Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Dec 2.
Published in final edited form as: Inorg Chem. 2019 Nov 11;58(23):15872–15879. doi: 10.1021/acs.inorgchem.9b02293

Carboxylate Structural Effects on the Properties and Proton Coupled Electron Transfer Reactivity of [CuO2CR]2+ Cores

Courtney E Elwell , Mukunda Mandal , Caitlin J Bouchey †,, Lawrence Que Jr , Christopher J Cramer †,*, William B Tolman ‡,*
PMCID: PMC7208659  NIHMSID: NIHMS1582412  PMID: 31710477

Abstract

A series of complexes {[NBu4][LCuII(O2CR)] (R = −C6F5, −C6H4(NO2), −C6H5, −C6H (OMe), −CH3, and −C6H2(iPr)3)} were characterized (with the complex R = −C6H4(m-Cl) having been published elsewhere (Mandal et al. J. Am. Chem. Soc. 2019, 141, 17236)). All feature N,N,N″-coordination of the supporting L2− ligand, except for the complex with R = −C6H2(iPr)3, which exhibits N,N,O- coordination. For the N,N,N″-bound complexes, redox properties, UV–vis ligand-to-metal charge transfer (LMCT) features, and rates of hydrogen atom abstraction from 2,4,6,-tri-t-butylphenol using the oxidized, formally Cu(III) compounds LCuIII(O2CR) correlated well with the electron donating nature of R as measured both experimentally and computationally. Specifically, the greater the electron donation, the lower is the energy for LMCT and the slower is the reaction rate. The results are interpreted to support an oxidatively asynchronous proton-coupled electron transfer mechanism that is sensitive to the oxidative power of the [CuIII(O2CR)]2+ core.

Graphical Abstract:

graphic file with name nihms-1582412-f0001.jpg

INTRODUCTION

With the ultimate aim of using a mechanistic understanding to inform the development of more efficient and selective oxidation catalysts, copper complexes relevant to postulated catalytic intermediates have been targeted for detailed study.1 Species with [CuO]+, [CuOH]2+, [CuO2]+, and [CuOOR]+/2+ cores have garnered particular attention because of their hypothesized role as oxidizing intermediates in synthetic systems2 and in enzymes.3 Studies of complexes with such cores aim to understand spectroscopic and redox properties and to unravel structure/reactivity relationships. For example, in previous work, we prepared formally Cu(III) complexes LCuOH (L = bis(2,6-diisopropylphenylcarboxamido)pyridine and derivatives) (Figure 1), defined their structural and spectroscopic attributes, and examined their proton-coupled electron transfer (PCET) reactivity with C–H and O–H bonds in organic substrates.4,5 Rapid rates for these reactions were rationalized by the formation of a strong O–H bond in the product, LCu(OH2), and the effects of changing the electron-donating properties of the supporting ligand L2− were evaluated through variation of substituents on the aryl rings or reducing the pyridine moiety. Also, the PCET reactions of LCuOH and [LCuO2] with a series of para-substituted phenols were compared, revealing intriguing differences in mechanisms traversed as the properties of the phenol substrates were varied.5 These differences were traced to divergent thermodynamic properties of the two cores supported by the same ligand L2−.

Figure 1.

Figure 1.

Copper complexes supported by L2− (focus of this work in box).

More recently,6 we prepared a new formally Cu(III) complex LCuO2CR (R = −C6H4(m-Cl); Figure 1) and compared its rates of reactions with 9,10-dihydroanthracene (DHA) and 2,4,6,-tri-t-butylphenol (TTBP) to those of LCuOH and LCuOOC(Me2)Ph.7,8 Differences in the rates of PCET reactions were observed (LCuOH > LCuO2CR ≫ LCuOOC(Me2)Ph), and these differences were rationalized by invoking changes in the mechanisms followed. Importantly, in the current context, the reactions of LCuO2CR (R = −C6H4(m-Cl)) were found via theory to involve a concerted, yet asynchronous PCET process, with the proton migrating to the carbonyl oxygen atom that is likely not (or only weakly) bound to the copper ion, well after the electron is transferred to the copper center. Such a process represents an example of site-separated PCET, which has attracted significant attention, in part because it is likely to be important in biological contexts.9

Intrigued by the novel mechanistic findings for LCuO2CR (R = −C6H4(m-Cl)), we sought to probe the effects of changing R. Specifically, we asked how varying R might influence the structural, spectroscopic, redox, and reactivity properties of the [CuO2CR]2+ core and its precursor [CuO2CR]+. Herein, we report the results of such studies, which provide fundamental chemical insights, including into how thermodynamic attributes influence reactivity of formally Cu(III) species.

RESULTS AND DISCUSSION

Synthesis and Characterization of Complexes.

We targeted complexes [NBu4][LCuII (O2CR)] by treating a THF solution of [NBu4][LCuIIOH]5 with 1 equiv of RCO2H in the presence of molecular sieves, a method reported previously for the case of R = −C6H4(m-Cl). The complexes [NBu4][LCuII(O2CR)] were isolated as teal/green solids in good yields (53–66%) for the R groups indicated in Scheme 1. For the case of R = CH3, the complex was prepared by a reaction of [NBu4][O2 CCH3] with LCuII(MeCN) in THF (79%).4a The indicated formulations shown in Scheme 1 were supported by electrospray ionization mass spectroscopy (ESI-MS), elemental analysis, and UV–vis and EPR spectroscopy and by analogy to X-ray crystal structures reported here for R = −CH3 and –C6H2(iPr)3 and previously for R = −C6H4(m-Cl)6.

Scheme 1.

Scheme 1.

Syntheses of Complexes

The X-ray structures of the anionic portions of the complexes [NBu4][LCuO2 CR] (R = −CH3 and − C6H2(iPr)3 are shown in Figure 2. The topology of the complex with R = −CH3 (Figure 2a) is analogous to that reported for R = −C6H4(m-Cl)6 and features N,N,N″ coordination of L2− with a monodentate carboxylate completing the approximately square planar geometry (geometry index, τ4 = 0.2).10 The second carboxylate oxygen (O2) atom interacts weakly with the copper ion in the axial position (Cu–O2 = 2.545(2) Å). On the basis of spectroscopic similarities (vide infra), we propose analogous structures for all the other carboxylate complexes, with the exception of the complex with R = −C6H2(iPr)3. For this latter case (Figure 2b), while similar carboxylate coordination is observed, the supporting ligand binds differently, with one carboxamido unit flipped to bind via its O atom (N,N,O-coordination). Such isomerization of the carboxamide arms of L2− have been reported in complexes of Ru, Ni, and Fe.11 In contrast to previously reported O-coordination of a protonated carboxamide arm in Cu(II) complexes,12 the indicated RN═C(R)–O formulation (Scheme 1) is supported by the overall charge (−1) and the short C–N (N3–C17 = 1.290(2) Å vs N1–C23 = 1.347(2) Å) and long C–O bond distances (O4–C17 = 1.305(2) Å vs O3–C23 = 1.248(2) Å). We speculate that the steric bulk of the −C6H2(iPr)3 moiety is responsible for the observed N,N,O-coordination. Finally, we note a trend in Cu–O(carboxylate) distance that correlates with the steric profiles of the R groups; as R size decreases (−C6H2(iPr)3 > −C6H4(m-Cl) > −CH3), the Cu1–O1 bond shortens(1.945(2) Å > 1.929(3) Å > 1.906(1) Å). Put differently, the Cu⋯O2 distance in the most hindered complex (R = −C6H2(iPr)3, 2.919(1) Å) is significantly longer than in the other two structures (R = −C6H4(m-Cl), 2.516(3) Å; R = −CH3, 2.545(2) Å).

Figure 2.

Figure 2.

Representations of the anionic portions of the X-ray crystal structures of (a) [NBu4][LCuO2CCH3] and (b) [NBu4][LCuO2CC6H2(iPr)3] with all nonhydrogen atoms shown as 50% thermal ellipsoids and hydrogen atoms omitted for clarity. Selected bond distance (Å) and angles (deg): [NBu4][LCuO2CCH3]: Cu–O1, 1.9452(17); Cu–O2, 2.5450(19); Cu–N1, 2.0158(18); Cu–N2,1.9400(18); Cu–N3, 2.0163(19); N1–Cu–N3, 159.59(7); N1–Cu–N2, 80.44(8); N2–Cu–O1, 172.50(8). [NBu4][LCuO2CC6H2(iPr)3]: Cu–O1, 1.9055(12), Cu–O2, 2.9190(14); Cu–N1, 1.9906(13); Cu–N2, 1.9148(14); Cu–O4, 2.0020(12); N3–C17, 1.290(2); O4–C17, 1.3049(18); N1–C23, 1.347(2); O3–C23, 1.2470(19), N1–Cu–O4, 162.12(5); N1–Cu–N2, 81.06(6); N2–Cu–O1, 170.92(6).

The UV–vis and EPR spectra of the series [NBu4][LCuII(O2CR)] corroborate their structural features and distinguish between N,N,N″- and N,N,O-coordination. While all complexes exhibit typical d–d transitions in the UV–vis region (Figure S1, Table 1), the feature for the N,N,O-complex with R = −C6H2(iPr)3 is distinctly shifted toward longer wavelength/lower energy (λmax = 673 nm; νmax =14859 cm−1) compared to the other N,N,N″-complexes (λmax = 598–619 nm; νmax = 16155–16722 cm−1), consistent with the divergent coordination mode of L2−. Further supporting this assignment, the X-band EPR spectrum for this N,N,O- complex is distinct from the quite similar spectra for all the other complexes, which are also closely analogous to previously reported Cu(II) complexes4a,b,d featuring N,N,N″-coordination of L2− (Table S2 and Figure S2; illustrative spectra for R = −C6H5 and −C6H2(iPr)3 shown in Figure 3). Thus, while all spectra feature pseudoaxial signals typical for square planar Cu(II) compounds, the gz (2.228) and Cu hyperfine (AzCu = 545 MHz) values for the complex with R = −C6H2(iPr)3 differ significantly from other complexes (avg. gz = 2.201, avg. AzCu = 576 MHz), as illustrated for the specific example R = −C6H5 in Figure 3. Most notably, the N superhyperfine pattern for the complex with R = −C6H2(iPr)3 exhibits fewer lines (~12; Figure 3b, top) than for the other complexes (~17 for R = −C6H5, Figure 3b, bottom), consistent with fewer N atoms bound to the Cu(II) ion in the former complex and the postulated N,N,O- rather than N,N,N″-coordination. We thus conclude on the basis of UV–vis and EPR spectroscopy that the solid-state structures illustrated in Figure 3 are retained in solution.

Table 1.

Selected UV–Vis Dataa

complex R λmax (nm) νmax (cm−1) ε (M−1 cm−1)
[LCuII(O2CR)]b C6F5 598 16722 630
C6H4(NO2) 600 16667 300

C6H4(a)c 609 16420 400

C6H5 612 16340 420

C6H4(OMe) 617 16207 400

CH3 619 16155 350

C6H2(iPr)3 673 14859 300
[LCuIII(O2CR)]d C6FS 506 19763 5000
670 14925 10800

866 11547 9600

C6H4(NO2) 495 20202 5500

653 15314 13200

835 11976 10600

C6H4(C1)c 491 20367 5000

650 15385 13000

830 12048 10500

C6HS 493 20284 4700

646 15480 11300

827 12092 9200

C6H4(OMe) 485 20619 4700

636 15723 11200

824 12136 8800

CH3 467 21413 5800
624 16026 13700

809 12361 9700
a

Full data provided in Figures S1 and S5.

b

Only dd features listed. Conditions: 25 °C, THF.

c

Ref 6.

d

Conditions: −80 °C, THF.

Figure 3.

Figure 3.

X-band EPR data for [NBu4][LCuII(O2CR)] (R = −C6H2(iPr)3, red, top; R = −C6H5, black, bottom). (a) Stacked spectra with dashed lines drawn to highlight differences in alignment of gz and gx features. (b) Second-derivative data for the gy region with dashed lines indicating positive peaks to highlight the difference in the number of observable AN superhyperfine lines. See Table S2 for EPR parameters derived from spectral simulations for all complexes.

Cyclic voltammograms were measured for the carboxylate complexes dissolved in THF with 0.3 M Bu4NPF6 at room temperature using a glassy carbon electrode and the ferrocene/ferrocenium (Fc/Fc+) or the decamethyl ferrocene/decamethyl ferrocenium (Fc*/Fc*+) couples as internal standards (Figure S6). With the exception of the N,N,O-bound complex with R = −C6H2(iPr)3 (which exhibits irreversible features; see the Supporting Information), all of the [NBu4][LCuII(O2CR)] systems exhibited pseudoreversible CuIII/II redox couples (ipa/ipc ratios ranging from 1.00 to 1.34), which deviated from ideality to varying extents as reflected by peak separations (ΔEp) at 100 mV/s, ranging from 0.112 to 0.173 V (Table 2). Estimated E1/2 values within the series spanned a range of 150 mV (from 0.150 to 0.298 V vs Fc/Fc+) and were correlated with the electron donating nature of the bound carboxylate (less electron donating, higher E1/2). This correlation is illustrated in a plot of E1/2 versus the pKa (in H2O) of the free acid of the carboxylate ligand (Figure 4), which shows a linear dependence and a slope indicating how E1/2 increases with lower pKa (lower basicity of the carboxylate ligand). The E1/2 values for [NBu4][LCuII(O2CR)] are also compared to those of [LCuX] (X = OH, OOC(Me2)Ph) in Table 2; the lower redox potentials for the latter compounds are consistent with greater basicity of their X moieties compared to the carboxylates.

Table 2.

Cyclic Voltammetry Data for [LCu(O2CR)] and [LCuX] (X = OH, O C(Me2)Ph)a

complex R E1/2b ipa/ipcc ΔEpc
[LCu(O2CR)] C6F5 0.298(2) 1.0 0.15
C6H4(NO2) 0.239(5) 1.2 0.17

C6H4(Cl)d 0.228(8) 1.0 0.16

C6H5 0.169(5) 1.1 0.11

C6H4(OMe) 0.151(6) 1.3 0.14

CH3 0.150(3) 1.3 0.12
[LCuOH]e −0.074
[LCuO2CMe2Ph]f −0.154
a

Unless noted otherwise, all data were acquired in THF at room temperature using 0.3 M Bu4NPF6, glassy carbon electrode.

b

V vs Fc/Fc+.

c

Listed for carboxylate complexes only; for ΔEp, units are V at scan rate = 100 mV s−1.

d

Ref 6.

e

Ref 4b.

f

Ref 7.

Figure 4.

Figure 4.

Plot of E1/2 for the [LCu(O2CR)]0/−couple (THF) vs the pKa of RCO2H (H2O), with the data points labeled with the respective R groups. The red line is a linear fit to the data: E1/2 = −0.052pKa + 0.40 (R2 = 0.91).

Characterization of [CuIIIO2CR]2+ Complexes.

Treatment of solutions of [NBu4][LCuII (O2CR)] (all R except −C6H2(iPr)3 because of its irreversible electrochemistry) in THF with [AcFc][BArF4] at −80 °C resulted in an immediate color change from pale teal to deep blue. UV–vis spectra of the product solutions were similar and exhibited intense (ε ~ 5 × 103 − 1 × 104 M−1 cm−1) features between ~500 and 1000 nm (illustrative spectra for R = −C6F5 and −C6H4(OMe) in Figure 5; all spectra overlaid in Figure S5 with λmax/νmax data listed in Table 1). Consistent with formulation of the products as the 1-electron oxidized [CuIIIO2CR]2+ complexes, (a) for the case of R = C6F5, titration experiments showed that the maximum absorbance values were attained upon addition of 1 equiv of [AcFc][BArF4] (Figure S7), and (b) for the cases of R = C6H4(Y) (Y = NO2 and OMe), the addition of 1 equiv of Fc* bleached the solutions, and the UV–vis features reformed upon subsequent addition of 1 equiv of [AcFc][BArF4], with the reversibility indicated by closely similar results after 3 repeated cycles (Figure S8). Lastly, we note that similar spectra were observed (t1/2 ~20 min) when the oxidation reactions were performed at −25 °C in 1,2-difluorobenzene (DFB, Figure S5), conditions which were used in previous studies4b,d and were needed to observe some subsequent reactions with substrates (vide infra).

Figure 5.

Figure 5.

UV–vis spectra of [LCuIII(O2CR)] (R = −C6F5 and −C6H4(OMe)) in THF at −80 °C.

On the basis of computational results reported previously for the system with R = −C6H4(m-Cl)6, we attribute the UV–vis spectral features of all the N,N,N″-complexes reported herein to varying compositions of two different ligand-to-metal charge transfer (LMCT) transitions involving the supporting ligand (L2−) and the copper center, N-aryl π → Cu d and N-amide π → Cu d. These assignments were further confirmed by performing time-dependent density functional theory (TD-DFT) calculations on individual systems (see the Supporting Information). The absorption at >800 nm (<12300 cm−1) for the [CuO2CR]2+ cores occur at uniquely high wavelengths/low energies relative to those observed for other formally Cu(III) complexes supported by L2−.4 Differences along the series were observed that correlate with the electron-donating nature of the bound carboxylate ligand (longest wavelength, highest energy for most electron-withdrawing −C6F5: 506, 670, 866 nm). The trend may be viewed graphically by plotting the energy at the peak maximum near 800 nm (νmax) vs the pKa of the free acid of the corresponding carboxylate ligand (Figure 6a) or σ p13 for the case of R = −C6H4(Y) (Y = OMe, H, m-Cl, NO2) (Figure 6b). The indicated linear correlations illustrate how the observed νmax decreases as the bound carboxylate becomes increasingly electron-withdrawing (as pKa decreases and σp increases) and the metal center becomes more electrophilic. These trends are consistent with the LMCT assignment of the UV–vis features, as greater electrophilicity of Cu corresponds to lower d orbital (LUMO) energies, resulting in lower energies (νmax) for the LMCT transitions.

Figure 6.

Figure 6.

(a) Plot of νmax of the lowest energy feature in the UV–vis spectrum of [LCuIII(O2CR)] (THF, −80 °C) vs the pKa of RCO2H (H2O). The red line is the linear fit to νmax = 242pKa + 11000, R2 =0.96. (b) Plot of νmax for the lowest energy feature in the UV–vis spectrum of LCuIII(O2CC6H4(Y)) (THF, −80 °C) vs σp for the indicated Y. The red line is the linear fit to νmax = −149σp + 12100, R2 = 0.99. Note: both σp and σm values for the case of R = m-Cl are plotted; see ref 12.

Reactivity of [CuIIIO2CR]2+ Complexes.

We examined the reactions of LCuIII(O2CR) (all R except −C6H2(iPr)3 with TTBP and DHA in order to evaluate trends in reactivity with relatively weak O–H and C–H bonds. In the case of both substrates, reactions were indicated by a decay of the UV–vis spectroscopic features associated with LCuIII(O2CR), the observation of the aryloxyl radical by EPR spectroscopy in the reaction of LCuIII(O2CC6H4(Y))6 with TTBP, and the observation of the formation of anthracene by UV–vis spectroscopy in the reactions with DHA. Second-order rate constants (k2) for the reactions of LCuIII(O2CR) with TTBP (THF, −80 °C) and DHA (DFB, −25 °C) are listed in Table 3 (see Figures S9S13 for more detailed kinetic information). A clear trend in k2 with the electron donating properties of R is seen for the reactions with TTBP, which is illustrated in a plot of log k2 vs E1/2 for the [LCuIII,II(O2CR)]0/− couple in Figure 7. A similar linear relationship is seen in a plot of log k2 vs pKa of RCO2H (Figure S15), which follows from the dependence of E1/2 on the pKa of the free acid RCO2H (Figure 4). Together, these plots show that, the more electron withdrawing R, the more oxidizing is the complex and the faster is the rate of reaction with TTBP.

Table 3.

Second-Order Rate Constants (k2) for Reactions of [LCuIII(O2CR)] with the Indicated Substratesa

R TTBPb DHAc
C6F5 1.8(4) 0.014(3)
C6H4(m-Cl)d 0.3(1) 0.11(4)
C6H4(NO2) 0.23(3) 0.0079(4)
C6H5 0.052(8) 0.05(2)
C6H4(OMe) 0.09(1) 0.015(7)
CH, 0.025(3) 0.015(9)
a

Units: M−1 s−1.

b

THF, −80 °C.

c

DFB, −25 °C.

d

Ref 6.

Figure 7.

Figure 7.

Plot of log k2 for the reaction of LCuIIIO2CR with TTBP at −80 °C in THF vs E1/2 for the [LCu(O2CR)]0/− couple. Red line is the linear fit to log k2 = 10.7E1/2− 2.98, R2 = 0.90.

To gain insight into the experimentally characterized variation in reactivity as a function of the substituents on the benzoates in LCuIIIO2CR described above, density functional modeling of the PCET reaction coordinate was undertaken at the B3LYP-D3(BJ)/basis-II/SMD(THF)//B3LYP-D3(BJ)/basis-I level of theory1416 (basis-I: 6–31G(d)17 for light atoms and SDD18 for Cu; basis-II: def2-TZVP19 basis for nonmetals and SDD for Cu) along the broken-symmetry singlet potential energy surface (see the Supporting Information for details and coordinates file). Table 4 displays computed free energies of activation (ΔG) for reactions of the indicated complexes with the substrate 2,6-di-t-butylphenol (DTBP, lacking the third t-butyl group of TTBP for computational expediency). In fair agreement with the experimentally observed trend, increased electron-withdrawing ability of R correlates with an increase in the PCET rate, as quantified by the general decrease (within the error of DFT) in the computed ΔG values (Table 4).

Table 4.

Computed ΔG and Other Parameters for the Reactions of [LCuIIIO2CR] with DTBPa

R ΔGb Cu–O1 (Å) TS Cu–O1 (Å) 〈S2c
C6FS 2.3 1.868 1.984 0.853
C6H4(NO2) 3.7 1.864 1.966 0.827
C6H4(m-Cl)d 2.9 1.859 1.9S6 0.817
C6HS 7.9 1.861 1.948 0.730
C6H4(OMe) 7.2 1.860 1.943 0.700
a

SMD(THF)/B3LYP-D3(BJ)/basis-II//B3LYP-D3(BJ)/basis-I level of theory at 193.15 K.

b

Units: kcal mol−1.

c

Calculated at the UB3LYP-D3(BJ)/basis-I level of theory.

d

Ref 6.

In a related work,6 we have demonstrated that the reaction of LCuIII(O2CC6H4(m-Cl)) with DTBP proceeds via a concerted proton-coupled electron transfer pathway having a high degree of oxidative asynchronicity, meaning that the transition-state (TS) structure for the concerted event has more electron transfer-like character. Similar asynchronicity in electron/proton transfer may also be invoked for all the N,N,N″-carboxylates described herein, with the degree of asynchronicity increasing as the electron-withdrawing ability of R gradually increases (recall that the asynchronicity is oxidative in nature, and the Cu center systematically becomes more oxidizing). Along with other influencing factors, such increasing asynchronicity in electron/proton transfer has been documented to lower the free energy barrier of a concerted PCET process,20 so increased PCET reaction rates with more electron-withdrawing R is anticipated.

Furthermore, we note that, with increasing electron withdrawing nature of R, the cleavage of the Cu–O1 bond in individual TS structures (Figure 8) becomes easier, facilitating the formation of the product CuII species. Thus, while the equatorial Cu–O1 bond length in the CuIII reagent remains similar across the series (variation of ~0.009 Å, Table 4), it varies significantly in the respective TS structures (variation of ~0.041 Å, Table 4, Figure 8). This interpretation is supported by the comparison of the computed values of the 〈S2〉 operator, which gradually increase toward unity with increasing electron-withdrawing ability of R, consistent with increased separation of the developing radical loci in the products. The significant computational challenges associated with the broken-symmetry approach in the region of the transition state suggests that the quantitative variation as a function of substituent may not be particularly accurately predicted, but the consistency of the qualitative variation with the experiment suggests that the observed variations in geometric and electronic structures as a function of substituent are likely germane to the reactivity differences.

Figure 8.

Figure 8.

Representative transition-state structure for the reaction of LCuIIIO2CC6H5 with DTBP. Hydrogen atoms and the 2,6-diisopropylphenyl groups on the L2− ligand are omitted for clarity. Selected distances (Å): Cu–O1, 1.948; Cu–O2, 2.699; O2–H,1.178; H–O3, 1.211.

It is informative to compare the trends in LCuIII(O2CR)/TTBP reactivity to similar studies of the reactions of copper–oxygen species (supported by L2− and related ligands) with phenols or DHA.4d,5 Reactions of LCuIIIOH with a series of para-substituted phenols (XArOH, X = NO2, CF3, Cl, H, Me, OMe, NMe2) showed a general trend of increasing rate with more electron rich substrates (greater electron donating substituents, lower E1/2, higher pKa), albeit with outliers for the most electron poor phenols (e.g., X = NO2).5a These and other results were interpreted to indicate that these reactions involved concerted proton-coupled electron transfer but with asynchronicity favoring ET (X = NMe2) or favoring PT (X = OMe, Me, H, Cl).5 The rates of these reactions are significantly faster than those of LCuIII(O2CR) with TTBP, but the trends are analogous, with faster rates exhibited by the most electron-poor, oxidizing reagents (XArOH or LCuIII(O2CR), respectively). In another work,4d the reactivity of [CuOH]2+ cores supported by derivatives of L2− with perturbed electron donating capabilities was probed by incorporating −NO2 groups at the para positions of the 2,6-diisopropylaryl rings (NO2L2−, electron poor) or hydrogenating the central pyridine ring to generate a piperidine donor (pipL2−, electron rich). The observed trend in rates of PCET with DHA (NO2LCuOH > LCuOH > pipLCuOH) were rationalized by the same trend in product bond dissociation enthalpies (BDEs), which in turn resulted from greater influences of the ligand changes on redox potential than on hydroxide basicity. For the reactions of LCuO2CR, the trend in rate of reaction with TTBP is also dominated by the effect of R on the redox potential, as the rate increases as E1/2 increases and the basicity of the (free) carboxylate ion decreases. However, in these reactions, the trend in rates does not correlate with the free carboxylic acid O–H BDEs (Table S3), which notably are essentially identical for R = −NO2 and − OMe (107.1 and 106.8 kcal mol−1, respectively).21 We speculate that the absence of such a thermodynamic rationale reflects asynchronicity of the PCET reaction and/or perturbation of the BDEs through interactions of the carboxylate ion or product carboxylic acid with the copper ion (effects that we have not measured).

In contrast to the clear reactivity trend for the reactions of LCuIIIO2CR with TTBP (Figure 7), no such obvious trend is apparent for the (slower) reactions with DHA (Figure 9). Particularly striking are the essentially identical values of k2 for the reactions in the cases of R = −C6F5 and −C6H4(OMe) for which the [LCuIII,II(O2CR)]0,− E1/2 differs by ~0.15 V and the order of magnitude difference in rate constants for the cases of R = −C6H4(m-Cl) and − C6H4(NO2), which have very similar E1/2 values (~10 mV difference). DFT calculations also revealed no clear reactivity trend for reactions of [LCuIIIO2CR] with DHA (TTable S7). We speculate that the absence of a clear reactivity trend in the reactions with DHA may result from a change in mechanism across the series and/or steric effects that confound interpretation of the electronic influences of R.22

Figure 9.

Figure 9.

Plot of log k2 for the reaction of [LCuIIIO2CR] with DHA at −25 °C in DFB vs E1/2 for the [LCu(O2CR)]0/−couple.

CONCLUSIONS

A series of complexes {[NBu4][LCuII(O2CR)] (R = −C6F5, −C6H4(NO2), −C6H5, −C6H4(OMe), −CH3, and − C6H2(iPr)3)} were prepared, characterized, and compared to the data for the complex with R = −C6H4(m-Cl) published elsewhere.6 All adopted structures featuring N,N,N″-coordination of the supporting L2− ligand, except for the case where R = −C6H2(iPr)3, in which the supporting ligand isomerized to N,N,O-coordination as revealed by X-ray crystallography and UV–vis and EPR spectroscopy. For all complexes except the one with R = −C6H2(iPr)3, cyclic voltammograms exhibited pseudoreversible waves with E1/2 values that correlated with the electron donating properties of the carboxylate, as indicated by a plot of E1/2 vs the pKa of the carboxylic acid (Figure 4). Chemical oxidations of these complexes at −80 °C in THF or −25 °C in DFB resulted in the formation of LCuIII(O2CR), which exhibited appropriate LMCT features in their UV–vis spectra. These features as well as the rates of the reactions of LCuIII(O2CR) with the O–H bond in TTBP correlated well with the electron donating/withdrawing nature of R, as revealed by linear fits to plots of νmax with carboxylic acid pKa and (for the benzoates) σp (Figure 6), and a logarithmic plot of the second order rate constant for the reaction with TTBP vs the [LCuIII,II(O2CR)]0,− E1/2 (Figure 7). DFT calculations corroborated the experimentally observed reactivity trend during hydrogen atom abstraction from 2,6-di-t-butylphenol by N,N,N″-complexes. A systematic increase in the reaction rate is observed with the increasing electron withdrawing nature of R, quantified by, within error, a gradual decrease in the computed ΔG values, attributable to a gradual increase in the degree of asynchronicity in electron/proton transfer. No clear trend was observed in the slower reactions with DHA, pointing to possible mechanism changes across the series and/or complicating steric effects. Taken together, the evidence supports oxidatively asynchronous proton-coupled electron transfer pathways for the reactions of LCuIII(O2CR) with the hindered phenol, with the oxidizing power of the complexes being predominant in controlling the reaction rate.

Supplementary Material

SI xyz
SI pdf

ACKNOWLEDGMENTS

W.B.T. thanks the NIH for financial support (GM47365), and M.M. and C.J.C. acknowledge funding support as part of the Inorganometallic Catalyst Design Center, an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Basic Energy Sciences, under Award No. DESC0012702. We also thank Dr. V. G. Young, Jr. for assistance with X-ray crystallography. X-ray diffraction data were collected using a crystal diffractometer acquired through NSF-MRI Award No. CHE-1229400.

Footnotes

The authors declare no competing financial interest.

Supporting Information

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.9b02293.

Experimental details, X-ray crystallographic information, UV–vis spectra overlay, EPR spectral data, individual cyclic voltammograms, monitoring the oxidation of CuII to CuIII via UV–vis, UV–vis spectra decay traces, reactivity plot, computation protocol, computed UV–vis spectra, HOMO–LUMO energies, and energetics (PDF)

Cartesian coordinates of stationary points (XYZ)

Accession Codes

CCDC 1921670 and 1921671 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

REFERENCES

  • (1).Selected reviews:; (a) Mirica LM; Ottenwaelder X; Stack TDP Structure and Spectroscopy of Copper-Dioxygen Complexes. Chem. Rev 2004, 104, 1013–1045. [DOI] [PubMed] [Google Scholar]; (b) Lewis EA; Tolman WB Reactivity of Copper-Dioxygen Systems. Chem. Rev 2004, 104, 1047–1076. [DOI] [PubMed] [Google Scholar]; (c) Elwell CE; Gagnon NL; Neisen BD; Dhar D; Spaeth AD; Yee GM; Tolman WB Copper–Oxygen Complexes Revisited: Structures, Spectroscopy, and Reactivity. Chem. Rev 2017, 117, 2059–2107. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Liu JJ; Diaz DE; Quist DA; Karlin KD Copper(I)-Dioxygen Adducts and Copper Enzyme Mechanisms. Isr. J. Chem 2016, 56, 738–755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Selected examples:; (a) Trammell R; See YY; Herrmann AT; Xie N; Diaz DE; Siegler MA; Baran PS; Garcia-Bosch I Decoding the Mechanism of Intramolecular Cu-Directed Hydroxylation of sp3 C-H Bonds. J. Org. Chem 2017, 82, 7887–7904. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Trammell R; D’Amore L; Cordova A; Polunin P; Xie N; Siegler MA; Belanzoni P; Swart M; Garcia-Bosch I Directed Hydroxylation of sp2 and sp3 C–H Bonds Using Stoichiometric Amounts of Cu and H2O2. Inorg. Chem 2019, 58, 7584–7592. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ryland BL; Stahl SS Practical Aerobic Oxidations of Alcohols and Amines with Homogeneous Copper/TEMPO and Related Catalyst Systems. Angew. Chem., Int. Ed 2014, 53, 8824–8838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).(a) Solomon EI; Heppner DE; Johnston EM; Ginsbach JW; Cirera J; Qayyum M; Kieber-Emmons MT; Kjaergaard CH; Hadt RG; Tian L Copper Active Sites in Biology. Chem. Rev 2014, 114, 3659–3853. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Ciano L; Davies GJ; Tolman WB; Walton PH Bracing copper for the catalytic oxidation of C–H bonds. Nature Catal. 2018, 1, 571–577. [Google Scholar]; (c) Quist DA; Diaz DE; Liu JJ; Karlin KD Activation of dioxygen by copper metalloproteins and insights from model complexes. JBIC, J. Biol. Inorg. Chem 2017, 22, 253–288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (4).(a) Donoghue PJ; Tehranchi J; Cramer CJ; Sarangi R; Solomon EI; Tolman WB Rapid C–H Bond Activation by a Monocopper(III)–Hydroxide Complex. J. Am. Chem. Soc 2011, 133, 17602–17605. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Dhar D; Tolman WB Hydrogen Atom Abstraction from Hydrocarbons by a Copper(III)-Hydroxide Complex. J. Am. Chem. Soc 2015, 137, 1322–1329. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Gagnon N; Tolman WB [CuO]+ and [CuOH]2+ complexes: intermediates in oxidation catalysis? Acc. Chem. Res 2015, 48, 2126–2131. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Dhar D; Yee GM; Spaeth AD; Boyce DW; Zhang H; Dereli B; Cramer CJ; Tolman WB Perturbing the Copper(III)–Hydroxide Unit through Ligand Structural Variation. J. Am. Chem. Soc 2016, 138, 356–368. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Spaeth AD; Gagnon NL; Dhar D; Yee GM; Tolman WB Determination of the Cu(III)-OH Bond Distance by Resonance Raman Spectroscopy Using a Normalized Version of Badger’s Rule. J. Am. Chem. Soc 2017, 139, 4477–4485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (5).(a) Dhar D; Yee GM; Markle TF; Mayer JM; Tolman WB Reactivity of the copper(III)-hydroxide unit with phenols. Chem. Sci 2017, 8, 1075–1085. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Bailey WD; Dhar D; Cramblitt AC; Tolman WB Mechanistic Dichotomy in Proton-Coupled Electron-Transfer Reactions of Phenols with a Copper Superoxide Complex. J. Am. Chem. Soc 2019, 141, 5470–5480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).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]
  • (7).Neisen BD; Gagnon NL; Dhar D; Spaeth AD; Tolman WB Formally Copper(III)–Alkylperoxo Complexes as Models of Possible Intermediates in Monooxygenase Enzymes. J. Am. Chem. Soc 2017, 139, 10220–10223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (8).Analogous Ni-carboxylate species have been reported:; Pirovano P; Farquhar ER; Swart M; McDonald AR Tuning the Reactivity of Terminal Nickel(III)-Oxygen Adducts for C-H Bond Activation. J. Am. Chem. Soc 2016, 138, 14362–14370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (9).(a) Weinberg DR; Gagliardi CJ; Hull JF; Murphy CF; Kent CA; Westlake BC; Paul A; Ess DH; McCafferty DG; Meyer TJ Proton-coupled electron transfer. Chem. Rev 2012, 112, 4016–93. [DOI] [PubMed] [Google Scholar]; (b) Mayer JM Proton-coupled electron transfer: a reaction chemist’s view. Annu. Rev. Phys. Chem 2004, 55, 363–390. [DOI] [PubMed] [Google Scholar]; (c) Warren JJ; Tronic TA; Mayer JM Thermochemistry of Proton-Coupled Electron Transfer Reagents and its Implications. Chem. Rev 2010, 110, 6961–7001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (10).Yang L; Powell D; Houser R Structural variation in copper(I) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans 2007, 955–964. [DOI] [PubMed] [Google Scholar]
  • (11).(a) Wasilke J-C; Wu G; Bu X; Kehr G; Erker G Ruthenium Carbene Complexes Featuring a Tridentate Pincer-type Ligand. Organometallics 2005, 24, 4289–4297. [Google Scholar]; (b) Harrop TC; Tyler LA; Olmstead MM; Mascharak PK Change in the Spin State of an FeIII Center upon One N-to-O Switch in the Coordination of a 2,6-Pyridinedicarboxamido Unit: The Effect of Methyl Thioether and Methyl Ether Appendages at the Ligand Periphery. Eur. J. Inorg. Chem 2003, 2003, 475–481. [Google Scholar]
  • (12).Boyce DW; Salmon DJ; Tolman WB Linkage Isomerism in Transition-Metal Complexes of Mixed (Arylcarboxamido)-(arylimino)pyridine Ligands. Inorg. Chem 2014, 53, 5788–5796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).Because in the case of R = m-Cl the substituent occupies the meta rather than the para position, including this example in the plot is, strictly speaking, not appropriate. However, because the σp and σm values are similar (0.23 vs 0.37), we choose to include data points for both for this case in order to approximate its position in the Hammett plot.
  • (14).(a) Becke AD Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys 1993, 98, 5648–5652. [Google Scholar]; (b) Becke AD Density-functional Thermochemistry. IV. A New Dynamical Correlation Functional and Implications for Exact-exchange Mixing. J. Chem. Phys 1996, 104, 1040–1046. [Google Scholar]
  • (15).(a) 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]; (b) Johnson ER; Becke AD A Post-Hartree-Fock Model of Intermolecular Interactions: Inclusion of Higher-Order Corrections. J. Chem. Phys 2006, 124, 174104. [DOI] [PubMed] [Google Scholar]
  • (16).Marenich AV; Cramer CJ; Truhlar DG Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [DOI] [PubMed] [Google Scholar]
  • (17).Hehre WJ; Ditchfield R; Pople JA Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys 1972, 56, 2257–2261. [Google Scholar]
  • (18).Dolg M; Wedig U; Stoll H; Preuss H Energy-adjusted Ab Initio Pseudopotentials for the First Row Transition Elements. J. Chem. Phys 1987, 86, 866–872. [Google Scholar]
  • (19).Weigend F; Ahlrichs R Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys 2005, 7, 3297–3305. [DOI] [PubMed] [Google Scholar]
  • (20).Bím D; Maldonado-Domínguez M; Rulíšek 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]
  • (21).Tumanov VE; Denisov ET Estimation of the Dissociation Energies of O–O, C–O, and O–H Bonds in Acyl Peroxides, Acids, and Esters from Kinetic Data on the Degradation of Diacyl Peroxides. Pet. Chem 2005, 45, 237–248. [Google Scholar]
  • (22).To test the idea of steric effects, we measured the rate of reaction of the complex with R = −C6H4(m-Cl) with 1,4cyclohexadiene. More complicated kinetics than those found for the reaction with DHA were observed (an apparent induction period), but the subsequent second-order rate constant was ~5 × 10−2 M−1s−1, about 2 times smaller than the value for the reaction with DHA (see details in the Supporting Information Section 6, Figure S14). Simple steric arguments do not seem applicable, and more detailed studies would be necessary to understand these rate differences.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

SI xyz
SI pdf

RESOURCES