H2 Binding, Splitting, and Net Hydrogen Atom Transfer at a

Jan 23, 2019 - US State of the Union disappoints science advocates. US President Donald J. Trump nodded to science and women in his second State of...
0 downloads 0 Views 905KB Size
Subscriber access provided by Karolinska Institutet, University Library

Communication

H2 Binding, Splitting, and Net Hydrogen Atom Transfer at a Paramagnetic Iron Complex Demyan E. Prokopchuk, Geoffrey M. Chambers, Eric D. Walter, Michael T. Mock, and R. Morris Bullock J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 23 Jan 2019 Downloaded from http://pubs.acs.org on January 23, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

H2 Binding, Splitting, and Net Hydrogen Atom Transfer at a Paramagnetic Iron Complex Demyan E. Prokopchuk,‡,† Geoffrey M. Chambers,‡ Eric D. Walter,§ Michael T. Mock,‡ and R. Morris Bullock*,‡ Center for Molecular Electrocatalysis, Pacific Northwest National Laboratory, Richland, WA 99352, United States § Environmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, WA 99352, United States ‡

ABSTRACT: While diamagnetic transition metal complexes that bind and split H 2 have been extensively studied, paramagnetic complexes that exhibit this behavior remain rare. The square planar S = ½ FeI(P4N2)+ cation (FeI+) reversibly binds H2/D2 in solution, exhibiting an inverse equilibrium isotope effect of KH2/KD2 = 0.58(4) at -5.0 °C. In the presence of excess H2, the dihydrogen complex FeI(H2)+ cleaves H2 at 25 °C in a net hydrogen atom transfer reaction, producing the dihydrogen-hydride trans-FeII(H)(H2)+. The proposed mechanism of H2 splitting involves both intra- and intermolecular steps, resulting in a mixed first- and second-order rate law with respect to initial [FeI+]. The key intermediate is a paramagnetic dihydride complex, trans-FeIII(H)2+, whose weak FeIII-H bond dissociation free energy (calculated BDFE = 44 kcal/mol) leads to bimetallic H-H homolysis, generating trans-FeII(H)(H2)+. Reaction kinetics, thermodynamics, electrochemistry, EPR spectroscopy, and DFT calculations support the proposed mechanism.

The coordination and reactivity of H2 ligands in diamagnetic transition metal complexes has been intensely studied for decades.1 Recent efforts in sustainable energy have focused on using dihydrogen (H2) or protons/electrons (H+/e-) as energy carriers that are interconverted using molecular electrocatalysts.2 The thermodynamic bias for electrocatalytic H2 production or the reverse reaction, H2 oxidation, can be strongly influenced by modification of the ligand. Reactivity of these diamagnetic dihydrogen complexes is easily probed by NMR spectroscopy.3 Reactions of paramagnetic metal complexes with H2 have important implications for biological processes involving hydrogenase4 and nitrogenase5 enzymes, yet detailed reports of H2 binding/splitting with paramagnetic complexes are rare.1a,6 H2 binding to paramagnetic metal complexes has been postulated in a few instances,7 and three paramagnetic H2 complexes have been thoroughly characterized using advanced EPR spectroscopic techniques (Figure 1A-B).8 The S = ½ FeI(η2-H2)(SiPiPr3) and Co0(η2-H2)(BPiPr3) complexes (A) reversibly bind H2. The recently reported complex [FeI(H2)(depe)2]+ (B) reacts irreversibly with H2 to generate trans-[FeII(H)(H2)(depe)2]+. In terms of H2 cleavage by open-shell complexes, early kinetic studies of [CoII(CN)5]3- 9 and Co2(CO)810 provide evidence for a bimetallic transition state, [M---H---H---M]‡, where H2 homolysis furnishes diamagnetic metal hydride products.11 For example, the porphyrin complex RhII(TMP) (TMP = tetramesitylporphyrinato) cleaves H2 to generate diamagnetic RhIIIH(TMP), as demonstrated by clean second-

Figure 1. Selected paramagnetic M-H2 complexes (A, B) and terminal M-H complexes (C, D).

order kinetics with respect to [Rh].11d,11e Reduction of the three-coordinate complex LtBuFeIICl (L = bulky βdiketiminate ligand) by 1 equiv. KC8 followed by exposure to H2 yields the bridging dihydride complex (LtBuFeIIH)2,7e which is in equilibrium with the terminal hydride complex LtBuFeIIH.12 However, a discrete dihydrogen complex was not observed in any of these cases. Open-shell terminal metal hydride complexes are often unstable at room temperature, with M-H cleavage occurring by release of a proton or hydrogen atom.6a,13 In some instances, electrochemically induced hydrogen atom loss has been observed, producing H2 and diamagnetic products (Figure 1C-D).14 The anion RuIIH(OEP)- (C)14a,14b is electrochemically oxidized to generate an open-shell metal hydride complex that undergoes second-order M-H homolysis to release H2. Recently, the thermally sensitive S = ½ FeIIIH(N2)(PPSi-thiolate) complex (D) was

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

characterized, with kinetic studies of H2 loss revealing a second-order dependence on [Fe].14c Notwithstanding the examples of open-shell dihydrogen or hydride complexes mentioned above, the binding of H2 and its reactivity in paramagnetic complexes remains poorly understood. We previously reported the relationship between N2 binding affinity, metal oxidation state, and catalytic N2 silylation activity for a series of [Fen(P4N2)]n+ complexes (n = 0, 1, 2), where P4N2 is a tetradentate ligand.15 We now report studies of H2 binding and splitting at d7 [FeI(P4N2)]+ (FeI+) using kinetic, thermodynamic, spectroscopic, and computational studies. Our data support a mechanism involving intramolecular H2 cleavage at FeI and net hydrogen atom transfer via bimetallic H-H coupling at FeIII to generate trans-[FeII(H)(H2)(P4N2)]+ (transFeII(H)(H2)+). Exposing deep purple [FeI(P4N2)][B(C6F5)4] (FeI+)15 to H2 or D2 (1 atm) in fluorobenzene results in an equilibrium with the S = ½ adduct FeI(H2)+ (FeI(D2)+). Monitoring the temperature-dependent equilibria by UV-vis spectroscopy from 268-288 K (Figures 2A and S17) leads to a van’t Hoff analysis of FeI+/FeI(H2)+ (ΔG268 = -0.20(7) kcal/mol, ΔH = 2.48(7) kcal/mol, ΔS = -8.5(2) cal/mol・K) and FeI+/FeI(D2)+ (ΔG268 = -0.49(2) kcal/mol, ΔH = -5.49(2) kcal/mol, ΔS = 18.6(5) cal/mol・K; Figure 2B). Binding of D2 is thermodynamically more favorable than H2 (ΔΔG268 = 0.29(7) kcal/mol, ΔΔH = -3.01(7) kcal/mol, ΔΔS = -10.1(5) cal/mol・K), resulting in an inverse equilibrium isotope effect (EIE); KH2/KD2 = 0.58(4) at 268K. While inverse EIEs are well-established for diamagnetic H2 complexes,1b,16 this is the first reported EIE for H2 binding to a paramagnetic dihydrogen complex. Cooling a purple solution of FeI+ under H2 or D2 below 268K results in a color change to pale yellow, which we attribute to the exclusive formation of FeI(H2)+ (FeI(D2)+), based on the thermodynamic data. Consistent

Scheme 1. Reactivity of Fe complexes with H2.

with the assignment of H2 or D2 binding at Fe, experimental EPR spectra were simulated by changing only the coupling constant found for the FeI(H2)+ spectrum (AH = 35 MHz) and dividing by the gyromagnetic ratios to obtain γH/γD ≅ 6.51,17 yielding AD = 5.4 MHz for FeI(D2)+ (Figure 2C).18 Warming a fluorobenzene solution of FeI+/FeI(H2)+ above 288K under H2 slowly and irreversibly produces the d6 dihydrogen-hydride complex, trans-FeII(H)(H2)+ (Scheme 1). This reaction constitutes an unusual net hydrogen atom transfer reaction from H2 to FeI(H2)+. The 1H NMR spectrum of trans-FeII(H)(H2)+ at 25 °C shows characteristic resonances for the η2-H2 ligand at -10.49 ppm (T1 = 26 ms, 298K) and hydride at -15.53 ppm (T1 = 513 ms, 298K), akin to the trans-M(H)(H2)(diphosphine)2 (M = Fe, Ru, Os) family of complexes.19 The dihydrogen and hydride sites of transFeII(H)(H2)+ are fluxional in solution, undergoing intramolecular chemical exchange that can be observed by isotopic labeling experiments (Figure S11). The coupling constant JHD = 31.5 Hz for coordinated HD was used to calculate an H-H distance of 0.89 Å in transFeII(H)(H2)+.1a,1d,20 The H2 ligand in trans-FeII(H)(H2)+ is stable under argon, but N2 (1 atm) readily displaces H2, yielding trans-FeII(H)(N2)+ (see SI).

Figure 2. A: Binding equilibrium of H2/D2 to FeI+ with equilibrium constants KH2 and KD2. B: van’t Hoff plot for H2 (red circles) and D2 (blue squares). C: Experimental (black) and simulated (red) X-band EPR spectra (2-MeTHF glass) of FeI(H2)+ (80K) and FeI(D2)+ (90K).

ACS Paragon Plus Environment

Page 2 of 7

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Kinetic data for the reaction of FeI+ and H2 to generate trans-FeII(H)(H2)+ were obtained by monitoring the decay of FeI+ (λmax = 559 nm) by UV-vis spectroscopy under 1 atm H2 over several days. If H2 cleavage occurs through bimetallic H2 homolysis, a second-order dependence on [FeI+] would be expected; however, the data adheres to neither first- nor second-order kinetics (Figure S19). A loglog plot of the kinetic runs reveals a change in slope during the course of the reaction, which is characteristic of a mixed-order reaction where the second-order kinetic term is dominant at higher [FeI+] and the first-order term is prevalent at lower [FeI+] (Figure 3, top).11b,21 This kinetic behavior is analyzed by taking into account the total concentration of FeI+ and FeI(H2)+ ([FeI]tot) according to eq. 1.

Figure 3. Top: Log-Log plot of the kinetics of FeI+ under H2, showing the change in slope (reaction order) as [FeI+]tot decreases at longer reaction times (total reaction time of 114 h). Bottom: Kinetics to 75% completion under H2, fit to 𝐴 −𝐴 a mixed order integrated rate law where α = 𝑙𝑛 ( 𝑡 ∞ ) + 𝑘1 𝑡 (see eq. 1 and SI ). −

𝑑[𝐹𝑒 𝐼 ]𝑡𝑜𝑡 𝑑𝑡

= 𝑘1 [𝐹𝑒 𝐼 ]𝑡𝑜𝑡 + 𝑘2 [𝐹𝑒 𝐼 ]2𝑡𝑜𝑡

𝐴0 −𝐴∞

(1)

An excellent fit is achieved by plotting eα vs. Absorbance (Figure 3, bottom). The observed rate constant k1 is obtained by plotting the kinetic data as first-order dependent and measuring the slope at low [FeI+]tot (>75% completion), which gives k1 = 5.0(7) × 10-6 s-1. Analysis of this data using the integrated rate law gives the observed second-order rate constant k2 = 6(1) × 10-2 M-1 s-1 (see SI for derivation). Experiments with D2 follow clean first-order kinetics (k1D = 1.8 × 10-6 s-1), indicating that the first-order process is significantly slower than the second-order process under these conditions. Thus, a KIE of k1H/k1D = 2.8 is observed for the first-order process with respect to [FeI+]tot.

After formation of FeI(H2)+, we rationalized that the firstorder kinetic term could involve either pendant amineassisted H2 heterolysis2b or oxidative addition to generate a d5 FeIII(H)2+ intermediate (Figure 4A). DFT calculations indicate that oxidative addition is energetically preferred, proceeding through a proposed dihydride intermediate cisFeIII(H)2+ with ΔG‡calc = 22.7 kcal/mol. The computed barrier is consistent with our kinetic data, where ΔG‡exp(k1) = 24.7(1) kcal/mol, calculated using conventional transition state theory. Calculations also indicate that heterolytic H2 cleavage, forming an Fe-H bond and protonating the pendant amine to form cis-FeI(H)(NH)+ is possible (ΔG‡calc = 18.9 kcal/mol), but an energetically viable pathway to afford trans-FeIII(H)2+ was not found in our computations (Figure S22). Instead, intramolecular rearrangement of the cis-dihydride intermediate yields the intermediate transFeIII(H)2+. The wide bite angle in FeI+ (Ph2P-Fe-PPh2 = 108.71(3)°)15 may also facilitate intramolecular hydride rearrangement without dissociation of a Fe-P bond. The calculated k1H/k1D = 2.4 for oxidative addition (Table S6) is in good agreement with the aforementioned experimental KIE.

Figure 4. A: Computed mechanistic pathways to generate trans-FeIII(H)2+ with computed free energies relative to FeI(H2)+. B: Synthetic procedure yielding FeIII(H)2+. C: Molecular structure of FeII(H)2 with 50% probability ellipsoids; most H atoms are not shown. D: Experimental (black) and simulated (red) X-band EPR spectra (PhF, 105K) of trans-FeIII(H)2+.

We sought experimental support for the proposed d5 intermediate trans-FeIII(H)2+. The diamagnetic dihydride complex, trans-FeII(H)2, was easily prepared in 97% yield by exposing the d8 complex Fe0(N2)15 to 1 atm H2 in THF (Scheme 1 and Figure 4C). Cyclic voltammograms of transFeII(H)2 in fluorobenzene indicate that trans-FeIII(H)2+ is stable on the CV timescale, with E1/2(III/II) = -0.34 V vs. Cp2Fe0/+ at scan rates >30 V/s (Figure S15). Monitoring the chemical oxidation of trans-FeII(H)2 by EPR spectroscopy supports the formation of a thermally sensitive low spin d5 complex (g = 2.001, 2.079, 2.198)22 that we assign to trans-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

FeIII(H)2+ (Figures 4B and 4D). Therefore, the combined computational, kinetic, and spectroscopic data suggest that the first-order step in the H2 cleavage step involves formation of trans-FeIII(H)2+ by oxidative addition at FeI. We then turned our attention to understanding the second-order rate term that forms trans-FeII(H)(H2)+. When chemical oxidation of trans-FeII(H)2 is performed at 298K in a sealed tube under 1 atm H2, trans-FeII(H)(H2)+ is generated in 69% yield (Scheme 1, right). This observation provides compelling evidence that oxidation to generate trans-FeIII(H)2+ triggers bimetallic H-H coupling in the presence of excess H2 to generate trans-FeII(H)(H2)+. The calculated weak Fe-H bond dissociation free energy (BDFE) of trans-FeIII(H)2+ (44 kcal/mol) indicates spontaneous H2 release is thermodynamically favored (see SI). The FeIII/II redox couple is only reversible at fast scan rates (>30 V/s), however, which is inconsistent with rate constant k2. We suggest that a rapid electrochemically induced sidereaction produces an unknown intermediate, precluding us from obtaining meaningful kinetic information via electrochemical analysis (Figures S15, S16).14b,23 The proposed mechanism of H2 cleavage is shown in Scheme 2. Initially, dihydrogen coordination through the FeI+/FeI(H2)+ equilibrium generates the oxidative addition product, trans-FeIII(H)2+, with rate constant k1. Next, the mechanism could bifurcate into kinetically indistinguishable H atom transfer routes for the formation of trans-FeII(H)(H2)+ via rate constant k2, one of which has been discussed above and is shown in Scheme 2 (right). The second mechanism considered involves comproportionation through bimetallic hydrogen atom transfer from trans-FeIII(H)2+ to FeI(H2)+ (Scheme 2, left). To test this hypothesis, we treated trans-FeII(H)2 with Cp2Fe+ for 3 min at 25 °C under H2 to generate transFeIII(H)2+ in situ, followed by addition of FeI+ (forming FeI(H2)+ in situ) dissolved in fluorobenzene (Scheme 1, middle). Formation of trans-FeII(H)(H2)+ occurs in 65% yield, indicating that both bimetallic mechanistic pathways Scheme 2. Proposed mechanism of H2 splitting by FeI+. KH2 FeI+

H H

+

+

H

k1

+

H N

P P

=

FeIII

FeI

Ph

H

Ph

N

FeIII H

2+

HFe H H FeH

(H2)FeIIH+ +

FeIIH+

+ H2

H

FeII H H

+

k3 k-3

H H

FeII H

+

Ph

The Supporting Information is available free of charge on the ACS Publications website. Experimental details, syntheses, spectra (NMR, IR), DFT structures and Cartesian coordinates (PDF) Crystallographic data (CIF)

AUTHOR INFORMATION †Current

address: Department of Chemistry, Rutgers University, 73 Warren Street, Newark, NJ 07102, United States  Current address: Department of Chemistry and Biochemistry, Montana State University, Bozeman, MT 59717, United States

Corresponding Author

The authors declare no competing financial interest.

ACKNOWLEDGMENTS

k2 (H2)Fe H FeH

Supporting Information

Notes

FeIII(H)2+

2+

ASSOCIATED CONTENT

* [email protected]

PPh2 PPh2

BDFEFeH = 44 kcal/mol FeI(H2)+

(rate constant k2) appear to be possible. After bimetallic hydrogen atom transfer, one equivalent of transFeII(H)(H2)+ is produced, and H2 coordinates to the unobserved FeIIH+, generating the final product, transFeII(H)(H2)+. To further probe the intramolecular dihydrogen-hydride exchange in trans-FeII(H)(H2)+ mentioned earlier, high temperature 1H-1H-EXSY NMR experiments on transFeII(H)(H2)+ at 100 °C in C6D5Cl confirm the presence of chemical exchange cross-peaks with a rate constant k3(k-3) of ca. 7 s-1 (ΔG‡373 = 20 kcal/mol; Table S1). DFT calculations indicate that this process involves a transient sevencoordinate FeIV(H)3+ cation24 with a calculated free energy barrier of 20.5 kcal/mol for oxidative addition, in excellent agreement with experiment. (Scheme 2 and Figure S23). Complementary kinetic, spectroscopic, electrochemical, and computational evidence provide strong support for the mechanism of H2 cleavage at a paramagnetic FeI complex. Detailed kinetic analysis reveals a mixed first- and secondorder rate law that involves reversible dihydrogen coordination at FeI, FeI/FeIII oxidative addition of H2, and net hydrogen atom transfer involving an observable FeIII transdihydride intermediate with a weak Fe-H BDFE. We hope that these results provide a foundation for discovery of new reactivity of paramagnetic H2 complexes.

+

H N

via

P P

FeIV

N

H

PPh2 H PPh2

Ph DG‡exp = 20.0 kcal/mol DG‡calc = 20.5 kcal/mol

This research was supported as part of the Center for Molecular Electrocatalysis, an Energy Frontier Research Center funded by the U.S. Department of Energy (DOE), Office of Science, Office of Basic Energy Sciences. EPR experiments were performed using Environmental Molecular Sciences Laboratory, a national scientific user facility sponsored by the DOE’s Office of Biological and Environmental Research and located at Pacific Northwest National Laboratory (PNNL). PNNL is operated by Battelle for the U.S. DOE. Computational resources were provided by the National Energy Research Scientific Computing Center (NERSC) at Lawrence Berkeley National Laboratory. We thank Dr. Thilina Gunasekara and Dr. Andrew Preston for helpful discussions on the analysis of the kinetics. This work is dedicated to the memory of Prof. Jack Halpern

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society (1925-2018), whose ground-breaking studies of H2 and metal hydrides provided remarkable insights.

REFERENCES (1) (a) Jessop, P. G.; Morris, R. H. Reactions of transition metal dihydrogen complexes. Coord. Chem. Rev. 1992, 121, 155-284; (b) Heinekey, D. M.; Oldham, W. J., Jr., Coordination Chemistry of Dihydrogen. Chem. Rev. 1993, 93, 913-926; (c) Kubas, G. J. Fundamentals of H2 Binding and Reactivity on Transition Metals Underlying Hydrogenase Function and H2 Production and Storage. Chem. Rev. 2007, 107, 4152-4205; (d) Morris, R. H. Dihydrogen, dihydride and in between: NMR and structural properties of iron group complexes. Coord. Chem. Rev. 2008, 252, 2381-2394; (e) Crabtree, R. H. Dihydrogen Complexation. Chem. Rev. 2016, 116, 8750-8769. (2) (a) Bullock, R. M.; Appel, A. M.; Helm, M. L. Production of hydrogen by electrocatalysis: making the H-H bond by combining protons and hydrides. Chem. Commun. 2014, 50, 3125-3143; (b) Bullock, R. M.; Helm, M. L. Molecular Electrocatalysts for Oxidation of Hydrogen Using Earth-Abundant Metals: Shoving Protons Around with Proton Relays. Acc. Chem. Res. 2015, 48, 2017-2026; (c) Lee, K. J.; Elgrishi, N.; Kandemir, B.; Dempsey, J. L. Electrochemical and spectroscopic methods for evaluating molecular electrocatalysts. Nat. Rev. Chem. 2017, 1, 0039; (d) Robinson, S. J. C.; Heinekey, D. M. Hydride & dihydrogen complexes of earth abundant metals: structure, reactivity, and applications to catalysis. Chem. Commun. 2017, 53, 669-676; (e) Bellini, M.; Bevilacqua, M.; Marchionni, A.; Miller, H. A.; Filippi, J.; Grützmacher, H.; Vizza, F. Energy Production and Storage Promoted by Organometallic Complexes. Eur. J. Inorg. Chem. 2018, 2018, 4393-4412. (3) (a) Yang, J. Y.; Bullock, R. M.; Shaw, W. J.; Twamley, B.; Fraze, K.; DuBois, M. R.; DuBois, D. L. Mechanistic Insights into Catalytic H 2 Oxidation by Ni Complexes Containing a Diphosphine Ligand with a Positioned Amine Base. J. Am. Chem. Soc. 2009, 131, 5935-5945; (b) Liu, T.; Chen, S.; O'Hagan, M. J.; Rakowski DuBois, M.; Bullock, R. M.; DuBois, D. L. Synthesis, Characterization, and Reactivity of Fe Complexes Containing Cyclic Diazadiphosphine Ligands: The Role of the Pendant Base in Heterolytic Cleavage of H2. J. Am. Chem. Soc. 2012, 134, 62576272; (c) Liu, T.; DuBois, D. L.; Bullock, R. M. An iron complex with pendent amines as a molecular electrocatalyst for oxidation of hydrogen. Nat. Chem. 2013, 5, 228--233; (d) Hulley, E. B.; Welch, K. D.; Appel, A. M.; DuBois, D. L.; Bullock, R. M. Rapid, Reversible Heterolytic Cleavage of Bound H2. J. Am. Chem. Soc. 2013, 135, 11736-11739; (e) Hulley, E. B.; Helm, M. L.; Bullock, R. M. Heterolytic cleavage of H 2 by bifunctional manganese(I) complexes: impact of ligand dynamics, electrophilicity, and base positioning. Chem. Sci. 2014, 5, 4729-4741; (f) Zhang, S.; Appel, A. M.; Bullock, R. M. Reversible Heterolytic Cleavage of the H–H Bond by Molybdenum Complexes: Controlling the Dynamics of Exchange Between Proton and Hydride. J. Am. Chem. Soc. 2017, 139, 7376-7387. (4) (a) Lubitz, W.; Ogata, H.; Rüdiger, O.; Reijerse, E. Hydrogenases. Chem. Rev. 2014, 114, 4081-4148; (b) Schilter, D.; Camara, J. M.; Huynh, M. T.; Hammes-Schiffer, S.; Rauchfuss, T. B. Hydrogenase Enzymes and Their Synthetic Models: The Role of Metal Hydrides. Chem. Rev. 2016, 116, 8693-8749. (5) (a) Hoffman, B. M.; Lukoyanov, D.; Yang, Z.-Y.; Dean, D. R.; Seefeldt, L. C. Mechanism of Nitrogen Fixation by Nitrogenase: The Next Stage. Chem. Rev. 2014, 114, 4041-4062; (b) Lukoyanov, D.; Yang, Z.-Y.; Khadka, N.; Dean, D. R.; Seefeldt, L. C.; Hoffman, B. M. Identification of a Key Catalytic Intermediate Demonstrates That Nitrogenase Is Activated by the Reversible Exchange of N 2 for H2. J. Am. Chem. Soc. 2015, 137, 3610-3615; (c) Lukoyanov, D.; Khadka, N.; Yang, Z.-Y.; Dean, D. R.; Seefeldt, L. C.; Hoffman, B. M. Reductive Elimination of H 2 Activates Nitrogenase to Reduce the N≡N Triple Bond: Characterization of the E4(4H) Janus Intermediate in Wild-Type Enzyme. J. Am. Chem. Soc. 2016, 138, 10674–10683. (6) (a) Hoff, C. D. Thermodynamic and kinetic studies of stable low valent transition metal radical complexes. Coord. Chem. Rev. 2000, 206, 451-467; (b) Capps, K. B.; Bauer, A.; Kiss, G.; Hoff, C. D. The rate and mechanism of oxidative addition of H2 to the Cr(CO)3C5Me5 radical— generation of a model for reaction of H2 with the Co(CO)4 radical. J. Organomet. Chem. 1999, 586, 23-30. (7) (a) Baya, M.; Houghton, J.; Daran, J.-C.; Poli, R.; Male, L.; Albinati, A.; Gutman, M. Synthesis, Structure, and Electrochemical Properties of Sterically Protected Molybdenum Trihydride Redox Pairs: A

Paramagnetic “Stretched” Dihydrogen Complex? Chem. Eur. J. 2007, 13, 5347-5359; (b) Hetterscheid, D. G. H.; Hanna, B. S.; Schrock, R. R. Molybdenum Triamidoamine Systems. Reactions Involving Dihydrogen Relevant to Catalytic Reduction of Dinitrogen. Inorg. Chem. 2009, 48, 8569-8577; (c) Kinney, R. A.; Hetterscheid, D. G. H.; Hanna, B. S.; Schrock, R. R.; Hoffman, B. M. Formation of [HIPTN 3N]Mo(III)H− by Heterolytic Cleavage of H2 as Established by EPR and ENDOR Spectroscopy. Inorg. Chem. 2010, 49, 704-713; (d) Bart, S. C.; Lobkovsky, E.; Chirik, P. J. Preparation and Molecular and Electronic Structures of Iron(0) Dinitrogen and Silane Complexes and Their Application to Catalytic Hydrogenation and Hydrosilation. J. Am. Chem. Soc. 2004, 126, 13794-13807; (e) Dugan, T. R.; Holland, P. L. New routes to low-coordinate iron hydride complexes: The binuclear oxidative addition of H2. J. Organomet. Chem. 2009, 694, 2825-2830. (8) (a) Lee, Y.; Kinney, R. A.; Hoffman, B. M.; Peters, J. C. A Nonclassical Dihydrogen Adduct of S = ½ Fe(I). J. Am. Chem. Soc. 2011, 133, 1636616369; (b) Suess, D. L. M.; Tsay, C.; Peters, J. C. Dihydrogen Binding to Isostructural S = ½ and S = 0 Cobalt Complexes. J. Am. Chem. Soc. 2012, 134, 14158-14164; (c) Gunderson, W. A.; Suess, D. L. M.; Fong, H.; Wang, X.; Hoffmann, C. M.; Cutsail Iii, G. E.; Peters, J. C.; Hoffman, B. M. Free H2 Rotation vs Jahn-Teller Constraints in the Nonclassical Trigonal (TPB)Co-H2 Complex. J. Am. Chem. Soc. 2014, 136, 14998-15009; (d) Doyle, L. R.; Scott, D. J.; Hill, P. J.; Fraser, D.; Myers, W.; White, A. J. P.; Green, J.; Ashley, A. E. Reversible coordination of N 2 and H2 to a homoleptic S =½ Fe(I) diphosphine complex in solution and the solid state. Chem. Sci. 2018, 9, 7362-7369. (9) (a) Vries, B. D. Reactions of the pentacyanocobaltate(II) ion: II. Liquid phase hydrogenation and decomposition. J. Catal. 1962, 1, 489497; (b) Burnett, M. G.; Connolly, P. J.; Kemball, C. The reaction of potassium pentacyanocobaltate(II) with hydrogen. J. Chem. Soc. A 1967, 800-804; (c) Halpern, J.; Pribanic, M. Hydrogenation of pentacyanocobaltate(II) at high pressures. Inorg. Chem. 1970, 9, 26162618. (10) (a) Klingler, R. J.; Rathke, J. W. High-Pressure NMR Investigation of Hydrogen Atom Transfer and Related Dynamic Processes in Oxo Catalysis. J. Am. Chem. Soc. 1994, 116, 4772-4785; (b) Tannenbaum, R.; Dietler, U. K.; Bor, G.; Ungváry, F. Fundamental metal carbonyl equilibria: Reinvestigation of the equilibrium between dicobalt octacarbonyl and cobalt tetracarbonyl hydride under hydrogen pressure. J. Organomet. Chem. 1998, 570, 39-47. (11) (a) Simándi, L. I.; Budó-Záhonyi, É.; Szeverényi, Z. Effect of strong base on the activation of molecular hydrogen by pyridinebis(dimethylglyoximato)cobalt(II). Inorg. Nucl. Chem. Lett. 1976, 12, 237-241; (b) Chao, T.-H.; Espenson, J. H. Mechanism of hydrogen evolution from hydridocobaloxime. J. Am. Chem. Soc. 1978, 100, 129-133; (c) Van Voorhees, S. L.; Wayland, B. B. Formation of metallo hydride, formyl, and alkyl complexes of Rh(TMTAA). Organometallics 1987, 6, 204-206; (d) Wayland, B. B.; Ba, S.; Sherry, A. E. Reactions of hydrogen or deuterium molecule with a rhodium(II) metalloradical: kinetic evidence for a four-centered transition state. Inorg. Chem. 1992, 31, 148-150; (e) Cui, W.; Wayland, B. B. Activation of C−H / H−H Bonds by Rhodium(II) Porphyrin Bimetalloradicals. J. Am. Chem. Soc. 2004, 126, 8266-8274. (12) Smith, J. M.; Lachicotte, R. J.; and Patrick, L. H. N=N Bond Cleavage by a Low-Coordinate Iron(II) Hydride Complex. J. Am. Chem. Soc. 2003, 125, 15752-15753. (13) (a) Ryan, O. B.; Tilset, M.; Parker, V. D. Chemical and electrochemical oxidation of group 6 cyclopentadienylmetal hydrides. First estimates of 17-electron metal-hydride cation-radical thermodynamic acidities and their decomposition of 17-electron neutral radicals. J. Am. Chem. Soc. 1990, 112, 2618-2626; (b) Poli, R. In Recent Advances in Hydride Chemistry; Peruzzini, M., Poli, R., Eds.; Elsevier: Amsterdam, 2001, p 139-188; (c) Baya, M.; Dub, P. A.; Houghton, J.; Daran, J.-C.; Belkova, N. V.; Shubina, E. S.; Epstein, L. M.; Lledós, A.; Poli, R., "Investigation of the [Cp*Mo(PMe 3)3H]n+ (n = 0, 1) Redox Pair: Dynamic Processes on Very Different Time Scales," Inorg. Chem. 2009, 48, 209-220. (d) Tilset, M.; Fjeldahl, I.; Hamon, J.-R.; Hamon, P.; Toupet, L.; Saillard, J.-Y.; Costuas, K.; Haynes, A. Theoretical, Thermodynamic, Spectroscopic, and Structural Studies of the Consequences of One-Electron Oxidation on the Fe−X Bonds in 17- and 18-Electron Cp*Fe(dppe)X Complexes (X = F, Cl, Br, I, H, CH 3). J. Am. Chem. Soc. 2001, 123, 9984-10000; (e) Hu, Y.; Shaw, A. P.; Estes, D. P.; Norton, J. R. Transition-Metal Hydride Radical Cations. Chem. Rev. 2016, 116, 8427-8462; (f) Kuo, J. L.; Lorenc, C.; Abuyuan, J. M.; Norton,

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

J. R. Catalysis of Radical Cyclizations from Alkyl Iodides under H 2: Evidence for Electron Transfer from [CpV(CO)3H]−. J. Am. Chem. Soc. 2018, 140, 4512-4516. (14) (a) Collman, J. P.; Hutchison, J. E.; Wagenknecht, P. S.; Lewis, N. S.; Lopez, M. A.; Guilard, R. An unprecedented, bridged dihydrogen complex of a cofacial metallodiporphyrin and its relevance to the bimolecular reductive elimination of hydrogen. J. Am. Chem. Soc. 1990, 112, 8206-8208; (b) Collman, J. P.; Wagenknecht, P. S.; Lewis, N. S. Hydride transfer and dihydrogen elimination from osmium and ruthenium metalloporphyrin hydrides: model processes for hydrogenase enzymes and the hydrogen electrode reaction. J. Am. Chem. Soc. 1992, 114, 5665-5673; (c) Gu, N. X.; Oyala, P. H.; Peters, J. C. An S = ½ Iron Complex Featuring N2, Thiolate, and Hydride Ligands: Reductive Elimination of H2 and Relevant Thermochemical Fe–H Parameters. J. Am. Chem. Soc. 2018, 140, 6374-6382. (15) Prokopchuk, D. E.; Wiedner, E. S.; Walter, E. D.; Popescu, C. V.; Piro, N. A.; Kassel, W. S.; Bullock, R. M.; Mock, M. T. Catalytic N 2 Reduction to Silylamines and Thermodynamics of N2 Binding at Square Planar Fe. J. Am. Chem. Soc. 2017, 139, 9291-9301. (16) (a) Bender, B. R.; Kubas, G. J.; Jones, L. H.; Swanson, B. I.; Eckert, J.; Capps, K. B.; Hoff, C. D. Why Does D2 Bind Better Than H2? A Theoretical and Experimental Study of the Equilibrium Isotope Effect on H2 Binding in a M(η2-H2) Complex. Normal Coordinate Analysis of W(CO)3(PCy3)2(η2-H2). J. Am. Chem. Soc. 1997, 119, 9179-9190; (b) Bullock, R. M.; Bender, B. R. In Encyclopedia of Catalysis; Horváth, I., Ed.; John Wiley & Sons, Inc.: New York, 2002; Vol. 4, p 281-348; (c) Parkin, G. Temperature-Dependent Transitions Between Normal and Inverse Isotope Effects Pertaining to the Interaction of H−H and C−H Bonds with Transition Metal Centers. Acc. Chem. Res. 2009, 42, 315-325; (d) Gómez-Gallego, M.; Sierra, M. A. Kinetic Isotope Effects in the Study of Organometallic Reaction Mechanisms. Chem. Rev. 2011, 111, 48574963. (17) Stößer, R.; Herrmann, W. Isotope Effects in ESR Spectroscopy. Molecules 2013, 18, 6679-6722. (18) Stoll, S.; Schweiger, A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 2006, 178, 42-55. (19) (a) Morris, R. H.; Sawyer, J. F.; Shiralian, M.; Zubkowski, J. Two molecular hydrogen complexes: trans-[M(η2H2)(H)(PPh2CH2CH2PPh2)2]BF4 (M = Fe, Ru). The crystal structure determination of the iron complex. J. Am. Chem. Soc. 1985, 107, 55815582; (b) Bautista, M.; Earl, K. A.; Morris, R. H.; Sella, A. NMR properties of the complexes trans-[M(η2-H2)(H)(PEt2CH2CH2PEt2)2]+ (M = Fe, Ru, Os). Intramolecular exchange of atoms between η2-dihydrogen and hydride ligands. J. Am. Chem. Soc. 1987, 109, 3780-3782; (c) Bautista, M.; Earl, K.; Morris, R. NMR Studies of the Complexes trans-[M(η2H2)(H)(Ph2PCH2CH2PEt2)2]X (M=Fe, X = BPh4; M = Os, X = BF4): Evidence for Unexpected Shortening of the H-H Bond. Inorg. Chem. 1988, 27, 1124-1125; (d) Bautista, M. T.; Earl, K. A.; Maltby, P. A.; Morris, R. H.; Schweitzer, C. T.; Sella, A. Estimation of the hydrogenhydrogen distances of η2-dihydrogen ligands in the complexes trans[M(η2-H2)(H)(PR2CH2CH2PR2)2]+ [M = iron, ruthenium, R = Ph, M = osmium, R = Et] by solution NMR methods. J. Am. Chem. Soc. 1988, 110,

7031-7036; (e) Ricci, J. S.; Koetzle, T. F.; Bautista, M. T.; Hofstede, T. M.; Morris, R. H.; Sawyer, J. F. Single-crystal x-ray and neutron diffraction studies of an η2-dihydrogen transition-metal complex: trans-[Fe(η2H2)(H)(PPh2CH2CH2PPh2)2]BPh4. J. Am. Chem. Soc. 1989, 111, 88238827; (f) Bautista, M. T.; Cappellani, E. P.; Drouin, S. D.; Morris, R. H.; Schweitzer, C. T.; Sella, A.; Zubkowski, J. Preparation and spectroscopic properties of the η2-dihydrogen complexes [MH(η2H2)PR2CH2CH2PR2)2] + (M = iron, ruthenium; R = Ph, Et) and trends in properties down the iron group triad. J. Am. Chem. Soc. 1991, 113, 4876-4887; (g) Earl, K. A.; Jia, G.; Maltby, P. A.; Morris, R. H. η2Dihydrogen on the brink of homolytic cleavage: trans-[Os(H--H)H(PEt2CH2CH2PEt2)2]+ has spectroscopic and chemical properties between those of the isoelectronic complexes trans[OsH(PPh2CH2CH2PPh2)2(η2-H2)]+ and ReH3(PPh2CH2CH2PPh2)2. J. Am. Chem. Soc. 1991, 113, 3027-3039; (h) Hills, A.; Hughes, D. L.; JimenezTenorio, M.; Leigh, G. J.; Rowley, A. T. Bis[1,2bis(dimethylphosphino)ethane]dihydrogenhydridoiron(II) tetraphenylborate as a model for the function of nitrogenases. J. Chem. Soc., Dalton Trans. 1993, 3041-3049. (20) Law, J. K.; Mellows, H.; Heinekey, D. M. H−H Distances in Elongated Transition Metal Dihydrogen Complexes:  Effects of Temperature and Isotopic Substitution. J. Am. Chem. Soc. 2001, 123, 2085-2086. (21) (a) Espenson, J. H. Chemical Kinetics and Reaction Mechanisms; McGraw-Hill, New York, 1981; (b) Fettinger, J. C.; Kraatz, H.-B.; Poli, R.; Quadrelli, E. A.; Torralba, R. C. Electrophilic Addition vs Electron Transfer for the Interaction of Ag+ with Molybdenum(II) Hydrides. 1. Reaction with CpMoH(PMe3)3 and the Mechanism of Decomposition of [CpMoH(PMe3)3]+. Organometallics 1998, 17, 5767-5775. (22) (a) Walker, F. A.; Reis, D.; Balke, V. L. Models of the cytochromes b. 5. EPR studies of low-spin iron(III) tetraphenylporphyrins. J. Am. Chem. Soc. 1984, 106, 6888-6898; (b) Rakowsky, M. H.; More, K. M.; Kulikov, A. V.; Eaton, G. R.; Eaton, S. S. Time-Domain Electron Paramagnetic Resonance as a Probe of Electron-Electron Spin-Spin Interaction in Spin-Labeled Low-Spin Iron Porphyrins. J. Am. Chem. Soc. 1995, 117, 2049-2057. (23) (a) Olmstead, M. L.; Hamilton, R. G.; Nicholson, R. S. Theory of cyclic voltammetry for a dimerization reaction initiated electrochemically. Anal. Chem. 1969, 41, 260-267; (b) Fang, M.; Wiedner, E. S.; Dougherty, W. G.; Kassel, W. S.; Liu, T.; DuBois, D. L.; Bullock, R. M. Cobalt Complexes Containing Pendant Amines in the Second Coordination Sphere as Electrocatalysts for H2 Production. Organometallics 2014, 33, 58205833; (c) Lasia, A. Improvements in the determination of the kinetics of dimerization reactions by cyclic voltammetry. J. Electroanal. Chem. Interfacial Electrochem. 1983, 146, 413-416. (24) (a) Towarnicky, J. M.; Schram, E. P. Tetrakis(triphenylphosphine)trihydridoruthenium(II) cation, (PPh3)4RuH3+, its formation, spectroscopic characterization and chemical reactivity. Inorg. Chim. Acta 1980, 42, 33-36; (b) Ginsberg, A. P.; Tully, M. E. Stereochemical nonrigidity in seven-coordinate trihydridorhenium complexes. J. Am. Chem. Soc. 1973, 95, 4749-4751.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Table of Contents artwork

ACS Paragon Plus Environment

7