Atomically Dispersed Manganese Catalsyt For ORR

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Articles

https://doi.org/10.1038/s41929-018-0164-8

Atomically dispersed manganese catalysts for


oxygen reduction in proton-exchange membrane
fuel cells
Jiazhan Li1,2,11, Mengjie Chen2,11, David A. Cullen3, Sooyeon Hwang4, Maoyu Wang5, Boyang Li6, Kexi Liu6,
Stavros Karakalos   7, Marcos Lucero5, Hanguang Zhang2, Chao Lei8, Hui Xu8, George E. Sterbinsky9,
Zhenxing Feng   5, Dong Su   4, Karren L. More10, Guofeng Wang6, Zhenbo Wang   1* and Gang Wu   2*

Platinum group metal (PGM)-free catalysts that are also iron free are highly desirable for the oxygen reduction reaction (ORR)
in proton-exchange membrane fuel cells, as they avoid possible Fenton reactions. Here we report an efficient ORR catalyst that
consists of atomically dispersed nitrogen-coordinated single Mn sites on partially graphitic carbon (Mn-N-C). Evidence for
the embedding of the atomically dispersed MnN4 moieties within the carbon surface-exposed basal planes was established by
X-ray absorption spectroscopy and their dispersion was confirmed by aberration-corrected electron microscopy with atomic
resolution. The Mn-N-C catalyst exhibited a half-wave potential of 0.80 V versus the reversible hydrogen electrode, approach-
ing that of Fe-N-C catalysts, along with significantly enhanced stability in acidic media. The encouraging performance of the
Mn-N-C catalyst as a PGM-free cathode was demonstrated in fuel cell tests. First-principles calculations further support the
MnN4 sites as the origin of the ORR activity via a 4e− pathway in acidic media.

D
evelopment of cost-effective and high-performance platinum Kinetic losses primarily result from the dissolution of active metal
group metal (PGM)-free catalysts for the sluggish oxygen sites15,16. These problems may be caused by H2O2 oxidative attack
reduction reaction (ORR) is key to realizing the large-scale or by protonation of the doped N neighbouring the active metal
application of proton-exchange membrane fuel cells (PEMFCs)1–6. sites followed by anion adsorption14. Carbon corrosion of catalysts
A wide variety of carbon defects, including nitrogen doping and also promotes dissolution of the metal sites18,19 and induces signifi-
surface-exposed carbon basal plane edges and steps, are believed to cant charge- and mass-transport resistances due to the changes in
act as preferred catalytic sites, thus facilitating the ORR by directly the carbon lattice structures and morphologies20. Thus, a relatively
providing adsorption sites and/or by modifying the electronic prop- high degree of graphitization is vital for PGM-free catalyst stabil-
erties of carbon7,8. However, the activity of such metal-free active ity to enhance the corrosion resistance of the carbon. The Fe-N-C
sites in acidic media is not sufficient because they suffer from large catalysts are also criticized for their participation in and/or promo-
overpotential, mostly catalysing the ORR via a two-electron path- tion of the Fenton reactions (Fe2+ +​  H2O2), where dissolved Fe ions
way9. In contrast, the co-doped metal and nitrogen in the form of combine with H2O2, a byproduct of the two-electron ORR21. As a
MNxCy can tune the electronic and geometric properties of carbon result, a significant amount of active oxygen-containing hydroxyl
matrix more significantly. The MNxCy moieties themselves are also and hydroperoxyl radicals are generated that can degrade the iono-
believed to be acting as the active sites to directly adsorb O2 and mer within the electrode and the polymer membrane in PEMFCs
catalyse the subsequent O–O bond breaking in acidic media, there- (Supplementary Note 1). Therefore, high-performance PGM- and
fore significantly improving the ORR catalytic activity6,10,11. The Fe-free catalysts are highly desirable for PEMFC technologies. In
bonding energies of MNxCy sites with O2 and other ORR interme- principle, the Co-N-C catalyst is a logical alternative22, but it suf-
diates are dependent on the nature of the transition metal, which fers from substantial generation of H2O2 during the ORR in acidic
leads to significantly different activities and stabilities. Among the media, and is thus likely to participate in Fenton reactions23. The
studied transition metals, Fe or Co with N-doped carbon form the Co–N coordinated structures are even less stable in acids than
M-N-C catalysts, which are recognized to be the most promising Fe-N-C catalysts24. Unlike Fe and Co, Fenton reactions involving
PGM-free catalysts12. They exhibit encouraging ORR activities Mn ions are insignificant because of weak reactivity between Mn
and stabilities even in harsh acidic media13. Despite tremendous and H2O2 (ref. 23). We discovered that Mn doping promotes the sta-
efforts to develop Fe-N-C and Co-N-C catalysts, they still suffer bility of nanocarbon (NC) catalysts and catalyses graphitic struc-
from insufficient durability, especially at the desirable high voltages tures in catalysts25,26. Preliminary density functional theory (DFT)
(>​0.6 V), which limits their practical applications in PEMFCs14–17. calculations predicted that MnN4 moieties embedded in carbon

1
MIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and Storage, School of Chemistry and Chemical Engineering, Harbin
Institute of Technology, Harbin, China. 2Department of Chemical and Biological Engineering, University at Buffalo, The State University of New York,
Buffalo, NY, USA. 3Materials Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN, USA. 4Center for Functional Nanomaterials,
Brookhaven National Laboratory, Upton, NY, USA. 5School of Chemical Biological, and Environmental Engineering, Oregon State University, Corvallis, OR,
USA. 6Department of Mechanical Engineering and Materials Science, University of Pittsburgh, Pittsburgh, PA, USA. 7Department of Chemical Engineering,
University of South Carolina, Columbia, SC, USA. 8Giner Inc, Newton, MA, USA. 9Advanced Photon Source, Argonne National Laboratory, Argonne, IL,
USA. 10Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN, USA. 11These authors contributed equally: Jiazhan Li,
Mengjie Chen *e-mail: wangzhb@hit.edu.cn; gangwu@buffalo.edu

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 935


Articles Nature Catalysis

have comparable activity and enhanced stability relative to FeN4 (extent of graphitization) of carbon particles can be tuned by vary-
sites. Inspired by this theoretical prediction, we hypothesized that ing the Mn content in the first step. As shown in Supplementary
Mn-N-C catalysts are probably more stable than Fe-N-C in harsh Fig. 3, the carbon particle size is slightly increased by adding more
acidic media. However, a grand challenge is to increase the density Mn content. Large particles with an irregular shape appeared
of MnN4 active sites in catalysts because Mn atoms tend to form when the Mn content reached 30at%, indicating that excess Mn
unstable and inactive metallic compounds, oxides and carbides ions in the solution disturb the controlled growth of the ZIF nano-
during the heat treatment when simply increasing the amount of crystals. Raman results further indicated that the Mn content in
Mn precursors13,27. the first step can significantly increase the extent of graphitiza-
Knowledge gained through our previous studies on Fe and Co tion in the resulting carbon structures (Supplementary Fig. 4 and
catalysts22,27–31 provides guidelines for the synthesis of Mn-N-C cata- Supplementary Table 1). All three of the nMn-NC-first (n =​  10,
lysts with atomic dispersion of MnN4 active sites within a porous 20 and 30 at%) samples exhibited much smaller ratios of D and G
three-dimensional (3D) carbon framework. A zinc-containing peak areas (AreaD/AreaG) relative to the Mn-free 0Mn-NC sample.
zeolitic imidazolate frameworks (ZIF-8), a type of metal–organic The carbon structures remain nearly the same after the second
framework with flexible control of the structure and chemistry, adsorption step of synthesis.
enables the formation of atomically dispersed MN4 active sites for According to inductively coupled plasma-mass spectrom-
Fe- or Co-N-C catalysts22,32–34. The unique hydrocarbon networks etry (ICP-MS), the Mn content in the 20Mn-NC-second catalyst,
in ZIF precursors can be directly converted into N- and M-doped which exhibited the highest catalyst activity, can be up to 3.03 wt%
highly disordered carbon, while maintaining their original poly- (Supplementary Table 2), which is comparable to other atomically
hedral particle shape and porous structure during thermal activa- dispersed Fe and Co catalysts (Supplementary Table 3). As no sig-
tion35. Recently, synthesis based on ZIF-8 precursors with chemical nificant Mn clustering was observed by HR-STEM and XRD analy-
doping or ion confinement has resulted in atomically dispersed Fe ses, homogeneous dispersion of atomic Mn species is likely, which
or Co sites with increased density2,22,27,36. Motivated by these suc- was further examined by X-ray absorption spectroscopy (XAS)
cesses, we explore the ZIF approach to prepare atomically dispersed measurements. X-ray absorption near edge structure (XANES)
Mn-N-C catalysts. However, unlike Fe and Co ions, Mn ions can- spectroscopy and extended X-ray absorption fine structure (EXAFS)
not easily exchange the original Zn and form complexes with N spectroscopy analyses are based on our established experimental and
in the ZIF-8 precursor. Only a low density of atomic Mn sites are theoretical modelling methods22,27,28,37,38. Several standard Mn com-
introduced using a conventional one-step chemical doping. Due to pounds (for example, MnO, MnO2 and Mn phthalocyanine (Pc))
a wide variety of Mn valences from 0 to +​7, Mn aggregates easily were included in the XANES/EXAFS studies to compare with the
form during the high-temperature carbonization even at a low con- 20Mn-NC-second catalyst. MnPc has a well-defined MnN4 chemical
tent. Thus, realizing atomically dispersed MnN4 sites with increased structure (Supplementary Fig. 5), which provides a baseline for the
density is very challenging. bulk catalyst measurements and can be used to identify the possible
Here, we report a catalyst with atomically dispersed MnN4 sites active sites in the 20Mn-NC-second catalyst. As shown in Fig. 2a, the
obtained through a two-step synthesis strategy involving doping XANFS edge profile of the 20Mn-NC-second catalyst is close to that
and adsorption processes by leveraging the unique properties of of Mn(ii) in MnO but is far from that of Mn(iv) in MnO2, suggest-
ZIF-8 precursors, which has been shown to effectively increase the ing that the oxidation state of Mn in the 20Mn-NC-second catalyst
active-site density. In the first step of synthesis, Mn ions are com- is close to 2+​. Figure 2b shows the EXAFS spectrum of the MnPc,
bined with Zn ions to prepare Mn-doped ZIF-8 precursors. After which is well fitted with MnN4 coordination (CNMn–N =​  4  ±​  0.4,
carbonization and acid leaching, the derived porous carbon is used where CN refers to coordination number and Mn–N is the scatter-
as a host to adsorb additional Mn and N sources followed by a sub- ing path). The fit is also plotted in k-space in Supplementary Fig. 5.
sequent thermal activation. This atomically dispersed Mn-N-C As shown in Fig. 2c, the EXAFS of the 20Mn-NC-second catalyst
catalyst achieves promising activity and excellent stability for the is well represented by a combination of the Mn–N and Mn–C scat-
ORR in aqueous acids. tering paths (Supplementary Table 4). The first shell coordination
number of Mn–N is given by CNMn–N =​  3.2  ±​ 1.0, suggesting the
Results dominant Mn–N structure in the 20Mn-NC-second catalyst is likely
Atomically dispersed and N-coordinated Mn sites. A continu- to be MnN4. We also compared the Mn K-edge EXAFS spectra of
ous two-step doping and adsorption approach as shown in Fig. 1 20Mn-NC-second to manganese nitride (Mn4N) and metallic man-
has proved effective for gradually introducing more MnN4 active ganese (Mn foil) to exclude the formation of nitrides and metallic
sites in Mn-N-C catalysts. Supplementary Fig. 1 presents the over- Mn (Supplementary Fig. 6). Both Mn4N and Mn foil exhibit peaks at
all morphology of the best-performing 20Mn-NC-second catalyst ~2.3 Å assigned to Mn–Mn distance, corresponding to the metallic
(where 20 is molar percentage of Mn against the total metals in structure. The absence of a Mn–Mn peak in the 20Mn-NC-second
solutions durng the synthesis of Mn-doped ZIF-8 precursors; sec- spectra further verifies there is no such Mn clusters in the catalyst.
ond refers to the sample obtained after the second adsorption step) We note that the peak at around 1.9 Å deviates from the fitted curve
and shows a homogeneous distribution of polyhedral carbon par- for 20Mn-NC-second catalyst, which may be due to a weak Mn–O
ticles with a size of about 50 nm, which directly transforms from scattering path caused by a very small amounts of O, since MnPc
the 20Mn-ZIF-8 nanocrystal precursors with slight size reduction. does not have a Mn–O scattering path. We have attempted to add a
The dominant diffraction peaks corresponding to carbon in the Mn–O scattering path directly from the MnO structure, but this did
X-ray diffraction (XRD) spectra together with the broad D and G not provide us with representative structure information. Compared
bonds in the Raman spectra (Supplementary Fig. 2) are consistent with standard MnPc that has a well-defined and uniform MnN4 mol-
with a partially graphitized carbon structures (meso-graphitic), ecule structure, the 20Mn-NC-second catalyst exhibits disorganized
which is confirmed by high-resolution scanning transmission structures, especially in terms of the disordered nature and defective
electron microscopy (HR-STEM) images (Supplementary Fig. 1h). structures that comprise the edges of carbon planes and grain/domain
Both XRD patterns and STEM images verified the absence of any boundary at the surface of the carbon particles. Due to a low content
crystalline Mn-containing phases or clusters in the catalyst. The of Mn and a more disorganized, defective carbonaceous/meso-gra-
carbon particles exhibit a high degree of microporosity, an essen- phitic structure, the EXAFS data of the actual catalyst 20Mn-NC-
tial characteristic for increasing surface areas available for host- second exhibits more noise, which could also contribute to the peak
ing active sites. The morphology (shape and size) and structure deviation observed at around 1.9 Å. We also compared the Mn

936 Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal


Nature Catalysis Articles

Carbonization Thermal
Adsorption
acid leaching activation

Mn-ZIF precursor Mn-NC-first (step 1) Adsorption Mn-NC-second (step 2)

C N Mn Cyanamide Mn and cyanamide in micropores

Fig. 1 | Schematic of atomically dispersed MnN4 site catalyst synthesis. A two-step doping and adsorption approach can gradually increase the density
of the atomically dispersed and nitrogen-coordinated MnN4 sites into the 3D carbon particles. In the first step, Mn-doped ZIF-8 precursors are carbonized
and then leached with an acid solution to prepare a partially graphitized carbon host with optimal nitrogen doping and microporous structures. In the
second step, additional Mn and N sources were adsorbed into the 3D carbon host followed by a thermal activation to generate increased density of MnN4
active sites.

catalysts at different synthesis stages, which shows the evolution of Figure 3a–c present typical bright-field and medium-angle
the Mn local structures and demonstrates the formation of atomi- annular dark field (MAADF) STEM images of the carbon particles
cally dispersed Mn sites instead of Mn clusters. Although we tried in the 20Mn-NC-second catalyst that show highly disordered car-
the measurement for 20Mn-NC-first to study the possible structure bon structures with randomly oriented graphitic domains and less
changes of the active sites after the adsorption step, due to a very low dense, pore-like morphologies (dark contrast in Fig. 3c). Similar
Mn concentration (0.68 wt% by ICP-MS) the recorded spectrum was carbon phases were observed in other ZIF-8-derived catalysts22,27,28,
not sufficient for high-quality data analyses (Supplementary Fig. 7). but the Mn-N-C catalyst appears to contain domains with a higher
The elemental quantifications of different precursors and degree of graphitization. Low voltage, aberration-corrected STEM
Mn-N-C catalysts determined from X-ray photoelectron spectros- imaging coupled with electron energy loss spectroscopy (EELS)
copy (XPS) are summarized in Supplementary Table 5. The content was used to study the atomic dispersion and local environment
of N in the 20Mn-ZIF precursor is 31.2 at% and the N 1s spectrum of Mn at an atomic resolution. The MAADF detector extends
shows a symmetric and sharp peak at 398.6 eV corresponding to the range of the collection semi-angle to 54–200 mrad (versus
C=​N of 2-methylimidazole (Fig. 2d)39. After pyrolysis, the hydrocar- 86–200 mrad for high-angle annular dark field STEM), which can
bon networks of the precursors transformed into partially graphi- provide more contrast from lower-scattering, light elements for
tized carbon with a N content of 2.2 at%. The samples obtained at example, carbon and nitrogen. Distinguishable signals for C, N, O
the different processing steps exhibited nearly identical C 1s spectra and Mn are observed in the EELS maps across the carbon nanopar-
(Fig. 2e and Supplementary Table 6) and carbon content, indicating ticles (Fig. 3e–h), respectively, acquired for the particle shown in
that the second adsorption step does not induce any change to the Fig. 3d. The bright spots shown in the high-resolution MAADF-
carbon structure. Although the N content continuously decreased STEM images in Fig. 3i,j correspond to single heavy atoms that are
during processing to a level of 1.3 at% after the second adsorption uniformly dispersed across/within the carbon structure, indicating
step, the amount of N appears to be sufficient to form a significant a high density of Mn doping. Furthermore, an EEL point spectrum
fraction of MnN4 coordination relative to the low content of Mn (Fig. 3k) was obtained by placing the 1 Å electron probe directly
(approximately half that of N). In addition to two dominant peaks on a single atom (for example, as circled in red in Fig. 3j), which
that correspond to graphitic- and pyridinic-N, respectively, the N 1s is near the edge of a single or double graphene layer (basal planes)
XPS peak at 399.1 eV could be attributed to N atoms bonded to M that form the graphite domains of the carbon nanoparticles. Based
sites40, likely MnN4 in the catalyst. The percentage of MnN4 increased on the ångström resolution of the electron probe and thin sam-
from 5.8% to 11.2% after the second adsorption step (Supplementary ple area, the signal that contributes to the EEL point spectrum
Table 7); the percentage of pyridinic-N decreased by about 10%, (Fig. 3k) originates from the atom and its closest neighbouring
likely due to further coordination with Mn ions as evidenced by the atoms. The co-existence of N and a single Mn within this ångström
N 1s peak shift from the original pyridinic-N at 398.4 eV to 399.1 eV. region provides strong evidence for a Mn–N coordinate structure,
The XPS Mn 2p peak of 20Mn-NC-first (Fig. 2f) shows a 0.7 eV shift suggesting that single Mn ions are anchored by N within carbon.
to a higher bonding energy compared with the sample before acid The immediate vicinity of single Mn sites with N is confirmed by
leaching (20Mn-NC). It was reported that the interaction between similar EELS analyses acquired from different areas of the catalysts
metallic clusters and MNx could change the charge density of the (Supplementary Fig. 8).
central metal ions and lead to a shift in the binding energy12. Thus, Atomically dispersed Mn sites are also observed in MAADF-
the positive shift may be due to the removal of Mn clusters during STEM images acquired of the 20Mn-NC-first catalyst and the
acid leaching. Considering that 20Mn-NC-first and 20Mn-NC- EEL point spectra reveal a similar co-existence of Mn and N
second exhibit the same Mn 2p peak position, no additional Mn (Supplementary Fig. 8). We infer from this result that the state of
clusters are produced during the second adsorption step. Mn in the 20Mn-NC-first is similar to that in the 20Mn-NC-second,

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 937


Articles Nature Catalysis

a 1.6 b c
0.6
MnPc MnPc 20Mn-NC-second
1.6 Mn-N
Fit Fit
20Mn-NC-second Mn-N
1.2 MnO
Normalized Xμ (E)

1.2 0.4 Mn-C


MnO2

∣χ (R )∣ (Å–3)

∣χ (R )∣ (Å–3)
0.8 Mn-C
0.8

0.2
0.4 0.4

0.0
6,540 6,550 6,560 0 1 2 3 4 0 1 2 3 4
Energy (eV) Radial distance (Å) Radial distance (Å)

d e f
20Mn-NC-second N 1s 20Mn-NC-second Mn 2p 20Mn-NC-second C 1s
Mn-Nx

20Mn-NC-first 20Mn-NC-first 20Mn-NC-first


Graphitic Pyridinic
C–O
Intensity (a.u.)

Intensity (a.u.)

Intensity (a.u.)
20Mn-NC 20Mn-NC 20Mn-NC

C–C, C–N

20Mn-ZIF precursor 20Mn-ZIF precursor 20Mn-ZIF precursor


C=N
Positive shift C=C
~0.7 eV
Pyrrolic

406 404 402 400 398 396 660 654 648 642 636 296 292 288 284 280
Binding energy (eV) Binding energy (eV) Binding energy (eV)

Fig. 2 | Structural characterization by XANES, EXAFS and XPS. a, The experimental K-edge XANES spectra of 20Mn-NC-second catalyst
and reference samples (MnO, MnO2 and standard MnPc). b,c, Fourier transforms of Mn K-edge EXAFS data (open circle) and model-based
fits (red line) of standard MnPc (b) and the 20Mn-NC-second catalyst (c). d–f, High-resolution N 1s (d), C 1s (e) and Mn 2p (f) XPS data of
materials obtained from different processing steps. The 20Mn-NC was prepared through a pyrolysis of the 20Mn-ZIF-8 precursor. The 20Mn-
NC-first is obtained after an acid leaching and a heat treatment of 20Mn-NC. The 20Mn-NC-second catalyst is the final catalyst after two-step
doping and adsorption processes.

and is likely that of a MnN4 coordinated structure. The primary dif- ORR activity of atomically dispersed Mn-N-C catalysts. The
ference between the two catalysts is that the density of MnN4 active ORR activity and four-electron selectivity (H2O2 yield) of all cata-
sites is significantly increased after the second adsorption step. The lyst samples were evaluated using a rotating ring-disk electrode
signals of both Mn and N become much stronger after the second (RRDE) in a 0.5 M H2SO4 electrolyte. Unlike ZIF-derived Fe and
adsorption step as evidenced in the comparison of the EELS ele- Co catalysts, one-step chemical doping of Mn into ZIFs is not an
mental maps (Supplementary Fig. 9), indicating a higher density of effective method to generate high ORR activity. With Mn doping
MnN4, which is also verified by MAADF-STEM imaging, as shown content ranging from 0 to 30 at%, poor ORR activity is measured
in Supplementary Fig. 8. Although a number of bright spots are uni- for all Mn-NC-first samples exhibiting half-wave potential (E1/2) less
formly dispersed in the MAADF-STEM images for 20Mn-NC-first, than 0.65 V, which is likely due to an insufficient density of active
only a few Mn-N coordinated sites were confirmed by EELS. Many sites (Supplementary Fig. 10). Inspired by the concept of host–guest
of the brighter atoms probed showed N but not Mn by EELS and may strategy41, we introduced more MnN4 active sites into the promising
be residual Zn (0.1 at%, Supplementary Table 5). After the second Mn-NC-first host through the second adsorption step. The adsorp-
adsorption step, the co-existence of Mn and N becomes dominant tion step leads to significantly enhanced activity for all samples;
and is easily verified by EEL spectra acquired for many differ- the 20Mn-NC-second catalyst exhibited the most positive E1/2 up
ent areas, indicating a significant increase in the density of MnN4 to 0.8 V versus the reversible hydrogen electrode (RHE) (Fig. 4a).
sites in the 20Mn-NC-second catalyst. ICP-MS results also verify Note that there is no Fe contamination during this multistep pro-
that the Mn content increases from 0.68 wt% in the 20Mn-NC-first cessing procedure, as evidenced by negligible Fe content in the
to 3.03 wt% in the 20Mn-NC-second. The two-step Mn doping- catalysts determined by ICP-MS and XPS analyses (Supplementary
adsorption method effectively increases the density of MnN4 active Fig. 11) as well as poor activity of the Mn-free control sample after
sites without the concomitant formation of Mn clusters. multiple heating treatments. Considering the comparable carbon

938 Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal


Nature Catalysis Articles
a b c

10 nm 5 nm 5 nm

d e C f N

10 nm

g O h Mn i

1 nm

j k
C
Counts (a.u.)

Mn

2 nm
300 400 500 600 700
Energy (eV)

Fig. 3 | Morphology and atomic structure of the 20Mn-NC-second catalyst. a–d, Representative HR-TEM (a,b) and STEM (c,d) images of a carbon
particle in the catalyst. e–h, EELS elemental maps of C (e), N (f), O (g) and Mn (h) of the area in (d). i,j, Aberration-corrected MAADF-STEM images.
k, EEL point spectrum from the atomic sites circled in red in j. Image resolution in i and j is roughly 0.1 nm using an accelerating voltage of 60 kV.

structure and N dopant (Supplementary Tables 1 and 5) before and maximum activity during the adsorption step. The adsorption using
after the adsorption step of synthesis, the enhanced performance distinct Mn salts without N source still leads to an improvement in
originated from the increasing MnN4 concentration. A poison- the activity (Fig. 4c and Supplementary Fig. 13). This suggests that
ing experiment was performed by employing KSCN to block the Mn ions chemically bonded with the pre-doped N in the carbon
M–Nx sites42. The negative shift of 150 mV of E1/2 (Supplementary host and form MnN4 active sites through subsequent thermal acti-
Fig. 12) further verifies that the primary active sites in the vation. Co-adsorption of Mn and N sources yields improved activity
20Mn-NC-second catalyst are metal-based MnN4, instead of with an increased N content in the final 20Mn-NC-second catalyst.
metal-free CNx sites. The adsorption step also leads to an obvious (Supplementary Table 5). Due to significant reductions of micro-
enhancement of activity for both ketjenblack (KJ-black) and poly- pores and Brunauer–Emmett–Teller surface areas of the carbon
aniline-derived nanocarbon (PANI-NC) (Fig. 4b), indicating that host after adsorption process (Supplementary Fig. 14), we believe
the adsorption step is an effective strategy to introduce MnN4 active that N sources can stabilize Mn ions through possible coordination
sites, regardless of the carbon host. However, an optimal nitrogen in the micropores. During the pyrolysis, the N source creates addi-
doping level and the pore structure of the carbon host are crucial for tional N doping in the carbon host, which can further coordinate

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 939


Articles Nature Catalysis

a 0 b 0 c 0
0Mn-NC KJ-black

Current density (mA cm–2)


Current density (mA cm–2)

Current density (mA cm–2)


Before adsorption
0Mn-NC-second
No N sources
–1 20Mn-NC –1 PANI-NC –1
Cyanamide
20Mn-NC-second
Phenanthroline
0Mn-NC-
–2 –2 second –2 Dipicolylamine
Melamine

KJ-black-
–3 –3 second –3
PANI-NC-
second
0.5 M H2SO4 0.5 M H2SO4 0.5 M H2SO4
–4 –4 –4
0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Potential (V versus RHE) Potential (V versus RHE) Potential (V versus RHE)

d e 10 f 0.8
0 1.0 20Mn-NC-second (oxygen)
20Mn-NC-second 20Mn-NC-second
20Mn-NC-second (air) 0.7
Current density (mA cm–2)

20Co-NC-second 20Co-NC-second

Power density (W cm–2)


8 20Fe-NC-second (oxygen)
20Fe-NC-second 0.8
–1 20Fe-NC-second 20Fe-NC-second (air) 0.6
Pt/C
H2O2 yeild (%)
–2
Pt/C (60 μg cm ), 0.1 M HClO4

Voltage (V)
6 0.5
0.6
–2 0.4
4 0.4 0.3
H2-O2
–3 0.2
2 0.2
H2-air 0.1
0.5 M H2SO4
–4 0 0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 0.0 0.4 0.8 1.2 1.6 2.0 2.4
Potential (V versus RHE) Potential (V versus RHE) Current density (A cm–2)

Fig. 4 | ORR activity studied by using RRDE and fuel cell tests. a–d, Steady-state ORR polarization plots in 0.5 M H2SO4 electrolytes (Pt/C catalyst
reference was studied in 0.1 M HClO4) to study the effect of synthesis steps (that is, first doping and second adsorption) on Mn-N-C catalyst activity (a),
effect of various carbon hosts used for the second adsorption step on resultant activity (b), effect of different nitrogen sources used for the adsorption step
on final catalyst activity (c), and comparison of catalytic activity of Fe-, Co- and Mn-N-C catalysts prepared from identical procedures (d). e, Four-electron
selectivity (that is, H2O2 yields) of the ORR on Fe-, Co- and Mn-N-C catalysts. f, Fuel cell performance of the best-performing 20Mn-NC-second and 20Fe-
NC-second catalysts in both H2/O2 and H2/air conditions. Error bars represent the standard deviation from at least three independent measurements.

with Mn ions and form MnN4 active sites. Cyanamide exhibited the very low level (<​4%). This further indicates a 4e−​pathway, rather
best performance, compared with other N sources (for example, than the 2e− +​  2e− pathway. A Tafel slope of ~80 mV dec−1 for the
dipicolylamine, phenanthroline and melamine), probably due to its 20Mn-NC-second is comparable to that of Co- or Fe-based catalysts
size and unique C≡​N structures. The C≡​N structures decompose (Supplementary Fig. 22 and Supplementary Table 11), indicating a
at the elevated temperature and easily bond to unsaturated carbon similar rate-determining step involving mixed controls including
atoms to form stable pyridinic-N34. This is favourable for the forma- transfer of the first electron (118 mV dec−1) and the diffusion of
tion of MnN4 active sites. intermediates at catalyst surfaces (59 mV dec−1)44.
The influence of post-treatments on the Mn-NC host before The 20Mn-NC-second catalyst is further studied as a cathode in
adsorption was investigated. When directly using 20Mn-NC without membrane electrode assemblies (MEAs) for fuel cells. The open cir-
an acidic treatment as the host for the second adsorption step, the cuit voltage is 0.95 V using H2 and O2, suggesting a high intrinsic
activity showed an obvious decline (Supplementary Fig. 15). Acid ORR activity in fuel cell environments. The 20Mn-NC-second cath-
leaching effectively removes the MnO clusters from the 20Mn-NC ode is capable of generating current densities of 0.35 and 2.0 A cm−2
host, which likely creates pore structures and increases micropore at 0.6 and 0.2 V, respectively, at a reasonable 1.0 bar partial pressure
sizes (from 1.0 to 1.4 nm) and surface areas (from 699 to 821 m2 g−1) (Fig. 4f). The corresponding power density is up to 0.46 W cm−2. The
(Supplementary Figs. 16 and 17 and Supplementary Table 8), which is achieved performance exceeds the Fe-N-C catalyst derived from
favourable for the subsequent adsorption. As a result, the 20Mn-NC- PANI reported in 2010 (ref. 45); however, the performance is inferior
second catalyst exhibits the highest electrochemically accessible sur- to the 20Fe-NC-second catalyst especially in the kinetic range.
face area of 715 m2 g−1 (Supplementary Fig. 18 and Supplementary
Table 9), corresponding to the highest ORR activity. The correlation Enhanced catalyst stability in acids. The 20Mn-NC-second cata-
between the porous structure of the carbon host and the ORR activ- lyst exhibited excellent stability, as evidenced by a loss of only 17 mV
ity was established (Supplementary Figs. 19 and 20). in E1/2 after 30,000 potential cycles from 0.6 to 1.0 V in O2-saturated
The ORR activity of the best-performing 20Mn-NC-second cat- 0.5 M H2SO4 solution (Fig. 5a). The stability is largely enhanced
alyst, which is higher than that of a 20Co-NC-second catalyst and over traditional Fe-N-C catalysts derived from PANI (loss 80 mV
approaches that of a 20Fe-NC-second catalyst (Fig. 4d), is 60 mV after 5,000 cycles)45, as well as the ZIF-derived 20Fe-NC-second
less than of a Pt/C catalyst (~0.86 V). The atomically dispersed catalyst (29 mV loss) (Fig. 5b). The microstructure and morphology
Mn catalyst represents one of the best PGM- and Fe-free catalysts of carbon in the 20Mn-NC-second catalyst remain nearly the same
(Supplementary Table 10)43. The H2O2 yield of the 20Mn-NC- (Supplementary Fig. 23) after the potential cycling tests. Long-term
second catalyst is less than 2% indicating a four-electron reduction durability tests were conducted by holding at a constant potential of
pathway, which is comparable to that of the 20Fe-NC-second, and 0.7 V for 100 hours during the ORR (Fig. 5c). The 20Mn-NC-second
much lower than that of the 20Co-NC-second catalyst (Fig. 4e). catalyst retains 88% of its initial current density and exhibits a 29 mV
When the catalyst loadings vary within a wide range from 0.2 to loss of E1/2 after 100 hours (Fig. 5d). The identical cyclic voltam-
1.2 mg cm−2, the H2O2 yields (Supplementary Fig. 21) remain at a metry curves during the stability tests (inset of Fig. 5d) indicate

940 Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal


Nature Catalysis Articles
a b
0 20Mn-NC-second, initial, 0.795 V 0 20Fe-NC-second, initial, 0.822 V
20Mn-NC-second, after 30 K, 0.778 V 20Fe-NC-second, after 30 K, 0.793 V

Current density (mA cm–2)


Current density (mA cm–2)
–1 –1
After 30,000 cycles After 30,000 cycles
17 mV loss 29 mV loss
–2 –2

–3 –3

–4 –4
0.5 M H2SO4, 25 °C 0.5M H2SO4, 25 °C

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Potential (V versus RHE) Potential (V versus RHE)
c 120 d
20Mn-NC-second 20Mn-NC-second, initial, 0.795 V
0
20Mn-NC-second, after 100 h, 0.766 V
100
Relative current density (%)

87.5%

Current density (mA cm–2)


10
–1 5
80
0 29 mV loss
3% increase
60 –2 –5

–10
40 –15
@ E = 0.7 V versus RHE –3 0.0 0.2 0.4 0.6 0.8 1.0
200 r.p.m., 0.5 M H2SO4, 25 °C
20
–4
0.5 M H2SO4, 25 °C
0
0 10 20 30 40 50 60 70 80 90 100 0.0 0.2 0.4 0.6 0.8 1.0
Time (h) Potential (V versus RHE)
e f
20Mn-NC-second
100 0 20Mn-NC-second, initial, 0.795 V
20Fe-NC-second
Relative current density (%)

20Mn-NC-second, after 100 h, 0.777 V


Current density (mA cm–2)

90.2% 10
80 80.8% –1
78.3% 5
71.8% 18 mV loss
0
28% increase
60
–2 –5
57.2% –10
40 –15
@ E = 0.8 V versus RHE 38.9% –3
0.0 0.2 0.4 0.6 0.8 1.0
200 r.p.m., 0.5 M H2SO4, 25 °C
20
–4
0.5 M H2SO4, 25 °C
0
0 10 20 30 40 50 60 70 80 90 100 0.0 0.2 0.4 0.6 0.8 1.0
Time (h) Potential (V versus RHE)

Fig. 5 | Catalyst stability studied by using potential cycling and constant potentials. a,b, Steady-state ORR polarization plots before and after potential
cycling stability tests (0.6–1.0 V, 30,000 cycles) for the 20Mn-NC-second (a) and 20Fe-NC-second (b) catalysts. c,e, i–t curves (where i is current density
and t is time) at constant potentials of 0.7 V (c) and 0.8 V (e). d,f, Steady-state ORR polarization plots and cyclic voltammetry curves before and after
constant potential tests at 0.7 V (d) and 0.8 V (f) for 100 hours in O2-saturated 0.5 M H2SO4.

excellent resistance to carbon corrosion. To test the 20Mn-MC- 20Fe-NC-second catalyst (60%), suggesting an enhanced carbon
second catalyst under harsher conditions, we performed the dura- oxidation resistance. The catalyst degradation may be dependent on
bility test at 0.8 V (Fig. 5e). A larger decline of the current density total number of electrons passing through catalysts during the ORR.
of 57% after 100 hours is observed, relative to 0.7 V. The loss of cur- Therefore, we calculated the degradation rates against the total
rent density during the test can be partially recovered by potential electric charge (Supplementary Table 12). The degradation of E1/2
cycling between 0 to 1.0 V. For the first 20 hours, nearly 90% of the at 0.8 V is 0.237 mV C−1 for the 20Mn-NC-second catalyst, much
activity can be restored. However, the irreversible loss accumulates less than that for the 20Fe-NC-second catalyst (0.716 mV C−1).
and the current density only recovers 70% after 80 hours, indicating Stability tests at a constant current density (1.42 mA cm−2) further
that the activity loss is due to multiple factors. The reversible activ- verified that the 20Mn-NC-second catalyst is more stable than
ity loss may be associated with deactivation of active sites by anion the 20Fe-NC-second catalyst (Supplementary Fig. 25). Corrosion
binding or reversible adsorption of oxygen-containing functional resistance of carbon in catalysts was further assessed by cycling at
groups on local carbon atoms adjacent to MnN4 sites, which are a high potential range (1.0–-1.5 V) in N2-saturated 0.5 M H2SO4
considered to be part of the active site46. After the 100 hour test at (Supplementary Fig. 26). After 5,000 cycles, the capacitance of the
0.8 V, E1/2 showed a loss of 18 mV for the 20Mn-NC-second catalyst 20Mn-NC-second and 30Mn-NC-second increase 52% and 34%,
(Fig. 5f). In comparison, the 20Fe-NC-second catalyst only retains respectively, suggesting carbon oxidation in catalysts. However,
39% of its activity and a loss of E1/2 is up to 48 mV (Supplementary these values are much lower than those of 0Mn-NC-second (70%)
Fig. 24). The increase of capacitance for the 20Mn-NC-second cat- and the 20Fe-NC-second catalyst (94%) (Supplementary Table 9),
alyst (inset of Fig. 5f) is around 30%, much less than that of the indicating significantly enhanced carbon stability due to Mn doping

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 941


Articles Nature Catalysis

a b
2 U = 1.23 V
O2 *OOH
U = 0.80 V

Free energy (eV)


*O + *OH

*O + H2O *OH + H O 2H O
2 2
0

–1
Reaction coordinates

Fig. 6 | Fundamental understanding of possible Mn active sites by using DFT calculations. a, Atomistic structure of MnN4C12 active site in the 20Mn-
NC-second catalyst. b, Calculated free-energy evolution diagram for ORR through a 4e− associative pathway on the MnN4C12 active site under electrode
potential of U =​ 1.23 V and U =​ 0.80 V. c, Atomistic structure of initial state (left panel, that is, state *OOH in b), transition state (middle panel) and final
state (right panel, that is, state *O +​ *OH in b) for OOH dissociation reaction on MnN4C12 active site. Grey, blue, purple, red and white balls represent C, N,
Mn, O and H atoms, respectively.

into ZIF-8 precursors. Thus, the enhanced stability of the Mn-N-C calculated (Fig. 6c). We assumed a 4e− pathway, in which an O2 mol-
catalyst likely results from the improved corrosion resistance of the ecule will first adsorb on the top of central Mn and then O2 will be
carbon in the catalysts. protonated to form OOH. Next, the OOH will dissociate into O and
To evaluate catalyst stability under real fuel cell operating condi- OH, and finally both O and OH will be protonated to form the final
tions, we conducted 100 hour life tests at a challenging high volt- product H2O. We employed the computational hydrogen electrode
age of 0.7 V for MEAs made with the 20Mn-NC-second catalyst method developed by Nørskov et al.47 and computed the free ener-
(Supplementary Fig. 27). Although an initial degradation occurred gies of all elementary steps as a function of electrode potential U
at 0.7 V, enhanced durability of the Mn-N-C cathode was observed, with reference to the RHE. Figure 6b shows that, under the standard
compared with Co- and Fe-based catalysts15,22. The Mn-N-C catalyst potentials of the ORR, U =​ 1.23 V, that is, zero overpotential, some
prepared from the two-step doping and adsorption approach repre- of the elementary reactions along the 4e− ORR associative pathway
sents one of the most stable PGM-free catalysts in acidic electrolytes are endergonic and thus thermodynamically unfavourable on the
(Supplementary Table 10). In addition to the intrinsic stability of the MnN4C12 site. However, the free-energy change for these elemen-
20Mn-NC-second catalyst, non-electrochemical factors including tary reactions involving charge transfer will become negative (that
degradation of the catalyst/ionomer interfaces and water flooding is, exergonic reaction) when the electrode potential U is lower than
also influence the performance durability of PGM-free cathodes. a limiting potential of 0.80 V. Moreover, our computational results
(Fig. 6b) show that both the free-energy differences between O2*
DFT calculations. We identified a variety of possible MnNxCy and OOH* as well as between OH* and H2O* (* represents the
active sites and predicted their adsorption energies, free-energy adsorption of the molecule on the active site) are close to zero under
evolution and activation energies for the ORR in acids. As shown the limiting potential, suggesting that the MnN4C12 site could have
in Supplementary Fig. 28, nine possible active sites (denoted as an optimal activity for the ORR48. We performed the climbing image
MnN2C12, MnN3C9, MnN3C11, MnN4C8, MnN4C10, MnN4C12, nudged elastic band (CI-NEB) calculation to locate the transition
MnN5C10, Mn2N5C12 and Mn2N6C14) are carefully examined in the state and predict the activation energies for the OOH dissociation
computational study. The predicted adsorption energy of O2, OOH, reaction49, which is the crucial step of breaking the O–O bond for
O, OH and H2O on each of these potential MnNxCy active sites are the 4e− ORR associative pathway on the MnN4C12 site. Figure 6c
presented in Supplementary Table 13. The DFT adsorption energy shows the atomic details of this reaction. In the initial state, OOH
results indicate that MnN2C12 and MnN3C9 sites, which have fewer is adsorbed on the central Mn atom; in the final state, both the dis-
chelated N atoms around the central metal atom than the other sociated O and OH are co-adsorbed on the central Mn atom. The
sites, bind the final product H2O too strongly to be good active sites. reaction process from the initial state to the final state requires pass-
Moreover, MnN3C11, MnN4C8, and Mn2N5C12 sites were predicted to ing a transition state and overcoming an energy barrier of 0.49 eV,
bind the intermediate OH too strongly, which are not good active which is surmountable. Consequently, the DFT results predict that
sites as well. Computational screening identified that MnN4C10, it is both thermodynamically and kinetically favourable for the 4e−
MnN4C12, MnN5C10, and Mn2N6C14 sites have an appropriate ORR to occur on the MnN4C12 site. According to the XANES and
binding strength with the ORR species and can be active sites for EXAFS analysis (Fig. 2), the atomic Mn sites in the catalysts have
the 4e− pathway. an oxidation state of 2+​ and coordinate with N mainly in the form
XAS analyses suggest that MnN4 structures are likely. Among of MnN4. The MAADF-STEM images (Fig. 3) further proved the
studied MnN4Cy sites, the MnN4C12 site (a Mn–N4 moiety bridging atomic dispersion and N was detected in the immediate vicinity
over two adjacent zigzag graphitic edges, as shown in Fig. 6a) was of the single Mn sites by EEL point spectrum. Hence, the compu-
further predicted as the most optimal active sites (Supplementary tational model explains well our experimental findings. Although
Table 13). Its free-energy evolution for the ORR (Fig. 6b) and the multiple possible sites in nanocarbon are able to coordinate with
transition state for an OOH dissociation on the active site were also single metals sites50, their ORR activities and stabilities in acids are

942 Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal


Nature Catalysis Articles
not sufficient. In contrast, the MnN4 structure therefore is the most Physical characterization. The overall particle size and distribution of the
stable and the most active sites for the ORR in acids, likely due to the catalysts’ morphologies were analysed using scanning electron microscopy on
a Hitachi SU 70 microscope at a working voltage of 5 kV. The crystal phases
lone pair electrons in pyridinic-N, which can form strong coordina- in catalyst samples were studied by using powder XRD on a Rigaku Ultima IV
tion bonds with transition metals. diffractometer with Cu Kα​X-rays. XPS was carried out using a Kratos AXIS
We also performed DFT calculations to specifically examine the Ultra DLD XPS system equipped with a hemispherical energy analyser and a
influence of water solvent on the ORR in Mn-N-C catalysts, sug- monochromatic Al Kα​source operated at 15 keV and 150 W. Pass energy was fixed
at 40 eV for all of high-resolution scans. The N2 isothermal adsorption/desorption
gesting that the water solvent effect is not significant enough to
was recorded at 77 K on a Micromeritics TriStar II. Before measurements, samples
alter the reaction mechanism for the ORR on the Mn-N-C cata- were degassed at 150 °C for 5 h under vacuum. Atomic-resolution MAADF
lysts (Supplementary Fig. 29). The MnN4 sites with surrounding images of atomically dispersed Mn sties were captured in a Nion Ultra STEM
graphitic-N dopants would also be active to promote the 4e− ORR U100 operated at 60 keV and equipped with a Gatan Enfina electron energy loss
following the same OOH dissociation pathway (Supplementary spectrometer at Oak Ridge National Laboratory. HR-TEM and high-angle annular
dark field STEM were performed on JEOL JEM-2100F and Hitachi HD2700C
Fig. 30). The MnN4 site with immediately adjacent graphitic-N with a probe-corrector, respectively, at Brookhaven National Laboratory. Mn
dopants could have a lower limiting potential and lower activa- K-edge EXAFS experiments were carried out at beamline 9BM-C at Advanced
tion energy for O–O bond breaking than the MnN4 site without Photon Sources at Argonne National Laboratory. The EXAFS data were collected
graphitic-N dopants. in fluorescence mode due to low Mn concentration by using a Vertox ME4 silicon
drift diode detector. Data analysis and EXAFS fitting were performed with the
Athena, Artemis, and IFEFFIT software packages.
Conclusion
An atomically dispersed and nitrogen-coordinated Mn-N-C cata- Electrochemical measurements. An electrochemical workstation (CHI760b)
lyst has been developed through a two-step doping and adsorption was employed to perform electrochemical measurements in a three-electrode
approach, which is effective to significantly increase the density of cell, in which an Hg/HgSO4 (K2SO4-saturated) electrode and a graphite rod
active sites. Atomically dispersed MnN4 sites were verified using were used as the reference and counter electrodes, respectively. A RRDE (Pine,
AFMSRCE 3005) with a disk diameter of 5.6 mm was used as working electrode.
XAS experiments to determine their possible coordination as well The reference electrode was calibrated to a RHE in the same electrolyte before
as directly observed at the atomic scale using low voltage, aberra- each measurement. The calibration of reference electrode was performed in
tion-corrected STEM imaging coupled with EELS analysis. The a two-electrode system. In the studied electrolyte (that is, 0.5 M H2SO4), a Pt
high activity of the Mn-N-C catalyst was evidenced by a commend- wire coated with Pt black was used for the RHE. The Pt electrode is purged and
able E1/2 of 0.80 V versus RHE in acids, which results from intrinsic saturated with high purity H2 during calibrating. A two-probe multi-meter was
employed to measure the potential difference between the reference electrode and
activity of atomically dispersed MnN4 sites with increased density the Pt electrode, which is relative to RHE. The catalyst ink for the RRDE tests was
within a 3D porous ZIF-derived carbon. The remarkable stability is prepared by ultrasonically dispersing 5.0 mg catalysts and 30 µ​l Nafion (5 wt%) into
due to the robust MnN4 sites and the enhanced corrosion resistance 2.0 ml isopropanol solution. The ink was drop-casted on the disk electrode with a
of adjacent carbon derived from Mn doping. DFT calculations also controlled loading of 0.8 mg cm−2 and dried at room temperature to yield a thin-
further confirmed that the MnN4C12 site has a favourable binding film electrode. The electrocatalytic activity for the ORR was tested by steady-state
measurement using staircase potential control with a step of 0.05 V at an interval
energy with O2, OOH and H2O during the ORR as well as a sur- of 30 s from 1.0 to 0 V versus RHE in O2-saturated 0.5 M H2SO4 solution at 25 °C
mountable energy barrier to break O–O bonds for complete 4e− and a rotation rate of 900 r.p.m. Before recording the steady-state ORR polarization
reduction. Therefore, the reported atomically dispersed Mn-N-C plots, cyclic voltammetry in O2-saturated 0.5 M H2SO4 at a scan rate of 50 mV s−1
catalyst demonstrates an alternative concept to develop robust and was performed to activate the catalysts until repeatable capacitance of catalyst
samples is achieved. Four-electron selectivity during the ORR was determined by
highly active PGM-free catalysts as replacements for Fe catalysts in
measuring the ring current for calculating H2O2 yield. Catalyst stability was studied
future PEMFC technologies. by cycling potentials from 0.6 to 1.0 V in O2-saturated 0.5 M H2SO4 at a scanning
rate of 50 mV s−1. To determine the carbon corrosion of catalysts, potential cycling
Methods was conducted at high potentials from 1.0 to 1.5 V in N2-saturated 0.5 M H2SO4 at a
Catalysts synthesis. In the first doping step, Mn-doped ZIF-8 precursors scanning rate of 500 mV s−1. In addition, holding at constant potential (for example,
were synthesized in a dimethylformamide (DMF) solution. The precursors are 0.7 and 0.8 V) or constant current density for 100 h during the ORR was carried
designated as nMn-ZIF, where n is the molar percentage of Mn against total out. Pt/Vulcan XC-72 (20 wt% Pt) was used as the Pt/C reference, which was tested
metals (Mn and Zn) in the solutions. In this work, the n value varied from 0 in 0.1 M HClO4 solution with a Pt loading of 60 µ​g  cm−2.
(Mn free) to 30 at% and the concentration of 2-methylimidazle and Zn2+ were
controlled at 80 and 40 mmol l−1, respectively. Typically, a controlled amount Fuel cell tests. The best-performing Mn-N-C catalyst was used to prepare
of zinc(ii) nitrate hexahydrate and manganese(iii) acetate dihydrate were cathodes for MEA tests in real fuel cell environments. The cathode catalyst inks
dissolved into DMF to form a uniform solution in a round-bottom flask (A). were dispersed by ball mixing the catalyst in 1-proponal, de-ionized water, and
2-Methylimidazole was dissolved in a conical-flask with DMF solution (B). Nafion suspension for 2 d. The inks were brush painted on one side of a Nafion
After mixing solutions A and B, the reaction temperature was increased to 212 membrane until the cathode catalyst loading reached ~4.0 mg cm−2. An in-
120 °C, holding for 24 h to allow Mn-doped ZIF-8 nanocrystals to grow. After house-made Pt electrode decal (0.25 mg Pt cm−2) was used as the anode, and it was
cooling down to room temperature, the Mn-ZIF-8 nanocrystals were collected by transferred onto the other side of brush painted Nafion 212 membrane at ~150 °C
centrifugation, washed at least three times with ethanol, and then dried at 60 °C for 3 min. The full catalyst-coated membrane, which had an active geometric area
in a vacuum oven for 5 h. The Mn-ZIF-8 precursors were subsequently heated at a of 5.0 cm2, was inserted between two Toray H30 gas diffusion layers and assembled
temperature of 1,100 °C in a tube furnace under N2 flow for 1 h to obtain the Mn into a single cell with single-serpentine flow channels. The single cell was then
and N co-doped nanocarbon (NC). The samples were labelled as nMn-NC, where evaluated in a fuel cell test station (100 W, Scribner 850e, Scribner Associates). The
n is atomic percent of Mn against total metal (for example, 20Mn-NC). Next, cells were conditioned at 0.3 V, 100% relative humidity and 80 °C for at least 3 h
acid leaching treatment in 0.5 M H2SO4 solution at 80 °C for 5 h was carried out to and until the steady-state current was reached. Air/oxygen flowing at 200 ml min−1
remove Mn clusters and open porous structures of nMn-NC samples. Typically, and H2 (purity 99.999%) flowing at 200 ml min−1 were used as the cathode and
100 ml acid solution was used for 100 mg sample. An additional heat treatment anode reactants, respectively. The back pressures during the fuel cell tests are based
at 900 °C under N2 flow for 3 h is necessary to repair the carbon structure. on US Department of Energy protocols, that is 1.0 bar reactant gas. Because of
Such-prepared carbon samples labelled as nMn-NC-first are ready for the 100% relative humidity, vapour pressure is around 0.5 bar. Thus, the total pressure
second adsorption step. applied to MEAs is around 1.5 bar (150 KPa). Fuel cell polarization curves were
In the second adsorption step, the nMn-NC-first powder was dispersed into a recorded in a current control mode. All the cathode catalyst layers contain 35 wt%
mixed solution (isopropanol: water =​ 1:1) containing Mn(ii) chloride and nitrogen of Nafion.
sources (for example, dipicolylamine, cyanamide, phenanthroline or melamine).
After 2 h ultrasonication along with and 5 h magnetic stirring, the mixture was Computational methods. All the first-principles DFT calculations in this work
collected by centrifugation (13,552g) and then dried at 60 °C in a vacuum oven were performed using the Vienna ab initio simulation package. The projector
for 5 h. A subsequent thermal activation at temperature of 1,100 °C under N2 augmented wave pseudopotential was used to describe the core electrons of
atmosphere for 1 h was conducted to synthesize the final catalysts, labelled as nMn- the elements. The energy cut-off was set as 400 eV to expand wave function.
NC-second. As control samples, 20Co-NC-second and 20Fe-NC-second samples The electronic exchange-correlation was described by generalized gradient
were prepared through similar procedures. approximation with the Perdew, Burke and Ernzernhof functional. The brillouin

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 943


Articles Nature Catalysis
zone was sampled with Monkhorst-pack 4 ×​  4  ×​  1 k-point mesh for active sites 23. Zhong, Y. et al. The constraints of transition metal substitutions (Ti, Cr, Mn,
MnN2C12, MnN3C9, MnN3C11, MnN4C10, MnN5C10, Mn2N6C14 and Mn2N5C12, Co and Ni) in magnetite on its catalytic activity in heterogeneous fenton and
3 ×​  3  ×​  1 k-point mesh for site MnN4C12, and 4 ×​  3  ×​  1 k-point mesh for site UV/Fenton reaction: from the perspective of hydroxyl radical generation.
MnN4C8. A vacuum region of 14 Å thick was added in the direction normal to the Appl. Catal. B 150–151, 612–618 (2014).
carbon layer to ensure negligible interaction between the slab and its images. In 24. Wu, G. et al. Synthesis–structure–performance correlation for polyaniline–
the DFT structure optimization calculations, the atomic positions were allowed Me–C non-precious metal cathode catalysts for oxygen reduction in fuel cells.
to relax until the force on each ion was below 0.01 eV Å−1. The transition states of J. Mater. Chem. 21, 11392–11405 (2011).
chemical reaction were located using the CI-NEB method, in which the force along 25. Gupta, S. et al. Quaternary FeCoNiMn-based nanocarbon electrocatalysts for
and perpendicular to the reaction path were relaxed to less than 0.05 eV Å−1. Zero- bifunctional oxygen reduction and evolution: promotional role of Mn doping
point energy correction has been included in all the reported energies. in stabilizing carbon. ACS Catal. 7, 8386–8393 (2017).
26. Wang, X. et al. Size-controlled large-diameter and few-walled
Data availability carbon nanotube catalysts for oxygen reduction. Nanoscale 7,
The data that support the findings of this study are available from the primary 20290–20298 (2015).
corresponding author (G.Wu) upon reasonable request. 27. Zhang, H., Osgood, H., Xie, X., Shao, Y. & Wu, G. Engineering
nanostructures of PGM-free oxygen-reduction catalysts using metal–organic
frameworks. Nano Energy 31, 331–350 (2017).
Received: 16 May 2018; Accepted: 17 September 2018; 28. Pan, F. et al. Unveiling active sites of CO2 reduction on nitrogen-
Published online: 29 October 2018 coordinated and atomically dispersed iron and cobalt catalysts. ACS Catal. 8,
3116–3122 (2018).
References 29. Wang, X. X. et al. Ordered Pt3Co intermetallic nanoparticles derived from
1. Lefèvre, M., Proietti, E., Jaouen, F. & Dodelet, J.-P. Iron-based catalysts with metal–organic frameworks for oxygen reduction. Nano. Lett. 18, 4163–4171
improved oxygen reduction activity in polymer electrolyte fuel cells. Science (2018).
324, 71–74 (2009). 30. Qiao, Z. et al. 3D polymer hydrogel for high-performance atomic iron-rich
2. Proietti, E. et al. Iron-based cathode catalyst with enhanced power catalysts for oxygen reduction in acidic media. Appl. Catal. B 219, 629–639
density in polymer electrolyte membrane fuel cells. Nat. Commun. 2, (2017).
416 (2011). 31. Gupta, S. et al. Engineering favorable morphology and structure of Fe-N-C
3. Wu, G. & Zelenay, P. Nanostructured nonprecious metal catalysts for oxygen oxygen-reduction catalysts through tuning of nitrogen/carbon precursors.
reduction reaction. Acc. Chem. Res. 46, 1878–1889 (2013). ChemSusChem 10, 774–785 (2017).
4. Gewirth, A. A., Varnell, J. A. & DiAscro, A. M. Nonprecious metal catalysts 32. Wang, H., Zhu, Q.-L., Zou, R. & Xu, Q. Metal-organic frameworks for energy
for oxygen reduction in heterogeneous aqueous systems. Chem. Rev. 118, applications. Chem 2, 52–80 (2017).
2313–2339 (2018). 33. Chen, Y. et al. Isolated single iron atoms anchored on N-doped porous
5. Chung, H. T. et al. Direct atomic-level insight into the active carbon as an efficient electrocatalyst for the oxygen reduction reaction.
sites of a high-performance PGM-free ORR catalyst. Science 357, Angew. Chem. Int. Ed. 56, 6937–6941 (2017).
479–484 (2017). 34. Zhang, H. et al. Single atomic iron catalysts for oxygen reduction in acidic
6. Wu, G., More, K. L., Johnston, C. M. & Zelenay, P. High-performance media: particle size control and thermal activation. J. Am. Chem. Soc. 139,
electrocatalysts for oxygen reduction derived from polyaniline, iron, and 14143–14149 (2017).
cobalt. Science 332, 443–447 (2011). 35. Xia, B. Y. et al. A metal–organic framework-derived bifunctional oxygen
7. Yan, D. et al. Defect chemistry of nonprecious-metal electrocatalysts for electrocatalyst. Nat. Energy 1, 15006 (2016).
oxygen reactions. Adv. Mater. 29, 1606459 (2017). 36. Yang, L., Zeng, X., Wang, W. & Cao, D. Recent progress in MOF-derived,
8. Gong, K., Du, F., Xia, Z., Durstock, M. & Dai, L. Nitrogen-doped carbon heteroatom-doped porous carbons as highly efficient electrocatalysts
nanotube arrays with high electrocatalytic activity for oxygen reduction. for oxygen reduction reaction in fuel cells. Adv. Func. Mater. 28,
Science 323, 760–764 (2009). 1704537 (2018).
9. Masa, J., Xia, W., Muhler, M. & Schuhmann, W. On the role of metals in 37. Feng, Z. et al. Atomic-scale cation dynamics in a monolayer VOx/α​-Fe2O3
nitrogen‐doped carbon electrocatalysts for oxygen reduction. Angew. Chem. catalyst. RSC Adv. 5, 103834–103840 (2015).
Int. Ed. 54, 10102–10120 (2015). 38. Weng, Z. et al. Active sites of copper-complex catalytic materials for
10. Liu, J. et al. High performance platinum single atom electrocatalyst for electrochemical carbon dioxide reduction. Nat. Commun. 9, 415 (2018).
oxygen reduction reaction. Nat. Commun. 8, 15938 (2017). 39. Liu, Z. et al. Tuning the electronic environment of zinc ions with a ligand for
11. Zitolo, A. et al. Identification of catalytic sites for oxygen reduction dendrite-free zinc deposition in an ionic liquid. Phys. Chem. Chem. Phys. 19,
in iron- and nitrogen-doped graphene materials. Nat. Mater. 14, 25989–25995 (2017).
937–942 (2015). 40. Chen, Y. Z. et al. From bimetallic metal–organic framework to porous
12. Jiang, W. J. et al. Understanding the high activity of Fe-N-C electrocatalysts carbon: high surface area and multicomponent active dopants for excellent
in oxygen reduction: Fe/Fe3C nanoparticles boost the activity of Fe-Nx. J. Am. electrocatalysis. Adv. Mater. 27, 5010–5016 (2015).
Chem. Soc. 138, 3570–3578 (2016). 41. Lai, Q. et al. Metal–organic-framework-derived Fe-N/C electrocatalyst with
13. Wu, G. et al. Carbon nanocomposite catalysts for oxygen reduction and five-coordinated Fe-Nx sites for advanced oxygen reduction in acid media.
evolution reactions: from nitrogen doping to transition-metal addition. ACS Catal. 7, 1655–1663 (2017).
Nano Energy 29, 83–110 (2016). 42. Wang, Q. et al. Phenylenediamine-based FeNx/C catalyst with high activity
14. Banham, D. et al. A review of the stability and durability of non-precious for oxygen reduction in acid medium and its active-site probing. J. Am.
metal catalysts for the oxygen reduction reaction in proton exchange Chem. Soc. 136, 10882–10885 (2014).
membrane fuel cells. J. Power Sources 285, 334–348 (2015). 43. Sahraie, N. R. et al. Quantifying the density and utilization of active sites
15. Ferrandon, M. et al. Stability of iron species in heat-treated polyaniline–iron– in non-precious metal oxygen electroreduction catalysts. Nat. Commun. 6,
carbon polymer electrolyte fuel cell cathode catalysts. Electrochim. Acta 110, 8618 (2015).
282–291 (2013). 44. Li, Y. et al. An oxygen reduction electrocatalyst based on carbon nanotube-
16. Wu, G. Current challenge and pÿerspective of PGM-free cathode catalysts graphene complexes. Nat. Nanotech. 7, 394–400 (2012).
for PEM fuel cells. Front. Energy 11, 286–298 (2017). 45. Wu, G. et al. Titanium dioxide-supported non-precious metal oxygen
17. Wu, G. et al. Performance durability of polyaniline-derived non-precious reduction electrocatalyst. Chem. Commun. 46, 7489–7491 (2010).
cathode catalysts. ECS Trans. 25, 1299–1311 (2009). 46. Herranz, J. et al. Unveiling N-protonation and anion-binding effects on
18. Choi, C. H. et al. Stability of Fe-N-C catalysts in acidic medium Fe/N/C-catalysts for O2 reduction in pem fuel cells. J Phys. Chem. C 115,
studied by operando spectroscopy. Angew. Chem. Int. Ed. 54, 16087–16097 (2011).
12753–12757 (2015). 47. Norskov, J. K. et al. Origin of the overpotential for oxygen reduction at
19. Kramm, U. I., Lefevre, M., Bogdanoff, P., Schmeisser, D. & Dodelet, J. P. a fuel-cell cathode. J. Phys. Chem. B 108, 17886–17892 (2004).
Analyzing structural changes of Fe-N-C cathode catalysts in pem fuel cell by 48. Kulkarni, A., Siahrostami, S., Patel, A. & Norskov, J. K. Understanding
mossbauer spectroscopy of complete membrane electrode assemblies. J. Phys. catalytic activity trends in the oxygen reduction reaction. Chem. Rev. 118,
Chem. Lett. 5, 3750–3756 (2014). 2302–2312 (2018).
20. Goellner, V. et al. Degradation of Fe/N/C catalysts upon high polarization in 49. Henkelman, G., Uberuaga, B. P. & Jónsson, H. A climbing image nudged
acid medium. Phys. Chem. Chem. Phys. 16, 18454–18462 (2014). elastic band method for finding saddle points and minimum energy paths.
21. Walling, C. Fenton’s reagent revisited. Acc. Chem. Res. 8, 125–131 (1975). J. Chem. Phys. 113, 9901–9904 (2000).
22. Wang, X. X. et al. Nitrogen-coordinated single cobalt atom catalysts for 50. Zhang, C., Zhang, W. & Zheng, W. Pinpointing single metal atom anchoring
oxygen reduction in proton exchange membrane fuel cells. Adv. Mater. 30, sites in carbon for oxygen reduction: Doping sites or defects? Chin. J. Catal.
1706758 (2018). 39, 4–7 (2018).

944 Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal


Nature Catalysis Articles
Acknowledgements out electrochemical measurements. D.A.C., K.L.M, S.H. and D.S performed electron
G.Wu thanks the Research and Education in eNergy, Environment and Water (RENEW) microscopy analyses and data interpretation. S.K. conducted XPS analysis. M.W., M.L.,
program at the University at Buffalo, SUNY and National Science Foundation (CBET- G.E.S. and Z.F. recorded and analysed XAS data. C.L. and H.X. carried out fuel cell tests.
1604392, 1804326) for partial financial support. G.Wu, G.Wang and H.X. acknowledge B.L., K.L. and G.Wang conducted computational studies.
support from the US Department of Energy (DOE), Energy Efficiency and Renewable
Energy, Fuel Cell Technologies Office (DE-EE0008075). Electron microscopy research Competing interests
was conducted at Oak Ridge National Laboratory’s Center for Nanophase Materials The authors declare no competing interests.
Sciences of (D.A.C. and K.L.M) and the Center for Functional Nanomaterials at
Brookhaven National Laboratory (S.H. and D.S., under contract No. DE-SC0012704),
which both are US DOE Office of Science User Facilities. XAS measurements were Additional information
performed at beamline 9-BM at the Advanced Photon Source, a User Facility operated Supplementary information is available for this paper at https://doi.org/10.1038/
for the US DOE Office of Science by Argonne National Laboratory under Contract s41929-018-0164-8.
No. DE-AC02-06CH11357 (Z.F. and G.E.S.). Z.W. and J.L. thank the National Natural Reprints and permissions information is available at www.nature.com/reprints.
Science Foundation of China (Grant No. 21273058 and 21673064) for support.
Correspondence and requests for materials should be addressed to Z.W. or G.W.

Author contributions Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
G.Wu, Z.W. and J. L. designed the experiments, analysed the experimental data, and published maps and institutional affiliations.
wrote the manuscript. J.L., M.C. and H.Z. synthesized catalyst samples and carried © The Author(s), under exclusive licence to Springer Nature Limited 2018

Nature Catalysis | VOL 1 | DECEMBER 2018 | 935–945 | www.nature.com/natcatal 945

You might also like