T pseudo-selfexchange reactions of TEMPO and related alkyl aminoxyl radicals have been found to involve significant hydrogen tunneling (as do some cross reactions), in contrast to the related reactions of aryl aminoxyl radicals.74,75 The BDFE and BDE of TEMPOH will serve as benchmarks for some of the following discussion. We have CPI-455 web recently critically evaluated the BDE and BDFE of TEMPOH in MeCN and C6H6 solvents, using both reported calorimetric measurements76 and E?and pKa data (Table 3).40 The calorimetric measurements, for diphenylhydrazine + 2 TEMPO azobenzene + 2 TEMPOH, were reinterpreted using the recently revised heat of formation of azobenzene.77 The other noteworthy redox reaction of TEMPO is its oxidation to the corresponding nitrosonium cation. The nitrosonium cation has received attention for its superoxide dismutase-type BMS-791325MedChemExpress Beclabuvir reactivity78 and catalytic alcohol oxidations,79 both of which can be described as PCET processes. In water E?TEMPO?+) = 0.74 V (vs. NHE),80,81 and in MeCN E?TEMPO?+) = 0.61 V82 (vs. SCE; better: 0.24 V vs. Cp2Fe+/0 33). Several 4-substituted derivatives of TEMPO have been investigated, including 4-oxo-, 4methoxy-, Mirogabalin web 4-amino-, and 4-hydroxy-TEMPO. Bond strengths for these and other aminoxyls in hexane have been reported by Malievskii et al. from kinetic and equilibrium measurements,83 but little acidity or redox potential data are available for these other TEMPO derivatives. As noted above, the TEMPO(?H) 1H+/1e- couple is an excellent example of a PCET reagent that favors GGTI298 chemical information concerted H?transfer over stepwise ET-PT or PT-ET pathways. TEMPOH (pKa = 41 in MeCN) is a very poor acid and TEMPO (pKa -4) is a poor base. Likewise, it is difficult to oxidize TEMPOH to TEMPOH? (Ep,a = 0.71 V vs. Cp2Fe+/0) and quite difficult to reduce TEMPO to TEMPO- (Ep,c = -1.95 V). These data indicate that under typical conditions, TEMPO- and TEMPOH?, the species at the top right and bottom left of the TEMPO square scheme (see Scheme 4), are high-energy species. The preference for concerted transfer of H?in reactions of TEMPO and TEMPOH can be illustrated by examining the energetics for the different pathways for the TEMPOH + TEMPO self exchange reaction (Scheme 7). HAT from TEMPOH to TEMPO has G?= 0 because it is a degenerate process. In MeCN, initial PT from TEMPOH to TEMPO gives TEMPO- + TEMPOH?. This reaction has an equilibrium constant of 10-45 based on the pKas of 41 and -4 respectively (Table 3), indicating a very unfavorable free energy, G T +60 kcal mol-1. Initial ET from TEMPOH to TEMPO is uphill by the same amount ( 2.7 V from the redox potentials). Note that for the unique case of a self-exchange reaction XH + X, these two values must be the same, because initial PT and ET both make the same intermediate state, XH+ + X-.84 Thus, there is a very large (60 kcal mol-1) bias favoring concerted transfer of e- and H+. The self-exchange reaction occurs readily, proceeding on the stopped flow timescale with an Eyring barrier G = 16.5 kcal mol-1 in MeCN.38,74 On this basis, the self-exchange cannot be proceeding through an intermediate state that is 60 kcal mol-1 above the ground state; the two particles must transfer together. This type of thermochemical argument, probably first applied to PCET by Meyer and coworkers,1 is quite powerful and is discussed in more detail for cross reactions in Section 6.NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptChem Rev. Author manuscript; available in PMC 2011.T pseudo-selfexchange reactions of TEMPO and related alkyl aminoxyl radicals have been found to involve significant hydrogen tunneling (as do some cross reactions), in contrast to the related reactions of aryl aminoxyl radicals.74,75 The BDFE and BDE of TEMPOH will serve as benchmarks for some of the following discussion. We have recently critically evaluated the BDE and BDFE of TEMPOH in MeCN and C6H6 solvents, using both reported calorimetric measurements76 and E?and pKa data (Table 3).40 The calorimetric measurements, for diphenylhydrazine + 2 TEMPO azobenzene + 2 TEMPOH, were reinterpreted using the recently revised heat of formation of azobenzene.77 The other noteworthy redox reaction of TEMPO is its oxidation to the corresponding nitrosonium cation. The nitrosonium cation has received attention for its superoxide dismutase-type reactivity78 and catalytic alcohol oxidations,79 both of which can be described as PCET processes. In water E?TEMPO?+) = 0.74 V (vs. NHE),80,81 and in MeCN E?TEMPO?+) = 0.61 V82 (vs. SCE; better: 0.24 V vs. Cp2Fe+/0 33). Several 4-substituted derivatives of TEMPO have been investigated, including 4-oxo-, 4methoxy-, 4-amino-, and 4-hydroxy-TEMPO. Bond strengths for these and other aminoxyls in hexane have been reported by Malievskii et al. from kinetic and equilibrium measurements,83 but little acidity or redox potential data are available for these other TEMPO derivatives. As noted above, the TEMPO(?H) 1H+/1e- couple is an excellent example of a PCET reagent that favors concerted H?transfer over stepwise ET-PT or PT-ET pathways. TEMPOH (pKa = 41 in MeCN) is a very poor acid and TEMPO (pKa -4) is a poor base. Likewise, it is difficult to oxidize TEMPOH to TEMPOH? (Ep,a = 0.71 V vs. Cp2Fe+/0) and quite difficult to reduce TEMPO to TEMPO- (Ep,c = -1.95 V). These data indicate that under typical conditions, TEMPO- and TEMPOH?, the species at the top right and bottom left of the TEMPO square scheme (see Scheme 4), are high-energy species. The preference for concerted transfer of H?in reactions of TEMPO and TEMPOH can be illustrated by examining the energetics for the different pathways for the TEMPOH + TEMPO self exchange reaction (Scheme 7). HAT from TEMPOH to TEMPO has G?= 0 because it is a degenerate process. In MeCN, initial PT from TEMPOH to TEMPO gives TEMPO- + TEMPOH?. This reaction has an equilibrium constant of 10-45 based on the pKas of 41 and -4 respectively (Table 3), indicating a very unfavorable free energy, G T +60 kcal mol-1. Initial ET from TEMPOH to TEMPO is uphill by the same amount ( 2.7 V from the redox potentials). Note that for the unique case of a self-exchange reaction XH + X, these two values must be the same, because initial PT and ET both make the same intermediate state, XH+ + X-.84 Thus, there is a very large (60 kcal mol-1) bias favoring concerted transfer of e- and H+. The self-exchange reaction occurs readily, proceeding on the stopped flow timescale with an Eyring barrier G = 16.5 kcal mol-1 in MeCN.38,74 On this basis, the self-exchange cannot be proceeding through an intermediate state that is 60 kcal mol-1 above the ground state; the two particles must transfer together. This type of thermochemical argument, probably first applied to PCET by Meyer and coworkers,1 is quite powerful and is discussed in more detail for cross reactions in Section 6.NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptChem Rev. Author manuscript; available in PMC 2011.T pseudo-selfexchange reactions of TEMPO and related alkyl aminoxyl radicals have been found to involve significant hydrogen tunneling (as do some cross reactions), in contrast to the related reactions of aryl aminoxyl radicals.74,75 The BDFE and BDE of TEMPOH will serve as benchmarks for some of the following discussion. We have recently critically evaluated the BDE and BDFE of TEMPOH in MeCN and C6H6 solvents, using both reported calorimetric measurements76 and E?and pKa data (Table 3).40 The calorimetric measurements, for diphenylhydrazine + 2 TEMPO azobenzene + 2 TEMPOH, were reinterpreted using the recently revised heat of formation of azobenzene.77 The other noteworthy redox reaction of TEMPO is its oxidation to the corresponding nitrosonium cation. The nitrosonium cation has received attention for its superoxide dismutase-type reactivity78 and catalytic alcohol oxidations,79 both of which can be described as PCET processes. In water E?TEMPO?+) = 0.74 V (vs. NHE),80,81 and in MeCN E?TEMPO?+) = 0.61 V82 (vs. SCE; better: 0.24 V vs. Cp2Fe+/0 33). Several 4-substituted derivatives of TEMPO have been investigated, including 4-oxo-, 4methoxy-, 4-amino-, and 4-hydroxy-TEMPO. Bond strengths for these and other aminoxyls in hexane have been reported by Malievskii et al. from kinetic and equilibrium measurements,83 but little acidity or redox potential data are available for these other TEMPO derivatives. As noted above, the TEMPO(?H) 1H+/1e- couple is an excellent example of a PCET reagent that favors concerted H?transfer over stepwise ET-PT or PT-ET pathways. TEMPOH (pKa = 41 in MeCN) is a very poor acid and TEMPO (pKa -4) is a poor base. Likewise, it is difficult to oxidize TEMPOH to TEMPOH? (Ep,a = 0.71 V vs. Cp2Fe+/0) and quite difficult to reduce TEMPO to TEMPO- (Ep,c = -1.95 V). These data indicate that under typical conditions, TEMPO- and TEMPOH?, the species at the top right and bottom left of the TEMPO square scheme (see Scheme 4), are high-energy species. The preference for concerted transfer of H?in reactions of TEMPO and TEMPOH can be illustrated by examining the energetics for the different pathways for the TEMPOH + TEMPO self exchange reaction (Scheme 7). HAT from TEMPOH to TEMPO has G?= 0 because it is a degenerate process. In MeCN, initial PT from TEMPOH to TEMPO gives TEMPO- + TEMPOH?. This reaction has an equilibrium constant of 10-45 based on the pKas of 41 and -4 respectively (Table 3), indicating a very unfavorable free energy, G T +60 kcal mol-1. Initial ET from TEMPOH to TEMPO is uphill by the same amount ( 2.7 V from the redox potentials). Note that for the unique case of a self-exchange reaction XH + X, these two values must be the same, because initial PT and ET both make the same intermediate state, XH+ + X-.84 Thus, there is a very large (60 kcal mol-1) bias favoring concerted transfer of e- and H+. The self-exchange reaction occurs readily, proceeding on the stopped flow timescale with an Eyring barrier G = 16.5 kcal mol-1 in MeCN.38,74 On this basis, the self-exchange cannot be proceeding through an intermediate state that is 60 kcal mol-1 above the ground state; the two particles must transfer together. This type of thermochemical argument, probably first applied to PCET by Meyer and coworkers,1 is quite powerful and is discussed in more detail for cross reactions in Section 6.NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptChem Rev. Author manuscript; available in PMC 2011.T pseudo-selfexchange reactions of TEMPO and related alkyl aminoxyl radicals have been found to involve significant hydrogen tunneling (as do some cross reactions), in contrast to the related reactions of aryl aminoxyl radicals.74,75 The BDFE and BDE of TEMPOH will serve as benchmarks for some of the following discussion. We have recently critically evaluated the BDE and BDFE of TEMPOH in MeCN and C6H6 solvents, using both reported calorimetric measurements76 and E?and pKa data (Table 3).40 The calorimetric measurements, for diphenylhydrazine + 2 TEMPO azobenzene + 2 TEMPOH, were reinterpreted using the recently revised heat of formation of azobenzene.77 The other noteworthy redox reaction of TEMPO is its oxidation to the corresponding nitrosonium cation. The nitrosonium cation has received attention for its superoxide dismutase-type reactivity78 and catalytic alcohol oxidations,79 both of which can be described as PCET processes. In water E?TEMPO?+) = 0.74 V (vs. NHE),80,81 and in MeCN E?TEMPO?+) = 0.61 V82 (vs. SCE; better: 0.24 V vs. Cp2Fe+/0 33). Several 4-substituted derivatives of TEMPO have been investigated, including 4-oxo-, 4methoxy-, 4-amino-, and 4-hydroxy-TEMPO. Bond strengths for these and other aminoxyls in hexane have been reported by Malievskii et al. from kinetic and equilibrium measurements,83 but little acidity or redox potential data are available for these other TEMPO derivatives. As noted above, the TEMPO(?H) 1H+/1e- couple is an excellent example of a PCET reagent that favors concerted H?transfer over stepwise ET-PT or PT-ET pathways. TEMPOH (pKa = 41 in MeCN) is a very poor acid and TEMPO (pKa -4) is a poor base. Likewise, it is difficult to oxidize TEMPOH to TEMPOH? (Ep,a = 0.71 V vs. Cp2Fe+/0) and quite difficult to reduce TEMPO to TEMPO- (Ep,c = -1.95 V). These data indicate that under typical conditions, TEMPO- and TEMPOH?, the species at the top right and bottom left of the TEMPO square scheme (see Scheme 4), are high-energy species. The preference for concerted transfer of H?in reactions of TEMPO and TEMPOH can be illustrated by examining the energetics for the different pathways for the TEMPOH + TEMPO self exchange reaction (Scheme 7). HAT from TEMPOH to TEMPO has G?= 0 because it is a degenerate process. In MeCN, initial PT from TEMPOH to TEMPO gives TEMPO- + TEMPOH?. This reaction has an equilibrium constant of 10-45 based on the pKas of 41 and -4 respectively (Table 3), indicating a very unfavorable free energy, G T +60 kcal mol-1. Initial ET from TEMPOH to TEMPO is uphill by the same amount ( 2.7 V from the redox potentials). Note that for the unique case of a self-exchange reaction XH + X, these two values must be the same, because initial PT and ET both make the same intermediate state, XH+ + X-.84 Thus, there is a very large (60 kcal mol-1) bias favoring concerted transfer of e- and H+. The self-exchange reaction occurs readily, proceeding on the stopped flow timescale with an Eyring barrier G = 16.5 kcal mol-1 in MeCN.38,74 On this basis, the self-exchange cannot be proceeding through an intermediate state that is 60 kcal mol-1 above the ground state; the two particles must transfer together. This type of thermochemical argument, probably first applied to PCET by Meyer and coworkers,1 is quite powerful and is discussed in more detail for cross reactions in Section 6.NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptChem Rev. Author manuscript; available in PMC 2011.