You are on page 1of 4

Communication

pubs.acs.org/JACS

Reactive Fe-Sites in Ni/Fe (Oxy)hydroxide Are Responsible for


Exceptional Oxygen Electrocatalysis Activity
Michaela Burke Stevens, Christina D. M. Trang, Lisa J. Enman, Jiang Deng, and Shannon W. Boettcher*
Department of Chemistry and Biochemistry, University of Oregon, Eugene, Oregon 97403, United States
*
S Supporting Information

the films using surface-interrogation scanning electrochemical


ABSTRACT: Fe is a critical component of record-activity microscopy and Mössbauer spectroscopy, respectively. Despite
Ni/Fe (oxy)hydroxide (Ni(Fe)OxHy) oxygen evolution the large difference in the proposed mechanism or active site
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

reaction (OER) catalysts, yet its precise role remains and the Fe/Ni oxidation state, all the materials in the cited
unclear. We report evidence for different types of Fe studies are similarly efficient OER catalysts.
Downloaded via NORTHWESTERN UNIV on February 23, 2021 at 04:10:43 (UTC).

species within Ni(Fe)OxHy those that are rapidly The various reports discussed above have all demonstrated
incorporated into the Ni oxyhydroxide from Fe cations changes in the electrochemical profile, e.g. Ni redox potential,
in solution (and that are likely at edges or defects) and are redox peak size and shape, etc., as a function of Fe
responsible for the enhanced OER activity, and those incorporation into the NiOxHy film. Theoretical/computational
substituting for bulk Ni that modulate the observed Ni studies have also provided insights into the role of Fe.15
voltammetry. These results suggest that the exceptional
However, it remains unclear where the Fe incorporates into the
OER activity of Ni(Fe)OxHy does not depend on Fe in the
film to result in highly active Fe−O, Ni−O, or Ni−O−Fe sites.
bulk or on average electrochemical properties of the Ni
It is also unclear if the changes in bulk electronic structure, i.e.,
cations measured by voltammetry, and instead emphasize
the role of the local structure. as interrogated by measurements of conductivity, redox peak
position, and average e− per Ni in the redox wave, are directly
responsible for the increased activity or simply a byproduct of
the Fe incorporation unrelated to the enhanced catalytic
U nderstanding the oxygen evolution half reaction (OER, in
alkaline media: 4OH− → 2H2O + 4e− + O2) is important
for an array of energy-conversion technologies.1,2 Fe-doped Ni-
activity. Answering these questions is important to develop a
mechanistic picture of OER and a structural picture of probable
active sites.
and Co-based oxide/oxyhydroxide catalysts are among the
Here we report the characterization of highly OER-active
most-active OER catalysts in alkaline media.3,4 Fe increases the
Ni(Fe) (oxy)hydroxides that are nearly identical in reduction
intrinsic activity of both Co and Ni (oxy)hydroxides by roughly
potential, conductivity, and electrochemical accessibility to their
30- and 1000-fold, respectively.3,5−8 For electrodeposited Ni or
Fe-free parent materials (that have low catalytic activity).
Co (oxy)hydroxide, Fe can be stabilized in a solid solution (i.e.,
Furthermore, we describe two Ni(Fe)OxHy films that have
it does not form phase segregated islands) apparently up to
∼25 and ∼50%, respectively.6,9 Further, Klaus et al. have shown incorporated Fe differently and have significantly different
that active Ni(Fe)OOH phases form regardless of catalyst electronic/electrochemical properties, while still maintaining
deposition method.10 high activity. These results indicate the presence of different
Despite the similarity of phases formed under a wide variety types of active sites within Ni(Fe) (oxy)hydroxides and
of preparation conditions, there have been many different emphasize the role of the local geometric structure on activity
mechanistic hypotheses regarding the role that Fe plays in as opposed to bulk electrochemical behavior.
(oxy)hydroxide OER activity.9,11,12 Li et al. have hypothesized An electrolyte permeable (i.e., disordered/porous) Ni(OH)2
that increased Ni−O hybridization as a result of the Lewis film was cathodically deposited from an aqueous Ni(NO3)2
acidity of Fe3+ resulted in a more-active material.11 Gorlin et al. solution onto a conductive substrate (Au or Pt).16 As Ni(OH)2
coupled quasi-in-situ X-ray absorption spectroscopy (XAS) and is cycled, a sharp precatalytic redox feature is visible (Figure
differential electrochemical mass spectrometry and reported S1). This redox wave is typically associated with the nominal 2+
that Ni remains in a 2+ oxidation state; they observe no → 3+ oxidation of Ni(OH)2 to NiOOH; however, as has been
evidence for increased hybridization as a function of Fe.12 previously shown, this redox feature typically accounts for
Friebel et al. show that the Fe−O bond length in Ni(Fe)OxHy between 1.3 and 1.7 e− per Ni.17,18 The extra 0.3−0.7 e− per Ni
is different from the Fe−O bond length in FeOxHy and that the could be due to further oxidation of the Ni3+ to an average
overpotential on this Fe−O site was lower than the Ni∼3.5+ with possible Ni4+ sites and/or to the oxidation of
overpotential on the Ni−O sites.9 oxygen in the lattice (i.e., anion redox).19 Due to the dynamic
It is similarly debated what the oxidation state of Ni and Fe changes in oxidation state, hydration level, counterion content
are during catalysis and what role this plays in determining the during electrochemical cycling or catalysis, we typically refer to
OER activity. Friebel et al.9 and Gorlin et al.12 see no change in
the Fe oxidation state (Fe3+) during the OER; however, Ahn Received: July 8, 2017
and Bard13 and Chen et al.14 have shown evidence for Fe4+ in Published: August 8, 2017

© 2017 American Chemical Society 11361 DOI: 10.1021/jacs.7b07117


J. Am. Chem. Soc. 2017, 139, 11361−11364
Journal of the American Chemical Society Communication

these (oxy)hydroxide catalysts as MOxHy (where M is the overpotential at 1 mA cm−2 decreases only ∼20 mV between
metal cation). cycles 5 and 100 and >50% of that change occurs in the first 10
During the first four voltammetry cycles under Fe-free cycles after Fe-spiking (Figure 1, cycles 5−15, orange and
conditions the NiOxHy anodic peak potential (Epa) shifts green). Electron microscopy shows that the film morphology
cathodically 12−15 mV and the anodic peak integrated does not dramatically change as a function of cycling (Figure
intensity decreases by 2−5%. From cycle 5−104 the Epa and S3).
e− per Ni do not change significantly (Figure S1). The inconsistency between observed changes in activity and
When Fe(NO3)3 at 1 mM is spiked into the alkaline solution electrochemical profile as a function of Fe incorporation
(Figure 1), there is an immediate increase in activity shown by suggests that there is initial Fe incorporation at easily accessible
edge/defect sites, forming Ni−O−Fe species that dramatically
enhance catalysis, but that this has minimal effect on the
majority of the Ni redox species (Figure 1a inset). In fact, this
large effect on the catalysis is evidence that the catalysis is not
particularly influenced by Fe in the NiOxHy bulk (i.e., Fe sitting
on a Ni site fully coordinated by bridging O(H) with other
metal cations) or by the bulk electronic structure. This is
consistent with previous suggestions of active Fe edge sites.20
To probe the effect of incorporating Fe species, the effective
electrical conductivity of Fe-spiked NiOxHy was measured as a
function of applied potential before and after the addition of
Fe(NO3)3 at 1 mM to the initially Fe-free electrolyte. Figure 2

Figure 1. Cyclic voltammetry demonstrating the role of Fe


incorporation into NiOxHy from electrolyte solution. (A) NiOxHy
cycled (10 mV s−1) initially in Fe-free aq. 1 M KOH (red, cycle 1−4) Figure 2. Steady-state effective conductivity of NiOxHy on an
then moved to a 1 M KOH solution with 1 mM Fe(NO3)3 (cycle 5− interdigitated array (IDA) electrode as a function of Fe incorporation
104). The inset depicts a possible schematic of Fe incorporation into a from 1 mM Fe(NO3)3 in a 1 M aq. KOH solution. The effective
NiOxHy platelet as a function of cycling in Fe saturated solution. From conductivity was measured before Fe spiking (pre-Fe), immediately
the cyclic voltammetry, (B) the percent change in e− per Ni (triangles) after Fe spiking (1), and after cycling (2−3). The film was held at each
and (C) the change in overpotential (mV) at 1 mA cm−2 (open step for 2 min. Cyclic voltammetry is overlaid. There were two cycles
circles) and position of Epa (mV) (squares) as a function of cycling (10 mV s−1) between each steady-state conductivity measurement; the
relative to the last cycle in Fe-free conditions. Fe content listed is from second of each series is shown.
ICP-OES analysis on samples cycled in identical conditions for 2
(orange), 10 (green), 30 (blue), and 100 cycles (violet) after Fe- shows no correlation between conductivity and activity; the
spiking. The error bars are based on samples in triplicate with ∼3−5 activity increases dramatically with Fe incorporation while there
μg cm−2 loading.
is only a small decrease in the effective conductivity. The
decrease as a function of cycling is also seen in the Fe-free
the 130−150 mV decrease in overpotential. This dramatic NiOxHy (Figure S4).7 As the film is cycled, the onset potential
change in activity is accompanied by a minimal increase in for conductivity shifts with the shift in Epa indicating a change
electrochemical accessibility to Ni (2−4%, as measured by the in bulk electronic structure. This result is consistent with the
integrated charge in the redox wave) and anodic shift in the Ni edge and defect sites dominating the catalytic response of the
Epa (2−5 mV). During 100 cycles in Fe3+ spiked solution, Epa system while the bulk electronic properties dominate the
shifts a maximum of 30 mV and the peak integration shrinks to electrical transport properties. This result is different from
reach 30% less e− per Ni than the Fe-Free NiOxHy. A similar coelectrodeposited Ni−Fe (oxy)hydroxides where the addition
trend is seen with the cathodic peak size and potential (Epc) of Fe increased the conductivity; the difference is likely
(Figure S2). After electrochemical analysis, the film has 23− explained by a higher degree of crystallinity for the Ni−Fe
25% Fe as determined by ICP-OES analysis (Table S1). relative to the pure Ni oxy(hydroxides) when both are
Despite the large changes in apparent redox wave and peak electrodeposited under similar conditions.7
potential of the Ni species, there is only a minimal increase in Small amounts of Fe are also known to induce large OER
activity after the initial cycle in the Fe-spiked solution. The activity changes on Au substrates likely due to FeOxHy sites
11362 DOI: 10.1021/jacs.7b07117
J. Am. Chem. Soc. 2017, 139, 11361−11364
Journal of the American Chemical Society Communication

integrated with Au−O surface species.21,22 Fe-spiked control Within 2 cycles in the Fe-spiked base, the NiOxHy film picks
experiments on blank Au and Pt electrodes (Figure S5) up an average of 11% Fe (the example in Figure 3a has 10% Fe)
demonstrate that although the activities of both “bare” and the activity increases by ∼150-fold; however, there is only
electrodes are significantly impacted by Fe spiking, the majority an ∼10 mV anodic shift in Epa and ∼2% difference in e− per Ni
of activity enhancement discussed here occurs on the NiOxHy in the redox wave. The Epa for the codeposited Ni(Fe)OxHy
rather than the substrate. However, fundamentally the Fe- sample with similar Fe content (Figure 3a) is ∼26 mV anodic
activated Au−O electrodes may be catalyzing the OER in a of the Fe-free NiOxHy wave and has ∼7% more e− per Ni. The
similar fashion to the Fe-activated NiOxHy with both utilizing Fe-spiked sample has twice the TOFtm (Figure 3c) than the
surface (in the case of Au−O) or edge (in the case of NiOxHy) codeposited Ni(Fe)OxHy. After 100 voltammetry cycles in the
Fe sites for the enhanced activity. It is possible that the Fe-spiked solution, the film incorporates ∼25% Fe. Epa is,
however, ∼55 mV cathodic of that observed for a codeposited
difference in active surface area between the NiOxHy film and
Ni(Fe)OxHy film with similar Fe content (Figure 3b). The e−
the Au surface is largely responsible for the difference in per Ni is ∼30% smaller for the Fe-spiked sample compared to
overpotential, although efforts to normalize and extract the the codeposited one. The activity of the two, however, is more
intrinsic activity suggest the Fe-activated NiOxHy is the faster similar than for the lower Fe concentration pair. For the
catalyst.22 codeposited 30% Fe sample, the cathodic and anodic peaks are
With continued cycling of NiOxHy in 1 M aq. KOH with 1 relatively narrow (and split by Epa − Epc = 80 mV) indicating a
mM Fe(NO3)3 (Figure 1) the Epa shifts significantly anodically. sample that has homogeneously incorporated the Fe into the
It is known that the reduction potential of nominally Ni2+/3+ bulk of the Ni lattice. This is consistent with the literature in
shifts anodically with increasing Fe content when deposited as a which ∼25% Fe is soluble in NiOOH.9 In contrast, the 25% Fe-
mixed (oxy)hydroxide.7,23,24 This shift in reduction potential spiked sample has two distinct reduction and oxidation peaks
has been associated with a bulk change in the electronic yielding broader redox features; this suggests that Fe is
structure of the material and can be used to determine if Fe has inhomogeneously distributed in the bulk and on the edges of
incorporated into the Ni(Fe)OxHy film. Figure 3a and 3b the NiOxHy in the Fe-spiked sample, thus inhomogeneously
compare an Fe-free NiOxHy film, a spiked Fe−NiOxHy film affecting the Ni redox wave position.
cycled 2 and 100 times, and a codeposited Ni(Fe)OxHy film The data discussed above demonstrate (1) a Fe−NiOxHy
with similar Fe content. The OER activity (Figure 3c) is based catalyst nearly identical to its parent NiOxHy in terms of Ni
on the total-metal turnover frequency (TOFtm) or amount of redox behavior but is ∼100-fold more active and (2) a spiked
O2 produced per metal cation per second. Fe-NiOxHy and codeposited Ni(Fe)OxHy that have similar Fe
content and activities, but large differences in e− per Ni, Epa,
and redox wave shape. This supports a view in which FeOxHy
species initially incorporate into NiOxHy from solution at
surface/edge/defect sites, before moving into the bulk of the
NiOxHy nanosheet samples and significantly affecting the Ni
redox behavior. Co-deposited Ni(Fe)OxHy films appear to
incorporate Fe more uniformly during synthesis. These results
indicate that the Ni−Fe activity is more related to the local
active-site environment than the bulk electronic structure and
there are Fe-sites in different local environments (i.e., edge
versus bulk). Future studies should therefore search for
different Fe geometries directly in samples such as these, for
example using EXAFS at the Fe K-edge or infrared
specroscopy.8
In conclusion, upon the introduction of Fe into the NiOxHy
from solution the electronic properties, reduction potential,
redox peak size, peak shape, etc. of the Fe−NiOxHy film change
independently of the OER activity. Fe may initially incorporate
at edges/defects and further incorporate into the “bulk” of the
NiOxHy nanosheets (with Fe also remaining at the edge/defect
sites). The e− per Ni in the redox wave does slightly increase
with Fe addition, but the sample stays active even as the e− per
Ni decreases with continued Fe incorporation in the bulk of the
NiOxHy. The electrical conductivity of Fe−NiOxHy decays with
Figure 3. Cyclic voltammetry (10 mV s−1) of Fe-free NiOxHy film cycling in Fe-containing solution even as the sample becomes
(red) compared to (A) spiked Ni(Fe)OxHy cycled 2 (orange, 10% Fe) more active. Finally, the reduction potential (as monitored by
and (B) 100 (violet, 25% Fe) times in 1 mM Fe(NO3)3 and the Epa) does not shift to a potential expected of a codeposited
codeposited Ni(Fe)OxHy films with similar Fe content (green 8% Fe sample with similar Fe content. These experiments are all
and teal 30% Fe, respectively). (C) TOFtm at η = 300 mV based on
consistent with the view that the “bulk” (oxy)hydroxide
total metal content using the average of the forward and reverse sweep
from panel (A). (D) e− per Ni obtained by integration of the oxidation electronic structure is not the main factor influencing OER
wave. Metal content and ratios were determined after electrochemical activity.
analyses using ICP-OES. Open and closed symbols are for codeposited These results may help explain the recent conflicting
and Fe-spiked samples, respectively. Error bars are based on samples hypotheses on the role of Fe in Ni-based OER catalysts. The
measured in triplicate. deposition method and electrochemical conditioning appear to
11363 DOI: 10.1021/jacs.7b07117
J. Am. Chem. Soc. 2017, 139, 11361−11364
Journal of the American Chemical Society Communication

be directly responsible for the Fe location and local (16) Stevens, M. B.; Enman, L. J.; Batchellor, A. S.; Cosby, M. R.;
environment within the film, but do not change the catalyst Vise, A. E.; Trang, C. D. M.; Boettcher, S. W. Chem. Mater. 2017, 29,
activity in all cases. Depending on how the catalyst is deposited, 120−140.
it might have a higher density of Fe in edge/defect sites versus (17) Capehart, T. W.; Corrigan, D. A.; Conell, R. S.; Pandya, K. I.;
Hoffman, R. W. Appl. Phys. Lett. 1991, 58, 865−867.
bulk sites. This could determine the electrochemical profile of (18) Corrigan, D. A.; Conell, R. S.; Fierro, C. A.; Scherson, D. A. J.
the material, and even the average oxidation state of the Ni and Phys. Chem. 1987, 91, 5009−5011.
Fe sites, but leave the overall activity of the film unaffected.


(19) Grimaud, A.; Hong, W. T.; Shao-Horn, Y.; Tarascon, J.-M. Nat.
Mater. 2016, 15, 121−126.
ASSOCIATED CONTENT (20) Hunter, B. M.; Hieringer, W.; Winkler, J.; Gray, H. B.; Müller, A.
*
S Supporting Information
M. Energy Environ. Sci. 2016, 9, 1734−1743.
(21) Klaus, S.; Trotochaud, L.; Cheng, M. J.; Head-Gordon, M.; Bell,
The Supporting Information is available free of charge on the A. T. ChemElectroChem 2016, 3, 66−73.
ACS Publications website at DOI: 10.1021/jacs.7b07117. (22) Zou, S.; Burke, M. S.; Kast, M. G.; Fan, J.; Danilovic, N.;
Experimental methods and additional voltammetry and Boettcher, S. W. Chem. Mater. 2015, 27, 8011−8020.
materials characterization data (PDF) (23) Louie, M. W.; Bell, A. T. J. Am. Chem. Soc. 2013, 135, 12329−


12337.
(24) Corrigan, D. A. J. Electrochem. Soc. 1987, 134, 377−384.
AUTHOR INFORMATION
Corresponding Author
*swb@uoregon.edu
ORCID
Shannon W. Boettcher: 0000-0001-8971-9123
Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
This work was supported by the National Science Foundation
under Grant CHE-1566348. J.D. acknowledges support from
Zhejiang University and C.D.M.T. from the University of
Oregon Presidential Undergraduate Research Scholarship.

■ REFERENCES
(1) Burke, M. S.; Enman, L. J.; Batchellor, A. S.; Zou, S.; Boettcher, S.
W. Chem. Mater. 2015, 27, 7549−7558.
(2) Hunter, B. M.; Gray, H. B.; Müller, A. M. Chem. Rev. 2016, 116,
14120−14136.
(3) Trotochaud, L.; Ranney, J. K.; Williams, K. N.; Boettcher, S. W. J.
Am. Chem. Soc. 2012, 134, 17253−17261.
(4) McCrory, C. C. L.; Jung, S.; Peters, J. C.; Jaramillo, T. F. J. Am.
Chem. Soc. 2013, 135, 16977−16987.
(5) Burke, M. S.; Zou, S.; Enman, L. J.; Kellon, J. E.; Gabor, C. A.;
Pledger, E.; Boettcher, S. W. J. Phys. Chem. Lett. 2015, 6, 3737−3742.
(6) Burke, M. S.; Kast, M. G.; Trotochaud, L.; Smith, A. M.;
Boettcher, S. W. J. Am. Chem. Soc. 2015, 137, 3638−3648.
(7) Trotochaud, L.; Young, S. L.; Ranney, J. K.; Boettcher, S. W. J.
Am. Chem. Soc. 2014, 136, 6744−6753.
(8) Hunter, B. M.; Blakemore, J. D.; Deimund, M.; Gray, H. B.;
Winkler, J. R.; Muller, A. M. J. Am. Chem. Soc. 2014, 136, 13118−
13121.
(9) Friebel, D.; Louie, M. W.; Bajdich, M.; Sanwald, K. E.; Cai, Y.;
Wise, A. M.; Cheng, M.-J.; Sokaras, D.; Weng, T.-C.; Alonso-Mori, R.;
et al. J. Am. Chem. Soc. 2015, 137, 1305−1313.
(10) Klaus, S.; Louie, M. W.; Trotochaud, L.; Bell, A. T. J. Phys.
Chem. C 2015, 119, 18303−18330.
(11) Li, N.; Bediako, D. K.; Hadt, R. G.; Hayes, D.; Kempa, T. J.; von
Cube, F.; Bell, D. C.; Chen, L. X.; Nocera, D. G. Proc. Natl. Acad. Sci.
U. S. A. 2017, 114, 1486−1491.
(12) Gorlin, M.; Chernev, P.; De Araujo, J. F.; Reier, T.; Dresp, S.;
Paul, B.; Krahnert, R.; Dau, H.; Strasser, P. J. Am. Chem. Soc. 2016,
138, 5603−5614.
(13) Ahn, H. S.; Bard, A. J. J. Am. Chem. Soc. 2016, 138, 313−318.
(14) Chen, J. Y. C.; Dang, L.; Liang, H.; Bi, W.; Gerken, J. B.; Jin, S.;
Alp, E. E.; Stahl, S. S. J. Am. Chem. Soc. 2015, 137, 15090−15093.
(15) Fidelsky, V.; Toroker, M. C. Phys. Chem. Chem. Phys. 2017, 19,
7491−7497.

11364 DOI: 10.1021/jacs.7b07117


J. Am. Chem. Soc. 2017, 139, 11361−11364

You might also like