Abstract
Two series of RuII(acac)2(py-imH) complexes have been prepared, one with changes to the acac ligands and the other with substitutions to the imidazole. The PCET thermochemistry of the complexes has been studied in acetonitrile, revealing that the acac substitutions almost exclusively affect the redox potentials of the complex (|ΔE1/2| >> |ΔpKa|•0.059 V), while the changes to the imidazole primarily affect its acidity (|ΔpKa|•0.059 V ≫ |ΔE1/2|). This decoupling is supported by DFT calculations, which show that the acac substitutions primarily affect the Ru-centered t2g orbitals, while changes to the py-imH ligand primarily affect the ligand-centered π-orbitals. More broadly, the decoupling stems from the physical separation of the electron and proton within the complex and highlights a clear design strategy to separately tune the redox and acid/basic properties of H atom donor/acceptor molecules.
Graphical Abstract
Two series of bis(acetylacetonate) ruthenium(II) pyridine-imidazole complexes have been prepared that show independent modulation of the imidazole pKa and the ruthenium reduction potential E1/2. This unusual decoupling of the pKa and E1/2 values are accomplished using different substitution patterns. The decoupling stems from differences in the orbitals affected by each substitution pattern and highlights a promising strategy to tune the individual components of PCET systems.
Introduction
Many chemical species react by proton-coupled electron transfer (PCET) and hydrogen atom transfer (HAT).1, 2 This includes free radical intermediates, metal complexes, and active sites in enzymes and on surfaces. Such reactions are important in a variety of societal, biological, and industrial processes, including combustion, oxidations in the atmosphere and natural waters, the trapping of reactive oxygen species in cells, and complex organic syntheses.1, 3–5
The reactivity of a PCET reagent X/XH rests on its thermochemical properties, described by a square scheme (Scheme 1). The diagonal of the Scheme is the X-H bond dissociation free energy (BDFE), and the free energy of an HAT reaction is simply the difference between the reactant and product BDFEs (eq 1). Around the outside of the square are the thermochemistry of the proton transfer (PT) and electron transfer (ET) components, the pKa and the reduction potential E°.
Scheme 1.
Square scheme for an XH/X PCET reagent and a drawing of the class of ruthenium complexes examined here.
(1) |
Across a series of compounds, it has long been known that the BDFE is much less sensitive to changes in substituents than the pKa and E° values. For para-substituted toluenes, for example, the BDFEs are almost identical with electron-withdrawing or donating substituents. A p-cyano group increases the acidity of toluene by 10-11 units and renders the incipient anion ~660 mV harder to oxidize. Yet the BDFEs for PhCH2–H and p-CNPhCH2–H differ only by ~1 kcal mol−1.6 For direct comparison, the 0.66 V variation in E° and the shift in pKa both correspond to a change in ΔG° of 15 kcal mol−1 (ΔΔG° = FΔE°, and ΔΔG° = 2.303RTΔpKa).
This thermodynamic compensation between the pKa and E° arises because an electron withdrawing group, for instance, makes a molecule more acidic (easier to lose H+) but also less reducing (harder to lose e−). Most systems are not as closely compensated as toluenes, as summarized by Tolman7 and discussed at greater length elsewhere.1 Interestingly, the change in E° is nearly always greater than that in pKa, perhaps due to the greater extent of delocalization of the electron that participates in the PCET reaction than that of the proton. This may also be because comparative studies across a series often place the substituents far from the reactive site, in this case the proton. For instance, the series of ‘substituted toluenes’ mentioned above does not include substitution at the α position, PhCH2X.
While the equilibrium constants for PCET/HAT reactions such as eq 1 are determined by the BDFEs, the mechanisms are in large part determined by the pKa and E° values. PCET reactions can involve concerted transfer of H+ + e− (H•) or can proceed by stepwise PT-then-ET or ET-then-PT. A stepwise path is often followed when the relevant pKa or E° make an initial PT or ET favorable, because the intrinsic barriers for moving a single particle are typically smaller than for PCET.
For concerted HAT reactions, the rate constants are affected by both the ΔG° and the nature of the XH and YH bonds. At the same driving force, HAT to or from carbon atoms is substantially slower than HAT involving just OH and NH bonds.8 Additionally, the polarity of H• donors and acceptors can significantly affect rate constants. An electrophilic H• acceptor, for example, will show selectivity towards more hydridic X–H bonds in the presence of acidic X–H bonds of similar or a bit greater strength. This has been coined the ‘polar-matching effect’ in the organic free-radical literature.9, 10
In the last few years, there has been increasing interest in how rate constants and intrinsic barriers for HAT reactions could be affected by the mismatch of donor and acceptor pKa and E°.11–17 Such a mismatch has been suggested to lead to an ‘asynchrony’ or ‘imbalance’ in the H+ and e− components of a concerted HAT reaction. In particular, the basicity of enzymatic and biomimetic metal oxo H• acceptors are suggested to play an important role in their selective cleavage of strong C–H bonds.11, 14, 18–20
For all of these reasons, it would be valuable to be able to independently tune the acid/base, outer-sphere ET, and homolytic bond strengths of PCET reagents. This report develops an approach to that aim, placing different substituent patterns on a ruthenium complex active for PCET. A complementary approach is to combine separate redox and acid/base cofactors across a linker, as in the anilinium-substituted cobaltocene developed by the Peters group.21, 22
We report here the synthesis, characterization, and PCET properties of five new derivatives of the ruthenium complex Ru(acac)2(py-imH) (acac = 2,4-pentanedionato, py-imH = 2-(2’-pyridyl)imidazole, Scheme 1 and Figure 1). The parent compound (no substituents, 4) and the hexafluoro-acac derivative Ru(hfac)2(py-imH) (3, hfac =1,1,1,5,5,5-hexafluoro-2,4-pentanedionato) were described by our laboratory some years ago.23 The seven compounds form two series of complexes, one with substitutions to the acac ligands and the other with substitutions instead to the imidazole moiety. In short, substitutions to the py-imH ligand change predominantly the pKa of the complex with minimal change to the redox potential, while the opposite is true for changes to the acac ligands. These series are a rare (if not the first) example of a system in which the pKa and E° of a single molecule can be tuned with a large degree of independence.
Figure 1.
Drawings of RuII(acac)2(py-imH) complexes (top row) and molecular structures of their oxidized, deprotonated forms RuIII(acac)2(py-im) from single-crystal X-ray diffraction (bottom row). These show the substitution patterns in acac-substituted (1-4) and py-imH-substituted (4-7) series. Complex 3 (not shown) is the hexafluoro-acac derivative Ru(hfac)2(py-imH) (the same as 4 but with CF3 instead of CH3 groups).
Results
i). Synthesis
The substituted RuII(acac’)2(py-imH) complexes (acac’ = acac, hfac, or the bis-tBu or bis-Ph derivatives (in 1 and 2)) were synthesized similarly to the parent complex23 with some modifications; see the Supporting Information (SI) for complete details. The appropriately substituted tris(acac’) complex was either purchased or synthesized from hydrated RuCl3. [Ru(acac’)2(MeCN)2][OTf] (OTf = triflate)24 was prepared from the tris(acac’) and serves as an excellent precursor to our complexes due to the relative lability of the MeCN ligands. The corresponding Ru(II) precursor, Ru(acac’)2(MeCN)2, was prepared directly from the tris(acac’) complex by heating in MeCN with activated zinc dust under inert conditions.25 The Ru(acac’)2(MeCN)2 complexes are indefinitely air stable as solids and for hours to days in solution (except for ready decomposition in CH2Cl2 and CHCl3). The [Ru(acac’)2(MeCN)2][OTf] complexes are even more stable, with decomposition in CH2Cl2 only occurring on the timescale of days to weeks.
The final Ru(acac’)2(py-imH) complexes were prepared by reaction of the appropriately substituted py-imH ligand either with Ru(acac’)2(MeCN)2 or with [Ru(acac’)2(MeCN)2][OTf] followed by reduction over zinc dust. The ligand exchange reactions were performed at ca. 80 °C. The RuII(acac’)2(py-imH) complexes were all air sensitive, attributed to their rapid oxidation by O2 to form Ru(acac’)2(py-im),23 and were therefore handled and stored in an N2-filled glovebox. The deprotonated Ru(III) analogs were also prepared and are air-stable (see SI for more detailed synthetic procedures).
The seven complexes studied here fall into two series (compounds in Figure 1 plus the hfac derivative 3). Compounds 1-4 differ in the groups on the 1 and 5 positions of the acac: tBu, Ph, CF3, and CH3 (parent). The series 4-7 all have acac ligands and differ in the N–N chelating ligand: 2-(2-pyridyl)imidazole (4), 2-(2-pyridyl)benzimidazole (5), and 2-(2-pyridyl)-4-trifluoromethylimidazole. The last ligand surprisingly formed two different isomers. Reaction with the RuIII precursor forms only the complex with the CF3 group on the imidazole carbon adjacent to the metal (6), while reaction with the RuII precursor places the CF3 group adjacent to the NH bond (7). These isomers do not interconvert in solution at room temperature, neither in the RuIIimH nor the RuIIIim form. RuII and RuIII complexes with the deprotonated form of the pyridyl 4,5-dinitroimidazole ligand, Ru(acac)2(py-(NO2)2im) were prepared but decomposed rapidly upon attempted protonation of the distal imidazole nitrogen (see SI Figure S1).
ii). Crystallographic characterization
X-ray crystal structures were solved for the new complexes in their air-stable RuIIIim state (Figure 1). The structures of 6 and 7 confirm that they are isomers, with the CF3 group proximal (6) or distal (7) to the ruthenium center. In 6, the Ru–N(im) distance is 0.03 Å longer than that in 7 and in all of the other complexes. The full crystallographic and metrical data for RuIIIim forms of 1, 2, 5-7, and Ru(acac)2(py-(NO2)2im) are tabulated in the SI (the structure of 4-RuIIIim was previously reported, and 3-RuIIIim is unstable23). All six structures share a very similar distorted octahedral geometry, with all trans angles >170°. The Ru–O bond lengths vary only a small amount across the series, with an overall range of 1.983(2)–2.029(3) Å and no single Ru–O bond varying more than 0.021 Å across the series. Similarly, the Ru–N distances do not vary very much, ranging from 2.039(6)–2.072(3) Å for Ru–N(py) and 1.994(6)–2.018(3) Å for Ru–N(im) (not including the 2.047(3) Å distance in 6 for the imidazole N next to a CF3 or the minor disordered component in 5).
iii). Spectroscopic characterization
Each complex was characterized by 1H NMR and UV-vis spectroscopies, and by mass spectrometry. The NMR spectra of the diamagnetic RuIIimH and paramagnetic RuIIIim complexes are all similar to the previously described spectra of 4 in these forms.23 For instance, the 1H chemical shifts in the RuIII spectra range from 18.44 to −75.79 ppm, with sharper peaks for the protons farther from the metal. The RuII spectra are sharp when free of any of the RuIII complexes, broadening apparently due to rapid exchange by electron transfer. The 1H chemical shifts of the RuIIimH complexes are also sensitive to water content, shifting up to 0.2 ppm in its presence.
The isomeric CF3-imidazole complexes 6 and 7 have close but distinct 19F NMR resonances in the RuIIimH forms (differing by 3 ppm). In contrast, the broad 19F resonances for paramagnetic RuIIIim congeners are separated by ~80 ppm.
UV-vis spectra of the RuII(imH) complexes in MeCN contain metal-to-ligand charge transfer (MLCT) bands in the visible region (Figure 2, Table 1) consistent with similar ruthenium(II) polypyridyl complexes.26–29 The optical absorbances are considerably less intense for the corresponding Ru(III) complexes.
Figure 2.
Visible region of the optical spectra of the RuIIimH forms of 1-7, showing the characteristic MLCT bands.
Table 1.
MLCT energies and corresponding extinction coefficients for the major transitions of 1-7 in the visible region.a
RuII(imH) | RuIII(im) | ||
---|---|---|---|
λmax,1 (ε)b | λmax,2 (ε)b | λmax (ε)b | |
1 | 583 (8,600) | 426 (7,900) | 495 (2,500) |
2 | 568 (13,600) | 425 (3,200)c | 655 (600)c |
3 | 519 (10,000) | 481 (9,600)c | — d |
4 | 568 (7,000) | 428 (6,700) | 486 (1,600) |
5 | 625 (6,400) | 423 (6,400) | 507 (2,300) |
6 | 577 (6,700) | 408 (6,500) | 519 (1,800) |
7 | 593 (6,700) | 420 (7,000) | 517 (1,900) |
iv). DFT Calculations: Structure, Orbital Character, and Spectroscopy
Density functional theory (DFT) calculations were performed to gain insight into the electronic changes across the series of complexes, specifically through calculation of the energies of the relevant molecular orbitals (MOs) and their localization on the Ru center, py-imH and acac ligands. The optimized structures were in good agreement with the available solid-state crystal structures, and they were similar across the family (see SI for computational details and coordinates). The calculated MO diagrams for 1-7 in the RuIIimH form are shown in Figure 3, in which each horizontal line indicates the energy of a molecular orbital, and the coloring shows the fractional contributions from Ru (black), py-imH (blue), and acac (red). The principal component of each orbital is identified in the labels at right.
Figure 3.
Computed Molecular orbital diagrams for 1-7. Percental contributions are shown by color: black (Ru), blue (py-imH), and red (acac).
The UV-vis spectra of the RuIIimH complexes were computed using TD-DFT, and the results are in excellent agreement with the experimental data (see Discussion). The lower-energy MLCT is primarily a R→py-imH transition, whereas the higher-energy MLCT is primarily Ru→acac. The substituents primarily move the manifold of orbitals together, so that the HOMO-LUMO gap is quite similar for almost all of the complexes, with the exception of 3. This is seen in the computed energies, the computed spectra, and in the experimental spectra. The most important transitions for each complex, their relative contributions from the Ru and ligands, and the corresponding electron difference density maps (EDDMs) are summarized in the SI (Table S8, Figure S87).
Significant changes in the acac substituents – compounds 2-4 – strongly affect the Ru-centered t2g orbitals and the acac π-orbitals (πacac). The hfac complex 3 is shown at the far left of Figure 3 because is it the most different. The four CF3 groups strongly lower the energies of the acac π* and π orbitals (the latter fallen so much that they are below the bottom of the plot and not shown). The hfac ligands also lead to substantially lower Ru t2g orbitals than most of the other compounds (by ca. ½ - ¾ eV). The diphenyl-acac complex 2 has similar effects but smaller in magnitude.
Substitutions to the py-imH ligand – compounds 4-7 – have little effect on the energies of the Ru or acac orbitals. The MO diagrams show differences primarily in the π and π* orbitals of the ligand. The benzimidazole complex 5 has lower π* orbitals due to its more extensive π system. The CF3 substituent on the imidazole (proximal 6 and distal 7 isomers), also lowers the py-imH π and π* orbitals, without affecting the other orbitals significantly. These trends are paralleled in the oxidized deprotonated RuIIIim MO diagrams for these complexes (Figures S74–75) and are evident in the thermochemistry discussed in the next section.
v). Thermochemical measurements: pKa, E1/2, and BDFE
The pKas of 1-7 in their RuII(py-imH) and [RuIII(py-imH)]+ forms were determined in MeCN through titrations with an acid or base of similar pKa and monitoring the equilibration by UV-vis spectroscopy (Figure 4; see SI). For complexes 1 and 4, the pKa(RuIIimH) was not accessible due to the instability of the anion. Similarly, the instability of the oxidized forms of 3 prevented measurement of its pKa(RuIIIimH+). The pKa values of all titrants were obtained from self-consistent equilibration studies,30 so the relative pKa values are all ±0.1 unit. All the measured pKa values are tabulated below in Table 2, and the individual titration data sets are provided in the SI.
Figure 4.
Overlay of dilution-corrected UV-vis spectra from the titration of 1-RuIIIimH+ with 0.05 (blue) to 100 (pink) equivalents of 2,4,6-collidine (pKa = 15.0030), and after the addition of excess strong base (dashed black line). Inset shows the linear plot of [RuIIIim][collidine–H+]/RuIIIimH+] vs. [collidine], with slope = Keq.
Table 2.
Measured thermochemical parameters for the PCET of complexes 1-7.a
E 1/2 imH | E 1/2 im | pKared | pKaox | BDFE1 b | BDFE2 b | BDFECV c | |
---|---|---|---|---|---|---|---|
1 | −0.86 | −1.23 | 23.0 | 16.8 | 55.8 | — | 55.6 |
2 | −0.57 | −0.89 | 22.3 | 16.2 | 61.6 | 62.6 | 61.6 |
3 | 0.29 | −0.07 | 19.7 | 13.6 | — | 78.0 | 76.9 |
4 | −0.64 | −1.00 | 22.2 | 16.1 | 59.9 | — | 60.0 |
5 | −0.60 | −0.96 | 21.5 | 14.8 | 59.0 | 59.9 | 59.3 |
6 | −0.57 | −0.91 | 19.0 | 12.7 | 56.9 | 57.6 | 56.9 |
7 | −0.55 | −0.88 | 16.6 | 10.3 | 54.0 | 55.0 | 54.1 |
All values measured in MeCN. E1/2 values are in V vs Fc+/0 and ± 0.01 V, pKa values are ± 1.0, and BDFEs are ± 1.0 kcal mol−1. Values in italics are calculated from the other three E1/2/pKa values using Hess’s Law.1
Cyclic voltammograms (CVs) of 1-7 in MeCN containing 100 mM [n-Bu4N][PF6] all yield chemically reversible waves corresponding to the RuIII/II redox couple. At a scan rate of 100 mV/s, anodic and cathodic peak currents were typically within 15% of each other and separated by about 60 mV, roughly the same as the ferrocene couple under the same conditions. pKa data for 1 and CV data for 6 are shown in Figures 4 and 5 as representative examples (see SI for complete details). Upon in situ deprotonation of RuII-im with a strong base, the CV waves shifted negative by 330 - 370 mV, depending on the complex. The redox potentials for the deprotonated forms were confirmed by CV of the isolated RuIII(py-im) complexes. The RuIII/II E1/2s of the protonated and deprotonated forms of each complex are given in Table 2.
Figure 5.
(A) CVs of 6-RuIIIimH+ before (blue) and after addition of excess triethylamine (pKa = 18.83;30 red) in MeCN solution containing ferrocene and [n-Bu4N][PF6]. (B) CVs of 6-RuIIIim in buffered 2,4,6-collidine/2,4,6-collidine–H+ (pKa = 15.0030) in varying ratios. (C) Plot of the E1/2 vs Fc+/0 of the 1e−/1H+ proton-coupled RuIII/II redox couple vs logarithm of the buffer ratio, with slope of 59.0 ± 0.1 mV and y-intercept (dashed line) of −701 ± 5 mV. The colors of the points correspond to those of the corresponding CVs in part (B).
The BDFEs were calculated using the thermochemical square scheme in Scheme 1, using the appropriate pKa and E1/2 values in MeCN (eq 2). The constant CG is the free energy required to reduce a proton to H• (with ferrocene in MeCN).1 Appropriate pKa and E1/2 values are the top and right sides of the square scheme, or the left and bottom sides—both lead from XH to X, the BDFE.
(2) |
For each complex for which all four pKa/E1/2 values were measured, this analysis gives two independent measurements of the BDFE that agreed within 1.0 kcal mol−1 of one another. This indicates the precision of the individual measurements.
The BDFEs were also measured electrochemically to provide more precise values (an example is shown in Figure 5). This was done by dissolving either the RuIIimH or RuIIIim form of the complex in MeCN containing electrolyte. Then to this solution was added a buffer with a pKa in between the pKas of the reduced and oxidized forms of the complex, such that oxidation/reduction of the complex was coupled to a proton transfer. Using the Nernst equation (eq 3) and the pKa of the buffer, the standard state potential for this 1e−/1H+ (n = 1) reaction was measured and converted to the BDFE (eq 4).
(3) |
(4) |
The E1/2 shifted ~59 mV with the logarithm of the buffer ratio and was independent of the scan rate, indicative of a PCET process reversible on the CV timescale. Each BDFE obtained in this way was within error of the average value obtained through the individual pKa and E1/2 measurements. Since E1/2 is usually measured more accurately then pKas, these values are likely to be more accurate. This method of determining the BDFE is analogous to measuring the PCET open-circuit potential (OCP) but only applicable when the process is electrochemically reversible.31
BDFEs can often also be measured by equilibration with a compound with a known, similar BDFE.1 However, the scarcity of 1e−/1H+ PCET reagents with BDFEs < 60 kcal mol−1 limited this approach to the BDFE of 5, which was equilibrated with phenazine (BDFE = 58.7 kcal mol−1 31). However, the uncertainty of this measurement was larger than those measurements detailed above (see SI).
The primary uncertainty in the electrochemically measured BDFEs is the uncertainty in in the CG term, ± 1.0 kcal mol−1.1 Since these equations use the same CG, the relative uncertainty in ΔBDFE values is much lower, estimated as ± 0.2 kcal mol−1.
vi). DFT Calculations: E1/2 and pKa values
Relative E1/2 and pKa values were computed by DFT. These values were anchored to the parent complex (4) by shifting all the absolute values by a scalar, such that the average pKa and E1/2 of 4 match the experimental values. Doing so minimizes internal methodology errors and allows a more meaningful comparison of the thermochemical trends. Figure 6 pictorially shows the experimental and computed differences in these parameters from those for 4.
Figure 6.
Bar chart of the E1/2 and pKa changes from the parent complex 4 induced by substitutions at the acac (1-4) and py-imH (4-7) ligands, converted to free energies. Darker bars are experimental values and lighter bars are calculated.
The series with modifications to the py-imH ligand (4-7) shows excellent agreement between the experimental and calculated ΔpKa and ΔE1/2 values: all of the deviations are ≤1.0 kcal mol−1. For the series with modifications to the acac ligands (1-4), a few computed values show larger deviations, up to 0.2 eV/5 kcal mol−1 for the ΔE1/2im and ΔBDFE for compound 1.
Discussion
Across a series of related compounds, many PCET redox couples show small variations of BDFEs due to compensation of changes in the E1/2 by changes in the pKa. In contrast, the Ru(acac)2(py-imH) system described here exhibits significant ‘decoupling’ of the pKa and E1/2. Substitutions on the acac ligands cause E1/2 changes at least 5 times as large as the pKa changes, when described in the same free energy units (|FΔE1/2| = |ΔΔG°| = 2.303RT|ΔpKa|). The most pronounced decoupling is seen in moving from 4 to 2. Introducing phenyl groups into the acac ligands makes the RuII(py-imH) complex 110 mV less reducing and the deprotonated RuII(py-im) analog 70 mV less reducing. However, the pKa shift is only 0.1 (± 0.1) unit, equivalent to 6 ± 6 mV. Substituents on the py-imH ligand decouple in the opposite direction, with larger changes in pKa than in E1/2, to varying degrees. Moving from 4 to 7, the complex with the distal CF3 group, lowers the pKa values by 5.6 / 5.8 units (RuII / RuIII complexes, resp.; ≡ 340 mV) while the reduction potentials shift only 90 / 120 mV (py-imH / py-im, resp.). On the other hand, in moving from 4 to the benzimidazole 5, the changes in E1/2 values nearly compensate those in the pKas. The overall trend of this decoupling is illustrated in Figure 6.
The DFT calculations provide insight into the origins of the E1/2 / pKa decoupling. The MO diagrams in Figure 3 show that the ruthenium d orbitals (t2g orbitals) are significantly affected by changes to the acac ligands but negligibly by changes to the py-imH ligand. This explains why the E1/2 values change much more across complexes 1-4 than across 4-7. The UV-vis spectra offer experimental support for this explanation. For complexes 1-4, the wavelength (λmax) of the lowest-energy MLCT correlates with the E1/2 of the complex (see Figure 7, λmax,1). Because this transition is primarily R→py-imH in character, the transition energies reflect changes in the energies of the Ru-centered t2g HOMOs. In contrast, across complexes 4-7 the changes in t2g energies are small and the λmax changes predominantly with the π*py-imH orbital energies. Furthermore, the higher-energy MLCT around 420 nm (λmax,2), which is primarily Ru→acac in character, shifts negligibly across the py-imH-tuned series (< 20 nm), suggesting that the energies of the t2g (and π*acac) orbitals remain unchanged for these complexes.
Figure 7.
The experimental and calculated λmax of the two visible MLCT bands (left axis) and the E1/2 (right axis) for each of RuIIimH complexes 1-7. The plot highlights the correlation of the E1/2 with the energy of the lower-energy MLCT (purple) in complexes 1-4.
While substituents in the py-imH ligand do not strongly shift the reduction potentials, deprotonation to the anionic py-im ligand shifts the E1/2 of all of the complexes by 345 ± 25 mV. The consistency of this shift is perhaps surprising in light of the differences in electronic structure between the compounds. The E1/2 shift in part reflects the change in the overall charge of the complex upon deprotonation, and this shift is typical of imidazole-type complexes.1, 32
The acidity of the imidazole N–H bond likely reflects the overall electron-rich or electron-deficient character of the ligand, which is reflected in the energies of the πpy-imH orbitals. These are the frontier orbitals most affected upon deprotonation (see Figure S76) and that shift significantly across the py-imH-tuned series. The πpy-imH orbitals are, however, negligibly affected by the acac substitutions, except for 3 (for which these changes are still dwarfed by the t2g orbital energy shifts).
More broadly, the decoupling of the redox and acid/basic properties of these complexes is affected by the physical separation of the electron and proton within the molecule and the different distances from the substituent to the e− or H+. In particular, the four acac substituents are quite distant from the acidic proton in the imidazole N–H bond. With the imidazole substituents, the substitutions in 5 and 7 are slightly closer to the acidic N–H bond, while the CF3 group of 6 is three bonds away from both the N–H bond and the ruthenium center. This helps to explain why the decoupling is smaller in the py-imH-tuned series than in the acac-tuned series. However, the significant decoupling even in 6 underscores the primacy of the frontier orbitals in determining this decoupling.
While the data and computations in this report are all in acetonitrile solvent, the conclusions should be general in different media. The relative E° and pKa values depend only on the differences in the free energies to transfer the relevant species from one solvent to another, and these differences should be small. For instance, the ΔG°transfers should be very similar for Ru(acac)2(py-imH)+ and Ru(acac)2(py-CF3imH)+, even between protic and aprotic solvents such as MeCN to H2O, because the molecules have similar sizes, charges, and functionalities. Therefore, the difference between their E1/2 values should not change significantly between the solvents. Solvent effects on PCET reactions have been discussed in our recent review.1
The thermodynamic coupling of E1/2 and pKa in these Ru complexes can be compared with similar effects in organic molecules. The pKas and redox potentials of unconjugated organic PCET reagents such as alkylamines and hydroxylamines are much more tightly coupled, their BDFEs being more substitution-independent.1, 7 In aromatic organic PCET reagents, such as phenols, indoles, arylamines, and toluenes, substitutions on the aromatic ring typically result in a larger change in redox potential than in pKa. This likely stems from the greater delocalization of the electron into the ring than the proton—analogous to the HOMO in the Ru(acac)2(py-imH) system being delocalized over the whole complex. Substitutions closer to the acidic proton in these organic systems might feasibly flip this decoupling in favor of changing the acidity. This seems true at least for substituted toluenes; for example, the pKas of para-tolunitrile and 2-phenylacetonitrile in DMSO are 31 and 22,33, 34 while their gas phase ionization energies are very similar (NCPhCH3, 9.32 eV; vs. PhCH2CN, 9.39 eV).35
Conclusions
In a newly synthesized series of Ru(acac)2(py-imH) complexes, we have tuned the redox and acid/base properties independently of one another. The former is achieved through substitutions to the acac ligand and the latter through substitutions to the imidazole. DFT calculations reveal that this ‘decoupling’ of the E1/2 and pKa can be attributed to differences in the molecular orbitals that are affected by the different substitution patterns. More broadly, the decoupling arises from the placement of the substituents closer to the e− (formally located on the ruthenium center) versus closer to the H+ (on the imidazole). The ability to separately tune the component properties of a PCET system is likely to be valuable in various future applications.
Supplementary Material
ACKNOWLEDGMENT
The authors are indebted to Dr. Adam Wu, Joshua Masland, Rodney D. Swartz, and Werner Kaminsky for their prior studies on this system, as well as to the Chemical and Biophysical Instrumentation Center at Yale University for assistance in spectroscopic characterization. MC thanks CONICET (Argentina) and NIH for support during his research stay.
FUNDING SOURCES
Financial support from the U.S. National Institutes of Health grants 3R01GM050422 and 1R35GM144105 is gratefully acknowledged. B.D.G. acknowledges the Yale Dox Fellowship for additional support.
Footnotes
SUPPORTING INFORMATION
General experimental considerations; preparative routes and characterization; NMR and UV-vis spectra, high-resolution mass spectra, cyclic voltammograms, and crystallographic data; data and discussion for pKa and electrochemical BDFE measurements; measurement of BDFE of 6 by UV-vis equilibration of 6-RuIIIim with 5,10-dihydrophenazine; computational details: computed thermochemical data, MO energies, UV-vis transitions and spectra, and calculated structures of the complexes (PDF).
Accession Codes
CCDC numbers 2253082 (1-RuIIIim), 2253083 (2-RuIIIim), 2253084 (5-RuIIIim), 2253085 (6-RuIIIim), 2253086 (7-RuIIIim), and 2253087 (Ru(acac)2(py-(NO2)2im) 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.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), 1–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Darcy JW; Koronkiewicz B; Parada GA; Mayer JM, A Continuum of Proton-Coupled Electron Transfer Reactivity. Acc. Chem. Res 2018, 51 (10), 2391–2399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Miller DC; Tarantino KT; Knowles RR, Proton-Coupled Electron Transfer in Organic Synthesis: Fundamentals, Applications, and Opportunities. In Hydrogen Transfer Reactions: Reductions and Beyond, Guillena G; Ramón DJ, Eds. Springer International Publishing: Cham, 2016; pp 145–203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Huynh MH; Meyer TJ, Proton-coupled electron transfer. Chem Rev 2007, 107 (11), 5004–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Dempsey JL; Winkler JR; Gray HB, Proton-coupled electron flow in protein redox machines. Chem Rev 2010, 110 (12), 7024–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Sim BA; Griller D; Wayner DDM, Reduction potentials for substituted benzyl radicals: pKa values for the corresponding toluenes. Journal of the American Chemical Society 2002, 111 (2), 754–755. [Google Scholar]
- 7.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 (1), 356–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Mayer JM, Proton-coupled electron transfer: a reaction chemist’s view. Annu. Rev. Phys. Chem 2004, 55, 363–90. [DOI] [PubMed] [Google Scholar]
- 9.Roberts BP, Polarity-reversal catalysis of hydrogen-atom abstraction reactions: concepts and applications in organic chemistry. Chem. Soc. Rev 1999, 28 (1), 25–35. [Google Scholar]
- 10.Tedder JM, Which Factors Determine the Reactivity and Regioselectivity of Free Radical Substitution and Addition Reactions? Angewandte Chemie International Edition in English 1982, 21 (6), 401–410. [Google Scholar]
- 11.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 (36). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.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 (44), E10287–E10294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.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 (27), 10777–10787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.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 (9), 4051–4062. [DOI] [PubMed] [Google Scholar]
- 15.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 (11), 4173–4183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.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 (45), 17090–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.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 (40), 16982–16989. [DOI] [PubMed] [Google Scholar]
- 18.Borovik AS, Role of metal-oxo complexes in the cleavage of C-H bonds. Chem. Soc. Rev 2011, 40 (4), 1870–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.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 (28), 11142–11150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Green MT; Dawson JH; Gray HB, Oxoiron(IV) in chloroperoxidase compound II is basic: implications for P450 chemistry. Science 2004, 304 (5677), 1653–6. [DOI] [PubMed] [Google Scholar]
- 21.Derosa J; Garrido-Barros P; Peters JC, Electrocatalytic Reduction of C-C pi-Bonds via a Cobaltocene-Derived Concerted Proton-Electron Transfer Mediator: Fumarate Hydrogenation as a Model Study. J. Am. Chem. Soc 2021, 143 (25), 9303–9307. [DOI] [PubMed] [Google Scholar]
- 22.Garrido-Barros P; Derosa J; Chalkley MJ; Peters JC, Tandem electrocatalytic N(2) fixation via proton-coupled electron transfer. Nature 2022, 609 (7925), 71–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Wu A; Masland J; Swartz RD; Kaminsky W; Mayer JM, Synthesis and characterization of ruthenium bis(beta-diketonato) pyridine-imidazole complexes for hydrogen atom transfer. Inorg. Chem 2007, 46 (26), 11190–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Oomura K.-i.; Ooyama D; Satoh Y; Nagao N; Nagao H; Scott Howell F; Mukaida M, Redox behavior of a binuclear ruthenium complex having a di-μ-nitrosyl ligand, [{Ru(acac)2}2(μ-NO)2] (acac = acetylacetonato). Inorg. Chim. Acta 1998, 269 (2), 342–346. [Google Scholar]
- 25.Gupta AK; Poddar RK, Improved synthesis and reactivity of tris(acetylacetonato)ruthenium(III). Indian J. Chem 2000, 39A, 457–60. [Google Scholar]
- 26.Kar S; Chanda N; Mobin SM; Datta A; Urbanos FA; Puranik VG; Jimenez-Aparicio R; Lahiri GK, Diruthenium complexes [{(acac)2RuIII}2(μ-OC2H5)2], [{(acac)2RuIII}2(μ-L)](ClO4)2, and [{(bpy)2RuII}2(μ-L)](ClO4)4 [L = (NC5H4N-C6H4-N-(NC5H4)2, acac = acetylacetonate, and bpy = 2,2’-bipyridine]. Synthesis, structure, magnetic, spectral, and photophysical aspects. Inorg. Chem 2004, 43 (16), 4911–20. [DOI] [PubMed] [Google Scholar]
- 27.Kar S; Chanda N; Mobin SM; Urbanos FA; Niemeyer M; Puranik VG; Jimenez-Aparicio R; Lahiri GK, Unusual monodentate binding mode of 2,2’-dipyridylamine (L) in isomeric trans-(acac)2RuII(L)2, trans-[(acac)2RuII(L)2]ClO4, and cis-(acac)2RuII(L)2 (acac = acetylacetonate). Synthesis, structures, and spectroscopic, electrochemical, and magnetic aspects. Inorg. Chem 2005, 44 (5), 1571–9. [DOI] [PubMed] [Google Scholar]
- 28.Ghumaan S; Sarkar B; Patra S; Parimal K; van Slageren J; Fiedler J; Kaim W; Lahiri GK, 3,6-bis(2’-pyridyl)pyridazine (L) and its deprotonated form (L – H+)− as ligands for {(acac)2Run+} or {(bpy)2Rum+}: investigation of mixed valency in [{(acac)2Ru}2(μ-L – H+)]0 and [{(bpy)2Ru}2(μ-L–H+)]4+ by spectroelectrochemistry and EPR. Dalton Trans. 2005, (4), 706–12. [DOI] [PubMed] [Google Scholar]
- 29.Seddon EA; Seddon KR; Clark RJH, The Chemistry of Ruthenium. Elsevier Science: 2013. [Google Scholar]
- 30.Tshepelevitsh S; Kütt A; Lõkov M; Kaljurand I; Saame J; Heering A; Plieger PG; Vianello R; Leito I, On the Basicity of Organic Bases in Different Media. Eur. J. Org. Chem 2019, 2019 (40), 6735–6748. [Google Scholar]
- 31.Wise CF; Agarwal RG; Mayer JM, Determining Proton-Coupled Standard Potentials and X-H Bond Dissociation Free Energies in Nonaqueous Solvents Using Open-Circuit Potential Measurements. J. Am. Chem. Soc 2020, 142 (24), 10681–10691. [DOI] [PubMed] [Google Scholar]
- 32.Stupka G; Gremaud L; Williams AF, Control of Redox Potential by Deprotonation of Coordinated 1H-Imidazole in Complexes of 2-(1H-Imidazol-2-yl)pyridine. Helv. Chim. Acta 2005, 88 (3), 487–495. [Google Scholar]
- 33.Bordwell FG; Bares JE; Bartmess JE; McCollum GJ; Van der Puy M; Vanier NR; Matthews WS, Carbon acids. 12. Acidifying effects of phenyl substituents. The Journal of Organic Chemistry 2002, 42 (2), 321–325. [Google Scholar]
- 34.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, (6), 1125–1133. [Google Scholar]
- 35.Linstrom PJ; Mallard WG NIST Chemistry Webbook, NIST Standard Database Number 69.http://webbook.nist.gov/chemistry/ (accessed March 21, 2023). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.