Abstract
Rate constants for hydrogen atom transfer (HAT) reactions of substituted toluenes with tert-butyl, tert-butoxy, and tert-butylperoxyl radicals are re-analyzed here using the free energies of related proton transfer (PT) and electron transfer (ET) reactions, calculated from an extensive set of compiled or estimated pKa and E° values. The Eyring activation energies ΔG‡HAT do not correlate with the relatively constant ΔG°HAT, but do correlate close-to-linearly with ΔG°PT and ΔG°ET. The slopes of correlations are similar for the three radicals except that the tBu• barriers shift in the opposite direction from the oxyl radical barriers—a clear example of the qualitative “polar effect” in HAT reactions. When cast quantitatively in free energy terms (ΔG‡HAT vs. ΔG°PT/ET), this effect is very small, only 5–10% of the typical Bell-Evans-Polanyi (BEP) effect of changing ΔG°HAT. This analysis also highlights connections between polar effects and the concepts of “asynchronous” or “imbalanced” HAT reactions, in which the PT and ET components of ΔG°HAT contribute differently to the barrier. Finally, these observations are discussed in light of the traditional explanations of polar effects and the potential for a rubric that could predict the extent to which contra-thermodynamic selectivity may be achieved in HAT reactions.
Graphical Abstract

INTRODUCTION
Hydrogen atom transfer (HAT) has been one of the fundamental reactions of organic chemistry for over a century (eq 1). It is common in biological transformations, atmospheric chemistry, combustion, catalysis, and synthetic chemistry. In catalysis and synthesis, achieving the needed selectivity is often difficult. The rates of HAT reactions are primarily governed by the strengths of the bonds being made and broken, so reagents that make stronger bonds are more reactive, and weaker bonds are easier to break.
| (1) |
“Polar effects” have emerged as a way to achieve selectivities that differ from the normal radical selectivity based on bond strengths.1 Electrophilic radicals such as alkoxyl or fluoroalkyl radicals (RO•, RF•) show a kinetic preference for electron-rich X–H bonds, while nucleophilic radicals such as simple alkyl radicals (R•) react faster with electron-poor X–H bonds.2,3 For example, F3C• abstracts H from HCl ~70 times slower than H3C•, despite having an equilibrium constant ~30 times greater.3 A dramatic effect is also seen in HAT reactions between acylated amino acids and electrophilic radicals such as Cl• or HO•, which preferentially abstract from distal positions instead of from the more thermodynamically favorable α-carbon.4–6 However, the effect is often more subtle, as in the abstractions from substituted toluenes analyzed here.
While polar effects are valuable, they remain largely qualitative and incompletely understood. They have been attributed to the electronegativities of the X and Y groups involved, and rationalized by dipole-dipole (electrostatic) stabilization of transition states.3 Frontier-orbital and valence bond (VB) models have provided important insights, emphasizing interaction of the X• SOMO orbital with the H–Y HOMO or LUMO.7,8,9 Some of the semiempirical models developed to predict HAT barrier heights have related perspectives.9–13
However, the dipole or frontier molecular orbital pictures are not easily applied to all HAT reactions. In the widely studied H-abstractions by compound I of cytochrome P450s, the H• being removed is ‘separated’ into an H+ that binds to the Fe=O oxygen and an e− that transfers to a porphyrin/thiolate radical cation (eq 2).14 Polar effects are more complex to rationalize in such reactions since the SOMO is distant from the H being transferred. More broadly, HAT reactions are a subset of concerted proton-coupled electron transfer (PCET) reactions in which the proton and electron transfer “together,” and this “together” may mean into the same orbital, onto the same atom, or onto the same protein.15 A different approach to polar effects is needed to encompass this continuum.
![]() |
(2) |
Analyses of HAT reactions have recently been expanded to include ‘imbalances’ in the ET or PT components of a concerted e–/H+ (H•) transfer.16–24 Computational and experimental studies alike have found that the relationship between ΔG°PT and ΔG°ET can affect barrier heights in ways not captured by ΔG°HAT. These effects have sometimes been attributed to a so-called “asynchronous” or “imbalanced” transition state, wherein the PT and ET reaction coordinates have proceeded to different extents.16–24 Srnec in particular has developed theoretical models for the effects of asynchronicity and discussed the possible connection with polar effects.17,25 Others have incorporated these effects into empirical models through statistical analysis and machine learning.9,16,21,26,27
From one perspective, the proposal of asynchrony or imbalance in a PCET reaction harkens back to the classic physical-organic discussions of ‘asynchrony’ that were often illustrated as off-diagonal paths in More O’Ferrall-Jencks diagrams such as in Figure 1.28,29 In his Principle of Nonperfect Synchronization (PNS), Bernasconi related this imbalance in PT reactions to changes in their intrinsic barriers.30,31 The PNS has also been invoked in HAT reactions to account for an increased intrinsic barrier in abstractions from allylic and benzylic C–H bonds compared to saturated substrates.32 However, the electron and the proton should be treated quantum-mechanically, not as classical particles with well-defined positions. Current theory often describes HAT as a simultaneous e–/H+ double-tunneling event,33,34 as discussed below.
Figure 1.

Generic potential energy surface for an exoergic PCET from XH to Y• and its projection onto a schematic More O’Ferrall-Jencks diagram. The HAT reaction pathway is drawn over the saddle point between the barriers for initial PT (left) or ET (right).
The goal of this paper is to connect the extensive but qualitative literature of polar effects with the emerging discussions of imbalanced HAT based on the relevant reduction potentials (E°) and acidities (pKas). Specifically, we re-analyze the reported rate constants for HAT from substituted toluenes to the nucleophilic tert-butyl radical and to the electrophilic tert-butoxyl and tert-butylperoxyl radicals, using E° and pKa values that we have assembled. As shown below, this is almost an ideal system to explore non-BEP effects because the toluenes have very similar BDFEs but vary widely in their pKas and E°s.
We examine these experimental data using simple linear free energy relationships (LFERs), exploring how the HAT rates and barriers depend on ΔG°ET, ΔG°PT, and ΔG°HAT. Through the ΔG°ET/ΔG°PT lens, we draw an explicit, experimental, and quantitative connection between the classical polar effects and emerging discussions of imbalance in HAT reactions. Finally, we discuss the results within the context of (i) the traditional explanations of polar effects, (ii) the asynchronicity parameter η developed by Srnec and coworkers,17 and (iii) quantum-mechanical treatments of HAT reactions. We hope that the analysis presented here will move the field towards a rubric that could predict the extent to which contra-thermodynamic selectivity may be achieved in HAT reactions.
LINEAR FREE ENERGY RELATIONSHIPS, briefly
Relationships between rates and driving forces are a foundation of physical organic chemistry and have been applied to HAT reactions for close to a century.35–42 These can be written in terms of rate constants and equilibrium constants in the Brønsted catalysis ‘law’ (eq 3) or as LFERs between ΔG‡ and ΔG° (eq 4). The latter are often referred to as BEP relationships, for Bell-Evans-Polanyi or Brønsted-Evans-Polanyi. Equations (3) and (4) are conceptually the same and the same series of reactions will give the same value of α in either treatment.
| (3) |
| (4) |
The proportionality constant, the Brønsted α, captures the sensitivity of the barrier height to changes in driving force. α is almost always between 0 and 1 and is often close to ½ (Marcus theory pins α to ½ for isoergic reactions, and it is frequently taken as ½ in its electrochemical incarnation as the symmetry factor β in the Butler-Volmer equation). α is expected from the Hammond Postulate to be closer to 1 for highly endoergic reactions where the transition state is close to the products, and smaller for highly exergonic reactions. This has suggested a rough connection between α and the position of the transition state along the reaction coordinate.43,44
Traditional applications of BEP relationships to HAT reactions have used enthalpies such as bond dissociation enthalpies (BDEs), correlating Arrhenius activation energies Ea with reaction ΔH°. However, free energies are at the core of the original Evans and Polanyi papers,38–40 the Brønsted law, LFERs, and Marcus-type analyses, so BEP relationships should use free energies. For HAT reactions, bond dissociation free energies (BDFEs) are appropriate for the typical solution reactions of organic, inorganic, and biological chemistry (despite the dominance of BDEs in organic chemistry textbooks). BDFEs are discussed and tabulated in a recent review.45 The free energy of an HAT reaction is then the difference between the BDFEs of the reactant and product.
RESULTS AND DISCUSSION
I. Compilation of data
A. Kinetic data
The many studies of HAT from substituted toluenes to tert-butyl,46–49 tert-butoxy,50–57 and tert-butylperoxyl58–60 radicals differ in the solvent and temperature in which the reactions were performed, the techniques used, and whether the measured rates are relative or absolute. The kinetic data for the reactions of tBuO• and tBuOO• used here are from a 1994 compilation by Héberger that evaluated, normalized, and averaged reported rate constants for these reactions.61 The compiled rate constants were measured in various solvents, ranging in polarity from carbon tetrachloride to acetonitrile, and were normalized to 40 °C for tBuO• and 20 °C for tBuOO•. The relatively small differences in temperature should not affect the general conclusions. For the tBu• HAT reactions, we have instead analyzed data from a single study.48 This is because the trends discussed in this work hold for each individual study (see SI Section S2), but the reported rate constants shift from one report to another, introducing needless scatter (for our purposes). The kinetic data used here are listed in Table 1.
Table 1.
Kinetic and thermochemical data for HAT from substituted toluenesa
|
b | Thermochemical parameters g | |||||||
|---|---|---|---|---|---|---|---|---|---|
| σ c | tBu• d | tBuO• e | tBuOO• f | BDFEh | E°(R•/−)i | E°(RH•+/0)i | pKa(RH)j | pKa(RH•+)j | |
| p-OMe | −0.27 | 5.96 | −1.08 | 81.8 | −2.19 | 1.43 | 58.1 | −2.7 | |
| p-t-Bu | −0.20 | 1.25 | 84.5 | −2.09k | 1.54 | 58.5 | −2.6 | ||
| p-Me | −0.17 | 1.15 | 5.73 | −1.30 | 84.2 | −2.02 | 1.63 | 57.1 | −4.3 |
| m,m-Me2 | −0.14 | 1.17 | 5.59 | 84.6 | −2.03k | 1.61 | 57.5 | −3.8 | |
| m-Me | −0.07 | 1.18 | 5.66 | −1.46 | 84.7 | −1.90 | 1.70 | 55.4 | −5.2 |
| p-OPh | −0.03 | 5.73 | −1.18 | 85.4 | −1.92k | 1.68k | 56.2 | −4.3 | |
| p-Ph | −0.01 | 5.84 | −1.90k | 1.70k | |||||
| H | 0.00 | 1.16 | 5.51 | −1.22 | 84.9 | −1.83 | 1.87 | 54.4 | −7.9 |
| p-F | 0.06 | 1.15 | 83.9 | −1.90 | 1.76k | 54.9 | −6.7 | ||
| m-OMe | 0.12 | −1.57 | 85.7 | −1.76k | 1.52 | 53.8 | −1.4 | ||
| p-Cl | 0.23 | 1.45 | 5.49 | −1.52 | 84.9 | −1.80 | 1.86 | 53.9 | −7.7 |
| p-Br | 0.23 | 5.36 | 86.0 | −1.65k | 1.83 | 52.2 | −6.5 | ||
| m-F | 0.34 | 1.26 | 84.3 | −1.54k | 1.99k | 49.0 | −10.3 | ||
| m-Cl | 0.37 | 1.32 | 5.37 | −1.73 | 85.1 | −1.51k | 2.01k | 49.1 | −10.1 |
| p-CO2Me | 0.45 | −1.60 | 84.6l | −1.42k | 1.99 | 47.3 | −10.1 | ||
| p-C(O)Me | 0.50 | −2.10 | −1.11 | 2.24 | |||||
| m-CN | 0.56 | 5.33 | −1.96 | 83.9 | −1.51 | 2.21 | 48.3 | −14.3 | |
| m,p-Cl2 | 0.60 | 1.57 | 83.5 | −1.27k | 2.20k | 44.0 | −14.5 | ||
| p-CN | 0.66 | 1.67 | 5.18 | −1.85 | 83.2 | −1.17 | 2.24 | 42.1 | −15.3 |
| m-NO2 | 0.71 | 4.97 | −1.98 | 84.8 | −1.16k | 2.30 | 43.0 | −15.2 | |
| p-NO2 | 0.78 | 4.91 | −1.84 | 84.0 | −1.08k | 2.33 | 41.2 | −16.2 | |
| Range m | 1.05 | 0.52 | 0.93 | 0.92 | 4.2 | 1.10 | 0.90 | 17.3 | 14.8 |
An empty cell indicates a value not known.
Bimolecular rate constants in M−1 s−1.
Rate constants at 48 °C in neat substituted toluene from ref 48.
Rate constants at 40 °C in various systems and solvents, compiled in ref61.
Rate constants at 20 °C in various systems and solvents, compiled in ref61.
For the sources of the BDFEs, E½s, and pKas, see Section 2 of the text and Section S3 of the SI.
BDFEs are ± 2.0 kcal mol−1.
Potentials are ± 0.10 V vs Fc+/0.
pKas are ± 2.0 units.
Potentials extrapolated from Hammett correlation (see SI Section S3 for details).
BDFE given is that of p-CO2Et-C6H4Me.
Each range given is that of the column above.
B. Thermochemical data
The BDFEs, reduction potentials, and pKas needed for the analysis were obtained from multiple sources, as described here and in the Supporting Information (SI Section S3). The values are in Table 1.
The reversible potentials for the 1e– reduction of about half of the substituted benzyl radicals (E°(R•/–)) have been measured in acetonitrile (MeCN) using photomodulation voltammetry.64 The values correlate well with the corresponding para- and meta-substituent Hammett constants, allowing estimation of the unmeasured E°(R•/–) values (see Figure S5). The 1e− oxidation potentials of the toluenes, E°(RH•+/0), were calculated from the established linear correlation65,66 with the measured gas-phase vertical ionization potentials obtained from NIST67 or by extrapolation using the Hammett constants.
The bond dissociation free energies (BDFEs) were estimated from the measured gas-phase bond dissociation energies (BDEs)68 using previously described thermochemical schemes involving an entropic correction to the BDE and the free energies to solvate RH, R•, and H•.45 When multiple BDEs have been reported, the recommended or most recent values were used.
Finally, the pKas of the toluenes and their radical cations, pKa(RH) and pKa(RH•+), were calculated from the redox potentials and BDFEs using eq 5 (and its analogue with pKa(RH•+) and E°(RH•+/0); all terms in kcal mol−1).45 Similar values and trends were obtained upon converting from gas-phase acidities.69 These extrapolations are discussed in greater detail in Section S3 of the SI.
| (5) |
The thermochemical parameters of interest were also estimated for the abstracting radicals in MeCN (Table 2). The solution-phase BDFEs were calculated from the gas-phase BDEs as described above. The E°(Y•/–) values have been measured in MeCN for tert-butyl70 and tert-butoxy radicals.71 The pKa(YH) values of each were calculated from values in other solvents using established correlations (see SI Section S3 for details). Each E°(XH•+/0) value was estimated from the gas-phase vertical ionization potentials as described above. The E°(tBuOO•/–), as well as the pKa(YH•+) of each abstractor, were determined from the above thermochemical parameters using eq 5. It is worth noting that even generous error bars do not detract from the trends discussed below.
Table 2.
BDFE, E°, and pKa values for tBu•,tBuO•, and tBuOO•/– a
| BDFE | E°(Y•/-) | E°(YH•+/0) | pKa(YH) | pKa(YH•+) | |
|---|---|---|---|---|---|
| tBu• | 90.9 | −2.94 | 3.37 | 78 | −29 |
| tBuO• | 101.5 | −0.70 | 2.81 | 45 | −12 |
| tBuOO• | 79.4 | −0.87 | 2.80 | 34 | −28 |
For the sources of the BDFEs, E½s, and pKas, see Section 2 of the text and Section S3 of the SI. BDFEs are ± 2.0 kcal mol−1, potentials are ± 0.10 V vs Fc+/0, and pKas are ± 2 units.
II. Correlations kinetic and thermodynamic data
A. Comparing kHAT for ZC6H4Me + tBu•, tBuO•, and tBuOO•
The rate of HAT from the toluenes to tBu• is about 4 orders of magnitude slower than to tBuO• and 2 orders of magnitude faster than to tBuOO•. This is the result of two primary effects. First, there are large differences in driving forces due to the different strengths of the Y–H bonds formed: the tBuO–H BDFE is ~11 kcal mol−1 higher than the Me3C–H bond, which is another ~11 kcal mol−1 stronger than the tBuOO–H bond. However, the HAT rates of tBu• do not sit directly in between those of tBuO• and tBuOO• but nearly 2 orders of magnitude slower. This can be explained by a higher intrinsic barrier for the HAT reactions of C-centered radicals as compared to those of O-centered radicals.72,73 We note that the BEP driving force effect and the intrinsic slowness of C-centered radicals are large effects, orders of magnitude (~106 and ~102, respectively). These contrast with the smaller magnitude of the polar effects discussed below.
B. Correlating HAT rate constants with the toluene BDFEs, pKas, and E°s
With consistent sets of HAT rate constants and thermochemical values in hand, we have analyzed rate/driving force LFERs for changes in the toluene substituents. Rate constants were converted to free energy barriers (ΔG‡HAT) using the Eyring equation. The ΔG°HAT for each reaction is the difference in the BDFEs of the R–H and Y–H bonds involved (Scheme 1, eq 6).
Scheme 1.

Reactions and energies for initial HAT, PT, and ET
The plot of ΔG‡HAT vs. ΔG°HAT shows essentially no correlation within the dataset for each radical (Figure 2A). This is likely due to the very small range of toluene BDFEs, only 4.2 kcal mol−1, which is on the order of the uncertainties themselves (±2 kcal mol−1). One way to view the constancy of the BDFEs is that the substituents cause offsetting changes in the pKa and E°. For example, para-nitrotoluene is 13.2 orders of magnitude more acidic than toluene, but the nitrobenzyl anion is 12.5 orders of magnitude (750 mV) less reducing. These shifts mostly cancel, leaving the BDFE of nitrobenzene only 0.9 kcal mol−1 weaker than that of toluene.
Figure 2.

Correlation of the HAT barrier height relative to the parent toluene with (A) the overall driving force, (B) the driving force for initial PT, and (C) the driving force for initial ET. Black points are the reactions of tert-butyl radical; blue, tert-butoxy; and red, tert-butyl peroxyl, and m indicates the slope of each linear fit. Note that in B and C the horizontal axes for ΔG° values are 40 times larger than the vertical axis for ΔΔG‡HAT. R2 values of fits: (B) 0.70 (tBu•), 0.88 (tBuO•), 0.77 (tBuOO•); (C) 0.65 (tBu•), 0.90 (tBuO•), 0.73 (tBuOO•). See Figure S7 for plots with data points labeled with their corresponding toluenes.
The similarity of the toluene BDFEs and insensitivity of the ΔG‡HAT values to these BDFEs is key to this study. It means that the comparison made here highlights the factors other than the BDFEs that affect the HAT barriers (i.e., non-BEP effects).
Plots of the HAT barriers ΔG‡HAT vs. the free energies for initial PT or ET—ΔG°PT and ΔG°ET—are shown in Figures 2B and 2C. The reactions do not proceed through initial PT or ET (which are +50 to +120 kcal mol−1 uphill); they are all concerted HAT. Nevertheless, the ΔG°PT and ΔG°ET are valuable because they are the energies of the opposite corners of the More O’Ferrall-Jencks diagram (the green hills in Figure 1). In contrast to the roughly constant ZC6H4CH2–H BDFEs, the pKas and E°s vary substantially, by 1.1 V in E° and by 17 units in pKa (both are a range of ~25 kcal mol−1 in ΔG°PT and ΔG°ET).
Two immediate observations should be made from the plots of ΔΔG‡ vs. ΔG°PT and ΔG°ET in Figures 2B and 2C. First, the ΔG‡ are not very sensitive to ΔG°PT or ΔG°ET (or (ΔG°PT – ΔG°ET) in a Figure below). The x-axes in these plots cover a 40-fold wider range than the y-axes (15-fold in Figure 2A). Second, the ΔG‡ nonetheless correlate linearly with ΔG°PT or ΔG°ET. Moreover, the correlations of tBuO• and tBuOO• are opposite in sign of the tBu• correlations. These observations and their implications will be discussed in the following sections.
The scatter in the data is in part due to the assumptions made to obtain the pKas and E°s. The rate data being from a host of different systems, temperatures, solvents, etc. likely also contributes to the scatter (see SI Section S2). Still, without the removal of any outliers, it is striking that these trends hold quite well. Plots of ΔΔG‡ vs. pKa, E½ and Hammett parameters are similar to the ΔG°PT and ΔG°ET plots shown here (SI Section S1). The shallow and opposite slopes of these lines are discussed in the next section.
C. Variation of HAT rate constants with substituents: polar effects and ΔG°PT / ΔG°ET correlations.
As emphasized in the original reports, HAT to the electrophilic tBuO• and tBuOO• radicals are faster with more electron-rich toluenes, while HAT to the nucleophilic tBu• radical is faster when electron withdrawing groups are present. For example, tBu• reacts ~3 times faster with the electron-poor para-cyanotoluene than with unsubstituted toluene, while tBuO• and tBuOO• react ~3 times slower. This is a clear polar effect. Since the range of kHAT for each radical approaches a factor of ten, changing the character of the radical can significantly affect abstraction selectivity.
Here we analyze the same rate data from a ΔG°ET/ΔG°PT perspective. The nucleophilic tBu• reacts faster with substrates that are better acids and poorer electron donors (e.g., para-cyano-toluene). The electrophilic oxyl radicals are the opposite, preferring substrates with a lower E° (more easily oxidized) and higher pKa (less acidic). Thus, qualitatively, there are good analogies between the traditional polar effects and the ΔG°ET/ΔG°PT perspective.
The ΔG°ET/ΔG°PT analysis is more quantitative than the traditional analyses of polar effects because the E° and pKa are (in principle) measurable thermochemical parameters. The slopes of the fit lines in Figure 2 are all unitless (kcal mol−1/kcal mol−1) and can therefore be compared. As noted above, the Brønsted plot (Figure 2A, ΔΔG‡HAT vs. ΔG°HAT) shows that the variation in the barrier is not due to differences in BDFEs but to differences in ΔG°ET and/or ΔG°PT. While the other plots are also ΔΔG‡HAT vs. ΔG°, they are not true Brønsted plots or BEP correlations because the free energy changes (ΔG°PT, ΔG°ET) are not for the full HAT reaction.
As noted above, the first lesson from these plots is that the magnitude of the polar effects is very small from an LFER perspective. The rate constants vary by about one order of magnitude for all three radicals, even though the Kas vary by 17 orders of magnitude (and similarly, E°s by 1.1 V). The unitless slopes in Figure 2B–C are all between 0.03 and 0.06. Compared to a typical BEP relationship with α ≈ ½, the barriers respond far less sensitively to changes in pKa and E°, about an order of magnitude less sensitively.
This conclusion is substantiated by the global electrophilicity index (ω), which quantifies the nucleo- or electrophilicity of a radical based on its electron affinity and ionization potential.74,75 The ω values computed for the benzyl radicals formed in the present reactions are all very similar, ranging from “moderately nucleophilic” to “weakly nucleophilic.” Furthermore, local electrophilicity indices, which describe the electrophilicity condensed at the radical center, describe the benzyl radicals as even more similar to one another (see Figure S9),75 despite their large differences in pKa and E°. This parallels the ΔG°ET/ΔG°PT analysis, that large swings in pKa and E° only induce weak to moderate polar effects.
A closer look at the slopes of the ΔΔG‡HAT vs. ΔG°PT or ΔG°ET plots in Figure 2 reveals several deeper implications. First, the slopes of the tBuOO• and tBuO• data are remarkably similar. This is most evident in Figure 2C, where the blue and red lines almost overlay. This is surprising because the tBuO• reactions are highly exoergic (–17 kcal mol−1 on average) while the tBuOO• reactions are endoergic (+5 kcal mol−1). Still, their barriers exhibit the exact same sensitivity to changes in reactant pKa/E°. The Hammond postulate would suggest that the uphill tBuOO• reactions have much “later” transition states and should therefore be more sensitive to changes in substituents. Viewed another way, this is a violation of the reactivity-selectivity principle: tBuO•, which reacts 107 times faster, shows virtually the same selectivity pattern as tBuOO•. Thus, the polar effect on the barrier height is insensitive to the traditional measure of the position of the transition state along the reaction coordinate.
The slopes are also roughly the same between the oxyl radical reactions and those of tBu•, except opposite in sign. This is mirrored in the Hammett ρ values for the reactions as well (Figure S2). The similarity is rough due to the large uncertainty for the tBu• slope. Still, the magnitude of the polar effect here appears independent of whether the incipient bond is C–H or O–H. This is surprising given the marked differences between carbon- and oxygen-centered radicals in both their intrinsic HAT rate and ability to hydrogenbond.72,73 This similarity between carbon and oxygen radicals has not been previously noted, to our knowledge, and we hope that its origin can be examined in future experimental and computational studies.
A final observation is that the constancy of ΔG°HAT in these reactions necessitates that the substituent effects are the same in the forward and reverse directions; kf and kr must vary by the same factor upon changing Z because their ratio does not change (Figure 3, eqs 9, 10). This, too, is a violation of the reactivity-selectivity principle. For instance, tBuO• + RH reactions are very downhill and the reverse tBuOH + R• are very uphill which would normally imply very different selectivities. Viewed another way, the changes in reactivity of the ZC6H4CH3 substrate are identical to the changes in reactivity of the ZC6H4CH2• radical. Again, this perhaps counterintuitive result should be general to any system where substituents change the barrier but not the driving force.76,77
Figure 3.

Simple free energy surface for the radical reactions in equation 9, illustrating that variation of the aryl substituent Z changes the barrier height but not the driving force. Eq 10 shows the implication of the essentially constant ΔG°HAT across the series.
In summary, the plots in Figure 2 reveal that the HAT rates correlate well with ΔG°ET and ΔG°PT, and that this is not attributable to changes in the overall ΔG°HAT. Moreover, tBu• and oxyl radicals respond oppositely to these changes in ΔG°ET/ΔG°PT, the former preferring acidic toluenes and the latter preferring electron-rich toluenes. When compared through unitless slopes, these polar effects are miniscule in comparison to BEP effects. Interestingly, the two oxyl radicals respond quantitatively the same to changes in ΔG°ET/ΔG°PT, despite the reactions of tBuOO• being endoergic and those of tBuO• being highly exoergic—an apparent violation of the reactivity-selectivity principle. Finally, the magnitude of the polar effect appears to be the same between the oxyl and tBu• radicals, which is surprising given the differences between the intrinsic HAT reactivity of C–H vs. O–H bonds.
D. Other models: “Imbalance,” the Srnec “asynchronicity” parameter, and quantum models.
The historical picture of polar effects in HAT reactions starts from transition state theory, with a classical hydrogen atom crossing over a transition state. From valence bond (VB) theory, the transition state can be stabilized by mixing with excited state configurations.78,79 For HAT reactions, a “polar” transition state is one that is stabilized by an excited state charge transfer configuration, namely, by X– + HY•+ or by XH+ + Y–. In the More O’Ferrall-Jencks diagram in Figure 1, the off-diagonal corners represent these excited state configurations. This is analogous to considering charged resonance structures of the transition state (namely, [X− ··H··Y+]‡ or [X+··H··Y–]‡). Conceptually, these approaches all can give HAT transition states in which the PT component is more (or less) developed than the ET component; [X∂– ···H··Y∂+]‡ or [X∂+··H···Y∂–]‡). This is the backdrop for the following sub-sections on “asynchrony” or “imbalance” in HAT reactions and quantitative models which describe it. We finish by discussing these classical pictures in view of a quantum-mechanical treatment of HAT reactions.
1). Imbalance.
As described above, a number of recent studies have discussed the possibility of an “imbalanced” or “asynchronous” transition state, wherein PT and ET are unevenly developed.16–24 This would correspond to a classical off-diagonal path in a More O’Ferrall-Jencks diagram (Figure 1). Indicators of such a pathway are (i) a difference between Brønsted α values for changes that predominantly affect the pKa or the E° of a reactant,18 or (ii) rates that correlate better with the pKa or E° of the changing reactant than with its BDFE.16,19,24,80
The present reactions, viewed through a ΔG°ET/ΔG°PT lens, highlight links between this imbalance and polar effects. tBu• rates increase as PT becomes more favorable but decrease as ET becomes more favorable. This demonstrates the greater sensitivity of tBu• barriers to changes along the PT coordinate. The opposite is true of tBuO• and tBuOO•, whose barriers show a greater sensitivity to the ET coordinate. In other words, the HAT reactivities of all three radicals suggest an imbalanced transition state: the nucleophilic tBu• in favor of PT and the electrophilic oxyl radicals in favor of ET.
2). The Srnec “asynchronicity” parameter η.
Srnec and coworkers have developed a mathematical model to predict the effects of thermodynamic “asynchronicity” on the barrier height for an HAT reaction, using the asynchronicity parameter η.17,25 η is essentially ΔG°PT – ΔG°ET (eq 11; F is Faraday’s constant), so we will use η and (ΔG°PT – ΔG°ET) interchangeably in this discussion. (ΔG°PT – ΔG°ET) is the difference in energy between the tops of the green hills in Figure 1.
For an electrophilic radical and/or hydridic X–H bond, η > 0 (ΔG°PT > ΔG°ET), and the opposite is true for a nucleophilic radical and/or acidic X–H bond. Interestingly, Srnec predicted and computationally showed that barriers will be lowest when |η| or |ΔG°PT – ΔG°ET| is largest. In other words, the rates will be fastest when the polarity of the H atom donor and acceptor are most different. Their quantitative relationship is in eq 12.17,25
| (11) |
| (12) |
The dataset of rate constants and thermochemical values developed here provide a detailed experimental test of eq 12 across a series of reactions. For each of the three radicals in this study, the HAT rate constants correlate linearly with |η| (Figure 4A). For the electron-rich tert-butyl radical, |η| is largest and rates are fastest with electron-poor toluenes. For the electrophilic oxygen-centered radicals, |η| is largest and rates are fastest with electron-rich toluenes. This is perhaps illustrated most simply in the plot of ΔG‡HAT versus ΔG°PT – ΔG°ET (Figure 4A): the highest barrier is when ΔG°PT and ΔG°ET are most similar (smallest η). For the tBu• reactions, the data tentatively suggest a peak or plateau of the barriers around η = 0. These results agree with the predictions of the Srnec model. The Srnec model has also been shown by Barman et al. and by Schneider et al., to provide improved correlations with experimental data for C–H bond activation by metal-oxo complexes.16,21
Figure 4.

(A) Correlation of the relative HAT barriers with ΔG°PT – ΔG°ET (bottom axis), or equivalently, the Srnec η (top axis). Slopes are given for ΔΔG‡ vs. ΔG°PT – ΔG°ET. (B) The relative HAT barriers predicted from eq 12. The dashed line indicates perfect agreement between the predicted and experimental relative barriers. Black points are the reactions of tert-butyl radical; blue, tert-butoxy; and red, tert-butyl peroxyl. R2 values: (A) 0.68 (tBu•), 0.91 (tBuO•), 0.78 (tBuOO•); (B) 0.34 (tBu•), 0.94 (tBuO•), 0.79 (tBuOO•). See Figure S8 for plots with data points labeled with their corresponding toluenes.
Despite the good correlation of the barrier height with |η| or |ΔG°PT – ΔG°ET|, the magnitude of the experimental effect is much smaller than that predicted by eq 12. The range of η for each of the oxygen-centered radical reactions is 1.34 V (31 kcal mol−1). According to eq 12, this should give a span of almost 8 kcal mol−1 in ΔG‡ and a 106 span in rate constant (Figures 4B and S6). However, within each radical series, the rate constants are all within a single order of magnitude. Therefore, η or (ΔG°PT – ΔG°ET) are informative predictors of polar effects in the present system, but their effects are small. Reference 21 similarly concluded that adding η to their analysis provided only a small improvement. The conclusion here of a small effect of η for the toluene reactions parallels the small influence of changes in pKa and E° discussed above.
The radical HAT reactions that exhibit more pronounced kinetic polar effects seem to have much larger thermodynamic “asynchronicities,” i.e., larger |ΔG°PT – ΔG°ET|). For instance, the classic contra-thermodynamic faster HAT from HCl to CH3• vs. CF3• (an effect of 102–3 in rate constant) can be traced to the ~1027 lower Ka of methane (in DMF83,84). A similar effect has been reported for the (pseudo-) self-exchange reactions of CH3/CF3• + CH4.3 In each case, the effect of |ΔG°PT − ΔG°ET| on the barrier seems to be 5–10% that of changing the overall HAT driving force.
The relationship between |ΔG°PT – ΔG°ET| and the height and position of the barrier is illustrated in Figure 5. Such “3D” More O’Ferrall-Jencks diagrams have been used in non-PCET contexts for some time.28,29,85–88 Figure 5A depicts the pathway for an HAT reaction in which ET and PT are equally unfavorable and the corresponding “synchronous” transition state. Figure 5B then shows the shift of the saddle point when ΔG°PT >> ΔG°ET: lower and closer to the ET corner of the More O’Ferrall-Jencks diagram. Taken together with the observation that the present reactions violate the reactivity-selectivity principle, we might then say that polar effects are insensitive to the position of the transition state along the HAT reaction coordinate, but sensitive to its position along the perpendicular coordinate (the off-diagonal). However, we emphasize that the “ET” and “PT” reaction coordinates in Figure 5 are not the positions of the quantum mechanical e– and H+, as discussed in section II.C.4 below.
Figure 5.

Pictorial schematic of the relationship between the energies of the off-diagonal species (ΔG°ET and ΔG°PT) and the position and height of the HAT transition state, when ΔG°PT = ΔG°ET (A) and when ΔG°PT >> ΔG°ET (B). Red arrows highlight the changes from (A) to (B), and gray curves in (B) represent the corresponding black curves in (A).
3). Comparison of ΔG°PT/ΔG°ET with Hammett treatments.
Polar effects have historically been probed and understood through Hammett correlations (eq 13), as discussed above. This treatment relates changes in rate constants to standard substituent constants σ typically determined from quite different reactions. The use of standard constants allows comparisons of Hammett ρ values across a wide range of transformations. For the toluene HAT reactions discussed here, the ρ values are −0.8 for tBuO• and tBuOO• and +0.5 for tBu•. The negative ρ values for the oxyl radical reactions are interpreted as buildup of positive charge at the HAT transition state, in contrast to the negative charge buildup for the tBu• radical reactions (ρ > 0). These values are one of the traditional indications of a significant polar effect, that the oxyl radicals are electrophilic while tBu• is nucleophilic.
| (13) |
As emphasized by a reviewer, the ρ values for the tBuO•, tBuOO•, and tBu• HAT reactions have large magnitudes for reactions that do not have charged intermediates. This leads to the intuition that these are quite polar reactions given that only a neutral atom is being transferred. In contrast, the BEP plots above indicate that the barriers are not very sensitive to the ΔG°PT and ΔG°ET. The origin of this difference in intuitions is the extremely large ρ values for ΔG°PT and ΔG°ET, for instance, ρ = 16.5 for the change in pKa (a 1017.3 change in Ka over for a change in σ of 1.05 (Table 1)).
In our view, the ΔG‡/ΔG° correlations (BEP or Brønsted) are valuable because they compare two things that relate to the set of reactions in question, the ΔG° between minima on the free energy surface and the transition state energies. In other words, the ΔG‡/ΔG° correlations describe relationships between features on the free energy surface of a single reaction. The Hammett analysis here is complementary; while it fundamentally compares kHAT with benzoic acid pKa values, it contextualizes the HAT reaction in the canon of organic reactions.
4). Quantum mechanical treatments of PCET.
The discussion above has been purely classical, but current PCET theories33,34 treat the proton and the electron as quantum particles. The PCET step is described in a Marcus-like model, where the [classical] heavy atoms and solvent rearrange to allow simultaneous electron and proton tunneling to move the system from the reactant to the product vibronic free energy surface. In this model, the surface hopping must involve synchronous ET and PT (though recent results suggest a 24 fs lag in one system89).
The quantum nature of the proton and electron make the use of More O’Ferrall-Jencks diagrams conceptually complicated. The off-diagonal axes cannot represent the physical motion of the particles. In Marcus theory, the barrier to ET is in large part the reorganization of the solvent to enable the hopping of the electron. Such charge separation and solvent involvement does not appear to be involved in HAT, at least from hydrocarbons. Ingold et al. showed that rate constants for H-abstraction from cyclohexane by cumyloxyl radical were completely independent of solvent from cyclohexane to acetic acid (±10%), a large range of static dielectric constants.90,91 More generally, there is typically a close correspondence of hydrocarbon radical chemistry in gas-phase and in solution. In those cases, invoking asynchrony that involves substantial buildup of charge seems unlikely—although polar effects have been explained by invoking dipolar interactions at the transition state.3
Even though the e– and H+ do not have classical positions, versions of More O’Ferrall-Jencks diagrams for PCET can be envisioned where the off-diagonal corners represent collective motions of heavy atoms that facilitate PT or ET. In the Marcus theory treatment of electron transfer, the reaction coordinate is usually referred to as a solvent coordinate, describing the collective motion of solvent molecules to allow the movement of charge.92 The Kiefer-Hynes theory of proton transfer similarly emphasizes the solvent reorganization due to charge movement.93 HAT reactions, however, likely have less charge transfer because formally a neutral atom is being transferred. Therefore, for HAT the important heavy atom motions probably emphasize inner-sphere reorganizations within the donor and acceptor molecules.92 In the reactions examined here, this could be the rehybridization of the benzylic carbon or collective changes in the C–C bond lengths with changes in unpaired spin density. If such collective nuclear motions responded differently to changes in PT vs. ET thermochemistry, this could be a source of “imbalance” despite the simultaneous double tunneling of the e– and H+.
CONCLUSIONS AND IMPLICATIONS
H-atom transfer (HAT) rate constants from substituted toluenes to tert-butyl, tert-butoxy, and tert-butylperoxyl radicals have long been known to show polar effects. The electrophilic oxyl radicals abstract H• faster from electron-rich toluenes, while the tert-butyl radical reacts faster with electron-poor toluenes. This report analyzes the relative rate constants in terms of the free energies of the highly unfavorable initial electron- or proton-transfer steps from the same reactants (ΔG°PT, ΔG°ET). While the primary determinant of such rate constants are the relevant bond dissociation free energies (BDFEs) and free energies of reaction (ΔG°HAT), these do not vary significantly for substituted toluenes. Therefore, this system provides a good test of other effects on the HAT barriers.
The most striking observation is that changes in PT or ET energetics of ~17 kcal mol−1 shift ΔG‡HAT by little more than 1 kcal mol−1. Polar effects here are thus an order of magnitude smaller than the typical BEP effect of changing the overall ΔG°HAT of HAT reactions. Strongly polar radicals are needed to achieve mildly contra-thermodynamic selectivity.
Nevertheless, the variations of ΔG‡HAT in each series correlate linearly with the changes in ΔG°PT or ΔG°ET (at essentially constant ΔG°HAT). This connects polar effects and ‘imbalanced’ or ‘asynchronous’ HAT reactions. Moreover, the reaction barriers within each set of reactions correlate inversely with |ΔG°PT – ΔG°ET|, with the largest barriers for the smallest ‘asynchrony.’ This is as predicted by Srnec and coworkers,17 though the magnitude of the effect is significantly smaller than their first-order analysis.
For all three radicals, the plots of ΔG‡HAT vs. ΔG°PT or ΔG°ET have the roughly the same magnitude of slope. The same behavior is found for the highly exoergic reactions of tBuO• and the endoergic reactions of tBuOO•. This is a violation of the reactivity-selectivity principle: the traditional metric of the position of the transition state has no bearing on polar effects in these reactions. These may be general properties of reaction sets where the barrier varies but the driving force does not.
These results suggest that polar effects in an HAT reaction may be predicted from the relative thermodynamics of single ET and PT steps, though the origin(s) of these correlations are not fully resolved. The pKa and E° parameters are complimentary to traditional predictors such as Hammett parameters and electrophilicity indices, with the advantage that these thermodynamic parameters have an independent meaning. Furthermore, the ΔG°ET/ΔG°PT lens brings VB theory, multidimensional Marcus theory, and Bernasconi’s PNS to bear on polar effects in a concise and tractable way. Further studies are needed to explore these effects across reaction sets where the ΔG°HAT or ΔG°PCET are not constant.
Polar effects are one of a number of smaller effects that modulate HAT barriers—small compared to changing the ΔG°HAT. Radical reactivity can also be modulated with steric differences, with the differing reactivities of C–H vs. O/N–H, for allylic vs. saturated bonds,32 and by introduction of hydrogen-bonding additives.94 Though these are small effects, they can have a significant bearing on reaction selectivity. After all, the difference between 70:30 vs. 30:70 selectivity is just 1 kcal mol−1 in barrier height.
Supplementary Material
ACKNOWLEDGMENT
We are grateful for valuable discussions with Professors Massimo Bietti, Gino DiLabio, and Martin Srnec, and with Dr. Scott Coste and others in the Mayer laboratory.
FUNDING
This work was supported by the U.S. National Institutes of Health (2R01GM50422 and 1R35GM144105 to J.M.M.). B.K.K. was supported by an NSF Graduate Fellowship during part of this work. B.D.G. acknowledges additional support from the Yale Berson and Dox Fellowships.
Footnotes
- Alternative linear free energy relationships: log(kHAT) vs. pKa, E°, BDFE, and Hammett substituent constants
- Analysis of individual tBu• and tBuO• data sets for comparison
- Discussion on the origin and estimation of the tabulated thermochemical parameters
- Alternative formulations of Figure 4B incorporating ΔG° and Srnec’s frustration parameter “σ”25
- Derivation of eq 11
- Additional representations of data discussed in the main text, including tabulated ΔG°PT and ΔG°ET values, numbering of substrates, and a plot of electrophilicity indices of selected substrates
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.
REFERENCES
- 1.Ruffoni A; Mykura RC; Bietti M; Leonori D, The interplay of polar effects in controlling the selectivity of radical reactions. Nat. Synth 2022, 1, 682–695. [Google Scholar]
- 2.Roberts BP, Polarity-reversal catalysis of hydrogen-atom abstraction reactions: concepts and applications in organic chemistry. Chem. Soc. Rev 1999, 28, 25–35. [Google Scholar]
- 3.Tedder JM, Which Factors Determine the Reactivity and Regioselectivity of Free Radical Substitution and Addition Reactions? Angew. Chem. Int. Ed. Engl 1982, 21, 401–410. [Google Scholar]
- 4.Salamone M; Basili F; Bietti M, Reactivity and selectivity patterns in hydrogen atom transfer from amino acid C-H bonds to the cumyloxyl radical: polar effects as a rationale for the preferential reaction at proline residues. J. Org. Chem 2015, 80, 3643–50. [DOI] [PubMed] [Google Scholar]
- 5.Watts ZI; Easton CJ, Peculiar stability of amino acids and peptides from a radical perspective. J. Am. Chem. Soc 2009, 131, 11323–5. [DOI] [PubMed] [Google Scholar]
- 6.O’Reilly RJ; Chan B; Taylor MS; Ivanic S; Bacskay GB; Easton CJ; Radom L, Hydrogen abstraction by chlorine atom from amino acids: remarkable influence of polar effects on regioselectivity. J. Am. Chem. Soc 2011, 133, 16553–9. [DOI] [PubMed] [Google Scholar]
- 7.Flemming I, Radical Reactions. In Frontier Orbitals and Organic Chemical Reactions, Wiley, Ed. Wiley: 1976; pp 182–207. [Google Scholar]
- 8.For a succinct discussion on these explanations and their histories, see p. 2156 of the following reference.
- 9.Roberts BP; Steel AJ, An extended form of the Evans–Polanyi equation: a simple empirical relationship for the prediction of activation energies for hydrogen-atom transfer reactions. J. Chem. Soc., Perkin Trans 2 1994, 2155–2162. [Google Scholar]
- 10.Johnston HS; Parr C, Activation Energies from Bond Energies. I. Hydrogen Transfer Reactions. J. Am. Chem. Soc 1963, 85, 2544–2551. [Google Scholar]
- 11.Zavitsas AA, Activation energy requirements in hydrogen abstractions. Quantitative description of the causes in terms of bond energies and infrared frequencies. J. Am. Chem. Soc 1972, 94, 2779–2789. [Google Scholar]
- 12.Zavitsas AA; Melikian AA, Hydrogen abstractions by free radicals. Factors controlling reactivity. J. Am. Chem. Soc 1975, 97, 2757–2763. [Google Scholar]
- 13.Zavitsas AA; Chatgilialoglu C, Energies of Activation. The Paradigm of Hydrogen Abstractions by Radicals. J. Am. Chem. Soc 1995, 117, 10645–10654. [Google Scholar]
- 14.Yosca TH; Rittle J; Krest CM; Onderko EL; Silakov A; Calixto JC; Behan RK; Green MT, Iron(IV)hydroxide pKa and the role of thiolate ligation in C-H bond activation by cytochrome P450. Science 2013, 342, 825–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Darcy JW; Koronkiewicz B; Parada GA; Mayer JM, A Continuum of Proton-Coupled Electron Transfer Reactivity. Acc. Chem. Res 2018, 51, 2391–2399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Barman SK; Yang MY; Parsell TH; Green MT; Borovik AS, Semiempirical method for examining asynchronicity in metal-oxido-mediated C-H bond activation. Proc. Natl. Acad. Sci. U.S.A 2021, 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Bim D; Maldonado-Dominguez M; Rulisek L; Srnec M, Beyond the classical thermodynamic contributions to hydrogen atom abstraction reactivity. Proc. Natl. Acad. Sci. U.S.A 2018, 115, E10287–E10294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Darcy JW; Kolmar SS; Mayer JM, Transition State Asymmetry in C-H Bond Cleavage by Proton-Coupled Electron Transfer. J. Am. Chem. Soc 2019, 141, 10777–10787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Goetz MK; Anderson JS, Experimental Evidence for pKa-Driven Asynchronicity in C-H Activation by a Terminal Co(III)-Oxo Complex. J. Am. Chem. Soc 2019, 141, 4051–4062. [DOI] [PubMed] [Google Scholar]
- 20.Kotani H; Shimomura H; Ikeda K; Ishizuka T; Shiota Y; Yoshizawa K; Kojima T, Mechanistic Insight into Concerted Proton-Electron Transfer of a Ru(IV)-Oxo Complex: A Possible Oxidative Asynchronicity. J. Am. Chem. Soc 2020, 142, 16982–16989. [DOI] [PubMed] [Google Scholar]
- 21.Schneider JE; Goetz MK; Anderson JS, Statistical analysis of C-H activation by oxo complexes supports diverse thermodynamic control over reactivity. Chem. Sci 2021, 12, 4173–4183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Usharani D; Lacy DC; Borovik AS; Shaik S, Dichotomous hydrogen atom transfer vs proton-coupled electron transfer during activation of X-H bonds (X = C, N, O) by nonheme iron-oxo complexes of variable basicity. J. Am. Chem. Soc 2013, 135, 17090–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Mandal M; Elwell CE; Bouchey CJ; Zerk TJ; Tolman WB; Cramer CJ, Mechanisms for Hydrogen-Atom Abstraction by Mononuclear Copper(III) Cores: Hydrogen-Atom Transfer or Concerted Proton-Coupled Electron Transfer? J. Am. Chem. Soc 2019, 141, 17236–17244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Barman SK; Jones JR; Sun C; Hill EA; Ziller JW; Borovik AS, Regulating the Basicity of Metal-Oxido Complexes with a Single Hydrogen Bond and Its Effect on C-H Bond Cleavage. J. Am. Chem. Soc 2019, 141, 11142–11150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Maldonado-Dominguez M; Srnec M, H-Atom Abstraction Reactivity through the Lens of Asynchronicity and Frustration with Their Counteracting Effects on Barriers. Inorg. Chem 2022, 61, 18811–18822. [DOI] [PubMed] [Google Scholar]
- 26.Ma S; Wang S; Cao J; Liu F, Rapid and Accurate Estimation of Activation Free Energy in Hydrogen Atom Transfer-Based C-H Activation Reactions: From Empirical Model to Artificial Neural Networks. ACS Omega 2022, 7, 34858–34867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Liu F; Ma S; Lu Z; Nangia A; Duan M; Yu Y; Xu G; Mei Y; Bietti M; Houk KN, Hydrogen Abstraction by Alkoxyl Radicals: Computational Studies of Thermodynamic and Polarity Effects on Reactivities and Selectivities. J. Am. Chem. Soc 2022, 144, 6802–6812. [DOI] [PubMed] [Google Scholar]
- 28.Jencks WP, General acid-base catalysis of complex reactions in water. Chem. Rev 1972, 72, 705–718. [Google Scholar]
- 29.O’Ferrall RAM, Relationships between E2 and E1cB mechanisms of β-elimination. J. Chem. Soc. B 1970, 274–277. [Google Scholar]
- 30.Bernasconi CF, The principle of nonperfect synchronization: more than a qualitative concept? Acc. Chem. Res 1992, 25, 9–16. [Google Scholar]
- 31.Bernasconi CF, Intrinsic barriers of reactions and the principle of nonperfect synchronization. Acc. Chem. Res 1987, 20, 301–308. [Google Scholar]
- 32.Salamone M; Galeotti M; Romero-Montalvo E; van Santen JA; Groff BD; Mayer JM; DiLabio GA; Bietti M, Bimodal Evans-Polanyi Relationships in Hydrogen Atom Transfer from C(sp3)-H Bonds to the Cumyloxyl Radical. A Combined Time-Resolved Kinetic and Computational Study. J. Am. Chem. Soc 2021, 143, 11759–11776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Hammes-Schiffer S; Stuchebrukhov AA, Theory of coupled electron and proton transfer reactions. Chem. Rev 2010, 110, 6939–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Warburton RE; Soudackov AV; Hammes-Schiffer S, Theoretical Modeling of Electrochemical Proton-Coupled Electron Transfer. Chem. Rev 2022, 122, 10599–10650. [DOI] [PubMed] [Google Scholar]
- 35.Anslyn EV; Dougherty DA, Experiments Related to Thermodynamics and Kinetics. In Modern Physical Organic Chemistry, University Science: Sausalito, CA, 2006; pp 421–487. [Google Scholar]
- 36.Bell RP, The Brönsted Equation—Its First Half-Century. In Correlation Analysis in Chemistry: Recent Advances, Chapman NB; Shorter J, Eds. Plenum: New York, 1978; pp 55–84. [Google Scholar]
- 37.Brönsted JN; Pedersen K, Die katalytische Zersetzung des Nitramids und ihre physikalisch-chemische Bedeutung. Z. Phys. Chem 1924, 108U, 185–235. [Google Scholar]
- 38.Evans M; Polanyi M, Further considerations on the thermodynamics of chemical equilibria and reaction rates. Trans. Faraday Soc 1936, 32, 1333–1360. [Google Scholar]
- 39.Evans M; Polanyi M, On the introduction of thermodynamic variables into reaction kinetics. Trans. Faraday Soc 1937, 33, 448–452. [Google Scholar]
- 40.Evans M; Polanyi M, Inertia and driving force of chemical reactions. Trans. Faraday Soc 1938, 34, 11–24. [Google Scholar]
- 41.Lowry TH; Richardson KS, Linear Free-Energy Relationships. In Mechanism and Theory in Organic Chemistry, Harper & Row: 1976; pp 60–70. [Google Scholar]
- 42.Pross A, Bell-Evans-Polanyi Principle. In Theoretical and physical principles of organic reactivity, Wiley: New York, 1995; pp 139–142. [Google Scholar]
- 43.However, this does not always hold true, as suggested by the aforementioned examples of additional factors that affect the barrier, and therefore also the α. A particularly salient example of this is the classic “nitroalkane anomaly,” wherein Brønsted α’s outside of the 0 to 1 range have been observed.
- 44.Kresge AJ, Deviant Broensted relations. J. Am. Chem. Soc 1970, 92, 3210–3211. [Google Scholar]
- 45.Agarwal RG; Coste SC; Groff BD; Heuer AM; Noh H; Parada GA; Wise CF; Nichols EM; Warren JJ; Mayer JM, Free Energies of Proton-Coupled Electron Transfer Reagents and Their Applications. Chem. Rev 2022, 122, 1–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Davis WH; Pryor WA, Polar effects in radical reactions. 7. Positive ρ values for the reactions of isopropyl and tert-butyl radicals with substituted toluenes. J. Am. Chem. Soc 1977, 99, 6365–6372. [Google Scholar]
- 47.Lomas JS; Briand S; Fain D, Reactions of thermally generated tert-butyl and di(tert-alkyl) ketyl radicals in toluene: cage effects and hydrogen transfer. J. Org. Chem 1991, 56, 166–175. [Google Scholar]
- 48.Dütsch HR; Fischer H, Nucleophilic character of the tert-butyl radical. Absolute rate constants for the reactions with substituted toluenes. Int. J. Chem. Kinet 1982, 14, 195–200. [Google Scholar]
- 49.Pryor WA; Tang FY; Tang RH; Church DF, Chemistry of the tert-butyl radical: polar character, ρ value for reaction with toluenes, and the effect of radical polarity on the ratio of benzylic hydrogen abstraction to addition to aromatic rings. J. Am. Chem. Soc 1982, 104, 2885–2891. [Google Scholar]
- 50.Kennedy BR; Ingold KU, Reactions of Alkoxy Radicals: I. Hydrogen Atom Abstraction from Substituted Toluenes. Can. J. Chem 1966, 44, 2381–2385. [Google Scholar]
- 51.Walling C; McGuinness JA, Positive halogen compounds. XVI. Comparison of alkoxy radicals from different sources and the role of halogen atoms in hypohalite reactions. J. Am. Chem. Soc 1969, 91, 2053–2058. [Google Scholar]
- 52.Paul H; Small RD; Scaiano JC, Hydrogen abstraction by tert-butoxy radicals. A laser photolysis and electron spin resonance study. J. Am. Chem. Soc 1978, 100, 4520–4527. [Google Scholar]
- 53.Walling C; Jacknow BB, Positive Halogen Compounds. II. Radical Chlorination of Substituted Hydrocarbons with t-Butyl Hypochlorite. J. Am. Chem. Soc 1960, 82, 6113–6115. [Google Scholar]
- 54.Brook JHT, Reaction of hydrocarbons with tert-butoxy radicals. Trans. Faraday Soc 1957, 53, 327–332. [Google Scholar]
- 55.Gilliom RD; Ward BF Jr, Homolytic abstraction of benzylic hydrogen. J. Am. Chem. Soc 1965, 87, 3944–3948. [DOI] [PubMed] [Google Scholar]
- 56.Johnston K; Williams GH, Homolytic reactions of aromatic side chains. Part II. Relative rates of α-hydrogen abstraction by t-butoxy-radicals. J. Chem. Soc 1960, 1446–1452. [Google Scholar]
- 57.Sakurai H; Hosomi A, Polar and solvent effects on homolytic abstraction of benzylic hydrogen of substituted toluenes by t-butoxy radical. J. Am. Chem. Soc 1967, 89, 458–460. [Google Scholar]
- 58.Howard JA; Chenier JHB, Absolute rate constants for the reaction of tert-butylperoxy radicals with some m- and p-substituted toluenes. J. Am. Chem. Soc 1973, 95, 3054–3055. [Google Scholar]
- 59.Korcek S; Chenier JHB; Howard JA; Ingold KU, Absolute Rate Constants for Hydrocarbon Autoxidation. XXI. Activation Energies for Propagation and the Correlation of Propagation Rate Constants with Carbon–Hydrogen Bond Strengths. Can. J. Chem 1972, 50, 2285–2297. [Google Scholar]
- 60.Howard JA; Ingold KU, Absolute rate constants for hydrocarbon oxidation. XI. The reactions of tertiary peroxy radicals. Can. J. Chem 1968, 46, 2655–2660. [Google Scholar]
- 61.Héberger K, Linear free energy relationships in radical reactions. II. Hydrogen abstraction from substituted toluenes by tert-Butyl, tert-Butoxyl and tert-Butylperoxyl radicals. J. Phys. Org. Chem 1994, 7, 244–250. [Google Scholar]
- 62.Hansch C; Leo A; Taft RW, A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev 1991, 91, 165–195. [Google Scholar]
- 63.Hansch C; Leo A, Substituent constants for correlation analysis in chemistry and biology. Wiley: New York, 1979. [DOI] [PubMed] [Google Scholar]
- 64.Sim BA; Griller D; Wayner DDM, Reduction potentials for substituted benzyl radicals: pKa values for the corresponding toluenes. J. Am. Chem. Soc 1989, 111, 754–755. [Google Scholar]
- 65.Bard AJ, Inner-sphere heterogeneous electrode reactions. Electrocatalysis and photocatalysis: the challenge. J. Am. Chem. Soc 2010, 132, 7559–67. [DOI] [PubMed] [Google Scholar]
- 66.Howell JO; Goncalves JM; Amatore C; Klasinc L; Wightman RM; Kochi JK, Electron transfer from aromatic hydrocarbons and their π-complexes with metals. Comparison of the standard oxidation potentials and vertical ionization potentials. J. Am. Chem. Soc 1984, 106, 3968–3976. [Google Scholar]
- 67.Linstrom PJ; Mallard WG, NIST Chemistry WebBook, NIST Standard Database Number 69. National Institute of Standards and Technology: Gaithersburg, Maryland. [Google Scholar]
- 68.Luo YR, Comprehensive Handbook of Chemical Bond Energies. CRC Press: Boca Raton, FL, 2007; p 20, 41–42, 258, 268. [Google Scholar]
- 69.Koppel IA; Koppel J; Pihl V; Leito I; Mishima M; Vlasov VM; Yagupolskii LM; Taft RW, Comparison of Brønsted acidities of neutral CH acids in gas phase and dimethyl sulfoxide. J. Chem. Soc., Perkin Trans 2 2000, 1125–1133. [Google Scholar]
- 70.Wayner DDM; McPhee DJ; Griller D, Oxidation and reduction potentials of transient free radicals. J. Am. Chem. Soc 1988, 110, 132–137. [Google Scholar]
- 71.Donkers RL; Maran F; Wayner DDM; Workentin MS, Kinetics of the Reduction of Dialkyl Peroxides. New Insights into the Dynamics of Dissociative Electron Transfer. J. Am. Chem. Soc 1999, 121, 7239–7248. [Google Scholar]
- 72.Mayer JM, Proton-coupled electron transfer: a reaction chemist’s view. Annu. Rev. Phys. Chem 2004, 55, 363–90. [DOI] [PubMed] [Google Scholar]
- 73.Warren JJ; Mayer JM, Predicting organic hydrogen atom transfer rate constants using the Marcus cross relation. Proc. Natl. Acad. Sci. U.S.A 2010, 107, 5282–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Héberger K; Lopata A, Assessment of Nucleophilicity and Electrophilicity of Radicals, and of Polar and Enthalpy Effects on Radical Addition Reactions. J. Org. Chem 1998, 63, 8646–8653. [Google Scholar]
- 75.De Vleeschouwer F; Van Speybroeck V; Waroquier M; Geerlings P; De Proft F, Electrophilicity and nucleophilicity index for radicals. Org. Lett 2007, 9, 2721–4. [DOI] [PubMed] [Google Scholar]
- 76.This has been previously suggested by others. For a discussion on the success and breakdown of the reactivity-selectivity principle, see the following reference.
- 77.Buncel E; Wilson H, The reactivity selectivity principle: Should it ever be used? J. Chem. Educ 1987, 64, 475–480. [Google Scholar]
- 78.Pross A, Qualitative Valence Bond Theory. In Theoretical and physical principles of organic reactivity, Wiley: New York, 1995; pp 83–121. [Google Scholar]
- 79.Shaik S; Wu W; Dong K; Song L; Hiberty PC, Identity Hydrogen Abstraction Reactions, X• + H−X’ → X−H + X’• (X = X’ = CH3, SiH3, GeH3, SnH3, PbH3): A Valence Bond Modeling. J. Phys. Chem. A 2001, 105, 8226–8235. [Google Scholar]
- 80.In this case, stepwise mechanisms—which might account for these differences in correlations—must be ruled out.
- 81. For a derivation of this equivalence, see Section S5 of the SI.
- 82. Since their initial publication, Srnec and coworkers have extended this model to include a new “frustration” factor (σ), which reflects the combined thermodynamic accessibility of both stepwise PCET intermediates. We have found that its inclusion makes little difference to the conclusions presented here and have therefore excluded it for the sake of clarity. See the SI for analysis of the present data including this additional factor.
- 83.Butin KP; Beletskaya IP; Kashin AN; Reutov OA, Structure and solvent effects in organomercuric electrochemistry. Polarographical scale of C–H bond acidities. J. Organomet. Chem 1967, 10, 197–210. [Google Scholar]
- 84.Russell J; Roques N, Effective nucleophilic trifluoromethylation with fluoroform and common base. Tetrahedron 1998, 54, 13771–13782. [Google Scholar]
- 85.Grunwald E, Structure-energy relations, reaction mechanism, and disparity of progress of concerted reaction events. J. Am. Chem. Soc 1985, 107, 125–133. [Google Scholar]
- 86.Thornton ER, A simple theory for predicting the effects of substituent changes on transition-state geometry. J. Am. Chem. Soc 1967, 89, 2915–2927. [Google Scholar]
- 87.Grunwald E, Reaction mechanism from structure-energy relations. 2. Acid-catalyzed addition of alcohols to formaldehyde. J. Am. Chem. Soc 1985, 107, 4715–4720. [Google Scholar]
- 88.Guthrie JP, Multidimensional Marcus Theory: An Analysis of Concerted Reactions. J. Am. Chem. Soc 1996, 118, 12878–12885. [Google Scholar]
- 89.Arsenault EA; Guerra WD; Shee J; Reyes Cruz EA; Yoneda Y; Wadsworth BL; Odella E; Urrutia MN; Kodis G; Moore GF; Head-Gordon M; Moore AL; Moore TA; Fleming GR, Concerted Electron-Nuclear Motion in Proton-Coupled Electron Transfer-Driven Grotthuss-Type Proton Translocation. J. Phys. Chem. Lett 2022, 13, 4479–4485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Avila DV; Brown CE; Ingold KU; Lusztyk J, Solvent effects on the competitive β-scission and hydrogen atom abstraction reactions of the cumyloxyl radical. Resolution of a long-standing problem. J. Am. Chem. Soc 1993, 115, 466–470. [Google Scholar]
- 91.Litwinienko G; Ingold KU, Solvent effects on the rates and mechanisms of reaction of phenols with free radicals. Acc. Chem. Res 2007, 40, 222–30. [DOI] [PubMed] [Google Scholar]
- 92.Sutin N, Theory of Electron Transfer Reactions: Insights and Hindsights. Prog. Inorg. Chem 1983, 30, 441–499. [Google Scholar]
- 93.Kiefer PM; Hynes JT, Theoretical aspects of tunneling proton transfer reactions in a polar environment. J. Phys. Org. Chem 2010, 23, 632–646. [Google Scholar]
- 94.Jeffrey JL; Terrett JA; MacMillan DW, O-H hydrogen bonding promotes H-atom transfer from alpha C-H bonds for C-alkylation of alcohols. Science 2015, 349, 1532–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.

