aggregate 6b via TS-6b (Figure 3), kinetic study that supports alternative pathway involving mononuclear transition structure TS-5b has been reported by Reetz and Maus.48 The reaction of heptanal with MeTi(OiPr)3 (5b) in CH2Cl2 exhibited second-order kinetics, first order in the aldehyde and first order in 5b. The result led them to conclude that reacting species is not dinuclear aggregate [MeTi(OiPr)3]2 but mononuclear 5b itself. However, the product alkoxide of the reaction, [C6H13(Me)CHO]Ti(OiPr)3, closely resembles Ti(OiPr)4 and hence would undergo aggregation with 5b to form [MeTi(OiPr)3][{C6H13(Me)CHO}Ti(OiPr)3] (6b’). Therefore, the reported kinetics would be also in accord with a pathway involving dinuclear [MeTi(OiPr)3]2 and 6b’ as reacting species, provided that these species have similar reactivity and aggregation tendency. Enantioselectivity. The high enantioselectivity of chiral titanium complexes derived from BINOLs has attracted the elucidation of its origin. There is a common trends in enantioselectivity, that is R ligands preferentially gives the product derived from re-face attack of the aldehydes. The trend implies that there exists a common structure in the transition states.56 To clarify the structure, DFT calculations were carried out for possible isomeric transition structures for a model reaction of formaldehyde with Me((R)-BINOLate)Ti2(OMe)5 (7a). The core skeleton of the tricyclic transition structure (TS-6) is chiral (Scheme 3). Generally, enantioselectivities are higher for aromatic aldehydes (ArCHO) in comparison with sterically less demanding nonbranched aliphatic aldehydes (CH3(CH2)nCHO). When this transition-structure core is applied to the reaction of aldehydes, the carbonyl group, coordinating to Ti2 in a sterically favorable anti-manner, undergoes re-face addition. On the other hand, si-face addition takes place through enantiomeric transition-structure core ent-TS-6. There are each eleven distinguished cis coordination cites (i.e., a–b, a–c, a–e, a–f, b–c, b–d, b–f, b–g, c–d, c–e, and a–d) in TS-6 and in ent-TS-6 for bidentate ligands. The introduction of axially chiral (R)-BINOLate to TS-6 and entTS-6 at the same cis cites results in the formation a pair of diastereomeric transition structures. Of total twenty two candidates of (R)-BINOLate derived transition structures for the addition of 7a to formaldehyde, six are sterically infeasible. Transition state search for the others gave sixteen transition structures TS-7a(a) – TS-7a(p) (Figure 5 and S-2) in increasing order of zero-point corrected potential energy (Table 2).

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme 3 Enantiomeric Core Skeletons of Tricyclic Transition Structure

Figure 5 Calculated geometries of stable transitions structures in the addition of Me((R)BINOLate)Ti2(OMe)5 (7a) to formaldehyde. The relative zero-point corrected potential energy (in parentheses) obtained from the M06/BS1 calculations in CH2Cl2 (SMD) is given in kcal/mol. Table 2 Relative Potential Energy of Transition Structures for the Addition of Me((R)BINOLate)Ti2(OMe)5 (7a) to Formaldehyde structure chelating cite enantiofacial ΔEa

ACS Paragon Plus Environment

Page 10 of 19

Page 11 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

selectivity (kcal/mol) TS-7a(a) TS-6; a–e re 0.0 TS-7a(b) ent-TS-6; c–e si 0.9 TS-7a(c) TS-6; d–e re 0.9 TS-7a(d) ent-TS-6; d–e si 1.1 TS-7a(e) TS-6; c–e re 3.0 TS-7a(f) ent-TS-6; b–f si 3.2 TS-7a(g) TS-6; c–d re 3.2 TS-7a(h) ent-TS-6; b–d si 3.2 TS-7a(i) TS-6; a–f re 3.3 TS-7a(j) TS-6; b–g re 3.7 TS-7a(k) ent-TS-6; c–d si 3.8 TS-7a(l) ent-TS-6 ;a–c si 4.7 TS-7a(m) TS-6; b–c re 4.8 TS-7a(n) ent-TS-6; f–g si 5.8 TS-7a(o) TS-6; f–g re 8.5 TS-7a(p) TS-6; b–d re 11.7 a The relative zero-point corrected potential energy obtained from the M06/BS1 calculations in CH2Cl2 (SMD).

The potential energy differences (ΔE) of these transition structures are within 11.7 kcal/mol (Table 2). There are five relatively low energy structures within an energy window of 3.0 kcal/mol. The most stable transition structure TS-7a(a) is lower in energy than the second stable TS-7a(b) and TS-7a(c) by 0.9 kcal/mol. In TS-7a(a), (R)-BINOLate occupies coordination cite a–e of TS6. One oxygen atom of the ligand bridges between Ti1 and Ti2 at cite a and another is coordinated to Ti2 at cite e. In the second stable TS-7a(b), (R)-BINOLate takes coordination cite c–e of entTS-6, chelating only Ti2 without participating in μ-oxo bridge. Also in TS-7a(c), TS-7a(d) (ΔE = 1.1 kcal/mol), and TS-7a(e) (ΔE = 3.0 kcal/mol), Ti2 is chelated by the ligand at cite d–e of TS6, at cite d–e of ent-TS-6, and at cite c–e of TS-6, respectively.95 When applied as the transition structure model for the reaction of aldehydes, TS-7a(a), TS-7a(c), and TS-7a(e) induce experimentally observed re-face addition. On the other hand TS-7a(b) and TS-7a(d) leads to opposite si-face addition products. In considering the real reaction, the participation of the sterically demanding isopropoxy group in μ-oxo bridge might induce energy increase in transition states. Therefore, the potential energy of the isopropoxy derivatives are estimated to be increased for TS-7a(b), TS-7a(c), TS7a(d), and TS-7a(e), with two μ-OMe groups at coordination cite a and e, in comparison to TS7a(a) with one μ-OMe group. Indeed, potential energy of TS-7b(b) and TS-7b(c) (the isopropoxide derivative of TS-7a(b) and TS-7a(c), respectively) was calculated to be +8.4 and +3.4 kcal/mol relative to TS-7b(a) (the isopropoxide derivative of TS-7a(a)), respectively (Figure 6). Accordingly, TS-7a(a) serves as the most likely transition structure model for the catalytic enantioselective re-face addition to aldehydes.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6 Calculated geometries of transitions structure TS-7b(b) and TS-7b(c) in the addition of Me((R)-BINOLate)Ti2(OiPr)5 (7b) to formaldehyde. The zero-point corrected potential energy (in parentheses) relative to that of TS-7b(a) obtained from the M06/BS1 calculations in CH2Cl2 (SMD) is given in kcal/mol. At the present stage, it is difficult to propose a transition structure model for the addition to ketones that exhibit opposite si-facial selectivity. For ketones (RCOR’), TS-7a(a) is incompatible because smaller substituent R’ (e.g., methyl group) on the carbonyl group, syn to the coordinating Ti2, would cause unfavorable interaction with the naphthalene ring of the ligand, while the other four structures are sterically feasible (Figure 5). TS-7a(b) and TS-7a(d) are candidates of the transition structure model for ketones compatible with the observed selectivity. The present study showed that the enantioselectivity is not directly controlled by the unfavorable (or favorable) interaction between the chiral ligand (BINOLs) and substrates in transition states. The absolute steric course of the reaction is governed by a match and mismatch between the ligand and the chiral transition-structure core (TS-6 and ent-TS-6). As shown in Scheme 4, TS-6 well matches with (R)-BINOLate to accommodate the ligand at cite a–e without steric constraint. On the other hand, it is sterically infeasible to place (R)-BINOLate in ent-TS-6 at the same cite and the pair is mismatched. A matched pair of TS-6 and (R)-BINOLate results in TS-7a(a), through which re face addition of aldehydes might take place.

Scheme 4 Match and Mismatch in Chiral Ligand and Transition Structure Core Conclusion The mechanisms for the enantioselective carbonyl addition of organotitanium reagents RTi(OiPr)3 catalyzed by BINOLate titanium complex 3 have been investigated with the aid of DFT (M06) calculations. The study reveals the low-energy, dinuclear tricyclic transition structure for the addition of the organotitanium reagents in aggregate form and provides support for the following mechanistic understanding of the reaction (Scheme 5): (1) The direct racemic reaction of RTi(OiPr)3 proceeds through a pathway involving initial aggregation with Ti(OiPr)4 followed by carbonyl addition of the resulting dinuclear aggregate 10 via tricyclic transition structure TS11. (2) On the other hand, the enantioselective reaction takes place through a pathway involving initial ligand exchange of RTi(OiPr)3 with BINOLate titanium complex 3 followed by the addition of the resulting chiral dinuclear titanium species 12 via the corresponding chiral tricyclic transition structure TS-13. (3) The enantioselective pathway is favorable relative to the racemic pathway not because BINOLate ligands accelerate the carbonyl addition but because the ligands substantially stabilize chiral dinuclear species 12 against deaggregation through the chelating bridge, thus keeping its concentration higher than that of 10. (4) Chiral transition structure TS-13 serves as a model accounting for the re-face addition commonly observed in the reaction of aldehydes with (R)-BINOLs.

ACS Paragon Plus Environment

Page 12 of 19

Page 13 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

Scheme 5 Plausible Mechanism for the Enantioselective Carbonyl Addition of Organotitanium Reagents Catalyzed by BINOLate Titanium Complexes Supporting Information. Calculated structures of titanium complexes and key intermediates in the methylation of formaldehydes with 5b and 6b (Figure S-1). Calculated geometries of transition structures in the addition of 7a to formaldehyde (Figure S-2). Tables listing energies, the optimized Cartesian coordinates, and the imaginary frequencies of TSs (Table S-1 and S-3). Relative electronic and free energies of TS-5b and TS-6b calculated using M06-D3/B1 (Table S2). Acknowledgments We thank Prof. Hisayoshi Kobayashi of Kyoto Institute of Technology for helpful discussion. Thanks are also due to Professor Robert S. Paton (Colorado State University) for supporting with the QHA script. This work was supported by KAKENHI (No. 15K05500) from Ministry of Education,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Culture, Sports, Science, and Technology (MEXT), Japan.

References and Notes (1) Soai, K.; Niwa, S. Enantioselective Addition of Organozinc Reagents to Aldehydes. Chem. Rev. 1992, 92, 833–856. (2) Pu, L. Yu, H.-B. Catalytic Asymmetric Organozinc Additions to Carbonyl Compounds. Chem. Rev. 2001, 101, 757–824, (3) Pu, L. Asymmetric Alkynylzinc Additions to Aldehydes and Ketones. Tetrahedron 2003, 59, 9873–9886. (4) Hatano, M.; Miyamoto, T.; Ishihara, K. Recent Progress in Selective Additions of Organometal Reagents to Carbonyl Compounds. Curr. Org. Chem. 2006, 11, 127–157. (5) Schmidt, F.; Stemmler, R. T.; Rudolph, J.; Bolm, C. Catalytic Asymmetric Approaches towards Enantiomerically Enriched Diarylmethanols and Diarylmethylamines. Chem. Soc. Rev. 2006, 35, 454–470. (6) Trost, B. M.; Weiss, A. H. The Enantioselective Addition of Alkyne Nucleophiles to Carbonyl Groups. Adv. Synth. Catal. 2009, 351, 963−983. (7) Yus, M.; Gonzales-Gomez, J. C.; Foubelo, F. Catalytic Enantioselective Allylation of Carbonyl Compounds and Imines. Chem. Rev. 2011, 111, 7774−7854. (8) Pellissier, H. Enantioselective Titanium-Catalysed 1,2-Additions of Carbon Nucleophiles to Carbonyl Compounds. Tetrahedron, 2015, 71, 2487–2524. (9) Collados, J. F.; Sola, R.; Harutyunyan, S. R.; Macia, B. Catalytic Synthesis of Enantiopure Chiral Alcohols via Addition of Grignard Reagents to Carbonyl Compounds. ACS Cat. 2016, 1952–1970. (10) Throughout this paper, the italic “BINOL” denotes the parent BINOL and H8-BINOL as well as their derivatives. Non-italicized “BINOL” and “H8-BINOL” are used as they are. (11) Brunel, J. M. Update 1 of: BINOL: A Versatile Chiral Reagent. Chem. Rev. 2007, 107, PR1−PR45. (12) Mori, M.; Nakai, T. Asymmetric Catalytic Alkylation of Aldehydes with Diethylzinc Using a Chiral Binaphthol-Titanium Complex. Tetrahedron Lett. 1997, 38, 6233–6236. (13) Zhang, F.-Y.; Yip, C.-W.; Cao, R.; Chan, A. S. C. Enantioselective Addition of Diethylzinc to Aromatic Aldehydes Catalyzed by Ti(BINOL) Complex. Tetrahedron: Asymmetry, 1997, 8, 585–589. (14) Zhang, F.-Y.; Chan, A. S. C. Enantioselective Addition of Diethylzinc to Aromatic Aldehydes Catalyzed by Titanium-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-bi-2-naphthol Complex. Tetrahedron: Asymmetry, 1997, 8, 3651–3655. (15) Harada, T.; Kanda, K. Enantioselective Alkylation of Aldehydes Catalyzed by a Highly Active Titanium Complex of 3-Substituted Unsymmetric BINOL. Org. Lett. 2006, 8, 3817–3819. (16) Harada, T.; Ukon, T. Practical and Highly Enantioselective Alkylation of Aldehydes Catalyzed by a Titanium Complex of 3-Aryl H8-BINOL. Tetrahedron: Asymmetry 2007, 18, 2499–2502. (17) Ukon, T.; Harada, T. Catalytic Asymmetric Alkylation of Aldehydes by Using Trialkylboranes. Eur. J. Org. Chem. 2008, 4405–4407. (18) Kinoshita, Y.; Kanehira, S.; Hayashi, Y.; Harada, T. Catalytic Enantioselective Alkylation of Aldehydes by Using Organozinc Halide Reagents. Chem. Eur. J. 2013, 19, 3311–3314. (19) Kumar, R.; Kawasaki, H.; Harada, T. Enantioselective Alkylation of Aldehydes Using Functionalized Alkylboron Reagents Catalyzed by a Chiral Titanium Complex. Org. Lett. 2013, 15, 4198–4201. (20) Shono, T.; Harada, T. Catalytic Enantioselective Synthesis of Secondary Allylic Alcohols

ACS Paragon Plus Environment

Page 14 of 19

Page 15 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

from Terminal Alkynes and Aldehydes via 1-Alkenylboron Reagents. Org. Lett. 2010, 12, 5270– 5273. (21) Chan, A. S. C.; Zhang, F.-Y.; Yip, C.-W. Novel Asymmetric Alkylation of Aromatic Aldehydes with Triethylaluminum Catalyzed by Titanium·(1,1‘-bi-2-naphthol) and Titanium·(5,5‘,6,6‘,7,7‘,8,8‘-octahydro-1,1‘- bi-2-naphthol) Complexes. J. Am. Chem. Soc. 1997, 119, 4080–4081. (22) Wu, K.-H.; Gau, H.-M. Remarkably Efficient Enantioselective Titanium(IV)−(R)-H8BINOLate Catalyst for Arylations to Aldehydes by Triaryl(tetrahydrofuran)aluminum Reagents. J. Am. Chem. Soc. 2006, 128, 14808–14809. (23) Chen, C.-A.; Wu, K.-H.; Gau, H.-M. Highly Enantioselective Aryl Additions of [AlAr3(thf)] to Ketones Catalyzed by a Titanium(IV) Catalyst of (S)-Binol. Angew. Chem. Int. Ed. 2007, 46, 5373 –5376. (24) Zhou, S.; Wu, K.-H.; Chen, C.-A.; Gau H.-M. Highly Enantioselective Arylation of Aldehydes and Ketones Using AlArEt2(THF) as Aryl Sources. J. Org. Chem., 2009, 74, 3500– 3505. (25) Biradar, D. B.; Gau, H.-M. Highly Enantioselective Vinyl Additions of Vinylaluminum to Ketones Catalyzed by a Titanium(IV) Catalyst of (S)-BINOL. Org. Lett. 2009, 11, 499–502. (26) Kumar, R.; Kawasaki, H.; Harada, T. Catalytic Enantioselective Synthesis of α‐Substituted Secondary Allylic Alcohols from Terminal Alkynes and Aldehydes via Vinylaluminum Reagents. Chem. Eur. J. 2013, 19, 17707–17710. (27) Morimoto, H.; Harada, T. Catalytic Enantioselective Synthesis of Trisubstituted Secondary Allylic Alcohols Starting from Terminal Alkynes and Aldehydes. Eur. J. Org. Chem. 2015, 7378– 7383. (28) Muramatsu, Y.; Harada, T. Catalytic Asymmetric Alkylation of Aldehydes with Grignard Reagents. Angew. Chem. 2008, 120, 1088–1090. (29) Muramatsu, Y.; Harada, T. Catalytic Asymmetric Aryl Transfer Reactions to Aldehydes with Grignard Reagents as Aryl Source. Chem. Eur. J. 2008, 14, 10560–10563. (30) Da, C.-S.; Wang, J.-R.; Yin, X.-G.; Fan, X.-Y.; Liu, Y.; Yu, S.-L. Highly Catalytic Asymmetric Addition of Deactivated Alkyl Grignard Reagents to Aldehydes. Org. Lett. 2009, 11, 5578−5581. (31) Muramatsu, Y.; Kanehira, S.; Tanigawa, M.; Miyawaki, Y.; Harada, T. Catalytic Enantioselective Alkylation and Arylation of Aldehydes by Using Grignard Reagents. Bull. Chem. Soc. Jpn. 2010, 83, 19–32. (32) Liu, Y.; Da, C.-S.; Yu, S.-L.; Yin, X.-G.; Wang, J.-R.; Fan, X.-Y.; Li, W.-P.; Wang, R. Catalytic Highly Enantioselective Alkylation of Aldehydes with Deactivated Grignard Reagents and Synthesis of Bioactive Intermediate Secondary Arylpropanols. J. Org. Chem. 2010, 75, 6869−6878. (33) Itakura, D.; Harada, T. Enantioselective Arylation of Aldehydes by Using Functionalized Grignard Reagents Generated from Aryl Bromides. Synlett 2011, 2875–2879. (34) Wu, K.-H.; Zhou, S.; Chen, C.-A.; Yang, M.-C.; Chiang, R.-T.; Chen, C.-R.; Gau, H.-M. Instantaneous Room-Temperature and Highly Enantioselective ArTi(O-i-Pr)3 Additions to Aldehydes. Chem. Commun., 2011, 47, 11668–11670. (35) Li, Q.; Gau, H.-M. Room Temperature and Highly Enantioselective Additions of Alkyltitanium Reagents to Aldehydes Catalyzed by a Titanium Catalyst of (R)-H8-BINOL. Chirality 2011, 23, 929–939. (36) Wu, K.-H.; Kuo, Y.-Y.; Chen, C.-A.; Huang, Y.-L.; Gau, H.-M. Enantioselective Additions of Aryltitanium Tris(isopropoxide) to Ketones: Structure of [(i-PrO)2Ti{μ-(S)-BINOLate}(μ-O-iPr)TiPh(O-i-Pr)2], Study of Mechanistic and Stereochemical Insights. Adv. Synth. Catal. 2013, 355, 1001–1008. (37) Uenishi, A.; Nakagawa, Y.; Osumi, H.; Harada, T. Practical Enantioselective Arylation and Heteroarylation of Aldehydes with In Situ Prepared Organotitanium Reagents Catalyzed by 3‐

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Aryl‐H8‐BINOL‐Derived Titanium Complexes. Chem. Eur. J. 2013, 19, 4896–4905. (38) Hayashi, Y.; Yamamura, N.; Kusukawa, T.; Harada, T. Enhancement of Catalytic Activity of Chiral H8-BINOL Titanium Complexes by Introduction of Sterically Demanding Group at the 3 Position. Chem. Eur. J. 2016, 22, 12095–12105. (39) Matsuda, A.; Ushimaru, T.; Kobayashi, Y.; Harada, T. Catalytic Enantioselective Arylation and Heteroarylation of Ketones with Organotitanium Reagents Generated In Situ. Chem. Eur. J. 2017, 23, 8605–8609. (40) For nice summary of this phenomenon, see; Reference 35. (41) Boyle, T. J.; Barnes, D. L.; Heppert, J. A.; Morales, L; Takusagawa, F. Kinetics and Thermodynamics of Intra- and Intermolecular Rearrangement in Binaphtholate Complexes of Titanium(IV). Organometallics 1992, 11, 1112–1126. (42) Balsells, J.; Davis, T. J.; Carroll, P.; Walsh, P. J. Insight into the Mechanism of the Asymmetric Addition of Alkyl Groups to Aldehydes Catalyzed by Titanium−BINOLate Species. J. Am. Chem. Soc. 2002, 124, 10336–10348. (43) Waltz, K. M.; Carroll, P. J.; Walsh, P. J. Solid State Structures and Solution Behavior of Titanium(IV) Octahydrobinaphtholate Complexes. Examination of Nonlinear Behavior in the Asymmetric Addition of Ethyl Groups to Benzaldehyde. Organometallics 2004, 23, 127–134. (44) Pescitelli, G.; Di Bari, L.; Salvadori, P. Multiple Solution Species of Titanium(IV) 1,1‘-Bi2-naphtholate Elucidated by NMR and CD Spectroscopy. Organometallics 2004, 23, 4223–4229. (45) Corden, J. P.; Errington, W.; Moore, P.; Partridge, M. G.; Wallbridge, M. G. H. Synthesis of Di-, Tri- and Penta-nuclear Titanium(IV) Species from Reactions of Titanium(IV) Alkoxides with 2,2′-Biphenol (H2L1) and 1,1′-Binaphthol (H2L2); Crystal Structures of [Ti3(μ2OPri)2(OPri)8L1], [Ti3(OPri)6L13], [Ti5(μ3-O)2(μ2-OR)2(OR)6L14] (R = OPri, OBun) and [Ti2(OPri)4L22]. Dalton Trans. 2004, 1846–1851. (46) Tang, H. Z.; Boyle, P. D.; Novak, B. M. Chiroptical Switching Polyguanidine Synthesized by Helix-Sense-Selective Polymerization Using [(R)-3,3‘-Dibromo-2,2‘-binaphthoxy](di-tertbutoxy)titanium(IV) Catalyst. J. Am. Chem. Soc. 2005, 127, 2136−2142. (47) Weber, B.; Seebach, D. Ti-TADDOLate-Catalyzed, Highly Enantioselective Addition of Alkyl- and Aryl-Titanium Derivatives to Aldehydes. Tetrahedron 1994, 50, 7473–484. (48) Reetz, M. T.; Maus S. Kinetic Studies of the Addition of Methyltitanium Reagents to Carbonyl Compounds. Tetrahedron 1987, 43, 101–108. (49) For review on organotitanium reagents, see; (a) Reetz, M. T. Organotitanium Reagents in Organic Synthesis. A Simple Means to Adjust Reactivity and Selectivity of Carbanions. Top. Curr. Chem. 1982, 106, 1–54. (b) Weidmann, B.; Seebach, D. Organometallic Compounds of Titanium and Zirconium as Selective Nucleophilic Reagents in Organic Synthesis. Angew. Chem. Int. Ed. Engl. 1983, 22, 31–45. (c) Seebach, D.; Weidmann, B.; Widler, L. Titanium and Zirconium Derivatives in Organic Synthesis. Modern Synthetic Methods 1983, 3, 217–353. (d) Reetz, M. T. Organotitanium Reagents in Organic Synthesis 1986, Springer, Berlin. (50) Kitamura, M.: Suga, S.; Kawai, K.; Noyori, R. Catalytic Asymmetric Induction. Highly Enantioselective Addition of Dialkylzincs to Aldehydes. J. Am. Chem. Soc. 1986, 108, 6071–6072. (51) Yamamoto, Y.; Kurihara, K.; Miyaura, N. Me-bipam for Enantioselective Ruthenium(II)Catalyzed Arylation of Aldehydes with Arylboronic Acids. Angew. Chem. Int. Ed. 2009, 48, 4414–4416. (52) Yamamoto, Y.; Shirai, T.; Watanabe, M.; Kurihara, K.; Miyaura, N. Ru/Me-BIPAMCatalyzed Asymmetric Addition of Arylboronic Acids to Aliphatic Aldehydes and α-Ketoesters. Molecules 2011, 16, 5020–5034. (53) Berrisford, D. J.; Bolm, C.; Sharpless, K. B. Ligand-Accelerated Catalysis. Angew. Chem. Int. Ed., 1995, 14, 1059–1070. (54) Harada T. Development of Highly Active Chiral Titanium Catalysts for the Enantioselective Addition of Various Organometallic Reagents to Aldehydes. Chem. Rec. 2016 16, 1256–1273. (55) The rate ratio between catalytic and direct was estimated to be 103 – 104, depending on the

ACS Paragon Plus Environment

Page 16 of 19

Page 17 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

size of the substituent (reference 38). (56) Several transition structure model have been proposed to rationalize the experimentally observed enantioselectivity on the basis of chemical intuition. See, reference 32, 36, and 42. (57) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, 2011. (58) Zhao, Y.; Truhlar, D. G. A New Local Density Functional for Main-Group Thermochemistry, Transition Metal Bonding, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Phys. 2006, 125, 194101. (59) Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215−241. (60) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378−6396. (61) Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for the Transition Metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270−283. (62) Wadt, W. R.; Hay, P. J. Ab Initio Effective Core Potentials for Molecular Calculations. Potentials for Main Group Elements Na to Bi. J. Chem. Phys. 1985, 82, 284−298. (63) Hariharan, P. C.; Pople, J. A. Theor. Chim. Acta 1973, 28, 213−222. (64) Grimme, S. Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory. Chem. Eur. J. 2012, 18, 9955−9964. (65) See, also: Ribeiro, R. F.; Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Use of Solution-Phase Vibrational Frequencies in Continuum Models for the Free Energy of Solvation. J. Phys. Chem. B 2011, 115, 14556−14562. (66) The calculations were carried out with a program GoodVibes: Funes-Ardoiz, I.; Paton, R. S. GoodVibes v2.0.1. 2017, DOI: 10.5281/zenodo.884527. (67) Mammen, M.; Shakhnovich, E. I.; Deutch, J. M.; Whitesides, G. M. Estimating the Entropic Cost of Self-Assembly of Multiparticle Hydrogen-Bonded Aggregates Based on the Cyanuric Acid·Melamine Lattice. J. Org. Chem. 1998, 63, 3821−3830. (68) Baradel, N.; Mobian, P.; Khalil, G.; Henry, M. Titanium(IV)-Based Helicates Incorporating the Ortho-Phenylenediamine Ligand: a Structural and a Computational Investigation. Dalton Trans., 2017, 46, 7594–7602. (69) NBO Version 3.1, Glendening, E. D.; Reed, A. E.; Carpenter, J. E.; Weinhold, F. (70) Legault, C. Y. CYLview, 1.0b; Université de Sherbrooke: Sherbrooke, Canada, 2009; http://www.cylview.org. (71) For the reaction of Grignard reagents, see; Ashby, E. C.; Duke, R. B.; Newmann, H. M. Concerning the Addition of Grignard Reagents to Ketones. Evidence for a Termolecular Mechanism. J. Am. Chem. Soc. 1967, 89, 1964–1965. (72) For the reaction of BuLi, see; McGarrity, J. F.; Ogle, C. A.; Brich, Z.; Loosli, H. -R. A RapidInjection (RI) NMR Study of the Reactivity of Butyllithium Aggregates in Tetrahydrofuran. J.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Am. Chem. Soc. 1985, 107, 1810–1815. (73) For the reaction of organolithiums, see (a) Nakamura, M.; Nakamura, E.; Koga, N.; Morokuma, K. Theoretical Studies on the Reaction of Solvated Methyllithium Open Dimer with Aldehyde. J. Am. Chem. Soc. 1993, 115, 11016–11017. (b) Hæffner, F. E.; Sun, C.; Williard, P. G. Mechanistic Variations Due to the Solvation State in the Reaction of MeLi in Dimer and Trimer Aggregates with Formaldehyde. J. Am. Chem. Soc. 2000, 122, 12542–12546. (74) For the reaction of Grignard reagents, see; (a) Yamazaki, S.; Yamabe, S. A Computational Study on Addition of Grignard Reagents to Carbonyl Compounds. J. Org. Chem. 2002, 67, 9346– 9353. (b) Mori, T.; Kato, S. Grignard Reagents in Solution: Theoretical Study of the Equilibria and the Reaction with a Carbonyl Compound in Diethyl Ether Solvent. J. Phys. Chem. A 2009, 113, 6158–6165. (75) For the reaction of organozincates, see; Uchiyama, M.; Nakamura, S.; Ohwada, T.; Nakamura, M.; Nakamura, E. Mechanism and Ligand-Transfer Selectivity of 1,2-Addition of Organozincate Complexes to Aldehyde. J. Am. Chem. Soc. 2004, 126, 10897–10903. (76) Weingarten, H.; VanWazer, J. R. Exchange of Parts between Molecules at Equilibrium. VI. Scrambling on Titanium of the Alkoxyl, Dimethylamino, and Halogen Substituents. J. Am. Chem. Soc. 1965, 87, 724–730. (77) Bradley, D. C.; Mehrotra, R. C.; Gaur, D. P. Metal Alkoxides; Academic Press: New York, 1978. (78) Babonneau, F.; Doeuff, S.; Leaustic, A.; Sanchez, C.; Cartier, C.; Verdaguer, M. XANES and EXAFS study of titanium alkoxides. Inorg. Chem. 1988, 27, 3166–3172. (79) Davis, T. J.; Balsells, J.; Carroll, P. J.; Walsh, P. J. Org. Lett. 2001, 3, 699-702. (80) Cauro-Gamet, L. C.; Hubert-Pfalzgraf, L. G.; Lecocq Z. S. Syntheses and Molecular Structures of Titanium Derivatives with Polymerizable Ligands. Toward Extended Arrays. Anorg. Allg. Chem.2004, 630, 2071–2077. (81) Johnson, A. L.; Davidson, M. G.; Lunn, M. D.; Mahon, M. F. Synthesis, Isolation and Structural Investigation of Schiff-Base Alkoxytitanium Complexes: Steric Limitations of Ligand Coordination. Eur. J. Inorg. Chem. 2006, 3088–3098. (82) Grasset, F.; Cazaux, J.-B.; Magna, L.; Braunstein, P.; Oliver-Bourbigou H. New Bis(aryloxy)–Ti(IV) Complexes and their Use for the Selective Dimerization of Ethylene to 1Butene. Dalton Trans. 2012, 41, 10396–10404. (83) Sabat, M.; Gross, M. F.; Finn M. G. Structure and Bonding of Heterobimetallic Fischer Carbene Complexes. Organometallics, 1992, 11, 745–751. (84) Yang, H.-T.; Zhou, S.; Chang, F.-S.; Chen, C.-R.; Gau, H.-M. Synthesis, Structures, and Characterizations of [ArTi(O-i-Pr)3]2 and Efficient Room-Temperature Aryl-Aryl Coupling of Aryl Bromides with [ArTi(O-i-Pr)3]2 Catalyzed by the Economic Pd(OAc)2/PCy3 System. Organometallics, 2009, 28, 5715–5721. (85) A closely relevant tricyclic transition structure has been reported for the addition of MeMgCl by Yamazaki and Yamabe (reference 74a). In the more recent study by Mori and Kato (reference 74b), this transition structure is higher in energy by 6.7 kcal/mol than a ladder-like transition structure. (86) For the addition of (MeMgCl)2 and Me3ZnLi to formaldehyde, the forming C–C bond length of 2.470 Å (reference 74a) and 2.24 Å (reference 75) are reported respectively. (87) Bradley, D. C.; Holloway, C. E. Nuclear Magnetic Resonance and Cryoscopic Studies on Some Alkoxides of Titanium, Zirconium, and Hafnium. J. Chem. Soc. A 1968, 1316–1319. (88) C. E. Holloway, Carbon-13 Nuclear Magnetic Resonance Study of Some Alkoxides of Titanium. J. Chem. Soc., Dalton Trans., 1976, 1050–1054. (89) 0.91 (0.018M) and 0.94 (0.052M) in C6H6 are reported; Kuehlein, K.; Clauss, K. GrowthReaction of Ethylene to Methyltitanium Compounds. Makromol. Chem. 1972, 155, 145–68. 1.01 (0.003–0.0016M) in C6H6 is also reported (reference 87). (90) A reviewer suggested the inclusion of empirical dispersion correction (reference 91, 92, and

ACS Paragon Plus Environment

Page 18 of 19

Page 19 of 19 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Organic Chemistry

93) to the calculations, considering the size of the molecules. We performed the reoptimization of TS-5b, TS-6b, 5b, 6b, and HCHO with the dispersion corrected M06 functional (M06-D3). While the activation free energy for four-membered transition structure transition structure TS-5b (M06D3) changes only a little (–0.2 kcal/mol), that for tricyclic transition structure TS-6b is decreased by 1.3 kcal/mol (Table S-2). The result implies that dispersion interactions between relatively remote isopropyl groups operate in stabilizing the tricyclic transition structure. (91) Grimme, S.; Schreiner, P. R.; Steric Crowding Can Stabilize a Labile Molecule: Solving the Hexaphenylethane Riddle. Angew. Chem. Int. Ed. 2011, 50, 12639–12642. (92) Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C. Dispersion-Corrected MeanField Electronic Structure Methods. Chem. Rev. 2016, 116, 5105−5154. (93) Rösel, S.; Quanz, H.; Logemann, C.; Becker, J.; Mossou, E.; Cañadillas-Delgado, L.; Caldeweyher, E.; Grimme, S.; Schreiner, P. R. London Dispersion Enables the Shortest Intermolecular Hydrocarbon H···H Contact. J. Am. Chem. Soc. 2017, 139, 7428−7431. (94) The initial structure for the calculation were obtained from the X-ray crystallographic coordinates of Ph(BINOLate)Ti2(OiPr)5 (reference 36). (95) On the other hand, transition structures TS-7a(f), TS-7a(i),TS-7a(j), TS-7a(n), and TS-7a(o), in which Ti1 bearing the methyl group is chelated by (R)-BINOLate are of higher energy (ΔE > 3.1 kcal/mol), possibly owing to the electron withdrawing nature of the ligand.

ACS Paragon Plus Environment