Sei sulla pagina 1di 7

Research Article

pubs.acs.org/acscatalysis

Modeling Fe/N/C Catalysts in Monolayer Graphene


Xiao-Dong Yang, Yanping Zheng, Jing Yang, Wei Shi, Jin-Hui Zhong, Cankun Zhang, Xue Zhang,
Yu-Hao Hong, Xin-Xing Peng, Zhi-You Zhou,* and Shi-Gang Sun*
Collaborative Innovation Center of Chemistry for Energy Materials, College of Chemistry and Chemical Engineering, Xiamen
University, Xiamen, Fujian 361005, Peoples Republic of China
S Supporting Information
*

ABSTRACT: Pyrolyzed Fe/N/C is one of the most


promising non-precious-metal catalysts for the oxygen
reduction reaction (ORR), which is supposed to boost the
commercialization of proton exchange membrane fuel cells
(PEMFC). However, the nature of the active sites of Fe/N/C
is not clear and has long been debated. The challenges mainly
come from highly heterogeneous structures formed during the
pyrolysis process as well as no suitable surface probes. To
elucidate the active sites, the most eective approach is
building well-dened model catalysts as single-crystal planes in
surface sciences. Herein, we designed a single-atomic-layer Fe/
N/C model catalyst based on monolayer graphene (FeNMLG) to explore the active sites. The model catalyst was prepared by 950 C heat treatment of graphene with controlled defects
under an FeCl3(g)/NH3 atmosphere. The as-prepared model catalyst exhibits ORR activity and SCN suppressive eect
comparable to those of normal nanoparticle-like Fe/N/C catalysts, indicating that active sites are successfully created in the
model catalyst. The eect of defect density, the layer number of graphene, and nitrogen species on the ORR activity has been
investigated. The main content of nitrogen species on FeN-MLG is Nx-Fe, and quantitative correlation between Nx-Fe and ORR
activity demonstrates that Nx-Fe species are the active site of Fe/N/C catalysts. The proposed model catalyst serves to simplify
the catalyst structures and to simulate the topmost atomic layer of normal Fe/N/C, where ORR is catalyzed. This model system
opens an opportunity to further understand the highly heterogeneous Fe/N/C catalysts.
KEYWORDS: oxygen reduction reaction, iron-based catalysts, monolayer graphene, model catalysts, active site

INTRODUCTION
Today, platinum-based catalysts are exclusively used in proton
exchange membrane fuel cells (PEMFCs) due to the sluggish
kinetics of the oxygen reduction reaction (ORR). For largescale commercialization of PEMFCs, there is a considerable
motivation to search for Earth-abundant materials to replace
platinum-based catalysts. Iron- and nitrogen-doped carbon (Fe/
N/C) catalysts are generally considered to be some of the most
promising candidates. Recently, signicant progress in Fe/N/C
catalysts has been made to a level comparable with that of
platinum-based catalysts.1,2 However, the structure of the active
sites has remained contentious even after 50 years of research.
Various structures have been proposed as active sites in the
literature, such as pyridinic N,3 iron encapsulated in graphene
nanoshells,4,5 four-coordinated FeN4,1,6 Fe-N2+2 bridging two
graphene edges,7 and ve-coordinate N-FeN2+2.8 Challenges in
the exploration of active sites are discussed as follows.
Pyrolyzed Fe/N/C is inherently highly heterogeneous.
Classical Fe/N/C has at least ve types of nitrogens.9,10 In
addition, the environment of iron atom is more complicated,11
being is either linked (coordinated) to nitrogen or carbon or
incorporated as metal nanoparticles. Thus, recording the
spectroscopic ngerprint of active sites is prevented. It becomes
XXXX American Chemical Society

more dicult when considering the noncrystallographic


ordering of the metal and nitrogen atoms. This complicated
situation makes it dicult to determine the ORR activity of
special species, such as inorganic Fe. Fe/Fe3C nanocrystals
were suggested to boost the activity.4,5,12 However, in the work
of Jaouen13 and Kramm,14 they believed inorganic Fe species
were ORR inactive, because equal ORR activity was observed in
the catalysts with or without inorganic Fe species.
The ORR is catalyzed on the topmost atomic layer.
Consequently, regular analytical methods provide only bulk
information on Fe/N/C catalysts, including Mo ssbauer
spectra,7,13,14 X-ray absorption near-edge spectroscopy
(XANES),7,13,14 time-of-ight secondary ion mass spectrometry
(TOF-SIMS),15 and X-ray photoelectron spectrometry (XPS,
near-surface information).16,17 Without specic analysis of the
catalyst surface structures where ORR happens, inexact or
incorrect conclusions may be made. It is obvious that the
exploration of surface species is much more important than
bulk analysis. Eorts have been made to apply strong ligands as
Received: September 21, 2016
Revised: November 14, 2016
Published: November 16, 2016
139

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis
surface probes, such as CO (work at 193 K)18 and halide ions
and sulfur-containing species (e.g., SCN, SO2, and H2S).10
Nevertheless, to date no suitable surface probes for Fe/N/C
have yet been developed to discriminate moieties with a good
resolution (e.g., iron in atomic dispersion and iron clusters).
Hence, both the heterogeneous structures and the lack of
suitable characterization technologies limit a full understanding
of Fe/N/C catalysts.
In the history of electrochemical catalysts, model catalysts
have provided simplied systems, which lead to an atomic-level
understanding of the reactions. The classic model catalysts, in
the past, were metal single-crystal planes with well-dened
surface atomic arrangements. Model systems for nonmetal
catalysts have been developed recently in monolayer graphene.
Graphene is a one-atom layer of sp2 carbon. The well-dened
structure and interesting electronic/mechanical properties1922
make graphene as perfect model platform for fundamental
electrochemistry research. The Dryfe group measured the
electron transfer kinetics between redox mediators (Fe(CN)63,
Ru(NH3)63+, and IrCl62) and the carbon surface by modeling
in monolayer and multilayer graphene.23,24 Correlation
between electrochemical activity and the defect density of
carbon material was simulated in a model graphene with a
series of defect densities.25 Pyridinic nitrogen was proved to
create ORR activity in a model catalyst with well-controlled Ndoping species in highly oriented pyrolitic graphite (HOPG).26
In comparison with HOPG, graphene is a more ideal carbon
material for preparing Fe/N/C model catalysts, because all of
the structure information provided by characterization
technologies comes strictly from the topmost atomic layer,
where the ORR occurs. In addition, no metals were introduced
into the HOPG-based model catalyst.26 It is well-known that
doping Fe can increase the ORR activity greatly, especially in
acidic medium.
In this study, a single atomic layer of a Fe/N/C model
catalyst was achieved by treating defective monolayer graphene
(MLG) at 950 C with FeCl3/NH3 sources. The model catalyst
exhibits ORR behaviors similar to those of normal pyrolyzed
Fe/N/C catalysts, including kinetic activities and SCN
suppressive eects. We further found that model catalysts in
monolayer and multilayer graphene have similar ORR activities.
More importantly, Nx-Fe (>50%) is found to be the main
content of the nitrogen species in model catalysts. A
quantitative correlation between Nx-Fe and ORR activity is
demonstrated.

Figure 1. Building Fe/N/C model catalysts in monolayer graphene.


Prior to iron/nitrogen doping, defects in graphene were introduced by
Ar+ irradiation. A heat treatment procedure was performed at 950 C
in FeCl3/NH3 to achieve Fe/N doping.

The second step of the heat treatment was Fe/N doping. The
sample-heating zone was heated from 350 to 950 C, and the
Fe-supplying zone was heated to 310 C to sublimate solid
FeCl3. Both heating zones were set up to reach the target
temperatures at the same time. NH3 gas was applied as a
nitrogen source. When the temperatures and Fe/N sources
were ready, graphene samples were moved back into the
sample-heating zone to achieve iron and nitrogen codoping.
Finally, Fe/N/C model catalysts in graphene were obtained. In
addition, MLG doped with nitrogen alone was prepared
through the same procedure but only in an NH3 environment.
For abbreviation, defected MLG without heteroatom doping,
with nitrogen doping alone, and with iron/nitrogen codoping
are denoted as D-MLG, N-MLG, and FeN-MLG, respectively.
FeN-MLG_5 min and FeN-MLG_10 min represent iron/
nitrogen codoping samples achieved at 950 C in 5 and 10 min,
respectively. Also, a sample denoted as FeN-KJ600 was made
for comparison from KJ600 carbon black by the same
synthesizing procedure.
Single Atomic Layer of Model Catalyst. The as-prepared
Fe/N/C model catalyst retained the single-atomic-layer
structure of graphene, which was conrmed by atomic force
microscopy (AFM), transmission electron microscopy (TEM),
and selective area electron diraction (SAED), as shown in
Figure 2. Figure 2A demonstrates the AFM image of a folded
FeN-MLG, as schematized at Figure 2B. From left to right, line
1 is drawn from the bottom-layer graphene to the upper-layer
graphene, and line 2 is drawn from the upper-layer graphene to
the SiO2/Si substrate. The corresponding height prole in
Figure 2C reveals that one layer of FeN-MLG is about 0.5 nm.
This is slightly higher than the expected value (0.35 nm) of
MLG but is still in the morphology of a single atomic layer.
To perform TEM, MLG was transferred to a SiN TEM grid
with a 2.8 m i.d. hole pattern and then doped with Fe/N.
Figure 2D shows a bright-eld TEM image of FeN-MLG. It is
clean and uniform. Figure 2E shows SAED patterns, with the

RESULTS AND DISCUSSION


Synthesis of Fe/N/C Model Catalysts in Graphene.
Figure 1 gives a schematic of building Fe/N/C model catalysts
in MLG. The CVD monolayer graphene was transferred to a
SiO2/Si substrate. Prior to heteroatom doping, defects in
graphene were introduced by Ar+ irradiation. The irradiation
treatment was carried out at a power of 10 W and a pressure of
500 mTorr with an Ar ow rate of 20 sccm. An exposure time
was employed to control defect densities. Samples were then
processed by heat treating. As illustrated in Figure 1, there are
two heating zones in the tube furnace. One is for treating
samples (graphene), and the other is for evaporating FeCl3 to
supply Fe sources. Heat treatment was carried out in two steps.
The rst step was cleaning the graphene sample. Graphene was
heated to 350 C for 3 h under an Ar atmosphere to remove
any possible residue on the graphene surface. Thereafter, the
sample was moved out of the heating zone by a magnetic rod.
140

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis

Figure 3. Spectroscopic characterization of model catalysts, including


iron/nitrogen-codoped MLG (FeN-MLG), nitrogen-doped MLG (NMLG), defective MLG (D-MLG), and pristine monolayer graphene
(MLG). (A) Raman spectra of FeN-MLG, N-MLG, D-MLG, and
MLG, as the D band and G band upshift with heteroatom doping.
Conditions: laser, 638 nm; spot size, 500 nm; 100 objective. Highresolution XPS of FeN-MLG, N-MLG, D-MLG, and MLG for Fe 2p
(B) and N 1s (C). N 1s spectra were deconvoluted into pyridinic-type
nitrogen (BE = 398.8 eV), metalnitrogen (399.9 eV), and graphitictype nitrogen (401.2 eV).

Figure 2. Characterization of Fe/N/C model catalyst in monolayer


graphene (FeN-MLG). (A) AFM image of folded FeN-MLG. (B)
Bright spot morphology schematic of folded FeN-MLG. (C) Height
prole at the folded FeN-MLG marked by the solid lines 1 and 2,
respectively, showing that the thickness of FeN-MLG is about 0.5 nm.
(D) Bright-eld TEM image of the graphene. Scale bar: 200 nm. (E)
Inset SAED patterns. (F) Prole plots of the diraction peak
intensities along the direction marked by white arrows in (E).

zone axis of [0001], selected at the holes. The inner hexagonal


diraction pattern was assigned to {1100}, and the outer
pattern was assigned to {1120}. Figure 2F presents the prole
plots of the diraction peak intensities along the white arrows
in Figure 2E. Higher intensities of {1100} in comparison to that
of {1120} reveal that FeN-MLG has a single-atomic-layer
structure.27 Note that no nanoparticle species (such as Fe,
Fe3C, FeNx) with dark contrast can be observed in the TEM
image of FeN-MLG (Figure 2D). Supplying the Fe source by
sublimation of the FeCl3 salt during heat treatment may benet
the dispersion of Fe. As a result, the sample prepared by this
approach has an absence of nanoparticles. In contrast, when the
Fe source was provided by spin-coating diluted FeCl3 solution
onto MLG, many particle-like species could be observed after
heat treatment under an ammonia atmosphere. In addition,
poor ORR activity was observed on such nanoparticlecontaining samples.
Raman spectroscopy is a powerful tool to characterize
MLG.28 In Figure 3A, ve samples are presented: MLG, DMLG, N-MLG, FeN-MLG_5 min, and FeN-MLG_10 min.
The I2D to IG ratio of MLG is around 2, showing the good
quality of pristine MLG. The D band appears in defective
MLG, while the 2D band is almost absent. After nitrogen or/
and iron doping, Raman peaks blue-shifted due to the disorder
and stress coming from heteroatoms.29 Notice that defects are
restored through heat treatment, as evidenced by the restored
2D band and lower D band.
Unlike nanoparticle catalysts, the graphene-based model
catalyst has just a single atomic layer, resulting in the advantage
that the structure information provided by analytical
technologies (e.g., XPS) is strictly limited to the surface.
Figure 3B shows the high-resolution XPS of Fe 2p, which

presents direct evidence of successfully implanted iron into


graphene in FeN-MLG. Higher iron content is observed in
FeN-MLG_10 min in comparison to that of FeN-MLG_5 min
(2.2 vs 1.6 atom %). Almost the same C 1s spectra of the ve
samples can be found in Figure S1 in the Supporting
Information. The N 1s spectra are presented in Figure 3C.
Nitrogen implanted in N-MLG and FeN-MLG is observed.
Here, the N 1s spectrum was deconvoluted into pyridinic-type
nitrogen (BE = 398.8 eV), metalnitrogen (399.9 eV), and
graphitic-type nitrogen (401.2 eV).30,31 With an increase in iron
content, namely from N-MLG to FeN-MLG (either 5 or 10
min), the content of pyridinic-N decreases, while that of metalnitrogen increases.
ORR Activity of FeN-MLG Model Catalysts. Figure 4A
shows a dramatic increase in ORR activity from the pristine
graphene to heteroatom-doped graphene. Fe/N-codoped
catalysts possess the highest ORR activity. The polarization
curves were obtained by normalization of the catalyst surface
area and subtraction of background currents, as shown in
Figure S2 in the Supporting Information. The Fe/N/C
nanoparticle catalyst (FeN-KJ600) was synthesized in the
same way as a comparison. The polarization curve of FeNKJ600 was measured at a rotating disk electrode (RDE) and
then normalized by BrunauerEmmettTeller (BET) surface
area as well as corrected diusion eecst by the Koutecky
Levich equation as presented in Figure S3 in the Supporting
Information. The kinetic currents of FeN-MLG_10 min and
FeN-KJ600 at 0.7 V are of the same order of magnitude:
0.133 and 0.389 A cm2, respectively. Currents at 0.7 V
were selected for comparison because it was in the kinetic
region for FeN-KJ600 (see Figure S3). Note that the catalytic
141

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis

Figure 5. (A) Raman spectra of pristine monolayer, three- to ve-layer


and six- to eight-layer graphene, normalized by the G band. (B)
Polarization curves of Fe/N/C model catalysts derived from
monolayer, three- to ve-layer, and six- to eight-layer graphene.
Figure 4. ORR behavior of FeN-MLG. (A) ORR polarization curves
of FeN-MLG, N-MLG, D-MLG, MLG, and FeN-KJ600 obtained in
O2-saturated 0.1 M H2SO4 solution. The polarization curve of FeNKJ600 was measured at a rotating disk electrode at 900 rpm. (B)
SCN ion suppressed ORR activity of FeN-MLG, which is similar to
that of FeN-KJ600 (C).

This result indicates that the same active moiety species are
formed in monolayer and multilayer graphene.
By further consideration of XPS and ORR results, the
conclusion could be reached that the higher activity of FeNMLG does not stem from the minor content of pyridinic-N.
For example, the pyridinic-N content decreases from 2.4 atom
% in N-MLG to 0.77 atom % in FeN-MLG_10 min, while the
activity is increased. To screen out a direct correlation between
the special moiety and ORR activity, catalysts with a series of
nitrogen contents were designed by controlled defect densities
on MLG. MLGs were exposed to Ar+ irradiation from 1 to 25 s
and then Fe/N codoping was processed in a batch in 10 min.
Figure 6A shows a series of Raman spectra of FeN-MLGs
with various defect densities. Raman spectra were normalized

activity of FeN-MLG is much lower than that of commercial


Pt/C (Figure S3C).
Similar ORR behavior was observed from SCN ion
suppression eects for FeN-MLG (Figure 4B) and for FeNKJ600 (Figure 4C). SCN ion is believed to form a strong
ligand with metal-center moieties, which thus suppresses the
ORR.10 Loss of 45% and 80% ORR activity in FeN-MLG and
FeN-KJ600, respectively, was demonstrated when 5 mM SCN
was added. After SCN suppression, the ORR current of FeNMLG was even lower than that of N-MLG, as evidenced by
their polarization curves in Figure S4 in the Supporting
Information. The reason for the higher current loss of FeNKJ600 may be due to the fact that the carbon nanoparticle has
higher ratio of edge carbon, in which active sites are preferably
speculated to form.32 The higher density of iron-center active
sites in FeN-KJ600 in comparison to FeN-MLG can also be
used to explain the higher ORR activity of FeN-KJ600. In
addition to ORR activity, the H2O2 yield or electron number is
also an important parameter to evaluate ORR catalysts.
Unfortunately, it is hard to transfer the FeN-MLG supported
on Si substrate to a rotating ring-disk electrode (RRDE) for the
measurement of electron numbers. Future work will be needed
to measure this parameter, such as by design of a ow cell.
Practical Fe/N/C catalysts are composed of carbon nanoparticles with multilayer graphene. Some active sites, such as
encapsulated metal nanoparticles/clusters on carbon, are
proposed in a structure with multilayer graphene. Is there an
activity dierence between monolayer and multilayer graphenebased catalysts? To answer this question, we then extended our
method to prepare model catalysts in multilayer graphene. The
preparation is similar to that mentioned above, using just fewlayer graphene instead of monolayer graphene. Figure 5A
shows the Raman spectra of monolayer and multilayer pristine
graphene. Broader peak width, lower intensity, and a blue shift
of the 2D band reveal the multilayer structure.33 Figure 5B
shows the ORR polarization curves of model catalysts derived
from graphene with dierent layer numbers. The kinetic
currents of model catalysts in monolayer (FeN-MLG), three- to
ve-layer (FeN-35LG), and six to eight-layer graphene (FeN68LG) at 0.6 V are 4.67, 6.26, and 7.53 A cm2, respectively.
Although the ORR activity is enhanced slightly with an increase
in the layer number of graphene, the dierence is not signicant
(only increasing 60% from a monolayer to six to eight layers).

Figure 6. (A) Raman spectra of FeN-MLGs with a series of defect


densities, normalized by the G band. (B) Correlation between the
ORR current at 0.6 V (j) and the mean distance between defects (LD).
(C) Correlation between the ORR current at 0.6 V and the percentage
of Nx-Fe moiety species.

by the G band. With an increase in defects, the D band


increases rst and then decreases, while the D band increases
and then merges with the G band, and the 2D band gradually
disappears. A quantitative formula has been proposed to
correlate the mean distance between defects in graphene (LD,
nm) with the intensity ratio ID/IG as shown in eq 1:34,35
2
2
2
(r 2 rS2) rS2 / LD2
ID
[e
= CA 2A
e(rA rS )/ LD]
2
IG
(rA 2rS )

(1)

where rS (1 nm) and rA (3.1 nm) are the radii of the


structurally disordered area and the activated area around
the ion-induced defects, respectively. The factor CA is dened
142

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis
by the electronphonon matrix elements. We used CA = 4.3 at
the red line excitation (see Figure S5 in the Supporting
Information).
Figure 6B presents the dependence of ORR activity on the
mean distance of defects. A larger LD value denotes a lower
density of defects. The currents of FeN-MLGs at 0.6 V increase
with decreasing LD: namely, with an increase in the defect
density. The original polarization curves can be found in Figure
S6 in the Supporting Information. Notice that the shortest LD
we made is 1.65 nm; it is dicult to create more defects due to
the easily disconnected carbon matrix or irradiation out of all of
the carbon atoms. The activity would decrease because of the
increased resistance with high heteroatom doping. Since Fe/N
heteroatoms are supposed to occupy defective positions, a
linear relationship is observed between the nitrogen content
and mean defect distance (Figure S7 in the Supporting
Information). High-resolution N 1s spectra of XPS (Figure S8
in the Supporting Information) show similar peak patterns in
the series of FeN-MLGs with dierent defect densities. They
were deconvoluted into pyridinic-type nitrogen (BE = 398.8
eV), metalnitrogen (399.9 eV), and graphitic-type nitrogen.
In all spectra, metal-nitrogen, which is Nx-Fe in this case, is the
main content (>50%). The ORR activity correlated on Nx-Fe
and graphitic-N are plotted as shown in Figure 6C and Figure
S9 in the Supporting Information, respectively. A linear
relationship between ORR activity and Nx-Fe is observed,
which reveals that the Nx-Fe moiety species contributes to the
ORR activity. However, in the case of graphitic-N, none of the
dependence relationship with activity can be found (Figure S9).

graphene-coated Cu foil, followed by Cu etching and PMMA


removal.37
X-ray Photoelectron Spectroscopy. X-ray photoelectron
spectroscopy (XPS) measurements were carried out on an
Omicron Sphera II hemispherical electron energy analyzer
(monochromatic Al K with 1486.6 eV operating at 15 kV and
300 W). The base pressure of the systems was 5.0 109 mbar.
XPS spectra were referenced to the carbon C 1s peak at 284.6
eV. The N 1s spectra were deconvoluted into pyridinic-type N
(398.8 eV), metal-N (399.9 eV), and graphitic-type N (401.2
eV).30,31 The tted curves were represented by 80% Gaussian/
20% Lorentzian line shape after Shirley-type background
subtraction. The full width at half-maximum (fwhm) for each
type of nitrogen component used for tting was limited in the
range of 1.11.50 eV, which was suggested by the literature.30
The nal fwhm of 1.5 eV was selected because the best tting
results were obtained with this parameter. N 1s spectra of a
series of FeN-MLG_10 min samples with various defect
densities were tted with the same parameter, including peak
position, fwhm, Gaussian/Lorentzian line shape, and asymmetry.
ORR Measurement. The graphene-based model catalysts
were cut into a rectangular shape. A copper wire was stuck to
the graphene by conductive silver paint to contact with an
external circuit. The electrical contacts were then covered with
epoxy. The exposed area of the model catalyst was measured
with a ruler. All electrochemical measurements were performed
in traditional three-electrode cells. Model catalysts were used as
the working electrodes. A homemade reversible hydrogen
electrode (RHE) was used as the reference electrode.
The electrolyte was 0.1 M H2SO4 and was bubbled with
high-puritye oxygen (99.999%) at 40 sccm. The working
electrode was subjected to potential cycling between 1.0 and
0.4 V (RHE) at a scan rate of 10 mV s1. Unlike normal Fe/N/
C catalysts, graphene samples had a low surface area, and the
absolute current was less than 10 A in this study. That is, the
ORR process is fully controlled by instinct kinetics, not by O2
mass transfer in a wide range potential window. Therefore, a
rotating disk electrode (RDE) was not needed during the ORR
test. The geometric area was used to calculate the current
density due to the low roughness of graphene. Solution ohmic
drop (i.e., iR drop) was compensated. The background
capacitive current was recorded in the same potential range
and scan rate, but in Ar-saturated electrolyte. The current
recorded in O2-saturated solution was corrected for the
background current to yield net ORR current (Figure S2 in
the Supporting Information). After the Ar background
(capacitive current) was subtracted, the forward and backward
ORR polarization curves were averaged. Thus, polarization
curves were obtained.
For FeN-KJ600, an electrochemical test was carried out on a
glassy-carbon RDE. FeN-KJ600 ink was prepared in a solution
mixture (0.50 mL of water, 0.50 mL of ethanol, and 0.05 mL of
5% Naon) in a concentration of 6 mg mL1. A 25 L portion
of the ink was pipetted onto the RDE, resulting in a catalyst
loading of 0.6 mg cm2. A glassy-carbon plate was used as the
counter electrode.
Raman, AFM, and TEM. Raman spectra were measured on
an XploRA Raman instrument (Jobin Yvon-Horiba, France),
which was equipped with 532 and 638 nm lasers. A 100
objective and 1200 lines mm1 grating were used. About 10
spectra were measured for each sample, and a typical spectrum
was selected. Raman spectra were normalized with the G band.

CONCLUSIONS
To sum up, single-atomic-layer Fe/N/C species were
successfully prepared as model catalysts. The good linear
dependence of ORR activity on defect density and the amount
of the Nx-Fe moiety strongly indicates that the active sites of
Fe/N/C are the Nx-Fe formed at graphene defects. The layer
number of graphene has little eect on the ORR activity. This
study paves the way to explore Fe/N/C in model catalysts
designed from monolayer graphene.

EXPERIMENTAL SECTION

Transferring Graphene to SiO2/Si Substrate. Here,


monolayer and multilayer graphene (three- to ve-layer
graphene), purchased from Xiamen G-CVD Material Technology Co., Ltd., was grown on a 25 m thick copper foil (Alfa
Aesar, item No. 13382) by chemical vapor deposition (CVD)
in a ow-type low-pressure reactor. For transfer, a poly(methyl
methacrylate) (PMMA) solution was spin-coated onto the top
side of the sample at 5000 rpm. Then the PMMA coat was kept
at room temperature for 2 h to dry. Typically, graphene will
grow on both sides of Cu foil and the back side graphene will
hinder the Cu etching process. The back side graphene was
removed by oxygen plasma etching for 3 min at a power of 100
W with a pressure of 315 mTorr. The Cu was etched in a
potassium persulfate solution (15 mg/500 mL) for 24 h and
then further etched in a H2O2/HCl solution (H2O:H2O2:HCl
ratio is 20:1:1) for another 24 h. After it was cleaned six times
in deionized water, graphene with PMMA lm was transferred
to SiO2 (300 nm)/Si substrate and PMMA was removed by
keeping it in chloroform overnight.36 A six- to eight-layer
graphene lm was prepared by the transfer of a PMMA-coated
three- to ve-layer graphene lm onto a three- to ve-layer
143

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis

(9) Wang, Y. C.; Lai, Y. J.; Song, L.; Zhou, Z. Y.; Liu, J. G.; Wang, Q.;
Yang, X. D.; Chen, C.; Shi, W.; Zheng, Y. P.; Rauf, M.; Sun, S. G.
Angew. Chem., Int. Ed. 2015, 54, 99079910.
(10) Wang, Q.; Zhou, Z. Y.; Lai, Y. J.; You, Y.; Liu, J. G.; Wu, X. L.;
Terefe, E.; Chen, C.; Song, L.; Rauf, M.; Tian, N.; Sun, S. G. J. Am.
Chem. Soc. 2014, 136, 1088210885.
(11) Li, W.; Wu, J.; Higgins, D. C.; Choi, J. Y.; Chen, Z. ACS Catal.
2012, 2, 27612768.
(12) Jiang, W. J.; Gu, L.; Li, L.; Zhang, Y.; Zhang, X.; Zhang, L. J.;
Wang, J. Q.; Hu, J. S.; Wei, Z.; Wan, L. J. J. Am. Chem. Soc. 2016, 138,
35703578.
(13) Zitolo, A.; Goellner, V.; Armel, V.; Sougrati, M. T.; Mineva, T.;
Stievano, L.; Fonda, E.; Jaouen, F. Nat. Mater. 2015, 14, 937942.
(14) Kramm, U. I.; Herrmann-Geppert, I.; Behrends, J.; Lips, K.;
Fiechter, S.; Bogdanoff, P. J. Am. Chem. Soc. 2016, 138, 635640.
(15) Lefevre, M.; Dodelet, J.; Bertrand, P. J. Phys. Chem. B 2002, 106,
87058713.
(16) Jaouen, F.; Herranz, J.; Lefevre, M.; Dodelet, J. P.; Kramm, U. I.;
Herrmann, I.; Bogdanoff, P.; Maruyama, J.; Nagaoka, T.; Garsuch, A.;
Dahn, J. R.; Olson, T.; Pylypenko, S.; Atanassov, P.; Ustinov, E. A.
ACS Appl. Mater. Interfaces 2009, 1, 16231639.
(17) Artyushkova, K.; Serov, A.; Rojas-Carbonell, S.; Atanassov, P. J.
Phys. Chem. C 2015, 119, 2591725928.
(18) Sahraie, N. R.; Kramm, U. I.; Steinberg, J.; Zhang, Y.; Thomas,
A.; Reier, T.; Paraknowitsch, J. P.; Strasser, P. Nat. Commun. 2015, 6,
8618.
(19) Mayorov, A. S.; Gorbachev, R. V.; Morozov, S. V.; Britnell, L.;
Jalil, R.; Ponomarenko, L. A.; Blake, P.; Novoselov, K. S.; Watanabe,
K.; Taniguchi, T.; Geim, A. K. Nano Lett. 2011, 11, 23962399.
(20) Balandin, A. A. Nat. Mater. 2011, 10, 569581.
(21) Lee, C.; Wei, X.; Kysar, J. W.; Hone, J. Science 2008, 321, 385
388.
(22) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang,
Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Science 2004, 306,
666669.
(23) Velicky, M.; Bradley, D. F.; Cooper, A. J.; Hill, E. W.; Kinloch, I.
A.; Mishchenko, A.; Novoselov, K. S.; Patten, H. V.; Toth, P. S.;
Valota, A. T.; Worrall, S. D.; Dryfe, R. A. W. ACS Nano 2014, 8,
1008910100.
(24) Valota, A. T.; Kinloch, I. A.; Novoselov, K. S.; Casiraghi, C.;
Eckmann, A.; Hill, E. W.; Dryfe, R. A. W. ACS Nano 2011, 5, 8809
8815.
(25) Zhong, J. H.; Zhang, J.; Jin, X.; Liu, J. Y.; Li, Q.; Li, M. H.; Cai,
W.; Wu, D. Y.; Zhan, D.; Ren, B. J. Am. Chem. Soc. 2014, 136, 16609
16617.
(26) Guo, D.; Shibuya, R.; Akiba, C.; Saji, S.; Kondo, T.; Nakamura,
J. Science 2016, 351, 361365.
(27) Zhou, H.; Yu, W. J.; Liu, L.; Cheng, R.; Chen, Y.; Huang, X.;
Liu, Y.; Wang, Y.; Huang, Y.; Duan, X. Nat. Commun. 2013, 4, 2096.
(28) Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri,
M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S.; Geim,
A. K. Phys. Rev. Lett. 2006, 97, 187401.
(29) Ferrari, A. C. Solid State Commun. 2007, 143, 4757.
(30) Artyushkova, K.; Kiefer, B.; Halevi, B.; Knop-Gericke, A.;
Schlogl, R.; Atanassov, P. Chem. Commun. 2013, 49, 25392541.
(31) Gottfried, J. M.; Flechtner, K.; Kretschmann, A.; Lukasczyk, T.;
Steinruck, H.-P. J. Am. Chem. Soc. 2006, 128, 56445645.
(32) Jia, Q.; Ramaswamy, N.; Hafiz, H.; Tylus, U.; Strickland, K.; Wu,
G.; Barbiellini, B.; Bansil, A.; Holby, E. F.; Zelenay, P. ACS Nano 2015,
9, 1249612505.
(33) Malard, L.; Pimenta, M.; Dresselhaus, G.; Dresselhaus, M. Phys.
Rep. 2009, 473, 5187.
(34) Lucchese, M. M.; Stavale, F.; Ferreira, E. H. M.; Vilani, C.;
Moutinho, M. V. O.; Capaz, R. B.; Achete, C. A.; Jorio, A. Carbon
2010, 48, 15921597.
(35) Cancado, L. G.; Jorio, A.; Ferreira, E. H.; Stavale, F.; Achete, C.
A.; Capaz, R. B.; Moutinho, M. V.; Lombardo, A.; Kulmala, T. S.;
Ferrari, A. C. Nano Lett. 2011, 11, 31903196.

Atomic force microscopy (AFM) was performed on a NTMDT system (NTEGRA Spectra) in tapping mode. Images
with 10 10 m2 area were obtained at a scan rate of 0.5 Hz
using VIT_P tip.
TEM images were obtained with a TECNAI F-20 transmission electron microscope operating at 200 kV. The
accelerating voltage used here (200 kV) is higher than the
critical energy predicted for severe knock-on damage of
graphene.38 Therefore, TEM images were quickly taken to
avoid highly destroying the graphene with the electron beam.
Prior to the focusing process, several selective areas were
labeled. The position was switched to the labeled area after
focusing was done, and then a diraction pattern image was
also quickly taken.

ASSOCIATED CONTENT

S Supporting Information
*

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acscatal.6b02702.
Additional XPS, electrochemical characterization results,
Ar adsorption/desorption isotherm of FeN-KJ600, and
further discussion (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail for Z.-Y.Z.: zhouzy@xmu.edu.cn.


*E-mail for S.-G.S.: sgsun@xmu.edu.cn.
ORCID

Zhi-You Zhou: 0000-0001-5181-0642


Author Contributions

These authors contributed equally.

Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
This work was supported by the National Basic Research
Program of China (2015CB932303), the Natural Science
Foundation of China (21373175 and 21361140374), and the
Fundamental Research Funds for the Central Universities
(20720150109). We thank Teng-Xiang Huang for help with the
AFM measurements.

REFERENCES

(1) Lefevre, M.; Proietti, E.; Jaouen, F.; Dodelet, J. P. Science 2009,
324, 7174.
(2) Wu, G.; More, K. L.; Johnston, C. M.; Zelenay, P. Science 2011,
332, 443447.
(3) Gong, K.; Du, F.; Xia, Z.; Durstock, M.; Dai, L. Science 2009, 323,
760764.
(4) Chung, H. T.; Won, J. H.; Zelenay, P. Nat. Commun. 2013, 4,
1922.
(5) Deng, D.; Yu, L.; Chen, X.; Wang, G.; Jin, L.; Pan, X.; Deng, J.;
Sun, G.; Bao, X. Angew. Chem., Int. Ed. 2013, 52, 371375.
(6) Kramm, U. I.; Herranz, J.; Larouche, N.; Arruda, T. M.; Lefevre,
M.; Jaouen, F.; Bogdanoff, P.; Fiechter, S.; Abs-Wurmbach, I.;
Mukerjee, S.; Dodelet, J. P. Phys. Chem. Chem. Phys. 2012, 14,
1167311688.
(7) Kramm, U. I.; Lefevre, M.; Larouche, N.; Schmeisser, D.;
Dodelet, J. P. J. Am. Chem. Soc. 2014, 136, 978985.
(8) Wei, P. J.; Yu, G. Q.; Naruta, Y.; Liu, J. G. Angew. Chem., Int. Ed.
2014, 53, 66596663.
144

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Research Article

ACS Catalysis
(36) Liang, X.; Sperling, B. A.; Calizo, I.; Cheng, G.; Hacker, C. A.;
Zhang, Q.; Obeng, Y.; Yan, K.; Peng, H.; Li, Q. ACS Nano 2011, 5,
91449153.
(37) Ji, H.; Zhao, X.; Qiao, Z.; Jung, J.; Zhu, Y.; Lu, Y.; Zhang, L. L.;
MacDonald, A. H.; Ruoff, R. S. Nat. Commun. 2014, 5, 3317.
(38) Zobelli, A.; Gloter, A.; Ewels, C.; Seifert, G.; Colliex, C. Phys.
Rev. B: Condens. Matter Mater. Phys. 2007, 75, 245402.

145

DOI: 10.1021/acscatal.6b02702
ACS Catal. 2017, 7, 139145

Potrebbero piacerti anche