Abstract
A practical approach is introduced for the rapid determination of 13C kinetic isotope effects that utilizes a “designed” reactant with two identical reaction sites. The mechanism of the Buchwald–Hartwig amination of tert-butylbromobenzene with primary and secondary amines is investigated under synthetically relevant catalytic conditions using traditional intermolecular 13C NMR methodology at natural abundance. Switching to 1,4-dibromobenzene, a symmetric bromoarene as the designed reactant, the same experimental 13C KIEs are determined using an intramolecular KIE approach. This rapid methodology for KIE determination requires substantially less material and time compared to traditional approaches. Details of the Buchwald–Hartwig amination mechanism are investigated under varying synthetic conditions, namely a variety of halides and bases. The enantioselectivity-determining step of the l-proline catalyzed aldol reaction is also evaluated using this approach. We expect this mechanistic methodology to gain traction among synthetic chemists as a practical technique to rapidly obtain high-resolution information regarding the transition structure of synthetically relevant reactions under catalytic conditions.
Keywords: intramolecular kinetic isotope effects, Buchwald–Hartwig amination, transition-state analysis, aldol reaction, reaction mechanisms
Graphical Abstract

INTRODUCTION
Many organic reactions involve bonding changes at a carbon atom. Determination of 13C kinetic isotope effects (KIEs) is a powerful tool that provides intimate details about the nature of bonding changes that occur at a carbon atom during the transition state (TS) of the rate-determining step (RDS) of a reaction. A landmark report by Singleton in 1995 introduced the use of quantitative 13C NMR methodology for the high-precision determination of 13C KIEs at natural abundance.1 An important addition to this methodology was the use of density functional theory (DFT) calculations for the quantitative interpretation of experimental 13C KIEs, using conventional transition-state theory with a one-dimensional tunneling correction to predict KIEs.2 Consequently, a correlation between the transition-state geometry and experimental KIEs has been well established.3 This combined experimental and theoretical approach has led to the resolution of several highly debated reaction mechanisms in organic chemistry.4
There are two complementary approaches for the determination of intermolecular 13C KIEs at natural abundance—starting material or product analysis. Taking the Buchwald–Hartwig amination of aryl halides as an example,5 Scheme 1 illustrates the design of the experiment for these two approaches. In both approaches, the “KIE sample” is the minor component in the reaction mixture, which can make its isolation challenging. As a result, to obtain a sufficient sample for 13C KIE determination (1–2 mmol), these experiments are typically performed on a 10–20 mmol scale (Scheme 1 approaches #1 and #2). Consequently, the use of this powerful mechanistic probe has been limited to easily scalable reactions requiring relatively inexpensive reagents. Additionally, altering reaction conditions (solvent, ligand, catalyst, etc.) can change the RDS and/or mechanism, and measurement of 13C KIEs on large-scale reactions for each set of conditions can be prohibitive.
Scheme 1.

Intermolecular Approaches for 13C KIE Determination at Natural Abundance (Singleton Method)
A third approach for the determination of 13C KIEs at natural abundance finds application in reactions that utilize a symmetric molecule as a reactant. When a symmetric molecule, with two identical reaction sites (distinguishable only by the isotope at each site), is desymmetrized during the course of a reaction, then the ratio of the two isotopomeric products formed reflects the intramolecular KIE for the first step that irreversibly desymmetrizes the molecule in the reaction. At natural abundance, this ratio of isotopomers is obtained by the relative integration of peaks of the two relevant carbon atoms in the 13C spectrum of the product (Ca and Ca‡, Scheme 2).
Scheme 2.

Utilizing Symmetric Molecules for Determination of 13C KIE at Natural Abundance
Singleton introduced the use of intramolecular 13C KIEs at natural abundance with an elegant study of the mechanism of the Baeyer–Villiger oxidation of cyclohexanone to ε-caprolactone (Scheme 2).6 In this example, the symmetry-breaking step probed by intramolecular KIEs occurred after the rate-limiting formation of the initial tetrahedral intermediate. In the other examples shown in Scheme 2, Singleton exploited intramolecular 13C KIE measurements to evaluate phenomenon such as tunneling and energy labeling in organic reactions.7,8 In reactions where the rate-limiting step coincides with the symmetry-breaking step, intramolecular 13C KIEs provide the exact same information as intermolecular KIE experiments.9 In these cases, performing an intramolecular KIE experiment has several advantages over an intermolecular KIE experiment that include: (1) analysis of a single sample for KIE determination, (2) insensitivity to fractional conversion since KIE is obtained as an intramolecular measurement, (3) KIE can be determined even when reactants are used in excess, and (4) requires significantly smaller-scale reactions since product from a high-conversion reaction can be used for analysis.
One key disadvantage is that this approach is limited to reactions utilizing symmetric reactants. We sought to exploit this by developing a simple but powerful strategy that extends the benefits of intramolecular KIE determination to reactions that do not typically use a symmetric substrate, using a purposely designed symmetric reactant with two identical reaction sites (Scheme 2). The only caveat is that only one position may undergo reaction (i.e., no double addition product), so the “standard” carbon used to determine the KIE is not compromised. In this study, we benchmark this strategy against the traditional intermolecular product KIE methodology to study the mechanism of the Buchwald–Hartwig amination. We then apply this strategy to rapidly obtain valuable mechanistic information for this reaction under different reaction conditions.
RESULTS AND DISCUSSION
We chose the Buchwald–Hartwig amination to benchmark this intramolecular 13C KIE method since the mechanism of this reaction is well known, providing a reliable mechanistic platform for the interpretation of experimental KIEs. Additionally, this reaction can be performed using a variety of different aryl halide reactants, solvents, ligands, and bases. This allows us to utilize the novel intramolecular KIE method as a rapid probe of potential changes in the reaction mechanism/RDS with varying reaction parameters.
A series of mechanistic studies in the early 2000s by Buchwald, Hartwig, and Blackmond established the currently accepted catalytic cycle for the Buchwald–Hartwig amination (Figure 1).10 According to this catalytic cycle, oxidative addition (OA) of 1 occurs to the κ2-Pd0-BINAP complex (Pd0L) in the RDS. A PdII OA complex (4) presumably binds the amine (2) followed by deprotonation of the amine proton by the alkoxide to form a new PdII complex (5). Reductive elimination (RE) from 5 yields amination product 3 and regenerates the Pd0L catalyst. Based on this catalytic cycle, the KIE at the carbon atom attached to the halide (KIEC–X) reports on the TS of the OA step (RDS).
Figure 1.

Catalytic cycle for the Buchwald–Hartwig amination reaction.
Traditional 13C Intermolecular Product KIE Study.
We chose the prototypical Buchwald–Hartwig amination of 4-tert-butylbromobenzene (1a) with two different amines, n-hexylamine (2a) and N-methylpiperazine (2b) for the determination of intermolecular 13C KIEs by analysis of the product. Following the protocol shown in Scheme 1, Traditional Approach # 2, duplicate reactions were conducted on a 15 mmol scale for both amines (2a and 2b) and taken to low conversion. Products 3aa/3ab were isolated (as the minor component) from each reaction mixture, and intermolecular 13C KIEs were determined in a standard way, by comparative 13C NMR analysis with samples of 3aa/3ab isolated from a 100% conversion reaction (no isotopic fractionation).1 Resulting KIEC–Br for 1a, from two independent measurements for each amine, is ~2.5% for both amines (Figure 2). Each of these four KIE measurements required 18 mmol of 1a (15 mmol for the KIE sample (20% conversion) +3 mmol for the Standard sample (100% conversion)) and about 9 h of NMR spectrometer time.
Figure 2.

Experimental 13C KIEs using the traditional intermolecular product KIE approach. Also shown is the transition structure for the widely accepted rate-determining oxidative addition of 1a, along with predicted KIEC–Br.
For the quantitative interpretation of experimental KIEs, the TS of the OA step of the Buchwald–Hartwig amination was investigated using four different DFT methods—(a) a B3LYP-GD3(BJ)11,12 functional and a Def2-TZVP basis set and ECP for Pd and I, and a Def2-SVP basis set for all other atoms, (b) an M06–2X13 functional with a LANL2DZ basis set and ECP for Pd and I, and a 6–311G** basis set for all other atoms, (c) a B3LYP functional with a SDDAll basis set and ECP for Pd and I, and a 6–311G** basis set for all other atoms, and (d) an M06-L14 functional with a Def2-TZVP basis set and an SDD ECP for Pd and I, and a Def2-SVP basis set for all other atoms—with a polarizable continuum model (PCM)15 solvent model for toluene on all methods as implemented in Gaussian 09.16 A comparative analysis shows that KIE predictions and geometries are similar across all methods (see the Supporting Information). The geometries and predicted KIEs in the manuscript are from the DFT method (a). From the scaled vibrational frequencies of the transition structures, predicted 13C KIEs were obtained using the program ISOEFF98.17,18 An infinite parabola tunneling correction was applied to all predicted KIEs.19
Predicted 13C KIE for the TS for OA of 1a to the κ2-Pd0-BINAP complex (TS-OA-1a, Figure 2) is 1.025, which is in excellent agreement with experimental measurements. Therefore, our experimental and theoretical 13C KIE study supports the currently accepted mechanism of the Buchwald–Hartwig amination reaction, where OA of aryl halide to a κ2-Pd0-BINAP complex is the first irreversible step (and likely RDS) in the catalytic cycle.
Proof-of-Principle 13C KIE Study Using Symmetric Bromoarene.
Having determined the KIEC–Br for 1a via the traditional product KIE approach, we next sought to determine KIEC–Br for the “designed” reactant 1,4-dibromobenzene (1asym). Equimolar reactions of 1asym and amine (2a or 2b) were each carried out on a 3 mmol scale under standard Buchwald–Hartwig amination conditions and quenched at 50–70% conversion of 1asym (to minimize the likelihood of double amination). Reaction mixtures were analyzed by 1H NMR and gas chromatography mass spectrometry (GC–MS) to verify that only mono-aminated products 3aadesym/3abdesym were formed. These mono-aminated products were isolated (~1–1.2 mmol) from the respective reaction mixtures via column chromatography and KIEC–Br determined from the relative integrations of Ca versus Ca‡ in the quantitative 13C NMR spectrum using well-established protocols for obtaining accurate relative integrations for two peaks within a single spectrum.20 The resulting 13C KIEs for 1asym, from two independent measurements for each amine, is shown in Figure 3. Each of these four KIE measurements required 3 mmol of 1asym and about 4 h of NMR spectrometer time. This corresponds to a >80% decrease in material required and a >50% decrease in spectrometer time to obtain the same mechanistic information as the traditional intermolecular methodology (vide supra).
Figure 3.

Experimental intramolecular 13C KIEs obtained using the symmetric bromoarene. Also shown is the transition structure for the widely accepted model for oxidative addition of 1asym, along with the predicted KIEC–Br.
For the quantitative interpretation of the KIEs for the reaction of 1asym, we located the TS for OA of 1asym to the κ2-Pd0-BINAP complex (TS-OA-1asym, Figure 3). The predicted 13C KIE for this TS is in excellent agreement with the experimental KIE. These results demonstrate that the intramolecular KIE approach using the symmetric bromoarene arrives at the same mechanistic conclusion as the traditional intermolecular product KIE approach, providing an important benchmark validating the use of this simple and efficient approach for the rapid determination of experimental 13C KIEs.
Effect of Varying Halide on the Mechanism of Buchwald–Hartwig Amination.
We decided to investigate the title reaction of aryl chlorides and aryl iodides using designed reactants 1,4-dichlorobenzene (1bsym) and 1,4-diiodobenzene (1csym) (shown in Figure 4). For direct comparison to the KIEs obtained with 1asym (Figure 3), reactions of 1bsym/1csym with 2a were run on a 3 mmol scale and taken to 50–70% conversion with respect to the aryl halide. The resulting mono-amination products 3badesym/3cadesym were isolated (1–1.2 mmol) from the reaction mixture, and KIEC–Cl and KIEC–I were determined as described previously. Following this, TSs for OA to the κ2-Pd0-BINAP complex for 1bsym (TS-OA-1bsym) and 1csym (TS-OA-1csym) were located to quantitatively evaluate the experimental KIEC–Cl and KIEC–I.
Figure 4.

Utilization of symmetric aryl halides to probe the effects of aryl halide on the mechanism of the Buchwald–Hartwig amination.
Upon changing from aryl bromide (KIEC–Br ~ 1.023) to aryl chloride (KIEC–Cl ~ 1.032), there is an increase in the magnitude of the experimental isotope effect. Gratifyingly, this increase is accurately predicted by DFT calculations; the predicted KIEC–X for the TS-OA-1asym and TS-OA-1bsym are 1.023 and 1.033, respectively (Figure 4). The magnitude of the 13C KIE prediction parallels the trend in C–X bond-length (rC–Br = 2.16 Å versus rC–Cl = 2.05 Å) at the respective TSs. These results indicate that the Buchwald–Hartwig amination of aryl chloride also proceeds via a mechanism involving OA of aryl chloride to a κ2-Pd0-BINAP complex in the first irreversible step (and likely RDS) in the catalytic cycle.
The experimental KIEC–I, determined from two independent experiments using 1csym as a designed reactant, yielded values that are very close to unity (Figure 4). This result is in variance with the previously discussed values for KIEC–Cl and KIEC–Br, both of which are consistent with OA being the first irreversible step for the aryl bromides/aryl chlorides. The predicted KIEC–I for transition structure TS-OA-1csym is 1.012, which is higher than the experimental KIEC–I of unity. Interpretation of this result is that the carbon atom attached to iodine is not involved in the first irreversible step in the catalytic cycle for 1csym; this rules out OA and RE (predicted RE KIEC–I of 1.028, structure in the Supporting Information) as the first irreversible step. Two possibilities emerge as an explanation for the observed KIEC–I—(a) binding of 1csym to a Pd0 complex is irreversible or (b) either amine binding or deprotonation of bound amine to form the pre-RE complex 5 (Figure 1) is the first irreversible step in the catalytic cycle. Further experiments are required to confirm the exact mechanism of the Buchwald–Hartwig amination of aryl iodides. However, the key takeaway is that vital information regarding changes in TS geometry and/or mechanism upon changing reactants is obtained from a single reaction performed under standard catalytic conditions.
Effect of Base on Oxidative Addition TS.
In 2001, elegant kinetic studies by Hartwig revealed two pathways for OA, in the Buchwald–Hartwig amination reaction of aryl chlorides and secondary amines catalyzed by Pd[P(t-Bu)3]2, that varied depending on the nature of the base used for the reaction (Figure 5).21 When NaOC(Me)3 was used as the base, the reaction was found to be first order in aryl chloride, zero order in amine, and positive order in base. This led to the proposal of a base-dependent pathway in which tert-butoxide displaces one of the ligands on the catalyst to form an anionic [P(t-Bu)3Pd0 (OC(Me)3)]− complex Pd0L1(OR)−, which adds aryl chloride in the turnover limiting step. When the bulkier NaOC(Et)3 was used, the reaction was zero order in base suggesting that OA of aryl halide occurred to a neutral Pd0P(t-Bu)3 complex (Pd0L1) in the turnover limiting step.
Figure 5.

Utilization of symmetric aryl chloride to probe the effects of base on the oxidative addition transition structure.
We sought to probe this base-dependent mechanism of OA using intramolecular 13C KIEs with 1bsym. Accordingly, separate reactions of 1bsym and 2b were conducted using in situ formed Pd[P(t-Bu)3]2 as the catalyst and either NaOC(Me)3 or NaOC(Et)3 as the base. Intriguingly, the experimental KIEC–Cl determined by intramolecular KIE analysis of 3bbdesym was drastically different depending on the base used (Figure 5). Duplicate measurements in the presence of NaOC(Me)3 as a base yielded a KIEC–Cl value of ~1.026. On the other hand, corresponding measurements in the presence of NaOC(Et)3 as a base resulted in KIEC–Cl values of ~1.041. To interpret the origin of this difference, we modeled the TS for OA to 1bsym using neutral PdL1 (TS-OAL1-1b sym) or anionic PdL1(OR)− with Na+ as the counterion (TS-OAL1+Base-1bsym). The resulting transition structures (Figure 5) are fundamentally different—while TS-OAL1–1bsym corresponds to a classic OA-type TS, the anionic TS-OAL1+Base-1bsym resembles an SNAr- type TS. The SNAr- type TS has a shorter forming Pd–C bond (rPd–C = 1.97 Å) and consequently gives a lower prediction of KIEC–Cl (1.021) compared to the OA-type TS, which has a longer forming Pd–C bond (rPd–C = 2.19 Å) and a larger predicted KIEC–Cl (1.039). Gratifyingly, both predictions are in good agreement22 with our experimental KIEC–Cl values and validate Hartwig’s mechanistic proposal that the less bulky base proceeds via the base-mediated anionic Pd0 pathway and the bulky base proceeds via OA to a neutral Pd0L1 complex. These results indicate that 13C KIEs are a sensitive probe of the ligand environment at a metal center during the KIE-determining step of a reaction, a finding that makes 13C KIEs a valuable addition to the tools available to probe transition metal-catalyzed reactions under synthetically relevant conditions.
Scope and Limitations.
Finally, to illustrate the utility of this methodology in other areas of catalysis, we studied the l-proline catalyzed aldol reaction23 of acetone (6) and aromatic aldehydes using terephthaldehyde (7) as a designed reactant (Figure 6). The aldol product 8, isolated from a carefully monitored reaction taken to ~50% conversion of 7, was used for the determination of intramolecular 13C KIEs. The resulting 13C KIEs for 7, from two independent measurements, is shown in Figure 6. Gratifyingly, the ~1.045 KIE observed on the aldehyde carbon atom is in excellent agreement with the predicted KIEs for the widely accepted Houk-List transition state for C–C bond formation (TS-CC-7, Figure 6).24 This result establishes C–C bond formation as the enantioselectivity-determining step of the reaction.
Figure 6.

Utilization of a symmetric aromatic aldehyde to probe the mechanism of the l-proline catalyzed aldol reaction.
The examples presented in this manuscript illustrate the advantages of using reactants with two identical reaction sites for the rapid determination of 13C KIEs at natural abundance. As mentioned earlier, an important caveat for the success of this approach is the careful exclusion of “double reaction.” Formation of even minor amounts of the product resulting from reaction at both sites will affect the 13C isotopic composition at the standard carbon (Scheme 2) and hence compromise the intramolecular KIE measurement. This limitation of intramolecular KIE determination can often be overcome using an excess of the symmetric reactant and/or terminating the reaction at low conversions. However, in reactions where the desymmetrized product is significantly more reactive than the symmetric starting materials, it might be difficult to avoid trace amounts of difunctionalization.
This novel approach has the potential to expand the scope of reactions amenable to KIE analysis. Intermolecular KIE approaches are not applicable to reactions where a non-symmetric reactant is used in large excess since high reaction conversion (80 or 100% required for starting material or product analysis, respectively) cannot be achieved. In these cases, the use of a designed symmetric reactant as a surrogate for the native reactant will allow for the determination of relevant KIEs by isolation of the desymmetrized product. Additionally, reactions where two processes run in competition to deliver two different products (such as SN2 versus E2 products in reactions of alkyl halides with a base) represent another class of reactions that cannot typically be studied using intermolecular KIE approaches. The use of a designed symmetric reactant and isolation of the mono-functionalized products from each reaction channel will allow for the determination of KIEs for both processes from a single reaction. This novel strategy is also expected to find application in studying dimerization reactions and reactions where the mechanism evolves as a function of time, two classes of reactions that are typically not suited for intermolecular KIE approaches. Finally, it is worthwhile to note that this approach can also be utilized to determine deuterium KIEs without explicit labeling, via 2H NMR at natural abundance, by the design of appropriate reactants with homotopic or enantiotopic hydrogen atoms. A detailed discussion of the full potential of this novel approach is included on page S20 of the Supporting Information.
CONCLUSIONS
In conclusion, we have benchmarked a novel method for the rapid determination of 13C KIEs at natural abundance. This method relies on the use of “designed” symmetric reactants and utilizes only a fraction of the material used for traditional 13C KIE experiments. Challenging problems in transition metal catalysis can be addressed using this method without arduous kinetic studies or synthesis of putative reaction intermediates, as evidenced by the rapid evaluation of the mechanism of the Buchwald–Hartwig amination under a variety of different reaction conditions. Additionally, the application of this approach to study an organocatalytic aldol reaction has been demonstrated. The potential application to other challenging mechanistic studies is also discussed. This approach will be particularly useful in reaction optimization to rapidly probe the effect of changing reaction parameters (solvent, ligand, catalyst, concentration, etc.) on the reaction mechanism. We expect this methodology to gain traction in the mechanistic study of organic reactions both in industry and academia.
Supplementary Material
ACKNOWLEDGMENTS
Financial support for this work was provided by NIGMS (R01 GM126283). M.J.V. and J.S.H. acknowledge support from the XSEDE Science Gateways Program (allocation IDs CHE160009 and CHE180061), which is supported by the National Science Foundation grant number ACI-1548562.
Footnotes
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.0c04752.
Experimental procedures, coordinates of all computed structures, and product characterization data; design of experiment for the determination of intermolecular KIEs for the Buchwald–Hartwig amination reaction via product analysis (Figure S1); design of experiment for the determination of intramolecular KIEs for the Buchwald–Hartwig amination reaction using purposely designed symmetric reactants (Figure S2); integrations for the intermolecular KIE experiment for the reaction of t-butylbromobenzene and n-hexylamine (Table S1); integrations for the intermolecular KIE experiment for the reaction of t-butylbromobenzene and N-methylpiperazine (Table S2) (PDF)
Complete contact information is available at: https://pubs.acs.org/10.1021/acscatal.0c04752
The authors declare no competing financial interest.
REFERENCES
- (1).Singleton DA; Thomas AA High-Precision Simultaneous Determination of Multiple Small Kinetic Isotope Effects at Natural Abundance. J. Am. Chem. Soc 1995, 117, 9357–9358. [Google Scholar]
- (2).Beno BR; Houk KN; Singleton DA Synchronous or Asynchronous? An “experimental” Transition State from a Direct Comparison of Experimental and Theoretical Kinetic Isotope Effects for a Diels-Alder Reaction. J. Am. Chem. Soc 1996, 118, 9984–9985. [Google Scholar]
- (3).Hirschi JS; Takeya T; Hang C; Singleton DA Transition-State Geometry Measurements From 13C Isotope Effects. The Experimental Transition State for the Epoxidation of Alkenes with Oxaziridines. J. Am. Chem. Soc 2009, 131, 2397–2403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Selected examples include; (a) Singleton DA; Merrigan SR; Liu J; Houk KN Experimental Geometry of the Epoxidation Transition State. J. Am. Chem. Soc 1997, 119, 3385–3386. [Google Scholar]; (b) DelM.onte AJ; Haller J; Houk KN; Sharpless KB; Singleton DA; Strassner T; Thomas AA Experimental and Theoretical Kinetic Isotope Effects for Asymmetric Dihydroxylation. Evidence Supporting a Rate-Limiting “(3 + 2)” Cycloaddition. J. Am. Chem. Soc 1997, 119, 9907–9908. [Google Scholar]; (c) Keating AE; Merrigan SR; Singleton DA; Houk KN Experimental Proof of the Non-Least-Motion Cycloadditions of Dichlorocarbene to Alkenes: Kinetic Isotope Effects and Quantum Mechanical Transition States. J. Am. Chem. Soc 1999, 121, 3933–3938. [Google Scholar]; (d) Frantz DE; Singleton DA Isotope Effects and the Mechanism of Chlorotrimethylsilane-Mediated Addition of Cuprates to Enones. J. Am. Chem. Soc 2000, 122, 3288–3295. [Google Scholar]; (e) Lee JK; Bain AD; Berti PJ Probing the Transition States of Four Glucoside Hydrolyses with 13C Kinetic Isotope Effects Measured at Natural Abundance by NMR Spectroscopy. J. Am. Chem. Soc 2004, 126, 3769–3776. [DOI] [PubMed] [Google Scholar]; (f) Zhu H; Clemente FR; Houk KN; Meyer MP Rate Limiting Step Precedes C–C Bond Formation in the Archetypical Proline-Catalyzed Intramolecular Aldol Reaction. J. Am. Chem. Soc 2009, 131, 1632–1633. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Edwards DR; Hleba YB; Lata CJ; Calhoun LA; Crudden CM Regioselectivity of the Rhodium-Catalyzed Hydroboration of Vinyl Arenes: Electronic Twists and Mechanistic Shifts. Angew. Chem., Int. Ed 2007, 46, 7799–7802. [DOI] [PubMed] [Google Scholar]; (h) Yoshikai N; Matsuda H; Nakamura E Ligand Exchange as the First Irreversible Step in the Nickel-Catalyzed Cross-Coupling Reaction of Grignard Reagents. J. Am. Chem. Soc 2008, 130, 15258–15259. [DOI] [PubMed] [Google Scholar]; (i) Bartelson KJ; Singh RP; Foxman BM; Deng L Catalytic Asymmetric [4 + 2] Additions with Aliphatic Nitroalkenes. Chem. Sci 2011, 2, 1940–1944. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Ryan SJ; Candish L; Lupton DW N-Heterocyclic Carbene-Catalyzed (4 + 2) Cycloaddition/Decarboxylation of Silyl Dienol Ethers with α,β-Unsaturated Acid Fluorides. J. Am. Chem. Soc 2011, 133, 4694–4697. [DOI] [PubMed] [Google Scholar]; (k) Lutz JP; Rathbun CM; Stevenson SM; Powell BM; Boman TS; Baxter CE; Zona JM; Johnson JB Rate-Limiting Step of the Rh-Catalyzed Carboacylation of Alkenes: C–C Bond Activation or Migratory Insertion? J. Am. Chem. Soc 2012, 134, 715–722. [DOI] [PubMed] [Google Scholar]; (l) Dell’Amico L; Rassu G; Zambrano V; Sartori A; Curti C; Battistini L; Pelosi G; Casiraghi G; Zanardi F Exploring the Vinylogous Reactivity of Cyclohexenylidene Malononi-triles: Switchable Regioselectivity in the Organocatalytic Asymmetric Addition to Enals Giving Highly Enantioenriched Carbabicyclic Structures. J. Am. Chem. Soc 2014, 136, 11107–11114. [DOI] [PubMed] [Google Scholar]; (m) Colletto C; Islam S; Juliá-Hernández F; Larrosa I Room-Temperature Direct β-Arylation of Thiophenes and Benzo[b]Thiophenes and Kinetic Evidence for a Heck-Type Pathway. J. Am. Chem. Soc 2016, 138, 1677–1683. [DOI] [PMC free article] [PubMed] [Google Scholar]; (n) Wang JY; Strom AE; Hartwig JF Mechanistic Studies of Palladium-Catalyzed Aminocarbonylation of Aryl Chlorides with Carbon Monoxide and Ammonia. J. Am. Chem. Soc 2018, 140, 7979–7993. [DOI] [PMC free article] [PubMed] [Google Scholar]; (o) Wendlandt AE; Vangal P; Jacobsen EN Quaternary Stereocentres via an Enantioconvergent Catalytic SN1 Reaction. Nature 2018, 556, 447–451. [DOI] [PMC free article] [PubMed] [Google Scholar]; (p) Kennedy CR; Zhong H; MacAulay RL; Chirik PJ Regio- and Diastereoselective Iron-Catalyzed [4+4]-Cycloaddition of 1,3-Dienes. J. Am. Chem. Soc 2019, 141, 8557–8573. [DOI] [PMC free article] [PubMed] [Google Scholar]; (q) Pannilawithana N; Yi CS Catalytic Carbon-Carbon Bond Activation of Saturated and Unsaturated Carbonyl Compounds via Chelate-Assisted Coupling Reaction with Indoles. ACS Catal. 2020, 10, 5852–5861. [Google Scholar]; (r) Sanders JN; Jun HJ; Yu RA; Gleason JL; Houk KN Mechanism of an Organocatalytic Cope Rearrangement Involving Iminium Intermediates: Elucidating the Role of Catalyst Ring Size. J. Am. Chem. Soc 2020, 142, 16877–16886. [DOI] [PubMed] [Google Scholar]
- (5).(a) Wolfe JP; Tomori H; Sadighi JP; Yin J; Buchwald SL Simple, Efficient Catalyst System for the Palladium-Catalyzed Amination of Aryl Chlorides, Bromides, and Triflates. J. Org. Chem 2000, 65, 1158–1174. [DOI] [PubMed] [Google Scholar]; (b) Surry DS; Buchwald SL Biaryl Phosphane Ligands in Palladium-Catalyzed Amination. Angew. Chem., Int. Ed 2008, 47, 6338–6361. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Surry DS; Buchwald SL Dialkylbiaryl Phosphines in Pd-Catalyzed Amination: A User’ s Guide. Chem. Sci 2011, 2, 27–50. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Wolfe JP; Buchwald SL Scope and Limitations of the Pd / BINAP-Catalyzed Amination of Aryl Bromides Catalytic Amination of Aryl Bromides Using. J. Org. Chem 2000, 65, 1144–1157. [DOI] [PubMed] [Google Scholar]
- (6).(a) Singleton DA; Szymanski MJ Simultaneous Determination of Intermolecular and Intramolecular 13C and 2H Kinetic Isotope Effects at Natural Abundance. J. Am. Chem. Soc 1999, 121, 9455–9456. [Google Scholar]; (b) Singleton DA; Hang C; Szymanski MJ; Meyer MP; Leach AG; Kuwata KT; Chen JS; Greer A; Foote CS; Houk KN Mechanism of Ene Reactions of Singlet Oxygen. A Two-Step No-Intermediate Mechanism. J. Am. Chem. Soc 2003, 125, 1319–1328. [DOI] [PubMed] [Google Scholar]
- (7).Gonzalez-James OM; Zhang X; Datta A; Hrovat DA; Borden WT; Singleton DA Experimental Evidence for Heavy-Atom Tunneling in the Ring-Opening of Cyclopropylcarbinyl Radical from Intramolecular 12C/13C Kinetic Isotope Effects. J. Am. Chem. Soc 2010, 132, 12548–12549. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Kurouchi H; Singleton DA Labelling and Determination of the Energy in Reactive Intermediates in Solution Enabled by Energy-Dependent Reaction Selectivity. Nat. Chem 2018, 10, 237–241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Roytman VA; Karugu RW; Hong Y; Hirschi JS; Vetticatt MJ 13C Kinetic Isotope Effects as a Quantitative Probe To Distinguish between Enol and Enamine Mechanisms in Amino-catalysis. Chem. - Eur. J 2018, 24, 8098–8102. [DOI] [PubMed] [Google Scholar]
- (10).(a) Alcazar-roman LM; Hartwig JF; Rheingold AL; Liable-Sands LM; Guzei IA Mechanistic Studies of the Palladium-Catalyzed Amination of Aryl Halides and the Oxidative Addition of Aryl Bromides to Pd(BINAP)2 and Pd(DPPF)2: An Unusual Case of Zero-Order Kinetic Behavior and Product Inhibition. J. Am. Chem. Soc 2000, 122, 4618–4630. [Google Scholar]; (b) Alcazar-Roman LM; Hartwig JF Mechanism of Aryl Chloride Amination: Base-Induced Oxidative Addition. J. Am. Chem. Soc 2001, 123, 12905–12906. [DOI] [PubMed] [Google Scholar]; (c) Alcazar-roman LM; Hartwig JF Mechanistic Studies on Oxidative Addition of Aryl Halides and Triflates to Pd(BINAP)2 and Structural Characterization of the Product from Aryl Triflate Addition in the Presence of Amine. Organometallics 2002, 21, 491–502. [Google Scholar]; (d) Singh UK; Strieter ER; Blackmond DG; Buchwald SL Mechanistic Insights into the Pd(BINAP)-Catalyzed Amination of Aryl Bromides: Kinetic Studies under Synthetically Relevant Conditions. J. Am. Chem. Soc 2002, 124, 14104–14114. [DOI] [PubMed] [Google Scholar]; (e) Shekhar S; Ryberg P; Hartwig JF; Mathew JS; Blackmond DG Reevaluation of the Mechanism of the Amination of Aryl Halides Catalyzed by BINAP-Ligated Palladium Complexes. J. Am. Chem. Soc 2006, 128, 3584–3591. [DOI] [PubMed] [Google Scholar]; (f) Shekhar S; Ryberg P; Hartwig JF; V, Y. U. Oxidative Addition of Phenyl Bromide to Pd(BINAP) vs Pd(BINAP)(Amine). Evidence for Addition to Pd(BINAP). Org. Lett 2006, 8, 851–854. [DOI] [PubMed] [Google Scholar]
- (11).(a) Becke AD Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]; (b) Lee C; Yang W; Parr RG Condens. Matter Mater. Phys. Phys. Rev. B 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- (12).(a) Grimme S; Antony J; Ehrlich S; Krieg H A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys 2010, 132, No. 154104. [DOI] [PubMed] [Google Scholar]; (b) Grimme S; Ehrlich S; Goerigk L Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem 2011, 32, 1456–1465. [DOI] [PubMed] [Google Scholar]
- (13).Zhao Y; Truhlar DG 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. [Google Scholar]
- (14).Zhao Y; Truhlar DG A New Local Density Functional for Main-Group Thermochemistry, Transition Metal Bonding, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Phys 2006, 125, 19401. [DOI] [PubMed] [Google Scholar]
- (15).Amovilli C; Barone V; Cammi R; Cancès E; Cossi M; Mennucci B; Pomelli CS; Tomasi J Recent advances in the description of solvent effects with the polarizable continuum model. Adv. Quantum Chem 1998, 32, 227–261. [Google Scholar]
- (16).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Mennucci B; Petersson GA; Nakatsuji H; Caricato M; Li X; Hratchian HP; Izmaylov AF; Bloino J; Zheng G; Sonnenberg JL; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark M; Heyd JJ; Brothers E; Kudin KN; Staroverov VN; Kobayashi R; Normand J; Raghavachari K; Rendell A; Burant JC; Iyengar SS; Tomasi J; Cossi M; Rega N; Millam JM; Klene M; Knox JE; Cross JB; Bakken V; Adamo C; Jaramillo J; Gomperts R; Stratmann RE; Yazyev O; Austin AJ; Cammi R; Pomelli C; Ochterski JW; Martin RL; Morokuma K; Zakrzewski VG; Voth GA; Salvador P; Dannenberg JJ; Dapprich S; Daniels AD; Farkas O; Foresman JB; Ortiz JV; Cioslowski J; Fox DJ Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2009. [Google Scholar]
- (17).Anisimov V; Paneth P ISOEFF98. A Program for Studies of Isotope Effects Using Hessian Modifications. J. Math. Chem 1999, 26, 75–86. [Google Scholar]
- (18).Scott AP; Radom L Harmonic Vibrational Frequencies: An Evaluation of Hartree-Fock, Møller-Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors. J. Phys. Chem. A 1996, 100, 16502–16513. [Google Scholar]
- (19).Bell RP The Tunnel Effect in Chemistry; Chapman & Hall: London, 1980; pp 60–63. [Google Scholar]
- (20). See Supporting Information for full details of this experiment.
- (21).Alcazar-Roman LM; Hartwig JF Mechanism of Aryl Chloride Amination: Base-Induced Oxidative Addition. J. Am. Chem. Soc 2001, 123, 12905–12906. [DOI] [PubMed] [Google Scholar]
- (22).We recognize that the predicted KIEC-Cl of 1.021 is slightly lower than the experimental KIEC-Cl of 1.026. However, when the average of all four functionals is considered, predicted KIEC-Cl for TS-OAL1+Base-1bsym increases to 1.025—which is in excellent agreement with experiment.
- (23).List B; Lerner RA; Barbas CF Proline-catalyzed direct asymmetric aldol reactions. J. Am. Chem. Soc 2000, 122, 2395–2396. [Google Scholar]
- (24).(a) Hoang L; Bahmanyar S; Houk KN; List B Kinetic and Stereochemical Evidence for the Involvement of Only One Proline Molecule in the Transition States of Proline-Catalyzed Intra- and Intermolecular Aldol Reactions. J. Am. Chem. Soc 2003, 125, 16–17. [DOI] [PubMed] [Google Scholar]; (b) Bahmanyar S; Houk KN Origins of Opposite Absolute Stereoselectivities in Proline Catalyzed Direct Mannich and Aldol Reactions. Org. Lett 2003, 5, 1249–1251. [DOI] [PubMed] [Google Scholar]; (c) Bahmanyar S; Houk KN; Martin HJ; List B Quantum Mechanical Predictions of the Stereoselectivities of Proline-Catalyzed Asymmetric Intermolecular Aldol Reactions. J. Am. Chem. Soc 2003, 125, 2475–2479. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
