Abstract
We present the development of ligands featuring the unconventional hydrogen bond donor, –CF2H, within a metal’s secondary coordination sphere. When metalated with palladium, o-CF2H functionalized 1,10-phenanthroline provides highly directed H-bonding interactions with Pd-coordinated substrates. Spectroscopic and computational analyses with a series of X-type ligand acceptors (–F, –Cl, –Br, –OR) establish the H-bonding interaction strength for the –CF2H group (~3 kcal/mol). Synthesis of Pd°/Ni° complexes and subsequent coupling (Ni) highlight the unique reductive and base compatibility of the –CF2H hydrogen bond donor group.
INTRODUCTION
Prearranged hydrogen bonding interactions that facilitate substrate capture and activation are ubiquitous in enzyme active sites. These secondary-sphere interactions are critical to many biological transformations including nitrogen fixation,1-3 hydrogenation,4-7 and CO2 reduction.8 To develop an understanding of how these noncovalent interactions facilitate small molecule activation, synthetic complexes featuring appended hydrogen bond (H-bond) donors are increasingly used in biomimetic design,9-12 and catalytic transformations.13-20 In almost all cases, these studies are limited to mid- to high- oxidation state transformations. H-bond donors are generally incompatible with low valent metal complexes because they undergo proton-transfer to the metal,21-23 which presents a fundamental challenge for evaluating H-bonding induced reactivity at highly reduced metal sites.24 Although biologically relevant H-bond donors (e.g. amines, thiols, or water) are often incompatible with reduced metal complexes,25-28 metalloenzymes have evolved to overcome this limitation by: (1) isolating their active-sites,29 (2) employing highly specific H+/e− transport systems,30 and/or (3) operating at or above the proton reduction limit through charge delocalization (e.g. [FeS]n clusters; E° > −0.8 V vs SHE).31-32 An alternative approach within synthetic systems, where common reducing agents used to access biological model complexes exceed −2 V (vs SHE),33 is to use H-bond donors that are resistant to proton transfer.
The –CF2H functional group is a robust34 bioisostere for amines and thiols, exhibiting comparable H-bond donor strength.35 Although H-bonding effects of this unit have been identified using metal-free systems,36-37 the –CF2H unit has not been incorporated within a metal’s secondary coordination sphere for H-bonding to a metal coordinated substrate. To address this gap, we targeted ligands that accommodate appended –CF2H groups as a key design element. Ortho- functionalized pyridines provide high directionality to enforce H-bonding to metal bound substrates,13-14, 38-47 especially when constrained within a rigid chelating ligand (Figure 1). To probe secondary coordination sphere effects of –CF2H H-bonding to a metal bound substrate, we targeted 2-(difluoromethyl)-1,10-phenanthroline48 (phenCF2H).43-49 Herein we disclose the –CF2H group as a viable H-bond donor when appended to a ligand scaffold to evaluate (1) directional H-bond strength and (2) reductive stability.
Figure 1.
2-Pyridyl ligand frameworks for –OH and –CF2H H-bonds.
RESULTS AND DISCUSSION
Synthesis and characterization of complexes featuring secondary-sphere –CF2H hydrogen bonds.
Treatment of phenCF2H with PdCl2 in refluxing acetonitrile afforded PdCl2(phenCF2H) (1-Cl) in 55 % yield as confirmed by X-ray crystallography (Figure 2). The primary coordination sphere of 1–Cl features a square-planar geometry (τ= 0.08) at Pd, with equivalent Pd-Cl distances (Pd-Cl1 = 2.290(16) Å; Pd-Cl2 = 2.299(17) Å) and inequivalent Pd-N distances (Pd-N1 = 2.005(9) Å; Pd-N2 = 2.149(11) Å). We attribute the ligand binding asymmetry to a steric interaction of the ortho- group of the phenCF2H ligand. Importantly, the crystallographic characterization of 1-Cl (Figure 2B) reveal the –CF2H proton is oriented toward the neighboring –Cl at a distance of 2.339 Å (C-Cl = 3.15(2) Å) 50 within the sum of the van der Waals radii(vdW),51-52 consistent with formation of a –CF2H⋯Cl H-bond. To assess the –CF2H group as a secondary-sphere H-bond donor,13-14, 38-44 we evaluated the interaction energy, solution behavior, and donor strength of the H-bond.
Figure 2.
(A) Synthesis of complexes 1-X and 2, and (B) crystal structure of 1-Cl; Cl1-H11 = 2.339, Cl1-C11 3.15(2) Å with side view (50% probability ellipsoids; protons not involved in H-bonds are removed for clarity).
To interrogate the extent to which the –CF2H H-bond is retained in solution, we examined a series of 1-X complexes (X= F, Cl, Br): congeners featuring coordinated halides of varied H-bond acceptor strength.53 We prepared PdBr2(phenCF2H) (1-Br) in 80 % yield, analogous to 1-Cl, while PdF2(phenCF2H) (1-F) was accessed by oxidation of in situ prepared Pd(dba)(phenCF2H) (2) (dba = dibenzylideneacetone) with XeF2.54-59,60 Relative to the free ligand, phenCF2H, the 1H NMR spectra of complexes 1-X exhibit shifts of the –CF2H group (2JFH = 52-55 Hz) downfield from 7.01 to ≥ 8.3 ppm, and corresponding modest 19F NMR shifts (≤ Δδ 4.5 ppm). For 1-F, the –CF2H 1H NMR signal at 8.38 ppm is a triplet of doublets, with coupling constants of 53 and 25 Hz, which we attribute to both –CF2H 19F 2JFH coupling as well as a through-space –CF2H⋯F interaction (1JFH). Selective 1H{19F} decoupling of the Pd-F signal at −383 ppm collapsed the triplet of doublets to a triplet (2JFH = 53 Hz), further supporting assignment of 1JFH coupling; a hallmark of –F⋯H H-bond formation. The magnitude of ∣1JFH∣ of 25 Hz is consistent with “weak” H-bond formation.61-62 In general, the 1H chemical shift of H-bond donors (E-H) provides a spectroscopic handle to interrogate H-bond interactions, where strong H-bond interactions generally have a large deshielding effect.63 However, for the 1-X series, the induced H-bond chemical shift of –CF2H (X = F, Cl, Br; 8.38, 8.73, 8.73 ppm), is inverted relative to the predicted H-bond acceptor strength (F > Cl > Br).53
To understand why the –CF2H proton resonance undergoes an upfield rather than downfield shift in the 1H NMR spectra as the H-bond acceptor strength for 1-X increases, we used computational methods. Structures of 1-X were optimized and analyzed at the M062x/6-311++(2d,2p) SDD//M062x/def2tzvp PCM=(CH2Cl2) level of theory. 1-X undergoes geometric distortion as the size of X increases, forcing the phenCF2H ligand out of the PdX2 plane (Figure 3). The distortion of the close-contact for the –CF2H⋯X proton is an additional factor that may convolute the effect of NMR shielding. To determine whether the -CF2H⋯X close-contact is attractive or repulsive and the relative magnitude of the interaction, we employed a noncovalent interaction (NCI)64 analysis of 1-X. The NCI plots (Figure 3) demonstrate that the X⋯H interaction is attractive (sign(λ2)ρ < 0) for all X, and largest in magnitude for 1-F (Figure 3) (sign(λ2)ρmin: F= −0.034; Cl = −0.021; Br = −0.019). For context, the interactions of H-bond donors H2O, NH3, and CH2F2 interacting with a reference acceptor molecule, NH3, are −0.028 (H2O), −0.016 (NH3), and −0.013 (CH2F2) (Figure 3C). To further characterize the –CF2H H-bond, we examined charge transfer via a natural bond orbital (NBO) analysis. A comparison of 1–F and 1–Br revealed the C-H bond of –CF2H has a higher H charge for 1–F (Δe− = 0.03) with a concomitant decrease in C charge (Δe− = −0.02). These results are consistent with increased polarization induced by H-bond formation to the stronger acceptor (F).65 The computational analysis of 1–X supports an H-bonding interaction of Pd-X with the –CF2H unit and underscores the importance of geometric constraints of the H-bond acceptor.
Figure 3.
DFT optimized structures with NCI analysis (isosurface s = 0.5) of the X-H interaction of (A) 1-F, 1-Cl, and 1-Br, (B) side-view of 1-X distortion of Pd-N1-C3 angle, and (C) NCI analysis of H-bond interactions of H2O, NH3, and CH2F2 with NH3.
The geometric distortions that convolute analysis of the H-bonding effects in 1-X can be mitigated by comparing isostructural H-bond acceptors. When substituted in the para position, phenoxides provide tunable H-bond acceptor strength, that can be quantified by the Hammett parameter (where H-bond acceptor strength tracks as NMe2 > OMe > tBu > H > Br > NO2).66 Salt metathesis of 1-Cl with 2 equiv. thallium phenoxides resulted in the formation of Pd(OArR’)2(phenCF2H) (3-OArR’, R’= NMe2, OMe, tBu, H, Br, NO2) (Figure 4A). 1H NMR analyses of 3-OArR’ feature small shifts of the phenCF2H ligand resonances and a pair of corresponding aryl resonances, consistent with an identical primary coordination sphere across the series. X-ray analysis of 3-OArNO2 and 3-OArH provided clear support that the primary coordination sphere is not altered (Figure 4B; primary sphere geometry overlay RMS = 0.111; see SI). Across the series, the 1H NMR –CF2H resonance shifted downfield with increasing H-bond acceptor strength of the phenoxide (Figure 5A); consistent with an H-bond strength induced chemical shift. When the Δδ of the –CF2H resonance is plotted as a function of Hammett parameter (σp), the data are linear up to σp = −0.268 (3-OArOMe), with deviations from linearity observed for σp = −0.83 (3-OArNMe2).
Figure 4.
(A) Synthesis of complexes 3-OArR’ and 5-OArR’, (B) crystal structures of 4-Cl, 3-OArNO2, and 3-OArH (50% probability ellipsoids; protons not involved in H-bonds are removed for clarity). Dihedral angles shown between H13-C13-C10-N2. (C) Steric (1 vdW radii) and electrostatic potential surface (at 0.004 au) comparisons for appended groups of phenCR2H.
Figure 5.
Hammett plots of (A) –CR2H Δδ and (B) calculated ΔνC-H of 3-OArR’ and 5-OArR’.
To clarify the reasons that govern the nonlinear behavior described above, we interrogated the solid state structures of 3-OArR’. Both structurally characterized products, 3-OArNO2 and 3-OArH, exhibit close –CF2H⋯O heavy-atom –C⋯O distances consistent with H-bonding interactions (2.820(5) and 2.93(2) Å respectively) (Figure 4B). For an sp3 carbon atom, the heavy atom positions (C-CF2) provide an accurate –H atom position and orientation (riding model) even without location from a difference electron map.67 Using this treatment, the H⋯O contacts of 3-OArNO2 and 3-OArH are 2.197 and 2.002 Å, respectively. Importantly, these distances are flipped relative to the heavy atom O⋯C treatment. The difference between the H⋯O contacts of the riding model protons is the result of –CF2H rotation (3-OArNO2 = 11.8°; 3-OArH ≤ 1°) about the Cpy-CF2H torsion angle, with a stronger H-bond interaction exhibiting a shorter H-bond contact (Figure 6). The consequence of –CF2H coplanarity for 3-OArH is that as the H-bond strength is increased, a closer contact is not possible and the 1H NMR response becomes entirely due to through-space C-H polarization and charge transfer. Computational analysis of the series supports the angular constraint of the H-bond distance on H-bond acceptor strength and that coplanar geometry (R’= –tBu, –OMe, –NMe2) affords an H-bond with a maximum directionality (∠O-H-C) of 158° (Figure 6).68
Figure 6.
Analysis of the impact of H-bond donor strength on directionality. (A) NCCH dihedral and H⋯O contact as a function of H-bond acceptor strength (p-OArR’) and (B) Hammett plots of calculated NCCH dihedral angle (°) and O---H contact of 3-OArR’.
There are two competing factors that could generate the short H⋯X contacts from the –CF2H group; (1) an attractive interaction or (2) the minimization of steric repulsion from the –F substituents. With few exceptions, the identification of H-bond interactions with –CF2H units rely exclusively on crystallographic analyses. Assessment of close contacts does not necessitate an attractive interaction. However, in many of these structures,36, 69-70,71 rotation of the –CF2H group is hindered due to the steric repulsion from the –F groups, a factor noted above to control H-bond strength. To examine the steric component of the –CF2H unit to influence close contacts, we targeted an isosteric analogue of –CF2H that has similarly hindered –CR2H rotation but does not contain an acidic C-H group. Although –CH3 is a common control group for –CF2H (Charton parameter72 (ν): –CF2H = 0.68; –CH3 = 0.52), it does not capture the steric bias contributing to C-H orientation.73 Furthermore, fast rotation of the –CH3 group averages the H-bond interactions with non-interacting protons on the NMR timescale, resulting in a diminished 1H NMR response. Instead, we targeted an iPr group (ν = 0.76) as a structurally faithful model74 (pheniPr) to assess the C-H polarization contribution to H-bond formation (Figure 4C).
Metalation of the control ligand, pheniPr, with PdCl2(PhCN)2 afforded PdCl2(pheniPr) (4-Cl) in 82 % yield. X-ray diffraction of 4-Cl (Figure 4C) revealed a CH⋯Cl contact of 2.496 Å (riding model treatment) and C-Cl contact of 3.256(3) Å), which are a longer contacts than for the –CF2H analogue (1-Cl: H⋯Cl = 2.339 Å; C-Cl = 3.15(2) Å). To rigorously compare the H-bond contributions of –C(CH3)2H group vs –CF2H, we prepared analogous PdII phenoxide complexes. Salt metathesis of 4-Cl afforded the phenoxide complexes (5-OArR’).75 The 1H NMR spectra of the complexes 5-OArR’ each contain a characteristic septet peak between 4.50 and 5.30 ppm (Δδ = 0.80) assignable to the methine proton (H-C(CH3)2).
In comparison to –CF2H (Δδ of 3-OArR’ = 0.95), the Δδ of the –C(CH3)2H resonance is similar for weak H-bond acceptors (R’= –NO2 – –tBu; ΔΔδ ≤ 0.08) and larger only for strong H-bond acceptors (R’= –OMe – –NMe2; ΔΔδ ≤ 0.15) where through-space interactions are expected to dominate over –CR2H rotation effects (Table 1). These results were supported by computational analysis of the 3-OArR’ and 5-OArR’ series, which exhibit a phenoxide-dependent –CR2H dihedral orientation for –CF2H, but not for –C(CH3)2H (see SI). This result is consistent with weaker attractive noncovalent interactions for the non-polar –C(CH3)2H, compared to –CF2H. Together, the combined 1H NMR/computational analyses support the notion that the unique ability of the –CF2H group to form H-bonds is an electronic rather than a steric effect.
Table 1.
H-bonding and primary-sphere comparisons between Pd–phenCR2H compounds. Xb: H-bond acceptor; Xa: non-interacting X-type ligand on Pd.
Compound |
1H –CR2H (ppm) |
Xb⋯H (Å) |
Pd-Xa (Å) |
Pd-Xb (Å) |
DFT: Xb⋯H (Å) |
DFT: Pd-Xa (Å) |
DFT: Pd-Xb (Å) |
DFT: νC-H (cm−1) |
---|---|---|---|---|---|---|---|---|
–CF2H | ||||||||
1-F | 8.38 | 1.829 | 1.947 | 1.956 | 3128 | |||
1-Cl | 8.73 | 2.339 | 2.290(16) | 2.299(17) | 2.364 | 2.294 | 2.304 | 3205 |
1-Br | 8.73 | 2.533 | 2.431 | 2.441 | 3210 | |||
3-OArNO2 | 8.02 | 2.197 | 1.983(3) | 1.993(3) | 2.088 | 2.000 | 2.006 | 3186 |
3-OArBr | 8.47 | 2.051 | 1.998 | 2.004 | 3184 | |||
3-OArH | 8.72 | 2.002 | 2.002(9) | 2.020(8) | 1.995 | 1.993 | 1.999 | 3173 |
3-OArtBu | 8.81 | 1.948 | 1.992 | 2.006 | 3145 | |||
3-OArOMe | 8.86 | 1.948 | 2.002 | 2.007 | 3141 | |||
3-OArNMe2 | 8.97 | 1.955 | 2.003 | 2.010 | 3135 | |||
6-I | 9.08 | 2.772 | 1.941 | 2.665 | 3149 | |||
6-F | 8.68 | 1.895a | 1.977(5)a | 1.957(7)a | 1.869 | 1.946 | 1.996 | 3110 |
–C(CH3)2H | ||||||||
4-Cl | 5.07 | 2.496 | 2.2796(7) | 2.2891(7) | 2.430 | 2.300 | 2.314 | 3172 |
5-OArNO2 | 4.49 | 2.118 | 2.005 | 2.012 | 3120 | |||
5-OArBr | 4.82 | 2.125 | 2.023 | 2.029 | 3118 | |||
5-OArH | 4.93 | 2.116 | 1.996 | 2.002 | 3116 | |||
5-OArtBu | 4.95 | 2.066 | 1.994 | 2.005 | 3115 | |||
5-OArOMe | 5.20 | 2.106 | 1.997 | 2.005 | 3116 | |||
5-OArNMe2 | 5.25 | 2.112 | 1.998 | 2.013 | 3121 | |||
7-I | 5.28 | 2.794 | 1.943 | 2.681 | 3106 | |||
7-F | 5.01 | 2.073 | 1.949 | 1.994 | 3102 |
Distances reported are for one individual molecule in the unit cell.
Although the computational and experimental studies above indicate that phenCF2H ligand engages in H-bond interactions, quantifying the interaction strength is intrinsically difficult. In contrast to classic H-bond donors (OH, NH) that undergo a hypsochromic (blue) shift of the νE-H mode upon formation of an H-bond, H-bonds to C-H donors may exhibit distinct shifts: the νC-H may either undergo a hypsochromic (blue), no, or bathochromic (red) shift.76 As a result, it is difficult to identify νC-H shifts for even simple systems,36 and there are no reported correlations between C-H H-bond energy and νC-H.
To ascertain the influence of the –CR2H group on the H-bond interaction, the vibrational dependence (ΔνC-H) of the donor was analyzed computationally for 3-OArR’ and 5-OArR’. We found that the νC-H for 3-OArR’ exhibits bathochromic (red) shifts with increasing H-bond acceptor strength.77 The important difference between the two modeled complexes is that for R = CF2H, the ΔνC-H tracks with H-bond acceptor strength, while there is no systematic trend for R = C(CH3)2H (Figure 5B).78
To experimentally establish the relative strength of the CR2H⋯X interaction (for CF2H and C(CH3)2H, we interrogated the through-space H-F coupling interaction by NMR spectroscopy. Despite efforts to prepare Pd difluoride compounds (1-F, Figure 2), we found that they were generally unstable toward further chemical manipulations including isolation and crystallization. To overcome this limitation, we targeted the isolation of palladium aryl fluoride complexes because they exhibit higher stability relative to palladium difluorides.54-59 The addition of phenCF2H or pheniPr to Pd(Ph)I(Py)2 in CH2Cl2 afforded Pd(Ph)I(phenCF2H) (6-I) or Pd(Ph)I(pheniPr) (7-I) in good yields (84, 76%) (Figure 7A). Sonication of 6-I and AgF in CH2Cl2 while cooling in an ice bath facilitated halide exchange to produce Pd(Ph)F(phenCF2H) (6-F). The major isomer observed (>95%) exhibited H⋯F H-bonding interactions, vide infra.
Figure 7.
(A) Synthesis of complexes 6-I, 7-I, 6-F, and 7-F, (B) crystal structure of 6-F (30% probability ellipsoids; for clarity a single molecule from the unit cell is shown and protons not involved in H-bonds are removed), and (C) 1H and selective decoupled 1H{19F} NMR of 6-F and 7-F.
1H NMR spectra of 6-I and 6-F exhibit –CF2H resonances at 9.08 and 8.68 ppm respectively.79 Similar to 1-F, the –CF2H⋯F interaction in 6-F exhibits a through-space coupling interaction; ∣1JFH∣ = 23 Hz. This value is smaller than ∣1JFH∣ coupling to stronger H-bond donors (–OH = 50 Hz;43 –NH2 = 50-64 Hz80). To examine a direct comparison using stronger H-bond donors within this system, we targeted bpyOH as a related bidentate ligand. In contrast to the stable complex obtained using phenCF2H, we found that metalation of this ligand under analogous conditions produced benzene (via protonation of Pd-Ph) as a primary product: a reaction liability introduced by acidic H-bond donors (–OH).
To further interrogate the electronic requirements needed to visualize a through space 1JFH coupling to Csp3-H units, we analyzed the isostructural complex with pheniPr. The 1H NMR spectrum of the analogously prepared Pd(Ph)F(pheniPr) (7-F) complex does not feature detectable through-space 1JFH coupling, but is broadened.81 Selective Pd-F decoupling 1H{19F} NMR collapses the broadened resonance to a sharp septet, which is consistent with a very weak through-space H⋯F interaction of ∣1JFH∣ ≤ 3 Hz. The contrast between the 1JFH of 6-F and 7-F provides further support that the C-H bond in –CF2H is uniquely polarized and capable of strong attractive interactions and electronic communication via H-bonding. Similar to 1-X and 3-OAr, the –CF2H vibrations of 6-I/F were unable to be identified even with isotopic labeling of the –CF2H(D) group. An NCI analysis of 6-F indicated a stronger attractive interaction than found in 7-F (sign(λ2)ρ for s=0.5: 6-F = −0.029; 7-F = −0.021), consistent with polarization serving as the dominant factor in forming the non-conventional H-bonds (Table S13). Although preorganization can serve to structurally enforce H-bonding interactions to unconventional H-bond donors, the results of our combined experimental/computational investigations indicate that, even with structurally similar environments, the –CF2H group imparts significantly stronger interactions, when compared to an –iPr group.
Determination of –CF2H hydrogen bond strength.
To determine the strength of the unconventional –CF2H⋯X H-bonds for the strongest H-bond acceptor (F), we applied computational methods.82 We approximated the H-bond enthalpy (ΔHH-F) of the –CR2H⋯F-Pd bond of 1-F as the difference between the formation enthalpy of the Pd-F bond (ΔHPd-F) of 1-F and a reference compound that cannot H-bond.83 The –CF2H substituent, when placed in the para- position of the pyridine ring (phen4-CF2H), removes the attractive H-bond interaction for enthalpy analysis, but also imparts a geometric distortion relative to the para- substitution (overlay RMS = 0.111, See SI page S82). As such, the ΔHPd-F difference between ortho- vs para- does not simply reflect the H-bond contribution of the group, but also a primary-sphere perturbation (2-CF2H vs 4-CF2H; ΔPd-N = 0.095 Å; ΔPd-F < 0.011 Å; ΔHPd-F < 1.8 kcal/mol). To minimize the primary-sphere distortion and assess the electronic effect of the –CF2H group, we compared the isostructural pair PdF2(phen4-CF2H) and PdF2(phen) (overlay RMS = 0.006). The minimal structural changes and small enthalpic differences between –H and –CF2H (4-H vs 4-CF2H; ΔPd-N = 0.002 Å; ΔPd-F < 0.002 Å; ΔHPd-F < 0.8 kcal/mol) demonstrates that the –CF2H group exerts a modest electron withdrawing effect on the Pd center, which we expect to be similar in the case of ortho- substitution. After establishing the electronic effect of the –CF2H group, we evaluated differences between phenCF2H and pheniPr. Importantly, phenCF2H and pheniPr exhibit identical primary coordination sphere environments (Figure 8; RMS = 0.0159),84 which indicates that an enthalpy comparison between –CF2H and –C(CH3)2H should appropriately model the H-bond interaction effects (2–CF2H vs 2–C(CH3)2H; ΔPd-N < 0.007 Å; ΔPd-F < 0.003 Å). From this method (Figure 8), the –CF2H⋯F H-bond strength is ~ −3.2 kcal/mol, which highlights the use of a preorganized ligand scaffold to bias cooperative interactions using the –CF2H group.85
Figure 8.
(A) Calculation of H-bond formation enthalpy ΔHHF and (B) optimized structure overlay of 1-F and PdF2(pheniPr).
Reductive stability of –CF2H H-bond donors.
The characterization of a stable Pd° intermediate en route to 1-F (vide supra) indicates compatibility between secondary-sphere –CF2H H-bond donors and a low valent metal center. To further evaluate the stability of the –CF2H group to low valent metal compounds, we investigated intermediate 2. Combining phenCF2H with the Pd° precursor, Pd2(dba)3, afforded a dark orange solution. The resulting 1H and 19F NMR spectra feature broad ligand phenCF2H resonances at 7.21 and 8.88 ppm, consistent with Pd(phenCF2H)(dba) (2). Although no coordinated dba resonances were observed at room temperature, these were resolved upon cooling (– 30 °C) at 4.87, 4.56, and 4.33 ppm, consistent with fluxional behavior.86 Similar preparations with dihydroxybipyridine (bpyOH) were not successful and resulted in formation of a black precipitate. This result indicates that phenCF2H is compatible with Pd° in a complex that would otherwise not tolerate an H-bond donor containing ligand.87
To complement the stability of phenCF2H in the presence of a reduced metal, described above, we performed electrochemical experiments. Electrochemical reduction of phenol has an onset potential of –1.8 V, while for Ph-CF2H, no reduction is observed within the electrolyte window < −2.65 V (0.1 M [NBu4][PF6], CH3CN vs Fc+/Fc).88 Although certain reductants (i.e. KC8) defluorinate the R-CF2H group, the electrochemical analysis indicated alternative reductants or reduced metal precursors are expected to be compatible with –CF2H containing ligands.89 In support, solutions containing either Co(Cp*)2 or KFe(Cp)(CO)2 (E1/2 > −1.98 vs Fc)33 and Ph-CF2H are stable and do not undergo defluorination (see SI page S61).
Low valent metal complexes are often precursors and/or intermediates during common organometallic reactions, such as cross coupling. To assess the ability of the appended –CF2H containing ligand to tolerate reduced metal complexes competent for cross-coupling we investigated metalation of Ni°. Addition of phenCF2H to Ni(COD)2 (COD = 1,4 cyclooctadiene) resulted in an immediate color change to dark purple. The UV-vis spectrum of the resultant complex features absorbances at 434, 545, and 686 nm, consistent with previously reported bipyridine Ni(COD) complexes,90-91 which supports formation of Ni(phenCF2H)(COD) (8).92-93 Similar to related phenanthroline complexes, 8 promotes phenyl halide coupling.94 Addition of 1 equivalent PhI affords biphenyl in 90 % yield. A new paramagnetic species also forms, which was identified as NiI2(phenCF2H)(THF) (9-THF) by an X-ray diffraction experiment as well as independent synthesis (Figure 9).95 Formation of biphenyl by Ni° requires stability of the ligand with both the low valent metal center, and strongly basic groups (e.g. Ni-Ph).94 Solutions of 8 (THF 25°C) were stable for ~ 20 h, while similar preparations with a protic appended H-bond donor ligand (bpyOH) decompose and form a mirror on reaction vessel walls over < 1 h. Although the in situ reaction to form biphenyl proceeds with many aprotic polypyridyl ligands, when bpyOH is used the reaction proceeds in very low yield (< 10 % biphenyl). These results indicate incompatibility between protic ligands and the phenyl halide coupling reaction, even for stoichiometric reactivity, and contrasts with the results obtained for phenCF2H. Overall, the combined electrochemical measurements and compatibility with reactions involving low valent complexes demonstrate an unexplored important property of –CF2H units as H-bonding groups: reductive stability.
Figure 9.
(A) Phenyl halide coupling with in situ generated Ni° polypyridyl complexes and (B) isolated inorganic side product of phenCF2H reaction (9-THF) with crystal structure (50% probability ellipsoids; protons not involved in H-bonds are removed for clarity).
CONCLUSION
The investigation of the H-bond donor/acceptor interaction within the phenCR2H ligands revealed an unusual dependence on the H-bond acceptor. In addition to electronic considerations, the size contributes to geometric distortions of the primary coordination sphere as summarized by comparing 1H NMR shifts and calculated bond metrics in Table 1. Key differences between more common H-bond donors (e.g. 2-hydroxypyridine) and phenCF2H ligands are illustrated by their geometric requirements. In contrast to coplanar H-bonds formed with even large acceptors (X = Cl, Br, I) in similar geometries,44-45 the weaker H-bond donor –CF2H counterparts require a larger X⋯H contact which is accommodated by phenCF2H binding distortions. Smaller acceptor groups, X = F, OAr, maintain similar primary-sphere bond distances and exhibit trends with increasing H-bond acceptor strength for the –CR2H⋯X contact and 1H shift. Exploiting the unique H-bond properties of the –CF2H group, namely base and reductive tolerance, may require appending the group to more flexible polypyridyl ligands or ligands of smaller ring size to better facilitate coplanar H-bond donor interactions.
In summary, we have demonstrated that appended –CF2H groups, such as those positioned in the 2-position of phen (phenCF2H) can engage in intramolecular H-bonding interactions to a metal-coordinated substrate with moderate strengths (~ 3 kcal/mol). This unit is stable, even when incorporated into highly reducing and/or basic organometallic reactions that are incompatible with most traditional H-bond donor groups. The ability of the –CF2H group to form H-bonds, in addition to tolerating reducing and highly basic conditions, provides unique opportunities to develop and study synthetic systems capable of H-bonding in the presence of highly reduced metal centers, areas that we are currently pursuing.
Supplementary Material
ACKNOWLEDGMENT
This work was supported by the NIH (Grant No. 1R01GM111486-01A1) and the NSF (Grant No. CHE0840456) for X-ray instrumentation. This research was supported in part through computational resources and services provided by Advanced Research Computing at the University of Michigan, Ann Arbor. N.K.S. is a Camille Dreyfus Teacher–Scholar. J. P. S. acknowledges a Margaret & Herman Sokol Graduate Research Fellowship. We thank Froylan Omar Fernandez for fruitful discussions.
Footnotes
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website at DOI: https://pubs.acs.org/doi/10.1021/jacs.0c01718.
X-ray data for compounds (CIF)
Experimental and computational details (PDF)
The authors declare no competing financial interests.
REFERENCES
- 1.Spatzal T; Perez KA; Einsle O; Howard JB; Rees DC, Ligand binding to the FeMo-cofactor: Structures of CO-bound and reactivated nitrogenase. Science 2014, 345 (6204), 1620–1623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Hoffman BM; Lukoyanov D; Yang Z-Y; Dean DR; Seefeldt LC, Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage. Chem. Rev 2014, 114 (8), 4041–4062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Kim CH; Newton WE; Dean DR, Role of the MoFe Protein α-Subunit Histidine-195 Residue in FeMo-Cofactor Binding and Nitrogenase Catalysis. Biochemistry 1995, 34 (9), 2798–2808. [DOI] [PubMed] [Google Scholar]
- 4.Shima S; Pilak O; Vogt S; Schick M; Stagni MS; Meyer-Klaucke W; Warkentin E; Thauer RK; Ermler U, The crystal structure of Fe -hydrogenase reveals the geometry of the active site. Science 2008, 321 (5888), 572–575. [DOI] [PubMed] [Google Scholar]
- 5.Hiromoto T; Warkentin E; Moll J; Ermler U; Shima S, The Crystal Structure of an Fe -Hydrogenase-Substrate Complex Reveals the Framework for H-2 Activation. Angew. Chem. Int. Ed 2009, 48 (35), 6457–6460. [DOI] [PubMed] [Google Scholar]
- 6.Yang X; Hall MB, Monoiron Hydrogenase Catalysis: Hydrogen Activation with the Formation of a Dihydrogen, Fe-Hδ−•••Hδ+-O, Bond and Methenyl-H4MPT+ Triggered Hydride Transfer. J. Am. Chem. Soc 2009, 131 (31), 10901–10908. [DOI] [PubMed] [Google Scholar]
- 7.Huang GF; Wagner T; Wodrich MD; Ataka K; Bill E; Ermler U; Hu XL; Shima S, The atomic-resolution crystal structure of activated Fe -hydrogenase. Nat. Catal 2019, 2 (6), 537–543. [Google Scholar]
- 8.Jeoung J-H; Dobbek H, Carbon Dioxide Activation at the Ni,Fe-Cluster of Anaerobic Carbon Monoxide Dehydrogenase. Science 2007, 318 (5855), 1461. [DOI] [PubMed] [Google Scholar]
- 9.Han Z; Horak KT; Lee HB; Agapie T, Tetranuclear Manganese Models of the OEC Displaying Hydrogen Bonding Interactions: Application to Electrocatalytic Water Oxidation to Hydrogen Peroxide. J. Am. Chem. Soc 2017, 139 (27), 9108–9111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Ford CL; Park YJ; Matson EM; Gordon Z; Fout AR, A bioinspired iron catalyst for nitrate and perchlorate reduction. Science 2016, 354 (6313), 741. [DOI] [PubMed] [Google Scholar]
- 11.Yadav V; Gordon JB; Siegler MA; Goldberg DP, Dioxygen-Derived Nonheme Mononuclear FeIII(OH) Complex and Its Reactivity with Carbon Radicals. J. Am. Chem. Soc 2019, 141 (26), 10148–10153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Borovik AS, Bioinspired hydrogen bond motifs in ligand design: The role of noncovalent interactions in metal ion mediated activation of dioxygen. Accounts Chem. Res 2005, 38 (1), 54–61. [DOI] [PubMed] [Google Scholar]
- 13.Moore CM; Szymczak NK, 6,6 '-Dihydroxy terpyridine: a proton-responsive bifunctional ligand and its application in catalytic transfer hydrogenation of ketones. Chem. Commun 2013, 49 (4), 400–402. [DOI] [PubMed] [Google Scholar]
- 14.Dahl EW; Louis-Goff T; Szymczak NK, Second sphere ligand modifications enable a recyclable catalyst for oxidant-free alcohol oxidation to carboxylates. Chem. Commun 2017, 53 (14), 2287–2289. [DOI] [PubMed] [Google Scholar]
- 15.Nichols Eva M.; Derrick JS; Nistanaki SK; Smith PT; Chang CJ, Positional effects of second-sphere amide pendants on electrochemical CO2 reduction catalyzed by iron porphyrins. Chem. Sci 2018, 9 (11), 2952–2960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Costentin C; Drouet S; Robert M; Savéant J-M, A Local Proton Source Enhances CO2 Electroreduction to CO by a Molecular Fe Catalyst. Science 2012, 338 (6103), 90. [DOI] [PubMed] [Google Scholar]
- 17.Azcarate I; Costentin C; Robert M; Savéant J-M, Through-Space Charge Interaction Substituent Effects in Molecular Catalysis Leading to the Design of the Most Efficient Catalyst of CO2-to-CO Electrochemical Conversion. J. Am. Chem. Soc 2016, 138 (51), 16639–16644. [DOI] [PubMed] [Google Scholar]
- 18.Pegis ML; McKeown BA; Kumar N; Lang K; Wasylenko DJ; Zhang XP; Raugei S; Mayer JM, Homogenous Electrocatalytic Oxygen Reduction Rates Correlate with Reaction Overpotential in Acidic Organic Solutions. ACS Cent. Sci 2016, 2 (11), 850–856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Das S; Incarvito CD; Crabtree RH; Brudvig GW, Molecular Recognition in the Selective Oxygenation of Saturated C-H Bonds by a Dimanganese Catalyst. Science 2006, 312 (5782), 1941. [DOI] [PubMed] [Google Scholar]
- 20.Kuninobu Y; Ida H; Nishi M; Kanai M, A meta-selective C–H borylation directed by a secondary interaction between ligand and substrate. Nat. Chem 2015, 7 (9), 712–717. [DOI] [PubMed] [Google Scholar]
- 21.Connor GP; Lease N; Casuras A; Goldman AS; Holland PL; Mayer JM, Protonation and electrochemical reduction of rhodium– and iridium–dinitrogen complexes in organic solution. Dalton Trans. 2017, 46 (41), 14325–14330. [DOI] [PubMed] [Google Scholar]
- 22.Weiss CJ; Egbert JD; Chen S; Helm ML; Bullock RM; Mock MT, Protonation Studies of a Tungsten Dinitrogen Complex Supported by a Diphosphine Ligand Containing a Pendant Amine. Organometallics 2014, 33 (9), 2189–2200. [Google Scholar]
- 23.Morris RH, Bronsted-Lowry Acid Strength of Metal Hydride and Dihydrogen Complexes. Chem. Rev 2016, 116 (15), 8588–8654. [DOI] [PubMed] [Google Scholar]
- 24.Creutz SE; Peters JC, Exploring secondary-sphere interactions in Fe-NxHy complexes relevant to N2 fixation. Chem. Sci 2017, 8 (3), 2321–2328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Shanahan JP; Szymczak NK, Hydrogen Bonding to a Dinitrogen Complex at Room Temperature: Impacts on N2 Activation. J. Am. Chem. Soc 2019, 141 (21), 8550–8556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Delgado M; Gilbertson JD, Ligand-based reduction of nitrate to nitric oxide utilizing a proton-responsive secondary coordination sphere. Chem. Commun 2017, 53 (81), 11249–11252. [DOI] [PubMed] [Google Scholar]
- 27.Hartle MD; Delgado M; Gilbertson JD; Pluth MD, Stabilization of a Zn(II) hydrosulfido complex utilizing a hydrogen-bond accepting ligand. Chem. Commun 2016, 52 (49), 7680–7682. [DOI] [PubMed] [Google Scholar]
- 28.Horak KT; Agapie T, Dioxygen Reduction by a Pd(0)–Hydroquinone Diphosphine Complex. J. Am. Chem. Soc 2016, 138 (10), 3443–3452. [DOI] [PubMed] [Google Scholar]
- 29.Kinney RA; Saouma CT; Peters JC; Hoffman BM, Modeling the Signatures of Hydrides in Metalloenzymes: ENDOR Analysis of a Di-iron Fe(μ-NH)(μ-H)Fe Core. J. Am. Chem. Soc 2012, 134 (30), 12637–12647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Warren JJ; Mayer JM, Moving Protons and Electrons in Biomimetic Systems. Biochemistry 2015, 54 (10), 1863–1878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Liu J; Chakraborty S; Hosseinzadeh P; Yu Y; Tian S; Petrik I; Bhagi A; Lu Y, Metalloproteins Containing Cytochrome, Iron–Sulfur, or Copper Redox Centers. Chem. Rev 2014, 114 (8), 4366–4469. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Hosseinzadeh P; Lu Y, Design and fine-tuning redox potentials of metalloproteins involved in electron transfer in bioenergetics. Biochim. Biophys. Acta, Bioenerg 2016, 1857 (5), 557–581. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Connelly NG; Geiger WE, Chemical Redox Agents for Organometallic Chemistry. Chem. Rev 1996, 96 (2), 877–910. [DOI] [PubMed] [Google Scholar]
- 34.Geri JB; Wolfe MMW; Szymczak NK, The Difluoromethyl Group as a Masked Nucleophile: A Lewis Acid/Base Approach. J. Am. Chem. Soc 2018, 140 (30), 9404–9408. [DOI] [PubMed] [Google Scholar]
- 35.Zafrani Y; Yeffet D; Sod-Moriah G; Berliner A; Amir D; Marciano D; Gershonov E; Saphier S, Difluoromethyl Bioisostere: Examining the "Lipophilic Hydrogen Bond Donor" Concept. J. Med. Chem 2017, 60 (2), 797–804. [DOI] [PubMed] [Google Scholar]
- 36.Sessler CD; Rahm M; Becker S; Goldberg JM; Wang F; Lippard SJ, CF2H, a Hydrogen Bond Donor. J. Am. Chem. Soc 2017, 139 (27), 9325–9332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Malquin N; Rahgoshay K; Lensen N; Chaume G; Miclet E; Brigaud T, CF2H as a hydrogen bond donor group for the fine tuning of peptide bond geometry with difluoromethylated pseudoprolines. Chem. Commun 2019, 55, 12487–12490. [DOI] [PubMed] [Google Scholar]
- 38.Dahl EW; Kiernicki JJ; Zeller M; Szymczak NK, Hydrogen Bonds Dictate O2 Capture and Release within a Zinc Tripod. J. Am. Chem. Soc 2018, 140 (32), 10075–10079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Dahl EW; Dong HT; Szymczak NK, Phenylamino derivatives of tris(2-pyridylmethyl)amine: hydrogen-bonded peroxodicopper complexes. Chem. Commun 2018, 54 (8), 892–895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Dahl EW; Szymczak NK, Hydrogen Bonds Dictate the Coordination Geometry of Copper: Characterization of a Square-Planar Copper(I) Complex. Angew. Chem.-Int. Edit 2016, 55 (9), 3101–3105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Geri JB; Szymczak NK, A Proton-Switchable Bifunctional Ruthenium Complex That Catalyzes Nitrile Hydroboration. J. Am. Chem. Soc 2015, 137 (40), 12808–12814. [DOI] [PubMed] [Google Scholar]
- 42.Moore CM; Szymczak NK, Nitrite reduction by copper through ligand-mediated proton and electron transfer. Chem. Sci 2015, 6 (6), 3373–3377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Moore CM; Szymczak NK, Redox-induced fluoride ligand dissociation stabilized by intramolecular hydrogen bonding. Chem. Commun 2015, 51 (25), 5490–5492. [DOI] [PubMed] [Google Scholar]
- 44.Moore CM; Quist DA; Kampf JW; Szymczak NK, A 3-Fold-Symmetric Ligand Based on 2-Hydroxypyridine: Regulation of Ligand Binding by Hydrogen Bonding. Inorg. Chem 2014, 53 (7), 3278–3280. [DOI] [PubMed] [Google Scholar]
- 45.Tomas-Mendivil E; Diez J; Cadierno V, Palladium(II) complexes with symmetrical dihydroxy-2,2'-bipyridine ligands: Exploring their inter- and intramolecular interactions in solid-state. Polyhedron 2013, 59, 69–75. [Google Scholar]
- 46.Conifer CM; Taylor RA; Law DJ; Sunley GJ; White AJP; Britovsek GJP, First metal complexes of 6,6 '-dihydroxy-2,2 '-bipyridine: from molecular wires to applications in carbonylation catalysis. Dalton Trans. 2011, 40 (5), 1031–1033. [DOI] [PubMed] [Google Scholar]
- 47.Stahl SS; Thorman JL; Nelson RC; Kozee MA, Oxygenation of Nitrogen-Coordinated Palladium(0): Synthetic, Structural, and Mechanistic Studies and Implications for Aerobic Oxidation Catalysis. J. Am. Chem. Soc 2001, 123 (29), 7188–7189. [DOI] [PubMed] [Google Scholar]
- 48.Zhu S-Q; Xu X-H; Qing F-L, Silver-mediated oxidative C–H difluoromethylation of phenanthridines and 1,10-phenanthrolines. Chem. Commun 2017, 53 (83), 11484–11487. [DOI] [PubMed] [Google Scholar]
- 49.Related studies with 6,6’-bis(hydroxyl)-2,2’-bipyridine (bpyOH) (Refs 44-46), and phenanthroline deriatives (Ref 47) establish the viability of H-bonding within this ligand.
- 50.A riding model was used to model the –CF2H proton, however, the orientation toward –Cl is defined by Cl⋯C-F angles of 137.9(8)° and 88.0(6)°.
- 51.1-Cl exhibits a longer –E(H)⋯Cl contact, relative to analogous 2-hydroxypyridine derivatives (O-Cl = 2.856 – 2.861 Å) (Ref 44), consistent with a weaker attractive interaction from the –CF2H donor than traditional H-bond donors.
- 52.Bondi A, van der Waals Volumes and Radii. J. Phys. Chem 1964, 68 (3), 441–451. [Google Scholar]
- 53.Brammer L; Bruton EA; Sherwood P, Understanding the Behavior of Halogens as Hydrogen Bond Acceptors. Cryst. Growth Des 2001, 1 (4), 277–290. [Google Scholar]
- 54.Ball ND; Kampf JW; Sanford MS, Synthesis and reactivity of palladium(II) fluoride complexes containing nitrogen-donor ligands. Dalton Trans. 2010, 39 (2), 632–640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Fraser SL; Antipin MY; Khroustalyov VN; Grushin VV, Molecular fluoro palladium complexes. J. Am. Chem. Soc 1997, 119 (20), 4769–4770. [Google Scholar]
- 56.Grushin VV, Palladium fluoride complexes: One more step toward metal-mediated C-F bond formation. Chem.-Eur. J 2002, 8 (5), 1006–1014. [DOI] [PubMed] [Google Scholar]
- 57.Grushin VV, Thermal stability, decomposition paths, and Ph/Ph exchange reactions of [(Ph3P)2Pd(Ph)X] (X = I, Br, Cl, F, and HF2). Organometallics 2000, 19 (10), 1888–1900. [Google Scholar]
- 58.Grushin VV; Marshall WJ, trans-Difluoro Complexes of Palladium(II). J. Am. Chem. Soc 2009, 131 (3), 918. [DOI] [PubMed] [Google Scholar]
- 59.Nahra F; Brill M; Gomez-Herrera A; Cazin CSJ; Nolan SP, Transition metal bifluorides. Coord. Chem. Rev 2016, 307, 65–80. [Google Scholar]
- 60.Similar to related PdF2 complexes, 1-F was unstable to isolation (refs 49-54).
- 61.The coupling of 1-F is similar to that observed for HF⋯HF (Ref 62), 1JHF = 27-31 Hz.
- 62.Pecul M; Sadlej J; Leszczynski J, The 19F-1H coupling constants transmitted through covalent, hydrogen bond, and van der Waals interactions. J. Chem. Phys 2001, 115 (12), 5498–5506. [Google Scholar]
- 63.Del Bene JE; Perera SA; Bartlett RJ, Hydrogen Bond Types, Binding Energies, and 1H NMR Chemical Shifts. . J. Phys. Chem. A 1999, 103 (40), 8121–8124. [Google Scholar]
- 64.Johnson ER; Keinan S; Mori-Sánchez P; Contreras-García J; Cohen AJ; Yang W, Revealing Noncovalent Interactions. J. Am. Chem. Soc 2010, 132 (18), 6498–6506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Joseph J; Jemmis ED, Red-, blue-, or no-shift in hydrogen bonds: A unified explanation. J. Am. Chem. Soc 2007, 129 (15), 4620–4632. [DOI] [PubMed] [Google Scholar]
- 66.Garner DK; Allred RA; Tubbs KJ; Arif AM; Berreau LM, Synthesis and Characterization of Mononuclear Zinc Aryloxide Complexes Supported by Nitrogen/Sulfur Ligands Possessing an Internal Hydrogen Bond Donor. Inorg. Chem 2002, 41 (13), 3533–3541. [DOI] [PubMed] [Google Scholar]
- 67.Lubben J; Volkmann C; Grabowsky S; Edwards A; Morgenroth W; Fabbiani FPA; Sheldrick GM; Dittrich B, On the temperature dependence of H-Uiso in the riding hydrogen model. Acta Crystallogr., Sect. A 2014, 70 (4), 309–316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.This maximum angle is fixed based on the geometry of the intramolecular 6-member ring.
- 69.Lu N; Wei R-J; Lin K-Y; Alagesan M; Wen Y-S; Liu L-K, Weak hydrogen bonding and fluorous interactions in the chloride and bromide salts of 4-[(2,2,3,3-tetrafluoropropoxy)methyl]pyridinium. Acta Crystallogr., Sect. C 2017, 73 (4), 343–349. [DOI] [PubMed] [Google Scholar]
- 70.Arduengo Iii AJ; Calabrese JC; Davidson F; Rasika Dias HV; Goerlich JR; Krafczyk R; Marshall WJ; Tamm M; Schmutzler R, C–H Insertion Reactions of Nucleophilic Carbenes. Helv. Chim. Acta 1999, 82 (12), 2348–2364. [Google Scholar]
- 71.The steric effect is inherent and includes the compounds reported herein.
- 72.Charton M, Steric effects. I. Esterification and acid-catalyzed hydrolysis of esters. J. Am. Chem. Soc 1975, 97 (6), 1552–1556. [Google Scholar]
- 73.Additionally, –CH3 spectroscopically (νCH) convolutes the contribution of the H-bond donor H atom with the two non-interacting contributions via signal averaging of vibrational mode mixing.
- 74.Goldberg K; Groombridge S; Hudson J; Leach AG; MacFaul PA; Pickup A; Poultney R; Scott JS; Svensson PH; Sweeney J, Oxadiazole isomers: all bioisosteres are not created equal. MedChemComm 2012, 3 (5), 600–604. [Google Scholar]
- 75.iPr has a closer steric Charton parameter to –CF2H than –CH3 and does not have the complicating signal averaging (1H NMR) or vibrational mode mixing (IR) anticipated for –CH3.
- 76.Wang C; Danovich D; Shaik S; Mo Y, A Unified Theory for the Blue- and Red-Shifting Phenomena in Hydrogen and Halogen Bonds. J. Chem. Theory Comput 2017, 13 (4), 1626–1637. [DOI] [PubMed] [Google Scholar]
- 77.The νC-H for 1-X feature hypso- to bathochromic shift with increase H-bond strength (See SI).
- 78.Similar to the 1H NMR analysis of 3-OArR, the vibrational analysis shows distinct behavior for stronger H-bond acceptors, consistent with the angular dependence of the interaction.
- 79.These values diverge from typical H-bond behavior analogously to 1-X.
- 80.Lee D-H; Kwon HJ; Patel BP; Liable-Sands LM; Rheingold AL; Crabtree RH, Effects of a Nonligating Pendant Hydrogen-Bonding Group in a Metal Complex: Stabilization of an HF Complex. Organometallics 1999, 18 (9), 1615–1621. [Google Scholar]
- 81.Upon cooling 7-F to −80 °C no 1JFH through-space coupling is observed.
- 82.In contrast to intermolecular H-bond adducts where formation enthalpies can be easily calculated, intramolecular H-bonds have do not have a corresponding formation enthalpy for association.
- 83.We have previously used similar approximations for secondary-sphere H-bond interactions (see Ref. 43).
- 84.While the primary sphere is unperturbed, the –iPr and –CF2H groups orient in different ways, likely reflecting the difference between an attractive interaction preference.
- 85.This calculated value may be an underestimated due to the slight Pd-F2 elongation (0.010 Å) for phenCF2H which would result in a higher HPd-F.
- 86.At −70 °C, the 19F NMR spectrum of 2 is consistent with >3 distinct –CF2H environments.
- 87.Pd° complexes with appended –OH groups are known (Ref 28) and we attribute decomposition to both ligand type and close positioning of the –OH donor group to the metal center.
- 88.The observed reduction for phenol is a function of electrode and the Pt working electrode used is representative of a high potential for the reduction.
- 89.The free ligands, phenCF2H and 1,10 phenanthroline (phen), featured irreversible reductions at <-2.0 V (vs Fc/Fc+) respectively (See SI). The similarity in potential and irreversible behavior are consistent with a small electron withdrawing effect of the –CF2H group on a phenanthroline based reduction.
- 90.Abla M; Yamamoto T, Kinetic Study of Ligand Exchange Reactions of Bis(1,5-cyclooctadiene)nickel(0) with 2,2-Bipyridine, 4,4-Dimethyl-2,2-bipyridine, and 4,4,5,5-Tetramethyl-2,2-bipyridine. Bull. Chem. Soc. Jpn 1999, 72 (6), 1255–1261. [Google Scholar]
- 91.Powers DC; Anderson BL; Nocera DG, Two-Electron HCl to H2 Photocycle Promoted by Ni(II) Polypyridyl Halide Complexes. J. Am. Chem. Soc 2013, 135 (50), 18876–18883. [DOI] [PubMed] [Google Scholar]
- 92.The increase in intensity throughout the spectrum is indicative of the broad features of Ni(phenR)2 complexes and may indicate equilibration between the structures of Ni(phenCF2H)(COD) and Ni(phenCF2H)2(See previous reference).
- 93.The 1H NMR spectrum of 8 exhibits broad resonances at 5.48 and a triplet at 9.19 ppm, consistent with olefinic resonances of a COD ligand and the –CF2H in a dynamic coordination environment at Ni and formation of Ni(phenCF2H)(COD).
- 94.Mohadjer Beromi M; Brudvig GW; Hazari N; Lant HMC; Mercado BQ, Synthesis and Reactivity of Paramagnetic Nickel Polypyridyl Complexes Relevant to C(sp2)–C(sp3) Coupling Reactions. Angew. Chem. Int. Ed 2019, 58 (18), 6094–6098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.One of the two –CF2I⋯H contacts in the crystal structure is within the vdW radii, and may reflect H-bond formation.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.