Molecular Cobalt Catalysts for O2 Reduction: Low-Overpotential

Oct 17, 2017 - Molecular Cobalt Catalysts for O2 Reduction: Low-Overpotential Production of H2O2 and Comparison with Iron-Based Catalysts. Yu-Heng Wan...
1 downloads 12 Views 974KB Size
Subscriber access provided by Gothenburg University Library

Communication

Molecular Cobalt Catalysts for O2 Reduction: Low-Overpotential Production of H2O2 and Comparison with Iron-Based Catalysts Yu-Heng Wang, Michael L Pegis, James M. Mayer, and Shannon S. Stahl J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b09089 • Publication Date (Web): 17 Oct 2017 Downloaded from http://pubs.acs.org on October 17, 2017

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 free 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 accessible to all readers and 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.

Journal of the American Chemical Society 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 5

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

Molecular Cobalt Catalysts for O2 Reduction: Low-Overpotential Production of H2O2 and Comparison with Iron-Based Catalysts Yu-Heng Wang,† Michael L. Pegis,‡ James M. Mayer,‡ and Shannon S. Stahl*,† †Department of Chemistry, University of Wisconsin-Madison, 1101 University Avenue, Madison, WI 53706, United States ‡

Department of Chemistry, Yale University, New Haven, Connecticut 06520, United States

Supporting Information Placeholder ABSTRACT: A series of mononuclear pseudo-macrocyclic cobalt complexes have been investigated as catalysts for O2 reduction. Each of these complexes, with CoIII/II reduction potentials that span nearly 400 mV, mediate highly selective twoelectron reduction of O2 to H2O2 (93–99%) using decamethylferrocene (Fc*) as the reductant and acetic acid as the proton source. Kinetic studies reveal that the rate exhibits a firstorder dependence on [Co] and [AcOH], but no dependence on [O2] or [Fc*]. A linear correlation is observed between log(TOF) vs. E1/2(CoIII/II) for the different cobalt complexes (TOF = turnover frequency). The thermodynamic potential for O2 reduction to H2O2 was estimated by measuring the H+/H2 open-circuit potential under the reaction conditions. This value provides the basis for direct assessment of the thermodynamic efficiency of the different catalysts and shows that H2O2 is formed with overpotentials as low as 90 mV. These results are compared with a recently reported series of Fe-porphyrin complexes, which catalyze four-electron reduction of O2 to H2O. The data show that the TOFs of the Co complexes exhibit a shallower dependence on E1/2(MIII/II) than the Fe complexes. This behavior, which underlies the low overpotential, is rationalized on the basis of the catalytic rate law.

potentials. This correlation enables direct comparison of the rates and thermodynamic efficiencies of O2 reduction catalyzed by the present Co-based catalysts and a series of recently reported Fe-porphyrin catalysts.5 Molecular cobalt complexes bearing tetradentate N2O2based ligands, such as salen, salophen, bis-ketiminates, and related derivatives are common catalysts for aerobic oxidation reactions. 6 However, unlike Co-porphyrin and related N4ligated mononuclear Co complexes,7 such N2O2-ligated cobalt complexes have received comparatively little attention as ORR (electro)catalysts. 8 , 9 The 1,2-phenylene-linked bisketiminate-CoII complex 1, first reported by Jäger,10 served as a starting point for the present study (Figure 1). A cyclic voltammogram (CV) reveals a nearly reversible CoIII/CoII redox couple under anaerobic conditions in MeOH [Figure 1, black CV; E1/2(CoIII/II) = 0.31 V vs. Fc*+/0, Fc* = decamethylferrocene]. The CoIII/II redox potential does not change significantly under 1 atm O2 [E1/2(CoIII/II) » 0.28 V], although the CV trace is less reversible than under N2, which could arise from binding of O2 to the Co complex (see below).

The catalytic reduction of molecular oxygen (O2) is crucial to diverse processes, ranging from biological respiration and fuel cells to the selective oxidation of organic molecules. 1 Metalloporphyrins and other molecular transition-metal complexes have been widely studied as catalysts for the oxygen reduction reaction (ORR),2 typically forming H2O2 or H2O as the reaction product (eqs 1 and 2). Diverse complexes O2 + 4 H+ + 4 e- ® 2 H2O +

-

O2 + 2 H + 2 e ® 2 H2O2

E° = 1.23 V vs. NHE

(1)

E° = 0.68 V vs. NHE

(2)

have been identified that exhibit high rates and/or high selectively for H2O or H2O2, but an unmet challenge is the identification of molecular catalysts that operate with low overpotential (i.e., at potentials close to the thermodynamic potentials associated with eqs 1 and 2). A TEMPO/NOx electrocatalyst system was recently reported with an overpotential for O2 reduction to water of ~300 mV;3 however, transition-metal catalysts typically exhibit overpotentials of 500–1000 mV. 4 Here, we report pseudomacrocyclic cobalt complexes that catalyze the reduction of O2 to H2O2 with high selectivity and overpotentials as low as 90 mV. The latter behavior arises from an unusually shallow dependence of the catalytic turnover frequencies (TOFs) on the CoIII/II reduction

Figure 1. (a) Structure of the cobalt catalyst 1. (b) CVs of 0.50 mM 1 in MeOH with 0.1 M [NBu4][ClO4], and 25 mM AcOH under 1 atm O2 (red) or 1 atm N2 (black). Working electrode: 3.0 mm glassy carbon, reference electrode: Ag/Ag+, counter electrode: Pt wire. Scan rate: 10 mV/s.

Efforts to study O2 reduction by cyclic voltammetry were unsuccessful due to the slow reaction rate;11 however, use of decamethylferrocene (Fc*) as a chemical reductant allowed the reaction in eq 3 to be monitored by UV-visible spectroscopy (Figure 2). The increase in absorbance at 780 nm arises from the formation of Fc*+ during the course of the O2 + 2 Fc* + 2 AcOH

cat. [LCo]

H2O2 + 2 Fc*+ + 2 AcO–

2.303RT [H2O2][AcO]2 EO2/H2O2 = E°O2/H2O2 log( ) nF pO2[AcOH]2

ACS Paragon Plus Environment

(3) (4)

Journal of the American Chemical Society

The rates of ORR mediated by complexes 1–8 were probed under a uniform set conditions, buffered with AcOH/NBu4OAc (25 mM each). TOFs for the conversion of O2 to H2O2 were determined from the initial rates of Fc*+ formation by UV-vis (Figure S10-11).11 The data revealed that complexes with lower CoIII/II redox potential exhibit higher rates of ORR, and a nearly linear free energy relationship (LFER) is evident between the log(TOF) and E1/2(CoIII/II) (Figure 3). 1.5 8

Figure 2. UV-visible spectral changes in the two-electron reduction of O2 (10 mM) by Fc* (0.5 mM) with 1 (1.25 × 10−2 mM) in the presence of 25 mM each of AcOH and [NBu4][OAc] in O2-saturated MeOH at 298 K. Inset: Absorbance changes at 780 nm are due to formation of Fc*+.

reaction, and the TOF is calculated to be 0.027 s-1 with respect to O2 consumption.11 Iodometric titration of the final reaction mixture shows that the reaction in eq 3 proceeds with nearly complete selectivity for H2O2 (96% yield).7j,11, 12 Initial-rate kinetic studies reveal a rate law with a first-order dependence on [Co] and [AcOH], but zero-order with respect to pO2 and [Fc*] (eq 5 and Figure S6).11 In order to probe the lack of an O2 dependence in the rate law, the CoII complex 1 was added to O2-saturated MeOH and the dissolved O2 was monitored using a Clark electrode. Addition of 1 to the solution resulted in consumption of 1.1 ± 0.1 equiv of O2 (Figure S7-S9).11 This result, which supports a Co–O2 adduct13 as the catalyst resting state, accounts for the zero-order pO2-dependence of the rate. rate = kcat[Co(N2O2)][AcOH]

(5)

A series of pseudo-macrocyclic N2O2-ligated cobalt complexes 2-8 (Chart 1) were prepared in order to compare their ORR reactivity with that of 1. Cyclic voltammograms of complexes 1–8 show (quasi-)reversible redox couples, with E1/2(CoIII/II) values ranging from -0.085 – 0.31 V (Chart 1, Figure S4). Analysis of their reactivity with O2 in the presence of Fc* and AcOH (cf. eq 3) show that H2O2 is formed with high selectivity (93-99%) in all cases (Chart 1, Figure S5). Chart 1. Molecular N2O2-ligated cobalt complexes tested for O2 reduction.

0.5 Log(TOF)

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

Page 2 of 5

7 4

-0.5

5 3 1

6

-1.5

R² = 0.9221 -2.5 -150

-50

50

150

2

250

350

E1/2 (CoIII/II) (mV vs Fc*+/0)

Figure 3. Linear free energy relationship between the log(TOF) for Co-catalyzed ORR and E1/2(CoIII/II) for complexes 1–8. Reaction conditions: 0.9 mM Fc*, 0.02-25 µM Co, 25 mM each of AcOH and [Bu4N][AcO] in O2 saturated MeOH solution, and 1 mM urea×H2O2.

These data may be analyzed further to assess the overpotential for the reactions using recently reported methodology for overpotential determination under nonaqueous conditions.5, 14 The effective overpotential (heff) is defined as the difference between the thermodynamic potential for O2 reduction to hydrogen peroxide (EO2/H2O2) under the reaction conditions and the E1/2(CoIII/II) values (Figure 4). The standard potential for O2/H2O2 in methanol may be estimated from i) the standard aqueous cell potential for O2 + H2 ® H2O2, ii) the measurement of the open circuit potential (OCP) for H+/H2 (EH+/H2) under the reaction conditions (Figure S12)15 and iii) the free energy to transfer H2O2 from H2O to MeOH.16 The solvation energy of H2O2 in MeOH is expected to be nearly identical that in H2O and, therefore, the transfer free energy is negligible. 17 The thermochemical cycle associated with i–iii (see Supporting Information, section VIII) corresponds to the standard O2/H2O2 potential in methanol with acetic acid/acetate buffer (EO2/H2O2). This value is obtained by adding 0.68 V to the EH+/H2 measured by OCP (–0.31 V vs. Fc*), which corresponds to 0.37 V versus Fc*+/0. This value is adjusted to 0.46 V to account for the non-standard-state background concentration of 1 mM H2O2 using the Nernst equation (cf. eq 4).18

Figure 4. Redox potential scale showing E1/2(CoIII/II) for 1–8 and thermodynamic reduction potentials for H+/H2, O2/H2O2 and O2/H2O vs Fc*+/0 under the catalytic reaction conditions.

The thermodynamic analysis above yields heff values that range from 150-550 mV. Further analysis of the O2/H2O2

ACS Paragon Plus Environment

2

Page 3 of 5

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

ORR catalyzed by 1 showed that the rate is not altered in the presence of higher [H2O2], ranging from 1–100 mM.11 With a background [H2O2] of 100 mM, the O2/H2O2 heff is only 90 mV with 1. Direct comparison of this result with those of other catalysts is difficult because most prior studies employed unbuffered conditions, and OCP measurements were not performed. Both factors prevent accurate assessment of overpotentials. Nonetheless, a 90 mV overpotential is much lower than is typically associated with ORR.3,4,19,20 One catalyst system for which direct comparison is possible is a series of Fe-tetraphenylporphyrin [Fe(porphyrin)] derivatives, which catalyze selective conversion of O2 to water.5 The Fe(porphyrin)-catalyzed ORR reactions follow a rate law with a first order dependence on [Fe(porphyrin)], [H+] and [O2] 21 (eq 6). Overpotential analysis, similar to that rate = kcat[Fe(porphyrin)][H+][O2]

(6)

described above, was performed for these catalysts. The logarithm of the ORR rates measured for the Fe and Co catalysts are plotted relative to their respective heff values for formation of H2O2 (red, Co) and H2O (blue, Fe) in Figure 5.

Figure 5. Correlations between log(TOF) versus heff for the O2 reduction to H2O2 or H2O, catalyzed by Co(N2O2) (red), and Fe(porphyrin) (blue). 22 The Fe(porphyrin) data are taken from the Table 1 in ref. 5. ORR reactions with Co(N2O2) complexes were conducted in the presence of urea×H2O2: 1 mM (•), 10 mM (■), and 100 mM (▲), respectively (see also Table S1).

The trendline for the Co complexes exhibit a significantly shallower slope than that for the Fe complexes. Thus, while Co catalysts are slower than Fe catalysts at high overpotentials, Co catalysts are much faster at low overpotentials. This distinction between Co and Fe is even more noteworthy in light of the production of H2O2 with Co, since the formation of H2O2 has a 550 mV lower driving force relative to the formation of water (E° = 0.68 versus 1.23 V; eqs 1 and 2). Both classes of ORR reactions in Figure 4 exhibit high selectivity (Co: H2O2; Fe: H2O) across the entire range of redox potentials. This sharp distinction in selectivity contradicts a recent suggestion that H2O2/H2O selectivity is directly correlated to ORR overpotential, irrespective of catalyst identity, 23 even while it highlights to the need for further insights into the factors that contribute to product selectivity for mononuclear complexes such as these. Such efforts are the focus of ongoing study.

Scheme 1. General catalytic cycle for the O2 reduction catalyzed by Co(N2O2) and Fe(porphyrin) catalysts

The different slopes for the Co- and Fe-catalyzed reactions in Figure 5 may be rationalized by the rate laws for the Coand Fe-based reactions and the simplified catalytic cycle in Scheme 1. Both Co- and Fe-catalyzed ORR are proposed to be initiated by O2 binding to the MII species, followed by a series of proton- and electron-transfer steps to afford H2O2 or H2O as the final product. The rate law for Fe(porphyrin)-catalyzed ORR in eq 6 arises from pre-equilibrium O2 binding to FeII, followed by protonation of the FeIII(O2•) adduct.21 Facile and favorable O2 binding to CoII results in a CoIII(O2•) adduct as the catalyst resting state and a rate law that is zero-order in pO2.11 Several proton-transfer steps could account for the [AcOH] dependence evident in the catalytic rate law in eq 5; however, previous studies support protonation of a CoIII(O2•) adduct as the rate-limiting step.7f,g,k,o Both the O2 binding and protonation steps are expected to be favored for complexes with lower E1/2(MIII/II). The steeper slope observed with the Fe(porphyrin) complexes, therefore, may be rationalized by the additive effect of two electronically sensitive steps (O2 binding and protonation), whereas Co(N2O2)-catalyzed ORR is affected by only one step (protonation). More quantitatively, the slope of the Co-catalyzed ORR correlation in Figure 5 (6.1 decade of TOF/V) is similar to the slope observed for Fe(porphyrin) catalyst when the overpotential is adjusted solely by changing the pKa of the acid source with a single Fe(porphyrin) catalyst (5.1 decade of TOF/V).24 In summary, this study shows that Co(N2O2) complexes serve as effective homogeneous ORR catalysts that form H2O2 as the near-exclusive product of the reaction. Determination of the thermodynamic potential for EO2/H2O2 by measuring the open circuit potential for H+/H2 under the non-aqueous reaction conditions reveals that Co(N2O2)-catalyzed formation of H2O2 proceeds with quite low overpotential (heff). The relatively slow ORR rates observed with these catalysts prove to be a consequence of this low overpotential. In fact, analysis of the LFER relationship between log(TOF) and the heff (for H2O2 or H2O) shows that the Co(N2O2) catalysts are considerably faster than analogous Fe(porphyrin) catalyst would be at similarly low overpotentials. The good ORR activity of the Co(N2O2) catalysts arises, in part, from the catalytic rate law, which is zero order in pO2 owing to facile and favorable formation of a CoIII(O2) adduct. Collectively, these results establish an important foundation for more thorough comparison of ORR catalysts, such as those composed of different metal ions and/or ligand classes. We anticipate that acquisition of such data will enable future design and development of molecular catalyst systems that combine high rates with low overpotentials.

ACS Paragon Plus Environment

3

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

ASSOCIATED CONTENT Supporting Information Experimental procedures, NMR spectroscopic data, reaction time courses, cyclic voltammograms, iodometric titrations. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author [email protected]

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We thank Drs. Colin W. Anson and James B. Gerken for helpful discussions. 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, Office of Science, Office of Basic Energy Sciences.

REFERENCES (1) For general references, see: (a) Babcock, G. T.; Wikstrom, M. Nature 1992, 356, 301-309; (b) Stahl, S. S. Science 2005, 309, 18241826. (c) Lee, J.; Jeong, B.; Ocon, J. D. Curr. Appl. Phys. 2013, 13, 309-321; (d) Katsounaros, I.; Cherevko, S.; Zeradjanin, A. R.; Mayrhofer, K. J. J. Angew. Chem. Int. Ed. 2014, 53, 102-121. (2) (a) Collman, J. P.; Wagenknecht, P. S.; Hutchison, J. E. Angew. Chem. Int. Ed. Engl. 1994, 33, 1537-1554; (b) Kobayashi, N.; Nevin, W. A. Appl. Organometal. Chem. 1996, 10, 579–590; (c) Anson, F. C.; Shi, C.; Steiger, B. Acc. Chem. Res. 1997, 30, 437-444; (d) Rosenthal, J.; Nocera, D. G. Acc. Chem. Res. 2007, 40, 543-553; (e) Scanlon, M. D. ChemCatChem 2013, 5, 1696-1697; (f) Zhang, W.; Lai, W.; Cao, R. Chem. Rev. 2017, 117, 3717-3797. (3) Gerken, J. B.; Stahl, S. S. ACS Cent. Sci. 2015, 1, 234-243. (4) Variations in reaction conditions make it difficult to assess the overpotentials associated with molecular ORR catalysts; however, estimated overpotentials for diverse catalyst systems are compiled in Table S2 of the Supporting Information for ref. 3. (5) Pegis, M. L.; McKeown, B. A.; Kumar, N.; Lang, K.; Wasylenko, D. J.; Zhang, X. P.; Raugei, S.; Mayer, J. M. ACS Cent. Sci. 2016, 2, 850-856. (6) For leading references, see: (a) Bäckvall, J. E.; Awasthi, A. K.; Renko, Z. D. J. Am. Chem. Soc. 1987, 109, 4750-4752; (b) Bäckvall, J. E.; Piera, J. Angew. Chem., Int. Ed. 2008, 47, 3506-3523. (c) Wendlandt, A. E.; Stahl, S. S. J. Am. Chem. Soc. 2014, 136, 1191011913. (7) (a) Alt, H.; Binder, H.; Sandstede, G. J. Catal. 1973, 28, 8-19. (b) Jahnke, H.; Schönborn, M.; Zimmermann, G., Organic dyestuffs as catalysts for fuel cells. In Physical and Chemical Applications of Dyestuffs, Schäfer, F. P.; Gerischer, H.; Willig, F.; Meier, H.; Jahnke, H.; Schönborn, M.; Zimmermann, G., Eds. Springer Berlin Heidelberg: Berlin, Heidelberg, 1976; pp 133-181 (c) Geiger, T.; Anson, F. C. J. Am. Chem. Soc. 1981, 103, 7489-7496; (d) Chan, R. J. H.; Su, Y. O.; Kuwana, T. Inorg. Chem. 1985, 24, 3777-3784. (e) Kobayashi, N.; Nishiyama, Y. J. Phys. Chem 1985, 89, 1167–1170. (f) Fukuzumi, S.; Mochizuki, S.; Tanaka, T. Chem. Lett. 1989, 18, 2730. (g) Fukuzumi, S.; Mochizuki, S.; Tanaka, T. Inorg. Chem. 1989, 28, 2459-2465; (h) Fukuzumi, S.; Mochizuki, S.; Tanaka, T. Inorg. Chem. 1990, 29, 653-659; (i) Kang, C.; Anson, F. C. Inorg. Chem. 1995, 34, 2771–2780; (j) Fukuzumi, S.; Okamoto, K.; Gros, C. P.; Guilard, R. J. Am. Chem. Soc. 2004, 126, 10441-10449; (k)

Page 4 of 5

Fukuzumi, S. Chem. Lett. 2008, 37, 808-813. (l) Kadish, K. M.; Shen, J.; Frémond, L.; Chen, P.; Ojaimi, M. E.; Chkounda, M.; Gros, C. P.; Barbe, J.-M.; Ohkubo, K.; Fukuzumi, S.; Guilard, R. Inorg. Chem. 2008, 47, 6726–6737; (m) Zagal, J. H.; Griveau, S.; Silva, J. F.; Nyokong, T.; Bedioui, F. Coord. Chem. Rev. 2010, 254, 2755-2791; (n) Honda, T.; Kojima, T.; Fukuzumi, S. J. Am. Chem. Soc. 2012, 134, 4196-4206; (o) Mase, K.; Ohkubo, K.; Fukuzumi, S. J. Am. Chem. Soc. 2013, 135, 2800-2808; (p) Mase, K.; Ohkubo, K.; Fukuzumi, S. Inorg. Chem. 2015, 54, 1808-1815. (8) N2O2-based ligands have been employed with molecular vana– dium ORR catalysis: (a) Yamamoto, K.; Oyaizu, K.; Tsuchida, K. J. Am. Chem. Soc. 1996, 118, 12665-12672; (b) Tsuchida, E.; Oyaizu K.; Dewi, E. L.; Imai, T.; Anson, F. C. Inorg. Chem. 1999, 38, 37043708. (c) Liu, Z.; Anson, F. C. Inorg. Chem. 2000, 39, 274-280. (d) Liu, Z.; Anson, F. C. Inorg. Chem. 2001, 40, 1329-1333. (9) For examples of electrocatalytic O2 reduction by immobilized Co(N2O2)-type catalysts, see the following: (a) Choi, Y.-K., Park, J.K. and Jeon, S. Electroanalysis 1999, 11, 134-138; (b) Okada, T.; Katou, K.; Hirose, T.; Yuasa, M.; Sekine, I. J. Electrochem. Soc. 1999, 146, 2562-2568; (c) Shamsipur, M.; Najafi, M.; Hosseini, M.R. M.; Sharghi, H. Electroanalysis 2007, 19, 1661-1667. (10) (a) Jäger, E.-G.; Knaudt, J.; Rudolph, M.; Rost, M. Chem. Ber. 1996, 129, 1041-1047; (b) Seidel, R. W.; Goddard, R.; Breidung, J.; Jäger, E.-G. Z. Anorg. Allg. Chem. 2014, 640, 1946-1952. (11) See Supporting Information for details. ( 12 ) For protocol of iodometric titration, see: (a) Mair, R. D.; Graupner, A. J. Anal. Chem. 1964, 36, 194-204; (b) Fukuzumi, S.; Kuroda, S.; Tanaka, T. J. Am. Chem. Soc. 1985, 107, 3020-3027. (13) Characterization of the O2 adduct of complex 5 in MeOH was recently reported: Anson, C. W.; Ghosh, S.; Hammes-Schiffer, S.;

Stahl, S. S. J. Am. Chem. Soc. 2016, 138, 4186-4193. (14) Appel, A. M.; Helm, M. L. ACS Catal. 2014, 4, 630-633. (15) Roberts, J. A. S.; Bullock, R. M. Inorg. Chem. 2013, 52, 38233835. (16) Pegis, M. L.; Roberts, J. A. S.; Wasylenko, D. J.; Mader, E. A.; Appel, A. M.; Mayer, J. M. Inorg. Chem. 2015, 54, 11883-11888. ( 17 ) For relevant considerations, see: (a) Ohgaki, K.; Sano, F.; Katayama, T. J. Chem. Eng. Data 1976, 21, 55-58; (b) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. B 2009, 113, 63786396. (18) Complexes 1–8 were found to decompose in the presence of a standard-state concentration of H2O2 (1 M), so the rate measurements in Figure 3 were conducted in the presence of 1 mM H2O2·urea to establish a stable reference state. This 103-fold difference in [H2O2] leads to a 90 mV Nernstian shift in the thermodynamic O2/H2O2 potential. Control experiments showed that the urea present in H2O2·urea does not influence the reaction (also see Supporting Information, section IX). (19) Onset potentials for the ORR to H2O with heterogeneous Pt catalysts exhibit overpotentials of ~300 mV: Gasteiger, H. A.; Kocha, S. S.; Sompalli, B.; Wagner, F. T. Appl. Catal., B 2005, 56, 9-35. (20) Heterogeneous catalysts for O2 reduction to H2O2 has received less attention, but it can have lower overpotentials (50-450 mV): Siahrostami, S.; Verdaguer-Casadevall, A.; Karamad, M.; Deiana, D.; Malacrida, P.; Wickman, B.; Escudero-Escribano, M.; Paoli, E. A.; Frydendal, R.; Hansen, T. W.; Chorkendorff, I.; Stephens, I. E. L.; Rossmeisl, J. Nat. Mater. 2013, 12, 1137-1143. (21) Wasylenko, D. J.; Rodríguez, C.; Pegis, M. L.; Mayer, J. M. J. Am. Chem. Soc. 2014, 136, 12544-12547. (22) Reaction conditions: 0.3 mM Fe(porphyrin) derivatives 1-7 were added in DMF solution containing 0.1 M electrolyte [Bu4N][PF6] in the presence of 20 mM [DMF-H]OTf under 1 atm O2. Scan rate: 100 mV/s. See ref. 5 for the details. (23) Passard, G.; Ullman, A. M.; Brodsky, C. N.; Nocera, D. G. J. Am. Chem. Soc. 2016, 138, 2925-2928. (24) Pegis, M. L.; Wise, C. F.; Koronkiewicz, B.; Mayer, J. M. J. Am. Chem. Soc. 2017, 139, 11000-11003.

ACS Paragon Plus Environment

4

Page 5 of 5

TOC Graphic Overpotential (O2/H2O, mV)

550

750

950

1150

1350 resting state of Fe(porphyrin)

MII

+ O2

Fe(porphyrin)

logTOF

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

MIII-O2

Co(N2O2)

+ H+

resting state of Co(N2O2)



MIII-OOH + 2 e -, 1 H +

h = 90 mV

or

+ 4 e-, 3 H+

Low Overpotential 0

200

H2O2 or 2 H2O

400

600

800

Overpotential (O2/H2O2, mV)

ACS Paragon Plus Environment

5