3 A Predictive Model for the [Rh2(esp)2]-catalyzed Intermolecular C(sp )- H Bond Insertion of b-carbonyl Ester : Interplay Between Theory and Experiment Brett D. McLarney,a Steven R. Hanna,a Djamaladdin G. Musaev,b,* and Stefan Francea,c,* a School of Chemistry and Biochemistry, Georgia Institute of Technology, Atlanta, Georgia 30332 b Cherry L. Emerson Center for Scientific Computation and Department of Chemistry, Emory University, 1515 Dickey Drive, Atlanta, Georgia 30322 c Petit Institute of Bioengineering and Bioscience, Georgia Institute of Technology, Atlanta, Georgia 30332

ABSTRACT: Intermolecular C(sp3)-H insertions of b-carbonyl ester dirhodium carbenes are extremely rare. Toward devel- oping efficient reactions of these carbenes, a model for their insertion into C(sp3)-H bonds is described using DFT calcula- tions. In this study, the relevant electronic and steric components of b-carbonyl ester dirhodium carbenes that affect inter- molecular C(sp3)-H activation energies are explored, parameterized, and used to construct an intuitive model for predicting propensity for C-H insertion. The resulting insights from the theoretical investigation are actualized through with experi- ments to establish reactivity trends for these species and reaction discovery. Based on these integrated computational and experimental efforts, examples of intermolecular C(sp3)-H insertion featuring secondary a--b-amide esters are re- ported. The resulting carbenes feature an intramolecular 1,6- bond that affords increased stability and enhanced reactivity for C(sp3)-H insertion as compared to other b-carbonyl ester dirhodium carbenes. The reactivity of these carbenes is also highlighted through (1) an example of a cyclopropanation reaction, and (2) the use of a chiral dirhodium catalyst system.

Keywords: C-H Functionalization, Predictive Model, Dirho- H EWG EWG EWG EDG EWG EDG EDG H EDG dium Acceptor-Acceptor carbenes, Density Functional The- M M M M M Acceptor-only Acceptor-Acceptor Donor-Acceptor Donor-Donor Donor-only ory, Parametrization of Stereoelectronic Effects A A-A D-A D-D D 1. Introduction EWG = ester, ketone, nitro, nitrile, phosphonate, sulfone EDG = aryl, alkeny, alkynyl Diazo-derived metal-carbenes have been a cornucopia of synthetically useful reactivity. The reactivity of these spe- Figure 1. Classifications of metal- compounds cies prominently features C–H and X–H bond activation, These challenges include but are not limited to: (a) high cyclopropanation, and a host of [3+2] cycloaddition chem- energy barriers for A-A diazo decomposition, or nitrogen istry.1-9 Metal-carbene compounds have been subdivided extrusion (requiring more active catalysts); (b) weak between five categories: acceptor only (A), acceptor-accep- metal-carbene interactions (resulting in low enantioselec- tor (A-A), donor-acceptor (D-A), donor-donor (D-D), and 10,11 tivities due to catalyst dissociation); and (c) relatively donor only (D) (Figure 1). To date, studies on D-A carbe- small singlet-triplet energy splitting in the A-A carbenes noids feature a panoply of selective C(sp3)–H functionali- (providing undesired conformational changes and differ- zation chemistry in which the donor group is credited with 12,13 ential reactivities). These challenges often preclude the stabilizing the metal-carbene interaction.10 On the other A-A carbenes from regio- and stereoselective addition to hand, whilst their functional diversity renders them a . Moreover, they make the A-A carbenoids promising source of synthetic building blocks, A-A diazo 14 prone to side reactions (e.g., Wolff rearrangements, ylide compounds have found little application in the budding 15 16 3 formation and rearrangement, beta-lactam formation, field of intermolecular selective C(sp )–H functionalization 17 H-atom abstraction, etc.) which create complex mixtures chemistry due to several challenges. in reaction vessels. For this reason, the vast majority of C- H functionalization reactions that feature A-A diazo com- pounds are performed intramolecularly.5,18 Previous stud- ies on their intramolecular C(sp3)–H functionalization and cyclopropanation revealed that chemo- and regioselectiv- In the spirit of the findings by Nakamura and co-work- ity of A-A carbene reactivity is highly sensitive to the iden- ers, herein, we hypothesized that the difference in C-H ac- tity of the acceptor groups.5,16,19,20,21 However, understand- tivation energies for various dirhodium-carbenes of b-car- ing the roles of the steric bulkiness and electronic nature bonyl esters could be explained by two factors. First, the of functional groups in the previously studied intramolec- ability of the carbonyl groups to stabilize negative charge ular reactions failed to offer generalizable rules/concepts accumulation at the carbene carbon center that is antici- for intermolecular C(sp3)–H functionalization. pated to have a stabilizing effect on the C(sp3)–H insertion transition state, consequently, lowering the activation en- Therefore, this paper investigates the reactivity of dirho- ergy. Second, allylic strain between the carbonyl substitu- dium(tetracarboxylate)-carbene complexes (derived from ent and ester in the C–H insertion transition state is, in A-A diazo compounds) with alkane substrates in an aim to contrast, expected to have a destabilizing effect. Encour- develop generalized predictive rules for their effective uti- aged by previously reported predictive reactivity models,26 lization in selective intermolecular C(sp3)–H functionaliza- in this paper, using DFT, we the divide barrier of the dir- tion reactions. Previous investigations of intermolecular 3 3 hodium b-carbonyl ester carbene insertion into C(sp )–H C(sp )–H functionalization reactions have focused solely 6 bonds into two orthogonal steric and electronic compo- on carbenes derived from dimethyl diazomalonate and nents (Figure 2B). These components of the C–H insertion thus similarly lack comprehensiveness.22,23 As such, little is barrier are directly related to the specific nature of the car- known on the influence of the nature of the acceptor 3 bonyl substituent. Therefore, the reactivity of the dirho- groups on A-A carbene C(sp )–H insertion in an intermo- dium b-carbonyl ester carbene was quantified by changing lecular dirhodium(tetracarboxylate)-catalyzed manifold. the carbonyl substituent and examining effects on the ste- Given the challenges regarding utilization of A-A dirho- reoelectronic components. This acquired knowledge was dium-carbenes in intermolecular C(sp3)–H functionaliza- then used to build a map of chemical space, as a predictive tion, we turned to density functional theory (DFT) studies model, highlighting subcategories of the b-carbonyl ester to systematically probe acceptor group modifications in an class of A-A carbenes in terms of their propensity for aim to guide our complex experiments. To simplify the dis- C(sp3)–H insertion. Ultimately, this predictive model is val- cussion and provide a reasonable starting point (with di- idated through experimental evidence and led to discovery rectly comparable parameters), the b-carbonyl ester class of the dirhodium secondary N-aryl a-diazo-b-amide esters of A-A dirhodium carbenes will serve as the focus of this carbene that is highly reactive toward C(sp3)-H bonds. study. In a seminal study, Nakamura et al. modeled the dir- hodium(tetraformate)-catalyzed C(sp3)–H functionaliza- 2. Computational Methods tion between alkanes and methyl diazoacetate or diazome- Geometry optimizations and frequency calculations for thane (Figure 2A).24 This work demonstrated that, in a C- all reported structures were performed at the M06L level H insertion transition state mediated by dirhodium-car- of density functional theory in conjunction with 6-31G(d,p) benes, positive charge accumulates on the C-H bond donor basis sets for H, C, N, O, F and Cl atoms and LANL2DZ ba- while negative charge accumulates on the carbene unit. In sis set and corresponding Hay-Wadt effective core poten- this sense, C-H insertions are heuristically viewed as con- tial for Rh.27 Each reported minimum has confirmed to certed hydride transfer and carbocation trapping process. have zero imaginary frequencies and each transition state This view has been supported by experimental evidence (TS) structure to have only one imaginary frequency asso- showing that reactions are faster for benzylic and alpha- ciated with the reaction coordinate. Intrinsic reaction co- heteroatom C-H bond donors than for those containing ordinate (IRC) calculations were performed for selected only aliphatic C-H bonds.25 transition state structures to confirm their true identity.

(A) Nakamura et al. (2002) Bulk solvent effects are incorporated using the self-con- R R R C sistent reaction field polarizable continuum model (IEF- C R R + H 28 H H δ R R H PCM) with dichloromethane as the solvent. The calcu- (HCO2)4Rh2 H CO Me H lated Gibbs free energies are corrected to a solution stand- CO2Me 2 (HCO2)4Rh2 C δ- CO2Me ard state of 1M at room temperature (298.15K). To mini- (B) this work R mize redundant discussion, below we discuss only the cal- C R R H R CO2Me R CO Me culated enthalpies. Enthalpy values are generally more re- R C 2 L4Rh2 C C(O)R H C(O)R liable than Gibbs free energies (which depends on concen- tration related parameters and more) and thus considered to be advantageous for establishing meaningful trends. Predictive Model R R Electronic energies (with and without zero-point-correc- C R H δ+ tions) and Gibbs free energies are provided in the Support- OMe O ing Information. L4Rh2 -C sterics δ R O ‡ 3. An Instructive Model Study electronics ∆H = f(sterics,electronics) To elucidate the governing factors of A-A carbene trans- Figure 2. Dirhodium-carbene insertion into C(sp3)-H bond as fer from the metal-carbene complex to the C–H bond of described by Nakamura (A) and this work (B). the substrate, we studied the (a) [Rh2(esp)2]-carbene com- plex (2g) derived from the reaction of a-diazo methyl acetoacetate (1g) and Rh2(esp)2, and the (b) transition state incoming propane; meanwhile, the ketone carbonyl pre- (3g) associated with the carbene transfer from 2g to the C– ferred a syn-periplanar (“up”) conformation. This orienta- H bond of propane (Figure 3). DFT calculations demon- tion presumably allows for optimal dipole alignment as strated that in carbene transfer transition state 3g, positive well as to eschew steric interactions between the ketone charge accumulates on the carbon of the activated C-H substituent and the propane due to increasing 1,2-allylic bond while negative charge accumulates on the C* center strain in the transition state. of carbene unit: the calculated NBO charges of the C* cen- Intrigued by these analyses and results, we set forth an ter are -0.01 |e| and -0.45 |e| for 2g and 3g, respectively in-depth investigation into the impact, both electronic and (Figure 3C). Notably, we found that the lone-pair (LP) of steric, of the ketone substituent (R) on the charge distribu- the carbene carbon strongly interacts with the p* orbital of tion at the C*-center, as well as disruption of the preferred the adjacent carbonyl, C =O, in the insertion transition a anti-periplanar orientation of the b-carbonyl ester carbe- state, which impacts the charge distribution at the C* cen- noid. Below, we developed the tools to decouple electronic ter. In 3g, compared to 2g, the ester carbonyl assumed an and steric effects, and study their impact independently. anti-periplanar (“down”) conformation with respect to the A) Overall C-H Carbene Transfer nitrogen extrusion to C–H bond

H H O O O O O O Rh (esp) Me Me 2 2 Me OMe C* Me OMe Me Ca OMe H Me N2 Rh2(esp)2 Me 1g 2g 4g

Ca C*

B) Rh-Carbene (2g): Rh O

C* Ca

Side view Top view C) C-H Transition State (3g): propane propane Me H Me C H δ+ OMe C* Ca [Rh2] Rh O - O δ C* C* 3g Ca O Ca Me

(LP) C* -> π*(Ca=O) NBO: Orbital Interaction Energy: 42 kcal/mol Side view Top view D) Tabulated Energies, Charges and Bond Distances

NBO Charges, in |e| Distances (in Å) ΔH‡/ΔG‡(a) O C* C=O Rh–C*

2g: [Rh2(esp)2]-Carbene -0.55 -0.01 1.228 1.953 3g: TSC–H -3.4/11.9 -0.63 -0.45 1.234 2.301

(a) Calculated Relative to [Rh2(esp)2]-carbene + Substrate, and given in kcal/mol

Figure 3. (A) Overall C-H insertion reaction studied in this paper, (B) [Rh2(esp)2]-carbene complex derived from the reaction of a-diazo methyl acetoacetate and [Rh2(esp)2]-catalyst, (C) the C–H insertion transition state 3g, and (D) Calcu- lated NBO charges of carbonyl oxygen and carbene carbon, selective bond distances, and relative energy (in kcal/mol) of the transition state 3g.

3.1 Quantifying Electronic Effects C–H bond of propane were modeled and calculated (rela- To quantify electronic effects, we studied the tive to the reactants, [Rh2(esp)2]-carbene + propane, Figure

[Rh2(esp)2]-carbene complexes (2a-e) of a series b-ketoes- 4). At first, we used the well-established Hammett coeffi- ters with p-substituted phenyl ketone groups. By choosing cient, sp, to assess the effect of negative charge accumula- p-substituted phenyl ketones, we essentially minimize any tion at the carbene carbon on the energy of the transition steric effects and isolate the electronic ones. The transition state.29 states (3a-e) associated with metal-carbene transfer to the

O O O O O O variety of methyl-substituted carbonyls 5 (as a mimic of the OMe OMe OMe C* carbon center of the metal-carbene complex 2) and Rh (esp) Rh (esp) Rh (esp) MeO 2 2 Me 2 2 H 2 2 2a 2b 2c compared the values to the C-H insertion favorability of = 0 σp = -0.268 σp = -0.170 σp the corresponding full transition states 3 (Figure 6B). ΔH = 3.1 kcal/mol ΔH = 2.4 kcal/mol ΔH = 1.9 kcal/mol O O O O Lower PAs correspond to electron-poor carbonyls while OMe OMe higher PAs suggest electron-rich carbonyls. Along these Rh (esp) Rh (esp) F 2 2 Cl 2 2 lines, the acetophenones (5a-5e) and acetaldehyde (5f) 2d 2e were among the most electronically deficient carbonyl ac- σp = 0.062 σp = 0.227 ΔH = 1.3 kcal/mol ΔH = 0.5 kcal/mol ceptor groups (with PAs = 232.6–237.8 kcal/mol) while am- ides (5r-5x) were among the most electron-rich (with PAs Figure 4. Hammett coefficients and C-H insertion energies for = 238.3-250.2 kcal/mol). Most of the simple ketones b aryl -ketoester carbenes 2a-e (5g-5l) had PA ranges between 237.9 and 239.1 kcal/mol, We found a direct correlation (R2 = 0.98342) between whereas higher PAs were observed for esters (5o-5q). Hammett coefficient and the activation barriers at the car- A. O O O PA O bene transfer transition states 3a-e (Figure 5). For instance, Cα ‡ H C* R C C*H R OMe with an activation enthalpy of ∆H = 3.1 kcal/mol, the p- α 3 R C C*H2 α Rh (esp) 5 5’ 2 2 2 methoxyphenyl ketone 2a has the smallest sp. In contrast, ‡ 5f (R = H) 234.0 5n (R = OH) 241.6 the p-chlorophenyl ketone 2e with ∆H = 0.5 kcal/mol has R = 5g (R = Me) 239.1 5o (R = OMe) 242.9 5h (R = Et) 238.0 5p (R = OEt) 242.6 the largest sp. In general, as the phenyl ring becomes elec- X 5i (R = neoPent) 237.9 5q (R = OCH2F) 234.4 5j (R = i-Pr) 238.1 5r (R = NMe ) 248.5 5a (X = OMe) 237.8 2 tron deficient, the C-H activation barrier decreases and 5k (R = Cy) 238.1 5v (R = NH2) 248.6 5b (X = Me) 236.8 5l (R = t-Bu) 237.9 5c (X = H) 234.9 5m (R = Vinyl) 233.6 Hammett coefficient increases. These findings allow us to 5d (X = F) 234.6 conclude that the electronic nature of the carbonyl [i.e. 5e (X = Cl) 232.6 5s 5t 5u 5w MeO Me X p*(Ca=O)–LP(C*) interaction] plays an important role in N N N N N Me H determining the C–H activation barriers at the carbene H transfer transition state. 250.2 246.8 242.8 250.3 5xa (X = H) 242.1 5xd (X = 4-CF3) 238.3 Since the Hammett coefficients only apply to substituted phenyl rings, other parameters had to be considered to quantify the general ability of a carbonyl substituent to sta- bilize negative charge accumulation on the carbene. A rep- resentative example is shown by comparing the dirhodium carbenes of the 4-chlorophenyl-containing b-ketoester 2e and methyl acetoacetonate 2g, the initial model system. Given that phenyl ketones are generally considered to be more electron poor than methyl ketones, one would expect the 4-chlorophenyl system 2e to have a lower C-H activa- tion barrier. However, calculations showed that its activa- ‡ tion barrier (∆H = 0.5 kcal/mol) is higher than that of the methyl acetoacetate 2g (∆H‡ = -3.4 kcal/mol). Therefore, we continued to search for a universal parameter that Figure 6. (A) Calculated proton affinities (PAs, in kcal/mol) would better represent electronic effects across phenyl ke- of the substrates examined and (B) a schematic presentation tones as well as amides, esters, and alkyl ketones. of relative activation enthalpy (in kcal/mol) and proton affin- ity for each substrate and class. Employing Nakamura’s hydridic C–H insertion mecha- nistic framework, one should expect that more electron- poor carbenes (with lower PAs) will react faster than the electron-rich ones (with higher PAs). Representatives of each carbonyl type were selected to establish a hierarchy of electronic properties of the b-carbonylester carbenes and their reactivity with C-H bonds as a function of the nature of the carbonyl substituent (Figure 7). For the selected molecules, the following trend was observed: (5f) < phenyl ketone (5c) < methyl ketone (5g) < methyl ester (5o) < dimethyl amide (5r). This order matches well with well-established carbonyl electrophilicity rules and di- Figure 5. Relative activation enthalpy (in kcal/mol) vs Ham- rectly associates with the C–H activation energy. Thus, if mett coefficient. we were to assume that the sterics of the system had no impact on the C-H alkylation transition state, then alde- Next, proton affinity (PA) was explored as a potential electronic parameter (Figure 6).30 We calculated PAs for a hydes would be expected to have the lowest C(sp3)-H al- Figure 9. A-Values and relative activation enthalpies of the C– kylation barriers and amides the highest. However, it H insertion transition states 3 for some members of the alkyl should be noted that the PA values within each carbonyl series classes have broader ranges, can be impacted by other A correlation (R2 = 0.6893) exists between A-value and groups of the metal-carbene complex, and may overlap. the calculated C–H activation energies for the selected se- For example, replacing a hydrogen with a fluorine on the ries of R groups – the larger A-value (i.e. larger steric effect) methyl ester (i.e. 5q, the CH3-to-CH2F substitution) re- the higher energy of the transition state (Figure 9). How- duces PA of carbene carbon by 8 kcal/mol and makes it ever, there are some inconsistencies that warrant further more reactive toward C–H bond. Therefore, PA should be discussion. In the isopropyl (3j) and cyclohexyl (3k) cases, used with caution because it does not solely explain energy dramatically different activation barriers are found despite trends within a carbonyl class or across all calculated sub- both having the same A value of 2.15. The isopropyl group strates. can rotate to relieve steric interactions with the dirhodium O O O O O catalyst. In contrast, the size and rigidity of the cyclohexyl Me < < < < N H Me MeO chair conformation results in destabilizing steric interac- Me 5f 5c 5g 5o 5r tions with the dirhodium catalyst that are not reflected in (234.0) (234.9) (239.1) (242.9) (248.5) the A-value determination. Another example of the incon- sistencies is shown by the t-butyl group (3l with A = 4.5). Increasing PA; Decreasing electrophilicity; Decreasing favorability for C-H functionalization Despite having a higher A-value than the 4-OMe-phenyl Figure 7. Representative substrates and their calculated PA group 3a, 3l has a slightly lower DH value (∆∆H = 0.5 parameters (in kcal/mol). kcal/mol). In this case, the relative destabilization of the C- 3.2 Quantifying Steric Effects H activation TS for 3l caused by the 1,5 R•••O repulsion is compensated by stabilizing interactions between the t-bu- In the literature, steric effects of organic molecules are tyl group and dirhodium catalyst in the [Rh2(esp)2]-car- typically parameterized by A-values, which are derived bene complex. from the 1,3-diaxial interactions on cyclohexane rings.31 Hy- pothesizing that the 1,5 R•••O steric interaction in the Thus, while A-value represents a good parameter to es- [Rh2(esp)2]-carbenoid is analogous to the 1,3-diaxial strain timate steric effects in some cases, it has limitations and is on cyclohexane rings, we examined the relationship be- not available for many R groups. To circumvent this issue, tween available literature A-values of alkyl groups and the we introduced a new and more general geometric parame- calculated C–H activation energies of the carbene transfer ter, r, defined as the distance between the carbonyl sub- transition state 3 (Figure 8). stituent, R, and the ester carbonyl oxygen in the C–H in- sertion transition state (Figure 10A). As seen in Figure 10B, + 2 [Rh2] δ H this parameter correlates well (R = 0.98) with the A-values O OMe of the R groups discussed above (see Supporting Infor- δ- R H R O 3 A-value mation Figure S6). A. Defining the parameter r Figure 8. Schematic presentation of the used A-value param- eter.

To assess impact of steric effects while eschewing com- + [Rh2] δ plexity from electronic effects, a sterically diverse set of al- H O OMe O kyl ketone carbenes with a narrow PA range (237.9 – 239.1 δ- C* C R O 3 a kcal/mol) was selected (Figure 9). To directly compare the r steric effects in the alkyl ketones systems with that of an r aryl ketone system, we chose the 4-methoxyphenyl ketone 2.89 Å carbene 2a as the representative aryl substrate given its 3g similar electronic nature (PA = 237.8 kcal/mol) to the se- B. Relationship between A-value and r + lected alkyl ketone carbenes. Since literature A-values for [Rh2] δ H R2 = 0.98 the 4-substituted phenyl substituents are not available, in O OMe δ- R H our analyses, we used the A-value of 3.00, which corre- R O 3 A-value sponds to the phenyl (i.e., unsubstituted) group. r extendable accessible δ+ [Rh2] H R = Me Me Figure 10. Defining the parameter r (A) and its relationship to O OMe Me - Me A-value (B). δ 3g 3h Me 3i R O 3 A = 1.70 A = 1.75 A = 2.00 ΔH = -3.4 kcal/mol ΔH = -3.1 kcal/mol ΔH = -1.9 kcal/mol These findings encouraged us to use r as a general pa- rameter for assessing the impact of steric effect on the tran- Me Me Me Me Me sition state of the metal-carbene insertion into the C–H bond. Using the same substrates examined for PA calcula- 3j 3k OMe 3a 3l A = 2.15 A = 2.15 A = 3.00* A = 4.50 tions (see above), r values were determined and grouped ΔH = -1.4 kcal/mol ΔH = 1.4 kcal/mol ΔH = 3.1 kcal/mol ΔH = 2.6 kcal/mol by type of carbonyl substituent (Figure 11). A steric hierar- azetidine group (3t, r = 2.96 Å), which had strain equal to chy was then constructed to represent the severity of the the isopropyl ketone (3j). Interestingly, the secondary am- 1,5 steric interactions in the C-H insertion transition state. ides (3v-3x) have r values (2.75-2.78 Å) that are 0.2-0.3 Thus, more severe 1,5-strain would result in a larger r, and smaller than the tert-amides, but on par with the alkyl ke- less strain would result in a smaller r. As a representative tones. For instance, upon replacing one N-methyl group in example, for the aliphatic ketones (3g-3l), we found that dimethyl amide 3r with a hydrogen (as in 3w), a 0.3 Å de- the t-butyl ketone (3l) had the most 1,5-strain in the C-H crease in r value is seen. Close analyses suggest the pres- insertion transition state and the largest r (3.25 Å), while ence of a 1,6-hydrogen bonding interaction (discussed in the methyl ketone (3g), with the smallest r, had the least detail in Section 5). 1,5-strain. Comparatively, as expected, the aldehyde (3f) had even lower steric strain with r = 2.62 Å. 4. A Predictive Model 4.1. Mapping Chemical Space A. Calculated r values 3f (R = H) 2.62 3n (R = OH) 2.67 Having introduced parameters allowing deconstruction + R = 3g (R = Me) 2.89 3o (R = OMe) 2.85 [Rh2] δ H 3h (R = Et) 2.90 3p (R = OEt) 2.86 of the steric and electronic effects impacting the X 3i (R = neoPent) 2.93 3q (R = OCH F) 2.79 O OMe 2 - 3j (R = i-Pr) 2.96 3r (R = NMe ) 3.05 δ 3a (X = OMe) 3.08 2 [Rh2(esp)2]-catalyzed b-carbonyl ester carbene insertion 3k (R = Cy) 2.96 3v (R = NH2) 2.78 3 R O 3b (X = Me) 3.13 3l (R = t-Bu) 3.25 into the C(sp )-H bond, we next aimed to develop a general r 3 3c (X = H) 3.11 3m (R = Vinyl) 2.96 3d (X = F) 3.10 3e (X = Cl) 3.02 predictive model. This model is expected to predict car- X bonyl substituents that will facilitate the C–H insertion. 3s 3t 3u 3w MeO Me N N N N N H For this purpose, we represent relative enthalpy, DH, of the Me H 3xa (X = H) 2.75 C-H insertion transition state 3 as a function of PA and r 3.09 2.96 2.97 2.77 3xd (X = 4-CF3) 2.75 parameters, DH = f(PA, r), and interpolate this function us- ing a two-dimensional spline.32 (see Supporting materials, Figure S11). The 2D representation of the resulting chemical-space map with different b-carbonyl ester substrates (R) is given in Figure 12. As seen from this Figure, the aryl ketones (3a- e) are located on the upper left part of the map and have DH values > 0. The alkyl/vinyl ketones (3f-m) are spread across the left half of the map and have varied negative DH values which depend on r. 3n and esters (3o-q) are located on the lower half of the map with DH < 0. For the most part, the amides (3r-x) are clustered on the right half of the map and display varied DH values. Inter- Figure 11. Calculated r values (in Å) for substrates examined estingly, the amides split into separate regions based on (A) and to relation to relative activation energies (B). substitution, with the tertiary amides (3r-u) in the upper Overall, the calculations have established that the phe- right and the secondary ones (3v-x) mainly in the bottom nyl ketones (3a-3e) had more steric strain when compared right. to the aliphatic ketones as their r values ranged between From these observations, the existence of four major 3.02-3.13 Å. With the ability to rotate steric bulk away from classes of carbenes becomes apparent. Each class is the ester in the transition state, the aliphatic ketones (3g- grouped together into specific locations on the surface, 3l) are expected to lower strain than the phenyl groups, which allows the map to be divided into four sections or which prefer coplanarity with the carbonyl. The esters (3o- quadrants. 3q) followed with r values between 2.79 and 2.85 Å. Steric Quadrant 1 (Q1) contains the most electronically rich bulk on the esters is distally located and thus has little con- ‡ tribution to steric strain in the transition state. tert-Amides (PA>242) and sterically hindered carbenes, with ∆H C-H > (3r-3u) overlapped with this range with the dimethyl am- 0. The tertiary amides are the representative compound ide (3r) having a larger r value of 3.05 Å. Tethering the am- class in Q1. ide alkyl groups reduces the steric strain as evinced by the

Figure 12. Contour map showing calculated relationships between steric (r) and electronic (PA) parameters, and enthalpy (DH, in kcal/mol) of the C–H insertion transition states 3. Quadrant 2 (Q2) contains sterically hindered but elec- Interpretation: The developed map of chemical space ap- tron poor (PA<242) carbenes such as phenyl ketones (3a- proach (i.e. predictivity model) is envisioned to aid chem- e) and alkyl ketones (3i-l) where the alkyl group is partic- ists in building b-carbonyl ester dirhodium carbenoids ularly bulky (e.g., R = neopentyl, i-Pr, Cy, t-Bu). Carbenes with better C-H insertion reactivity. Steric factors seem to ‡ in Q2 have ∆H C-H > -1 kcal/mol. more directly correlate with relative enthalpies, while elec- tronic factors can be used to tune and modulate reactivity. Quadrant 3 (Q3) contains electron poor and sterically The steric component determines whether a species is unhindered carbenes such as the aldehyde 3f, simple alkyl Q1/Q2 or Q3/Q4 (top or bottom half of map). Conversely, ketones (3g/h, R= Me, Et), carboxylic acid 3n, esters bear- the electronic component determines whether a species is ing electron-withdrawing substituents (i.e., 3q, where R = Q2/Q3 or Q1/Q4 (left or right half of map). Further OCH F), and secondary amides with electron-withdrawing 2 strength of the map lies in the ability to properly capture N-aryl groups (3xc-3xg). Due to either reduced steric bulk stereoelectronic trends within substrate classes. For exam- or that bulk being rotated away from the carbonyl oxygen ple, the parent N-phenyl secondary amide species 3xa cor- in the transition state, the aldehyde, simple alkyl ketones, responds to a Q4 system. By incorporating an electron and esters have smaller r values. In contrast, the carboxylic withdrawing group on the N-aryl group (as in 3xd, where acid and the secondary amides have smaller r values due to aryl = 4-CF -Ph), the change in PA properly reflects the the presence of 1,6-hydrogen bonding. This, ultimately, 3 change to a more electron-poor system, resulting in a shift leads to relative activation enthalpies that are negative from Q4 to Q3. A similar shift is observed with esters 3o/p (∆H‡ <0). C-H vs 3q. Thus, by performing calculations of PA and r for a Quadrant 4 (Q4) houses electron-rich and sterically unhin- proposed carbene then identifying where it would fall on dered carbenes such as esters 3o/p, the primary amide 3v, the map (Q1-Q4), chemists can gauge the carbene's feasi- and secondary amides with electron-rich N-aryl groups bility for C-H insertion and use it to guide further carbene (e.g., 3xa and 3xb). Q4 also has negative activation en- design. thalpies. While each of these species has a smaller r value 4.2. Making the Model Accessible for one of the reasons discussed above, all have higher PAs due to lone pair donation into the carbonyl. For this map to have utility beyond rationalizing observ- able trends (i.e. to be predictive), calculating both PA and r value must be straightforward and not time-consuming.

While PA can be easily and accurately calculated, access to of C-H insertion product 8e. Through modulation of elec- r values requires time-intensive and complicated calcula- tronics, the Wolff rearrangement can be suppressed, allow- tions of the large dirhodium carbene complexes and re- ing for more observed C-H insertion. lated C-H insertion transition states. Thus, unless a chem- H ist has access to high-level computing capabilities, this O MeO2C CO2Me O 7 model would not be useful. To address this limitation while CO2Me CO2Me neat, reflux making the proposed model more accessible, we developed Rh (esp) N 2 2 R another steric parameter, r' (Figure 13). r' represents the R- R 2 (1 mol%) 1 MeOH quench 8 R 9 O distance in the anion of the corresponding b-carbonyl % isolated yields ester (6). r' can be easily calculated using commercial mo- R = OMe (a): PA = 237.8 0 : 96 lecular modeling softwares (e.g., Spartan, Macromodel, Me (b): PA = 236.8 21 : 71 H (c): PA = 234.9 34 : 60 etc.). Taking the t-butyl ketone as a representative exam- F (d): PA = 234.6 49 : 47 ple, a close similarity is observed between A-value, r, and Cl (e): PA = 232.6 62 : 32 r', with r and r' differing by ~0.04 Å (see Supporting mate- 3 rials). Furthermore, we have shown that r and r' parame- Scheme 1. C(sp )-H Alkylation vs Wolff Rearrangement for ters are closely correlated (R2 = 0.96) for the above pre- 4-Substituted Aryl a-Diazo-b-Ketoesters 1 sented carbenes (see Supporting materials). It is important Next, the other carbonyl classes were evaluated (Scheme to note that while r' is a close approximation of r, it does 2). In contrast to the clean reactions of the aryl ketones, not account for any added interactions and influences due the reactions with other carbonyl classes did not give ap- to the dirhodium catalyst and/or the substrate in transition preciable amounts of Wolff rearrangement products and state 3. Given this fact, throughout the remainder of our other degradation pathways are evident. In the case of the studies, we continue to use r as the appropriate steric pa- alkyl ketones, the observed data, again, are consistent with rameter. the prediction of the model stating the less steric effects

+ the more C–H insertion product. Indeed, while methyl ke- [Rh2] δ O OMe H 2 2 tone 1g, with r value of 2.89Å, reacted favorably to form the R = 0.96 O OMe R = 0.98 R O δ- R H r’ 6 3 C-H insertion product 8g in 75% yield, no C-H insertion R O A-value r accessible product was obtained with t-butyl ketone 1j, with r value extendable, easily computed extendable of 3.25 Å. Instead, cyclobutyl ketone 10j was isolated as the major product resulting from intramolecular C-H insertion Figure 13. Relationship between A-value, and r and r’ param- into one of the t-buyl methyl C-H bonds.33 eters. H

5. Experimental Validation of the Proposed Model O O 7 CO Me R 2 To test the predictive power of the proposed reactivity CO Me neat, reflux R 2 Rh2(esp)2 model, we first studied the reaction of aryl a-diazo-b-ke- N2 1 (1 mol%) 8 MeOH quench toesters with Rh2(esp)2 in neat, refluxing cyclohexane (Scheme 1). Each reaction was quenched with to O O O CO Me (Q3) Me CO2Me Me CO Me Me 2 (Q3) 2 trap any intermediates resulting from Wolff rear- (Q2) Me rangements.12 One should mention that these substrates gave high conversions of diazo starting materials and the 8g (75%) r = 2.89 only products formed were either from C-H insertion or 8h (34%) r = 2.90 8i (trace) r = 2.93 O O Wolff rearrangement. Me CO Me (Q2) CO2Me 2 O CO2Me MeO Me As mentioned in Section 3.1, the developed model sug- Me Me (Q4) Me 10j gests that the carbenes with more electron deficient phenyl rings (i.e. with a smaller PA value) should be more selective 8j (NDP) r = 3.25 8o (73%) r = 2.85 O O for C-H insertion vs. Wolff rearrangement. The provided CO Me CO2Me N 2 N experimental data is in perfect agreement with the predic- H (Q1) (Q4) tions of the model. Indeed, for electron rich 4-MeO-phenyl 8s (NDP) r = 3.09 8xa (89%) r = 2.75 substituted diazo 1a (PA-value of 237.8 kcal/mol), we find only Wolff product 9a in 96% yield. Similarly, the tolyl ke- Scheme 2. Rh(II)-catalyzed C(sp3)-H Alkylations. r values tone 1b (PA-value of 236.8 kcal/mol) gave only 21% yield of are given in Å. C-H insertion product 8b, but 71% yield of Wolff rear- rangement product 9b. Further reduction of PA values, i.e. While the reaction tolerated the ethyl group (as in 1h) to decrease of the electronic property of the phenyl ring, in- form 8h in 34% yield, only trace amounts of 8i (from iso- crease selectivity for C-H insertion vs Wolff rearrange- propyl ketone 1i) could be detected by NMR analysis of the ment: phenyl ketone 1c (PA-value of 234.9 kcal/mol) pro- crude reaction material. These results are fully consistent vided 34% of 8c; 4-fluorophenyl ketone 1d (PA-value of with our predictive quadrant model. Ketones 1i and 1j are 234.6 kcal/mol) gave 49% yield of 8d; and 4-chlorophenyl both in Quadrant 2 (Q2) are expected to have poor C-H ketone 1e (PA-value of 232.6 kcal/mol) afforded 63% yield insertion reactivity due to sterics, whereas ketones 1g and

1h are in Quadrant 3 (Q3) are expected to have more favor- 6. Reactivity of the Secondary Amides: Scope of Carbenes able C-H insertion activity. and Substrates. Dimethyl diazomalonate 1o is the most commonly studied In the literature, reactions of carbenes derived from sec- b-carbonyl ester carbene precursor and its reactivity for in- ondary a-diazo-b-amide esters are exceedingly rare. While termolecular C-H functionalization has been established in one report demonstrates that these carbenes can undergo the literature.22 Consistent with the literature and for our X-H insertions in the presence of a dimeric Rh(II) cata- 34 own benchmark data, 1o formed C-H insertion product 8o lyst, to the best of our knowledge, examples of their util- in 73% yield (Scheme 2). ity for cyclopropanations (the standard carbene reaction- trapping of C=C double bond) and their capacity to per- One tertiary and one secondary amide also were used to form C-H insertion reactions (i.e. alkylation) have not been validate the model. Unsurprisingly, the sterically-encum- reported.35 Toward this end, we sought to probe the gen- bered pyrrolidinyl amide 1s, which is a Q1 substrate, eral reactivity potential of secondary a-diazo-b-amide es- yielded no C-H insertion product. On the other hand, N- ters. phenyl amide 1xa (a Q4 substrate) was predicted to have a low barrier to C-H insertion and furnished C-H insertion At first, we explored simple cyclopropanation of styrene product in 89% yield. with a-diazo-b-amide ester 1xa. Gratifyingly, we found that the 1xa reacts with styrene 11 (2 equiv) in the presence Intrigued by the favorable secondary amide carbene in- of Rh2(esp)2 catalyst to cleanly furnish cyclopropane 12 in sertion into the cyclohexane C-H bond, we launched in de- 78% yield as a 3:1 mixture of diastereomers (Scheme 3). tail comparable structural and energetic analyses of the Ph 11 O O tertiary N,N-dimethyl amide 2r and secondary N-methyl (2.0 equiv) Ph CO Me Ph CO2Me amide 2w carbenes (Figure 14). We found that the C-H in- N 2 N H H Rh2(esp)2 (1 mol %) N2 Ph sertion transition state lies higher in energy for N,N-dime- CH2Cl2, reflux 1xa 12 thyl amide carbene (∆H = 6.5 kcal/mol) than secondary 78% yield; 3:1 dr amide carbene (∆H = -2.9 kcal/mol). The finding that the Scheme 3. Rh(II)-catalyzed cyclopropanation with a-di- secondary amide carbene is more reactive toward C–H azo-b-amide ester 1xa bond than tertiary amide carbene is consistent with the re- activity model showing larger r value for N,N-dimethyl am- To highlight the future potential of the reaction, we em- ide carbene (3.08 Å) than the secondary amide (2.77 Å). ployed a chiral dirhodium catalyst complex in hopes of ob- However, close examination of the structures of corre- serving enantioselectivity (Scheme 4). When the C-H in- sponding C-H insertion transitions state, 3r and 3w, re- sertion reaction described above was performed with di- spectively, we also found the 1,6-hydrogen bonding with azoamide 1xa and cyclohexane in the presence of Rh2(S- the 1.99 Å bond distance between the ester oxygen and the PTAD)4 (1 mol %), the C-H alkylation product 8xa was ob- amide hydrogen of the secondary amide. Based on these tained in 26% ee (2:1 er), albeit with a lower yield (25%). analyses we conclude that 1,6-hydrogen bonding serves as While modest, the enantiomeric excess represents a prom- the key factor for decreasing of r value, and, consequently, ising starting point for optimization and catalyst screening. enables the C-H insertion. Despite convoluting sterics and H H-bonding in this instance, r succinctly captures reactivity motif by predicting 3w to have smaller relative energy than O O 3r. A similar change (Dr = 0.1-0.2) in r is observed when 5.0 equiv. Ph CO2Me Ph CO Me N N 2 H comparing the corresponding esters to the carboxylic acid H N Rh2(S-PTAD)4 3n (r = 2.67 Å), which is also capable of the 1,6-hydrogen 2 (1 mol %) CH2Cl2, reflux bonding. 1xa 8xa 25% yield O 26% ee (2:1 er)

N H O O O S-PTAD

Scheme 4. Asymmetric C-H insertion using a chiral dirho-

1.99 Å H1-Oester dium catalyst r = 3.08 Å r = 2.77 Å In an aim to expand scope of the reactive secondary a- 3w: ∆H‡ = -2.9 kcal/mol 3r: ∆H‡ = 6.5 kcal/mol diazo-b-amide esters we, again turned to the newly devel- PA = 250.3 kcal/mol PA = 248.5 kcal/mol oped reactivity model and calculated PA and r values for a Figure 14. Representation of 1,6-Hydrogen bonding interac- series of N-aryl-substituted secondary a-diazo-b-amide es- tion in 3w and comparison with 3r. ters such as 1xa-1xg (Figure 15). Across the compounds, PAs varied between 238 and 244 kcal/mol while r values remained fairly consistent (Dr < 0.03 Å). Therefore, based

the reactivity model, one may expect the secondary amides and C(3)-H insertion products 14 and 15 were formed in with lower PA to be more reactive with C–H bond. 87% yield with 5.8:1 regioselectivity, respectively. Product

F 14 was obtained as a 1:1 mixture of diastereomers. X F F O O O O H CO2Me CO2Me CO2Me CO Me N N F N N 2 H H H 3 N X N F N 2 H 2 2 2 CF3 H 13 O 1xa: X = H Q4 Q3 1xg (5.0 equiv) 14 (1:1 dr) 1xe: X = Cl CO Me r = 2.74 Å r = 2.72 Å N 2 r = 2.75 Å H Rh2(esp)2 (1 mol %) PA = 242.1 kcal/mol PA = 238.0 kcal/mol CF3 N2 PA = 239.2 kcal/mol CH2Cl2, reflux O 1xb: X = OMe 1xf 1xf: X = CF 87% yield CO2Me r = 2.75 Å 3 N r = 2.75 Å 5.8:1 C(2):C(3) rr H PA = 244.0 kcal/mol CF3 PA = 238.1 kcal/mol 1xc: X = Cl 15 Q3 r = 2.74 Å PA = 241.1 kcal/mol Scheme 6. C-H insertion into pentane using 1xf. 1xd: X = CF3 r = 2.74 Å PA = 238.3 kcal/mol Conclusion We have introduced accessible parameters such as PA Figure 15. Steric (r) and Electronic (PA) Parameters for Vari- (proton affinity of the carbene carbon), r (the R-O distance ous Secondary b-Amide Esters in the C-H insertion transition state), and r' (the R-O dis- To validate the predictions of the reactivity model, we tance in the anion of the b-carbonyl ester) allowing decon- synthesized N-aryl-substituted secondary a-diazo-b-am- struction of steric and electronic effects in the [Rh2(esp)2]- ide esters 1xa-1xg and subjected them to the reaction con- catalyzed b-carbonyl ester carbene insertion into the 3 ditions with 5 equiv of cyclohexane (Scheme 4). Once C(sp )-H bond. We used these parameters and developed again, general reactivity pattern predicted by the model is a chemical-space map and general reactivity model that fully consistent with experiments. Indeed, the 4-methoxy- predicts carbonyl R-substituents which can enhance the phenyl substrate 1xb with a larger PA = 244 kcal/mol of- C–H insertion. fered a lower yield (24%). In contrast, the N-(4-Cl-phenyl) Following experimental studies have fully validated the and N-(4-CF3-phenyl) substrates (1xc and 1xd) with PAs of predictive power of this model and allowed us to establish: = 240 and 238 kcal/mol, respectively, gave their respective (1) For the aryl ketones, selectivity for C-H insertion vs. insertion products 8xc and 8xd in 55% and 63% yield. Wolff rearrangement can be modified by tuning the elec- Higher yields (69% and 75%) were observed with the N-(2- tronics of the aryl ring; (2) For the alkyl ketones, linear Cl-phenyl) and N-(2-CF3-phenyl) substrates (1xe and 1xf), alkyl groups tolerate intermolecular C-H insertion, respectively. Finally, the N-perfluoro-phenyl amide 1xg whereas branched or other sterically-hindered alkyl groups with the lowest PA (of 238.0 kcal/mol) afforded the C–H do not; (3) Simple esters are both sterically and electroni- insertion product 8xg in 82% yield. Considering the recent cally amenable to intermolecular C-H insertion; (4) The advancements in amide-based directing group C-H func- 36 steric bulk of the tertiary amides render them poor per- tionalization chemistry, this series of results are highly formers in intermolecular C-H insertion, and (5) Second- auspicious and offer tremendous opportunity for orthogo- ary amides, in contrast, have the highest selectivity for C- nal and/or tandem reactions. H insertion. We showed that the 1,6-hydrogen bonding be- H tween the ester oxygen and the amide hydrogen of the sec- ondary amide serves as the key enabling factor for the O 7 3 O Ar CO2Me C(sp )–H alkylation. Ar CO Me 5.0 equiv. N N 2 H H N Rh2(esp)2 (1 mol %) Our combined experimental and reactivity predictive 2 CH Cl , reflux 2 2 model studies led to discovery of the [Rh (esp) ]-catalyzed 1xa-1xg 8xa-8xg 2 2 3 F both C(sp )–H bond alkylation and C=C double bond cy- X F F O O O clopropanation reactions by the secondary N-aryl a-diazo- CO Me CO Me CO Me N 2 N 2 F N 2 b-amide esters. We established generality of this new H H H 3 X F C(sp )–H bond alkylation reaction by expanding it to (a) other secondary N-aryl a-diazo-b-amide esters, and (b) 8xa X = H; 42% 8xe X = Cl; 69% 8xg 82% substrates like cyclohexane and pentane. Finally, a prom- 8xb X = OMe; 24% 8xf X = CF3; 75% 8xc X = Cl; 55% ising example of asymmetric induction was also shown us- 8xd X = CF ; 63% 3 ing a chiral catalyst for the cyclopropanation.

Scheme 5. Rh(II)-catalyzed C-H insertions of diazoamides ASSOCIATED CONTENT 1x Supporting Information To expand substrate scope, we also studied the reaction Experimental procedures, optimization tables, computational of the reactive N-(2-CF -phenyl) amide 1xf with pentane 3 methods used, analytical techniques employed, and copies of (13, instead of cyclohexane) in the presence of Rh2(esp)2 NMR spectra (PDF). The Supporting Information is available catalyst under the similar conditions. Gratifyingly, C(2)-H free of charge on the ACS Publications website.

AUTHOR INFORMATION (14) Gronert, S.; Keeffe, J. R.; More O’Ferrall, R. A. Stabilities of Carbenes: Independent Measures for Singlets and Triplets. J. Am. Corresponding Author Chem. Soc. 2011, 133, 3381. Email: [email protected] (15) Padwa, A.; Hornbuckle, S. F. Ylide Formation from the Re- Email: [email protected] action of Carbenes and Carbenoids with Heteroatom Lone Pairs. Chem. Rev. 1991, 91, 263. Author Contributions (16) (a) Watanabe, N.; Masashio, A.; Hashimoto, S.-I.; Ikegami, The manuscript was written through contributions of all au- S. Enantioselective Intramolecular C-H Insertion Reactions of N- thors. All authors have given approval to the final version of Alkyl-N-tert-butyl-α-methoxycarbonyl-α-diazoacetamides Cata- lyzed by Dirhodium(II) Carboxylates: Catalytic, Asymmetric Con- the manuscript. struction of 2-Azetidinones. Synlett 1994, 12, 1031; (b) Anada, M.; Hashimoto, S.-I. Enantioselective Synthesis of 4-Substituted 2- Notes Pyrrolidinones by Site-selective C-H Insertion of α-Methoxycar- The authors declare no competing financial interests. bonyl-α-diazoacetanilides Catalyzed by Dirhodium(II) Tetrakis[N-phthaloyl-(S)-tert-leucinate]. Tetrahedron Lett. 1998, ACKNOWLEDGMENT 39, 79; (c) Anada, M.; Hashimoto, S.-I. Enantioselective Intramo- This work was supported by NSF under the CCI Center for Se- lecular C-H Insertion Route to a Key Intermediate for the Synthe- lective C-H Functionalization (CHE-1700982). B.D.M. thanks sis of Trinem Antibiotics. Tetrahedron Lett. 1998, 39, 9063; (d) NSF for a graduate research fellowship (DGE-1148903). D.G.M. Anada, M.; Watanabe, N. Highly Enantioselective Construction of and B.D.M. gratefully acknowledge NSF-MRI-R2 grant (CHE- the Key Azetidin-2-ones for the Synthesis of Carbapenem Antibi- otics via Intramolecular C-H Insertion Reactions of α-Methox- 0958205) and the use of the resources of the Cherry Emerson ycarbonyl-α-diazoacetamides Catalyzed by Chiral Dirhodium(II) Center for Scientific Computation. Carboxylates. Chem. Commun. 1998, 1517. (17) Cui, X.; Xu, X.; Jin, L.-M.; Wojtas, L.; Zhang, X. P. Stereose- REFERENCES lective Radical C-H Alkylation with Acceptor/Acceptor-Substi- (1) Davies, H. M. L. and B. T. Parr in Contemporary Carbene tuted Diazo Reagents via Co(II)-Based Metalloradical Catalysis. Chemistry, John Wiley & Sons, Inc., 2013, pp. 363-403. Chem. Sci. 2015, 6, 1219. (2) Guy Bertrand. Carbene Chemistry: From Fleeting Intermedi- (18) H. M. L. Davies and A. M. Walji, in Modern rhodium-cata- ates to Powerful Reagents. CRC Press. 2002. lyzed organic reactions, ed. P. A. Evans, Wiley-VCH, Weinheim, (3) Brookhart, M.; Studabaker, W. B. Cyclopropanes from Re- 2005, pp. 301-337. actions of Transition Metal Carbene Complexes with Olefins. (19) Nani, R. R.; Reisman, S. E. α-Diazo-β-ketonitriles: Chem. Rev. 1987, 87, 411. Uniquely Reactive Substrates for Arene and Alkene Cyclopropa- (4) Doyle, M. P.; Forbes, D. C. Recent Advances in Asymmetric nation. J. Am. Chem. Soc. 2013, 135, 7304. Catalytic Metal Carbene Transformations. Chem. Rev. 1998, 98, (20) Doyle, M. P.; Hu, W.; Wee, A. G. H.; Wang, Z.; Duncan, S. 911. C. Influences of Catalyst Configuration and Catalyst Loading on (5) Doyle, M. P.; Duffy, R.; Ratnikov, M.; Zhou, L. Catalytic Car- Selectivities in Reactions of Diazoacetamides. Barrier to Equilib- bene Insertion into C-H Bonds. Chem. Rev. 2010, 110, 704. rium Between Diastereomeric Conformations. Org. Lett. 2003, 5, (6) Audia, J. E.; Droste J. J.; Gevorgyan, V.; Rubin, M.; Chan, W.; 407. Yu, W.; McLarney, B.; France, S. (2016) “Dimethyl Diazomalo- (21) Zhang, B.; Wee, A. G. H. Conformational, Steric and Elec- nate.” Encyclopedia of Reagents for Organic Synthesis. John Wiley tronic Effects on the Site- and Chemoselectivity of the Metal-cat- & Sons, Ltd. alyzed Reaction of N-Bis(trimethylsilyl)methyl, N-(2-Indolyl)me- (7) Lindsay, V. N. G.; Nicolas, C.; Charette, A. B. Asymmetric thyl α-Diazoamides. Org. Biomol. Chem. 2012, 10, 4597. Rh(II)-Catalyzed Cyclopropanation of Alkenes with Diacceptor (22) Muller, P.; Tohill, S. Intermolecular Cyclopropanation Ver- Diazo Compounds: p-Methoxyphenyl Ketone as a General Stere- sus CH Insertion in RhII-catalyzed Carbenoid Reactions. Tetrahe- oselectivity Controlling Group. J. Am. Chem. Soc. 2011, 133, 8972. dron 2000, 56, 1725. (8) McLarney, B. D.; Cavitt, M. A.; Donnell, T. M.; Musaev, D. (23) Davies, H. M. L.; Hansen, T. Asymmetric Intermolecular G.; France, S. Rh(II)-Catalyzed β-C(sp2)-H Alkylation of Enol Carbenoid C-H Insertions Catalyzed by Rhodium(II) (S)-N-(p- Ethers, Enamides and Enecarbamates with α-Diazo Dicarbonyl Dodecylphenyl)sulfonylprolinate. J. Am. Chem. Soc. 1997, 119, Compounds. Chem. Eur. J. 2017, 23, 1129. 9075. (9) England, D. B.; Eagan, J. M.; Merey, G.; Anac, O.; Padwa, A. (24) Nakamura, E.; Yoshikai, N.; Yamanaka, M. Mechanism of The Rhodium(II) Carbenoid Cyclization-Cycloaddition Cascade C-H Bond Activation/C-C Bond Formation Reaction between Di- of Alpha-Diazo Dihydroindolinones for the Synthesis of Novel azo Compound and Alkane Catalyzed by Dirhodium Tetracarbox- Azapolycyclic Ring Systems. Tetrahedron 2008, 64, 988. ylate. J. Am. Chem. Soc. 2002, 124, 7181 (10) Davies, H. M. L.; Morton, D. Guiding Principles for Site Se- (25) (a) Taber, D. F.; Ruckle, R. E., Jr. Cyclopentane Construc- lective and Stereoselective Intermolecular C-H Functionalization tion by Dirhodium Tetraacetate-mediated Intramolecular C-H In- by Donor/acceptor Rhodium Carbenes. Chem. Soc. Rev. 2011, 40, sertion: Steric and Electronic Effects. J. Am. Chem. Soc. 1986, 108, 1857. 7686; (b) Doyle, M. P.; Westrum, L. J.; Wolthuis, W. N. E.; See, M. (11) Davies, H. M. L.; Beckwith, R. E. J. Catalytic Enantioselec- M.; Boone, W. P.; Bagheri, V.; Pearson, M. M. Electronic and Steric tive C-H Activation by Means of Metal-Carbenoid-Induced C-H Control in Carbon-hydrogen Insertion Reactions of Diazoaceto- Insertion. Chem. Rev. 2003, 103, 2861. acetates Catalyzed by Dirhodium(II) Carboxylates and Carbox- (12) Kirmse, W. 100 Years of the Wolff Rearrangement. Eur. J. amides. J. Am. Chem. Soc. 1993, 115, 958; (c) Davies, H. M. L.; Ren, Org. Chem. 2002, 2002, 2193. P.; Jin, Q. Catalytic Asymmetric Allylic C-H Activation as a Surro- (13) Scott, A. P.; Platz, M. S.; Radom, L. Singlet−Triplet Split- gate of the Asymmetric . Org. Lett. 2001, 3, tings and Barriers to Wolff Rearrangement for Carbonyl Carbenes. 3587. J. Am. Chem. Soc. 2001, 123, 6069. (26) For relevant literature on predictive reactivity modeling studies, see: (a) Zhao, S.; Gensch, T.; Murray, B.; Niemeyer, Z. L.; Sigman, M. S.; Biscoe, M. R. Enantiodivergent Pd-catalyzed C–C

Bond Formation Enabled Through Ligand Parameterization. Sci- New Method Addressed to Study Molecular Solutes at the QM Ab ence 2018, 362, 670-674; (b) Metsänen, T. T.; Lexa, K. W.; Santiago, Initio Level. THEOCHEM 1999, 464, 211; (c) Pascal-Ahuir, J.L.; C. B.; Chung, C. K.; Xu, Y.; Liu, Z.; Humphrey, G. R.; Ruck, R. T.; Silla, E.; Tuñoń, I. GEPOL: An Improved Description of Molecular Sherer, E. C.; Sigman, M. S. Combining Traditional 2D and Mod- Surfaces. III. A New Algorithm for the Computation of a Solvent- ern Physical Organic-derived Descriptors to Predict Enhanced excluding Surface. J. Comput. Chem. 1994, 15, 1127. Enantioselectivity for the Key Aza-Michael Conjugate Addition in (29) Hammett, L. P. Effect of Structure upon the Reactions of the Synthesis of Prevymis™ (Letermovir). Chem. Sci. 2018, 9, 6922- Organic Compounds. Benzene Derivatives. J. Am. Chem. Soc. 1937, 6928; (c) Santiago, C. B.; Guo, J.-Y.; Sigman, M. S. Predictive and 59, 96. Mechanistic Multivariate Linear Regression Models for Reaction (30) PA represents the enthalpy change in the reaction between Development. Chem. Sci. 2018, 35, 2398-2412; (d) Gormisky, P. E.; an anion or neutral species and a proton. For a detailed review, White, M. C. Catalyst-Controlled Aliphatic C–H Oxidations with see: Maksic, Z. B.; Kovacevic, B.; Vianello, R. Advances in Deter- a Predictive Model for Site-Selectivity. J. Am. Chem. Soc. 2013, 135, mining the Absolute Proton Affinities of Neutral Organic Mole- 14052-14055; (e) Goetz, A. E.; Sarah M. Bronner, S. M.; Cisneros, J. cules in the Gas Phase and Their Interpretation: A Theoretical Ac- D.; Joshua M. Melamed, J. M.; Paton, R. S.; Houk, K. N.; Garg, N. count. Chem. Rev. 2012, 112, 5240-5270. K. An Efficient Computational Model to Predict the Synthetic (31) Anslyn, E. V. and Dougherty, D. A. in Modern Physical Or- Utility of Heterocyclic Arynes. Angew. Chem. Int. Ed. 2012, 51, ganic Chemistry. Sausalito, CA: University Science Books, 2006, 2758-2762; (f) Maldonado, A. G.; Rothenberg, G. Predictive Mod- pp. 104-105. eling in Homogeneous Catalysis: A Tutorial. Chem. Soc. Rev. 2010, (32) MATLAB and Statistics Toolbox Release 2017b, The Math- 39, 1891-1902. Works, Inc., Natick, Massachusetts, United States. (27) (a) Gaussian 09, Revision D.01, M. J. Frisch, G. W. Trucks, (33) (a) Cheng, Q.-Q.; Deng, Y.; Lankelma, M.; Doyle, M. P. Cy- H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. cloaddition Reactions of Enoldiazo Compounds. Chem. Soc. Rev. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, 2017, 46, 5425; (b) Pitts, C. R.; Lectka, T. Chemical Synthesis of β- M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Lactams: Asymmetric Catalysis and Other Recent Advances. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, Chem. Rev. 2014, 114, 7930. J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, (34) Trstenjak, U.; Ilas, J.; Kikelj, D. Studies Towards the Syn- T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. thesis of Alkyl N-(4-Nitrophenyl)-3/2-oxomorpholine-2/3-car- Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. boxylates. Helv. Chim. Acta 2013, 96, 2160. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, (35) While Zhang reported the Co-catalyzed metalloradical cy- S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, clopropenation of alkynes using a-cyanodiazoacetamides, the J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gom- method involves a radical mechanism and doesn't involve car- perts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. benes. For more detail, see: Cui, X.; Xu, X.; Lu, H.; Zhu, S.; Wojtas, Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. L.; Zhang, X. P. Enantioselective Cyclopropenation of Alkynes Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, with Acceptor/Acceptor-Substituted Diazo Reagents via Co(II)- A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, Based Metalloradical Catalysis. J. Am. Chem. Soc. 2011, 133, 3304. and D. J. Fox, Gaussian, Inc., Wallingford CT, 2013; (b) Chai, J.-D.; (36) (a) Wu, Q.-F.; Wang, X.-B.; Shen, P.-X.; Yu, J.-Q. Enanti- Head-Gordon, M. Long-range Corrected Hybrid Density Func- oselective C-H Arylation and Vinylation of Cyclobutyl Carboxylic tionals with Damped Atom-Atom Dispersion Corrections. Phys. Amides. ACS Catal. 2018, 8, 2577; (b) Zhu, R.-Y.; Saint-Denis, T. Chem. Chem. Phys. 2008, 10, 6615; (c) Wadt, W. R.; Hay, P. J. Ab G.; Shao, Y.; He, J.; Sieber, J. D.; Senanayake, C. H.; Yu, J.- Q. Lig- Initio Effective Core Potentials for Molecular Calculations. Poten- and-Enabled Pd(II)-Catalyzed Bromination and Iodination of tials for Main Group Elements Sodium to Bismuth. J. Chem. Phys. C(sp3)-H Bonds. J. Am. Chem. Soc. 2017, 139, 5724; (c) Daugulis, 1985, 82, 284; (d) Hay, P. J.; Wadt, W. R. Ab Initio Effective Core O.; Roane, J.; Tran, L. D. Bidentate, Monoanionic Auxiliary-Di- Potentials for Molecular Calculations. Potentials for the Transi- rected Functionalization of Carbon-Hydrogen Bonds. Acc. Chem. tion Metal Atoms Scandium to Mercury. J. Chem. Phys. 1985, 82, Res. 2015, 48, 1053; (d) Rouquet, G.; Chatani, N. Catalytic Func- 270. tionalization of C(sp2)-H and C(sp3)-H Bonds by Using Bidentate (28) (a) Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechan- Directing Groups. Angew. Chem. Int. Ed. 2013, 52, 11726. ical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999 and references cited therein; (b) Tomasi, J.; Mennucci, B. Cances, E. The IEF Version of the PCM Solvation Method: An Overview of a