Construction of porous N-doped graphene layer for ...

46
1 Construction of porous N-doped graphene layer for efficient oxygen reduction reaction Xiaofang Chen 1, 2 , Yan Liang 1 , Li Wan 1 , Zongli Xie 2 , Christopher D. Easton 2 , Laure Bourgeois 3 , Ziyu Wang 4 , Qiaoliang Bao 4 , Yonggang Zhu 5 , Shanwen Tao 1 , Huanting Wang 1 * 1 Department of Chemical Engineering, Monash University, Victoria 3800, Australia 2 CSIRO Manufacturing, Private Bag 10, Clayton South, Victoria 3169, Australia 3 Monash Centre for Electron Microscopy – Department of Materials Science and Engineering, Monash University, Victoria 3800, Australia 4 Department of Materials Science and Engineering, Monash University, Victoria 3800, Australia 5 School of Science, RMIT University Melbourne, VIC 3001, Australia *Corresponding author. E-mail: [email protected]

Transcript of Construction of porous N-doped graphene layer for ...

Page 1: Construction of porous N-doped graphene layer for ...

1

Construction of porous N-doped graphene layer for efficient oxygen

reduction reaction

Xiaofang Chen1, 2, Yan Liang1, Li Wan1, Zongli Xie2, Christopher D. Easton2, Laure

Bourgeois3, Ziyu Wang4, Qiaoliang Bao4, Yonggang Zhu5, Shanwen Tao1, Huanting Wang1*

1Department of Chemical Engineering, Monash University, Victoria 3800, Australia

2CSIRO Manufacturing, Private Bag 10, Clayton South, Victoria 3169, Australia

3Monash Centre for Electron Microscopy – Department of Materials Science and Engineering,

Monash University, Victoria 3800, Australia

4Department of Materials Science and Engineering, Monash University, Victoria 3800,

Australia

5School of Science, RMIT University Melbourne, VIC 3001, Australia

*Corresponding author. E-mail: [email protected]

Page 2: Construction of porous N-doped graphene layer for ...

2

Abstract

Graphitic carbon materials have shown great potential for use as high-performance catalysts

for electrochemical reactions and devices. In this work, we developed a simple and versatile

method for synthesis of porous N-doped graphene layers (NGS) by high-temperature treatment

of chitosan film deposited on the graphitic carbon nitride (g-C3N4) nanosheets. In the

sandwiched chitosan/g-C3N4/chitosan structure, the g-C3N4 nanosheet served as a substrate for

chitosan film. The pyrolysis of this substrate, g-C3N4 nanosheet, prevented the severe

agglomeration of as-carbonized chitosan sheets and resulted the porous structure. The BET

surface area, micropore volume, nitrogen content and graphitic level of result sample highly

depended on the heat-treatment temperature. The NGS synthesized at 1000 °C (NGS-1000)

exhibited an ultrahigh specific surface area (1183 m2 g-1) and high nitrogen content (4.12%).

Importantly, NGS-1000 exhibited a higher limiting current density (5.8 mA cm-2) and a greater

stability than the commercial Pt/C electrocatalyst in alkaline media for oxygen reduction

reaction (ORR). Such excellent electrocatalytic performance can be explained by a balanced

combination of appropriate nitrogen doping level, the degree of graphitization, porous

structure, and high specific surface area.

Keywords: Microporous N-doped graphene, oxygen reduction reaction, nitrogen doping,

electrocatalytic activity, g-C3N4

Page 3: Construction of porous N-doped graphene layer for ...

3

1. Introduction

Nanocarbon materials such as carbon nanotubes/fibers, graphene, and nanoporous carbons

have been extensively studied as electrode materials for fuel cells and other electrochemical

devices due to their low cost, abundant structural variety, chemical and thermal stability, and

conductivity (Matter et al., 2006; Zhang et al., 2011). Carbon-based metal-free electrocatalysts

show high electronic conductivity, strong tolerance to acidic/alkaline environments, and

tuneable molecular structures (Choi et al., 2017; Klingele et al., 2016). Among nanocarbon

materials, graphene is a 2D nanosheet of sp2-hybridized carbon, widely used as the basic

building block of 1D and 3D graphitic materials (He et al., 2014; Zhang et al., 2012; Peng et

al., 2014); long-range -conjugation in graphene endows it with extraordinary properties such

as high surface area (2630 m2 g-1), very fast charged carrier mobility (~ 200 000 cm2 V-1 s-1),

and high thermal conductivity (Avouris et al., 2007; Balandin et al., 2008; Zhu Y., 2011;

Terrones M., 2011; Wang et al., 2012). It was reported that the introduction of graphene highly

increased electrochemically active surface area, exposed more active sites, and enhanced the

charge transport (Zhuang et al., 2017). The However, the graphene sheets readily restack

because of either capillary or van der Waals forces, so that the practical surface area of

graphene-based 2D materials is far below the theoretical value, leading to limited external

accessible surface and the active sites for electrochemical reactions such as oxygen reduction

reaction (ORR) (Ferrero et al., 2016; Presser et al., 2011, Wang et al., 2017).

To improve the electrochemical performance, introducing pores with divers sizes and

shapes into carbon materials has been widely investigated in various electrochemical energy

storage and conversion systems (Peng et al., 2014; Qie et al., 2013; Wei et al., 2015). The

attractive feature of nanoporous carbon is that they can readily interact with various species

not only at the surfaces but also through their porous frameworks, which is significantly

important for electrochemical reactions (Cooper A. I., 2003; Davis M. E., 2002; Xiao et al.,

Page 4: Construction of porous N-doped graphene layer for ...

4

2011). Therefore, one effective strategy is to form a porous structure in graphene to reduce its

stacking and increase its surface area because well-organized pores can effectively facilitate

the diffusion of electrolyte and transport of various ions while providing a continuous electron

pathway to ensure good electrical contact (Zhang L. P. and Z. H. Xia,2011; Jin et al., 2014).

Among several methods reported so far, the template approach is most effective for

synthesizing porous graphene with controllable size and morphology of pores by using either

a hard or soft template. Typically, the in-situ growth of porous metal or metal oxide in presence

of graphene or graphene oxide could produce porous graphene via later removal of the hard

template (Han et al., 2014; Zheng et al., 2005, Niu et al.,). On the other side, small molecules

and block copolymers are used as soft templates to produce porous graphene materials via a

bottom-up process (Peng et al., 2014; Hao et al., 2013). However, in these approaches,

graphene oxide is still required, which is not a cost-effective and versatile strategy for the

fabrication of porous graphene (Wei et al., 2015; Han et al., 2014). In the absence of graphene

oxide, some surfactants and polymers can be used as templates, and they are usually converted

porous carbon hybridized with graphene sheets during the pyrolysis process (Wei et al., 2015;

Xiao et al., 2011) Moreover, the resulting carbon is largely amorphous and has poor electronic

conductivity, which is not suitable for use as electrocatalysts (Wei et al., 2015; Xu et al., 2016).

Another practical strategy is to increase the accessibility of pores and active sites by

improving the wettability of the electrode surface by nitrogen doping (Xu et al., 2016; Wen et

al., 2012, Meng et al., 2017). It has been revealed that the electrocatalytic ORR activity

originates from heteroatoms present in the graphitic framework, making the catalysts non-

electron-neutral and thus improving the oxygen molecular adsorption and its reduction (Wei et

al., 2015; Tang et al., 2009; Wu et al., 2015). For example, N-doping essentially introduces

nitrogen into the carbon structure, either at the edges or replacing one of the sp2-hybridized

carbon atoms in the graphitic structure. Such a doping process induces a high concentration of

Page 5: Construction of porous N-doped graphene layer for ...

5

active sites. Recently, Geng et al. (Geng et al., 2011) have reported a nitrogen-doped carbon

material that exhibited ORR performance superior to commercial Pt/C catalysts, representing

an exciting development for the N-doped metal-free catalyst. The high nitrogen doping level

by incorporating of g-C3N4 into a mesoporous carbon highly enhanced the electron transfer

efficiency of g-C3N4, resulting in competitive catalytic activity and superior methanol

tolerance compared to a commercial Pt/C catalyst (Zheng et al., 2011). Moreover, it has a high

durability, even under acidic conditions (He et al. 2014; Kurak K. A. and A. B. Anderson, 2009;

Guo et al., 2016). Therefore, the introduction of nitrogen atom into graphene offers an effective

strategy to improve the catalytic activity of ORR.

For the development of high-performance carbon catalysts for ORR, the challenge is how

to achieve the high specific surface area, desirable porosity and nitrogen-doping level, and high

electronic conductivity to increase the accessibility and active sites of the electrocatalysts.

Herein we report a versatile method to realize high nitrogen-doping level by converting the

chitosan into porous graphene layer with the ultrahigh surface area. Specifically, the synthesis

of porous NGS is achieved by replicating the 2D structure of g-C3N4 nanosheets with chitosan

as additional carbon and nitrogen sources. Chitosan, deriving from the chitin in insect and

crustacean skins, is an abundant and nitrogen-containing renewable biopolymer (Hao et al.,

2015; Primo et al., 2012). Because chitosan molecules tend to bond together to form a large-

area polymer film, it can form graphene materials during the pyrolysis process (Hao et al., 2015;

Primo et al., 2012). In this study, we used g-C3N4 nanosheet initially as a thermally removable

substrate, on whose surface chitosan would be bonded by electrostatic interactions, forming

chitosan film outside of each g-C3N4 nanosheet and thus resulting a sandwich-like structure of

chitosan/g-C3N4/chitosan (GCN-CS). The subsequent high-temperature treatment allowed the

chitosan film to maintain the 2D structure of the g-C3N4 nanosheet and convert into the porous

nitrogen-doped graphene layers (NGS), accompanied by the decomposition of the g-C3N4

Page 6: Construction of porous N-doped graphene layer for ...

6

nanosheets. The resulting porous NGS presented excellent ORR performance, comparable to

Pt/C.

2. Materials and Methods

2.1 Materials

All chemicals were purchased from Sigma-Aldrich (Australia) and used without further

purification. Dicyandiamide, chitosan (low molecular weight, 50,000-190,000), acetic acid

(99.7%), ethanol, potassium hydroxide, sulfuric acid, 2-methylimidazole.

2.2 Synthesis of heating-exfoliated g-C3N4

10 g of dicyandiamide was heated at 550 °C for 4h in a muffle furnace to obtain bulk g-

C3N4. 1 g of as-produced bulk g-C3N4 was then mechanically ground with 0.3 g of 2-

methylimidazole, followed by further annealing at 500 °C for 3 h to fabricate g-C3N4

nanosheets by the heat-etching process.

2.3 Synthesis of porous NGS

The exfoliated g-C3N4 powder (0.2 g) was firstly dispersed in 40 ml ethanol by

ultrasonication for 1 h. 0.2 g of chitosan was completely dissolved in acetic acid solution. The

solution dispersed with g-C3N4 was then added into chitosan solution. To reach an adsorption

equilibrium of chitosan on g-C3N4 nanosheets, the above mixture was stirred for 14 h. The g-

C3N4 adsorbed with chitosan was collected by centrifugation (8000 rpm for 13 min) and dried

at 60 °C for 5 h. NGS was finally synthesized by drying at 60 °C for 2h, and pyrolysis in argon

at different temperatures (580, 700, 800, 900, 1000, and 1100 °C) for 3 h.

2.4 Characterization

Morphology of all samples was recorded on a Nova NanoSEM 450 (FEI, USA) operated

at a voltage of 3 kV. Transmission electron microscopy (TEM) images were collected on an

FEI Tecnai G2 T20 and a JEOL JEM-2100F FEGTEM both operated at an accelerating voltage

of 200 kV. Electron energy loss spectroscopy (EELS) was performed on a Gatan Enfina

Page 7: Construction of porous N-doped graphene layer for ...

7

spectrometer installed on the JEOL 2100F. Fourier transform infrared spectra of the samples

were measured with an FT-IR spectrophotometer (PerkinElmer Spectrum 100, USA)

technique. Raman spectra were collected and processed by Topspin 2.1. Raman samples were

analyzed in a confocal microscope system (WITec alpha 300R) with an excitation by a 532nm

laser. X-ray photoelectron spectroscopy (XPS) analysis was performed using an AXIS Nova

spectrometer (Kratos Analytical Inc., Manchester, UK) with a monochromated Al Kα source

at a power of 180 W (15 kV,12 mA), a hemispherical analyser operating in the fixed analyser

transmission mode and the standard aperture (analysis area: 0.3 mm × 0.7 mm). The atomic

concentrations of the detected elements were calculated using integral peak intensities and the

sensitivity factors supplied by the manufacturer. Binding energies were referenced to the C 1s

peak at 284.4 eV (graphitic carbon). The measurements of the nitrogen adsorption-desorption

isotherms of all samples were conducted on a physisorption analyzer (Micromeritics ASAP

2020, USA) at liquid nitrogen temperature. The samples were degassed at 200 °C for 8 h, prior

to the measurement.

2.5 Electrochemical measurements

The electrochemical measurements were performed via a μAuto lab electrochemical

analyzer in the three-electrode electrochemical cell, using Pt wire auxiliary as an electrode and

a saturated Ag/AgCl (saturated with 3 M HCl) as a reference electrode. A rotating disk working

electrode (Pine Research Instrumentation) was used. The working electrode was prepared as

following steps: 5 mg of catalysts (or commercial Pt/C, 20%, from Sigma-Aldrich, Australia)

were dispersed in 1 mL of ethanol under ultrasonic sonication. Then 8 μL of the resulting

suspension was dropped on the surface of the pre-cleaned rotating disk electrode (5 mm

diameter, 0.196 cm2), followed by dropping the Nafion solution (ethanol solution containing

0.03 ml of 0.5% wt% Nafion), and dried at room temperature. A 0.1 M KOH (or 0.5 M H2SO4)

solution was used as the electrolyte for the CV and RDE studies. O2 and argon were saturated

Page 8: Construction of porous N-doped graphene layer for ...

8

to the solution to reach oxygen-rich and oxygen-free environments, respectively. The

electrolyte was saturated with argon (or O2) prior running the test. CV curves were recorded at

a scan rate of 50 mV s-1. The LSV curves were recorded at a scan rate of 10 mV s-1. To make

sure O2 saturation, a flow of O2 was maintained over the electrolyte during the CV and LSV

test. The electron transfer numbers in ORR were calculated by Koutecky-Levich equation, at

various electrode potentials:

1

J=1

JL+1

JK=

1

Bω0.5+1

JK

B = 0.62nFC0D02/3

ϑ−1/6

where J is the measured current density, JK is the kinetic-limiting current density, JL is the

diffusion-limiting current density, ωis the angular velocity, n is transferred electron number,

F is the Faraday constant (96485 C mol-1), C0 is the bulk concentration of O2 (1.2×10-6 mol cm-

3 for 0.1 M KOH, 1.26×10-6 mol cm-3 for 0.5 M H2SO4), D0 is the diffusion coefficient of O2

(1.9×10-5 cm2 s-1 in 0.1 M KOH, 2.1×10-5 cm2 s-1 in 0.5 M H2SO4), and ʋ is the kinematic

viscosity of the electrolyte (0.01 cm2 s-1 in KOH, 1.07 ×10-2 cm2 s-1 in 0.5 M H2SO4).

3. Results and Discussion

The synthesis of NGS is achieved by carbonization of chitosan film deposited on g-C3N4

nanosheets, schematically illustrated in Figure 1. Firstly, the heat-exfoliated g-C3N4

nanosheets were dispersed in ethanol aqueous and subsequently mixed with chitosan solution.

Because of the interactions between the amine groups in g-C3N4 and the groups (-OH and NH2

groups) in the chitosan molecular (Hao et al., 2015; Pandiselvi K. and S. Thambidurai, 2014;

Primo et al., 2012), chitosan could be easily attached on both sides of each g-C3N4 sheet,

forming a sandwich-like structure of GCN-CS. Due to the pyrolysis of g-C3N4 at the

temperatures over 700 °C, g-C3N4 nanosheets in the sandwich-like structure served as a spacer

to inhibit the severe agglomeration of as-carbonized chitosan sheets and possibly create or

Page 9: Construction of porous N-doped graphene layer for ...

9

enlarge the pores on as-carbonized chitosan during the high-temperature treatment.

Additionally, both g-C3N4 and chitosan served as nitrogen source for the nitrogen doping. As

a result, a porous nitrogen-doped graphene layer is obtained.

The g-C3N4 nanosheets were obtained by heat-etching the ground mixture of bulk g-C3N4

and 2-methylimidazole in the air. The resulted g-C3N4 nanosheets were characterized by the

Scanning Electron Microscopy (SEM) and Transmission Electron Microscopy (TEM) images

(Figure 2a and Figure S1), a typical nanosheet-like structure is observed. The g-C3N4

nanosheets were chosen as a soft substrate for the formation of chitosan film on its surface,

forming a sandwich-like structure of GCN-CS. Before carbonization, GCN-CS (shown in 2b)

is much thicker than the g-C3N4 nanosheets, due to full coverage of chitosan on g-C3N4

nanosheets. Fourier Transform Infrared Spectroscopy (FTIR) spectra measurement was

performed to illustrate the presence of chitosan in GCN-CS. As shown in Figure S2, the peaks

at 1082 and 1053 cm-1 assigned C-N stretching vibration in GCN-CS become stronger,

compared to the pure g-C3N4 nanosheets, which arises from the contribution of C-N bond in

chitosan molecules.

Such a soft substrate of g-C3N4 can effectively prevent the agglomeration of as-carbonized

chitosan film during high-temperature treatment, because the gas generated from the

decomposition process of the g-C3N4 nanosheets presumably drives its neighboring layers

apart, prevents the agglomeration of as-formed carbonized sheets, and even punches holes at

the defects on the carbonized sheets. Therefore, ultrathin sheets of NGS are observed in SEM

image (Figure 2c), which is consistent with that observed in the TEM image where the

nanosheets are distinctively visible and transparent (Figure 2d). In addition, the corresponding

selected area electron diffraction (SAED) pattern (Figure 3a) shows highly oriented pyrolytic

graphite, indicating that the chitosan film is successfully transformed to graphene layer. The

high-resolution TEM (HRTEM) image (Figure 3a) reveals that NGS are mainly composed of

Page 10: Construction of porous N-doped graphene layer for ...

10

single-/few-layer graphene nanosheets and graphitic-like structure with an interlayer spacing

of 0.35 nm that is almost corresponds to the (002) plane of graphite. Some micropores (~ 2.5

nm) are observed in partial nanosheets which may be created the during the processes of gas

evolution from the decomposition of g-C3N4 and carbonization of chitosan-derived carbon

(Figure 3b). However, without g-C3N4 nanosheets as the substrate, chitosan only could be

converted into bulk carbon rather than graphitic carbon sheets during the high-temperature

treatment, shown in Figure S3a. Moreover, when bulk g-C3N4 used as the substrate, some thick

plates randomly agglomerated are observed (Figure S3b), due to nonuniform deposition of

chitosan. These results illustrate that ultrathin film of chitosan is crucial for the formation of

graphene by chitosan during the high-temperature treatment as proved by Primo et al. (Primo

et al., 2012; Mateo et al., 2016).

To investigate the possible formation mechanism of porous NGS, the high-temperature

treatment process was traced. As illustrated by thermogravimetric analysis (TGA), chitosan

exhibits a major weight loss at 300-600 °C, whereas there is a huge weight loss (95 wt %)

ending at around 700 °C for GCN-CS (Figure S4). This is consistent with the previous studies

that g-C3N4 experiences sublimation at 650 °C and complete pyrolysis at around 725 °C (Li et

al., 2012; Chen et al., 2016). This result suggests that only 5 wt % remains of GCN-CS should

attribute to nitrogen doping. When carbonized at 580 °C (the resulting sample is denoted NGS-

580), chitosan film on the surface of g-C3N4 sheets is partially carbonized and g-C3N4

nanosheets start decomposing, forming very thick nanosheets with amorphous structure

(Figure S5a, S6a, and S7a), because this temperature is not high to drive the chitosan the

complete carbonization of chitosan and decomposition of g-C3N4. Increasing the temperature

to 700 °C (denoted NGS-700), the decomposition of g-C3N4 is not finished so that

comparatively thick nanosheets are observed (Figure S5b, S6b, and S7b). HRTEM images

show the amorphous carbon layer attached on g-C3N4 nanosheets in the samples produced at

Page 11: Construction of porous N-doped graphene layer for ...

11

lower temperatures such as NGS-580 and NGS-700. Therefore, the sandwiched GCN-CS is

then treated at higher temperatures (800, 900, 1000, and 1100 °C) in an argon atmosphere to

achieve complete decomposition of g-C3N4 and full conversion of the chitosan film into

isolated porous NGS. The as-synthesized porous N-doped graphene layer are referred to NGS-

800, NGS-900, NGS-1000, and NGS-1100, respectively. At a temperature higher than 800°C,

g-C3N4 nanosheets are thoroughly decomposed and the graphitic degree of the carbonized

chitosan is greatly improved. Therefore, compared with morphology of NGS-800 (Figure S5c,

S6c, and S7c), the nanosheets of NGS-900, NGS-1000 and NGS-1100 are ultrathin with

graphitization strips (Figure S5d-f, 2f, S6d-f, and S7d-f), which are composed of transparent

thin films with some thick ripples, implying that they are single and/or few layered wrinkled

graphene layers.

The porous structure of NGS was analyzed by nitrogen adsorption technique. Figure 4a

shows the isotherms of NGS samples and g-C3N4. The shape of the isotherm curves of NGS

samples and their sharp increase in nitrogen adsorption at low relative pressures reveal the

porous nature of NGS. The DFT pore size distribution (Figure 4b) indicates the presence of

many micropores which are located at around 1.1 nm and some mesopores. Some micropores

observed by HRTEM (Figure 3b) are not revealed by DFT results due to their low quantity.

Interestingly, the Brunauer-Emmett-Teller (BET) specific area of NGS (Table 1) increases

from 443.4 to 1118.3 m2 g-1 when the synthesis temperature rising from 800 to 1100 °C; but it

drops to 509.8 m2 g-1 when the temperature synthesis further raised to 1100 °C. A similar trend

is also observed for the micropore volume (Table 1). The BET surface area of NGS-1000 is

higher than that of the graphene synthesized by CVD (850 m2 g-1) (Chen et al., 2011) and

monolayer-patched graphene derived from glucose and dicyandiamide (820 m2 g-1) (Li et al.,

2012). The higher specific surface area is probably caused by the porous structure and less

stacked graphene layer.

Page 12: Construction of porous N-doped graphene layer for ...

12

he Raman spectra of the samples (Figure 5a) show two typical features of carbon, namely

D and G bands located at 1368 and1598 cm-1, respectively, for an excitation of 532 nm. The

intensity ratio of D and G bands is proportional to the degree of graphitization (Ferrari A. C.,

2007; Wu et al., 2014). A decrease (from 1.1 to 1.0) of ID/IG for NGS samples is observed when

the synthesis temperature increases from 800 to 1000 °C, indicating that the degree of

graphitization is increased with rising the synthesis temperature (Guo et al., 2010; Ajlani et al.,

2016; Jorio A. and L. Cançado, 2013). The presence of nitrogen in NGS-1000 is demonstrated

by the observation of the nitrogen K-edge in the electron energy loss spectrum (EELS) (see

Figure 5b). An important question is the location of this nitrogen in relation to the graphene

layers. If nitrogen atoms were present within the graphene layers, producing N-doped graphene

layer, one would expect a change in the fine structure of the carbon K-edge observed in the

EELS spectrum (Zhang L. P. and Z. H. Xia, 2011; Nicholls et al., 2013). The spectrum exhibits

typical sp2 coordinated carbon atoms: sharp peak * at 284.4 eV and a broad peak at the range

from 290.0 to 296.8 eV (Zhang L. P. and Z. H. Xia, 2011; Nicholl et al., 2013; Suenaga K.,

and M. Koshino, 2010). Nicholls et al. (Nicholls et al., 2013) demonstrated that the nitrogen

does influence the carbon K-edge, resulting in an extra feature at ~ 290 eV if N replaces some

carbon atoms in graphene. Such a feature is not visible in our EELS data, probably due to the

small amount of nitrogen compared with carbon in the sample.

X-ray photoelectron spectroscopy (XPS) was performed to probe the chemical composition

of NGS samples. The total N content displays a steady decrease from 19.21% to 4.12%

accompanied by the rise of synthesis temperature from 580 to 1100 °C based on the elemental

quantification from XPS (Table S1). Whereas the C content shows a contrary tendency,

increasing from 75.77% (580 °C) to 89.74% (1100 °C), indicating a higher degree of

carbonization. The carbon content of NGS-580 (75.77%) is much higher than that of g-C3N4

(42.28%), further confirming that chitosan is present on the surface of g-C3N4 nanosheets.

Page 13: Construction of porous N-doped graphene layer for ...

13

Figure 5c shows the high-resolution N 1s XPS spectrum of NGS-1000. Component fitting of

the N 1s spectra suggests that the N atoms are present in different structural and substitutional

forms. The high-resolution N 1s XPS spectra are fit using five tentative components. The major

component at 400.7 eV corresponds to the hexagonal N-substitution in the interior of a

graphitic unit (i.e. N bound to 3 sp2 carbon atoms); the major component at 398.1 eV is assigned

to pyridinic type N (i.e. N bond to two sp2 carbon atoms, either at the edge or next to a carbon

vacancy in a graphene structure.) (Hellgren et al., 2016; Jansen et al., 1995; Gammon et al.,

2003). The remaining components located at 399.2, 402.4, 405.4 eV are assigned to ionized

pyridine, quaternary N, and a unassigned contribution, respectively (Hellgren et al., 2016). The

peak assigned as ionized pyridine based on observations by Hellgren et. al. (Hellgren et al.,

2016), however, is commonly assigned as pyrrolic N (Scardamaglia et al., 2014). Quaternary

N, in this case, refers to graphitic N with a +1 charge, i.e. N+. The unassigned contribution at

high binding energy is acknowledged as being difficult to assign as the origin of the

contribution is poorly understood (Gammon et al., 2003, Scardamaglia et al., 2014).

Regardless, as the contribution to the total N species for this component is relatively low, the

accurate assignment is not critical for this study. X axis in Figure 5d and Figure S8 show the

relative fraction of N for each of the components assigned to the NGS samples. At the lowest

temperature (580 °C), the pyridinic N component dominates, with over 50% of the relative

fraction of N. With increasing temperature, the fraction of pyridinic N decreases while graphitic

N (and quaternary N) increases, where the graphitic N component dominates at temperatures

above 900°C. Jointly, as temperature increase, the total N concentration decreases as

demonstrated in Table S1 and right Y axis in Figure 5d. Thus sample fabrication at increasing

temperatures favors the formation of graphitic N over other species. The higher pyrolysis

temperatures induce decomposition of the g-C3N4 structure as well as the decrease the

incidence of pyridinic N.

Page 14: Construction of porous N-doped graphene layer for ...

14

The nitrogen-doped carbon materials were used as the metal-free catalysts for ORR, which

is an important cathodic reaction of fuel cells (Wen et al., 2012; Guo et al., 2016). Although

N-doped graphene materials exhibit high electrocatalytic activity for ORR, the identification

of active sites for these materials are still debated due to the presence of different types of

nitrogen species, morphology, and graphitization level.

The ORR performance of NGS samples in comparison with commercial 20% Pt/C catalyst

was evaluated by a rotating disc electrode (RDE) technique. Observed from cyclic voltammetry

(CV) curves (Figure 6a), pronounced cathodic ORR peaks are observed in O2-saturated

solution, while featureless voltammetric currents within -0.03 to + 1.2 V versus the reversible

hydrogen electrode (RHE) are detected in argon-saturated solution, highlighting outstanding

electrocatalytic activity of NGS. Notably, the reduction peak of NGS samples in O2-saturated

solution becomes more positive, as the annealing temperature increased, except for NGS-1100,

indicating that the catalytic activity is improved. Figure 6b shows the liner-sweep

voltammograms in O2-saturated KOH (0.1 M) of NGS samples treated at various synthesis

temperatures. Among them, NGS-1000 exhibits a more considerable electrocatalytic activity

than the other materials. It has a high onset potential and a well-defined limiting current plateau

for ORR. It is obviously observed that the limiting current density increases with the synthesis

temperature, except for NGS-1100, which means the ORR performance of NGS is strongly

dependent on the synthesis temperature.

In our work, the synthesis temperature crucially affects the graphitic level and the

component of the nitrogen dopant in NGS samples, which cause a significant difference

towards the ORR activity. As reported, the catalytic activity of carbon materials can be

correlated to its electronic properties and thus the amorphous carbon has limited

electrocatalytic activity due to its disordered structure and limited electroconductivity (Zhu Y.,

2011; Primo et al., 2012). Enhancing the annealing temperature can overcome the shortage of

Page 15: Construction of porous N-doped graphene layer for ...

15

low graphitic degree. In addition, the heat-treatment temperature can change the components

of the nitrogen dopant, due to the difference of the thermal stability for various nitrogen species.

Lai et al. (Lai et al., 2012) found that the ORR electrocatalytic activity of N-doped graphene-

based materials greatly depended on the graphitic N content which determined the limiting

current density, while the pyridinic N content improved the onset potential in alkaline

conditions. They also pointed out that the total N content in N-doped carbon materials did not

play a significant role in ORR. In our work, it is worth noting that the synthesis temperature

(>800 °C) slightly decreases total N content. Whereas graphitic N relative content gradually

increases and is even higher than that of pyridinic N when the synthesis temperature is higher

than 900 °C. Importantly, NGS samples exhibit more negative onset potential and higher

current density when the synthesis temperature is increased to 1000 °C. It is illustrated that the

atom ratio of nitrogen in NGS is not crucial for the improvement of ORR performance, while

graphitic N could contribute to the improvement of electrocatalytic activity for ORR and prefer

a 4e reduction pathway.

In addition, ORR performance of NGS-1000 outperforms most of the N-doped carbon

materials reported in the literature (Table S2). There is a noting fact that NGS-1000 has highest

specific surface area than that of graphitic carbons reported in the literature listed in Table S2

and micropore volumes, which implies a better exposure of active sites for ORR that could

enhance the catalytic activity (Guo et al., 2016; Hao et al., 2015; Parvez et al., 2012; Jeon et

al., 2011; Lee et al., 2010).

Besides, the value of limiting current density of NGS-1000 is 5.8 mA cm-2 (at 0.304 V),

which is higher than that of Pt/C (5.5 mA cm-2) (Figure 6c). It has been demonstrated that NGS

cannot be produced from bulk g-C3N4. The sample prepared by heat-treatment of bulk-g-C3N4

thereby presents a much lower ORR activity than NGS-1000. NGS-1000 shows higher

electrocatalytic activity than pure graphene (Cong et al., 2014; Sheng et al., 2011) and nitrogen-

Page 16: Construction of porous N-doped graphene layer for ...

16

doped graphene without nanopores (Liu et al., 2014) in terms of onset potential and limited

current under the similar test conditions due to the nitrogen dopants and the high specific

surface area of NGS-1000, which may facilitate fast mass transport and exchange in the

reaction. Moreover, the NGS-1000 presents a comparable ORR activity to Pt/C catalysts on

almost the same onset potential (0.98 V vs RHE).

The current density of NGS-1000 was analyzed by rotating disc electrodes. The linear

sweep voltammetry (LSV) was performed at different rotating speeds from 400 to 2500 rpm in

the O2-saturated 0.1 M KOH solution (Figure S9a). NGS-1000 exhibits well-defined plateaux

of diffusion-limiting currents below 0.4 V at all rotating speeds, implying it is an efficient

electrocatalyst. The Koutckey-Levich (K-L) plots (Figure S9b) have good linearity. Calculated

from the Levich constants, the electron transfer number of NGS-1000 is around 3.5 in alkaline

media, which is similar to that of NGS-800, 900, and 1100, indicating the 4 e- mechanism

(Figure S10). To further evaluate the electrocatalytic activity of NGS, the ORR was also

carried out in 0.5 M H2SO4 solution saturated with oxygen or argon. It is found that NGS-1000

exhibits a good ORR performance with 0.8 V vs RHE onset potential, and almost 3 electron

transfers under acidic condition (0.5 M H2SO4 solution) (Figure S11), which means that NGS

is chemically stable and can be used in both alkaline and acid solutions.

The long-term stability of NGS-1000 and Pt/C was examined and compared by

chronoamperometric measurement operated at 0.8 V (vs RHE) in an O2-saturated 0.1 M KOH

solution. Figure 6d shows that the commercial Pt/C exhibits a rapid decrease of current density

with only 60.6% retention, after 10000 s testing, indicating a poor stability of the Pt/C toward

ORR. In sharp contrast, 86.2% current density is retained with the catalyst NGS-1000 under

the same testing conditions. Such a stability may be attributed to the porous structure of NGS.

The pores of NGS facilitates the transport of reactant, intermediates, and products, meanwhile,

the conductivity of graphitic structure endows a fast electron transport.

Page 17: Construction of porous N-doped graphene layer for ...

17

4. Conclusions

In summary, porous N-doped graphene layer with a very high ORR performance was

synthesized by pyrolysis of g-C3N4 nanosheets absorbed with chitosan. During the

carbonization process, g-C3N4 served as a self-sacrificing template for carbonization of

chitosan into NGS, it also functioned as a nitrogen source. The XPS measurements revealed

there was a higher content of graphitic N detected in the NGS samples synthesized at higher

temperatures. The electrochemical measurements showed that the porous N-doped graphene

layer exhibited an apparent catalytic activity for ORR in 0.1 M KOH solution, and the sample

synthesized at 1000 °C (NGS-1000) exhibited the highest electrocatalytic performance. The

higher ORR performance and better durability than the commercial Pt/C can be attributed to

the N-doping and high specific surface area, and porosity.

Acknowledgment

This work was in part supported by the Australian Research Council (project no.

DP150100765). The authors acknowledge use of facilities within the Monash Centre for

Electron Microscopy and in particular equipment funded by the Australian Research Council

grant LE110100223.

References

Matter et al., 2006. The Role of Nanostructure in nitrogen-containing carbon catalysts for the

oxygen reduction reaction. Journal Catalysis 239, 83-96.

Zhang L. P. and Z. H. Xia, 2011. Mechanisms of oxygen reduction reaction on nitrogen-doped

graphene for fuel cells. Journal of Physical Chemistry C115, 11170-11176.

Choi I.; et al., 2017. Doped porous carbon nanostructure as non-precious metal catalysts

prepared by amino acid glycine for oxygen reduction reaction. Applied Catalysis B-

Environmental 211, 235-244.

Page 18: Construction of porous N-doped graphene layer for ...

18

Klingele et al., 2016. A review on metal-free doped carbon materials used as oxygen reduction

reaction catalysts in solid electrolyte proton exchange fuel cells. Fuel Cells 16, 522-529.

He et al., 2014. High-rate oxygen electroreduction over graphitic-N species exposed on 3D

hierarchically porous nitrogen-doped carbons. Angewandte Chemie International Edition 53,

9503-9507.

Zhang et al., 2012. Highly conductive and porous sctivated reduced graphene oxide films for

high-power supercapacitors. Nano Letter 12, 1806-1812.

Peng et al., 2014. Nanoarchitectured graphene/CNT@porous carbon with extraordinary

electrical conductivity and interconnected micro/mesopores for lithium- sulfur batteries.

Advanced Functional Materials 24, 2772-2781.

Avouris et al., 2007, Carbon-based electronics. Nature Nanotechnology 2, 605-615.

Balandin et al., 2008. Superior thermal conductivity of single-layer graphene. Nano Letter 8,

902-907.

Zhu Y., 2011. Carbon-based supercapacitors produced by activation of graphene. Science 332,

1537-1541.

Terrones M., 2011. Graphene and graphite nanoribbons: morphology, properties, synthesis,

defects and applications. Nano Today 5, 351-372.

Wang et al., 2012. Review on recent progress in nitrogen-doped graphene: synthesis,

characterization, and its potential spplications. ACS Catalysis 2, 781-794.

Zhuang et al., 2017. MoB/g-C3N4 Interface materials as a schottky catalyst to boost hydrogen

evolution. Angewandte Chemie International Edition 2018, 57, 496-500.

Ferrero et al., 2016. The influence of pore size distribution on the oxygen reduction reaction

performance in nitrogen doped carbon microspheres. Journal of Material Chemistry A 4, 2581-

2589.

Page 19: Construction of porous N-doped graphene layer for ...

19

Presser et al., 2011. Carbide-derived carbons-from porous networks to nanotubes and graphene.

Advanced Functional Materials 21, 810-833.

Wang et al., 2017. Metal-organic framework derived carbon-confined Ni2P nanocrystals

supported on graphene for an efficient oxygen evolution reaction. Chemical Communications,

2017, 53, 8372-8375.

Qie et al., 2013. Synthesis of functionalized 3D hierarchical porous carbon for high-

performance supercapacitors. Energy Environmental Science 6, 2497-2504.

Wei et al., 2015. Nitrogen-doped nanoporous carbon/graphene nano-sandwiches: synthesis and

application for efficient oxygen reduction. Advanced Functional Materials 25, 5768-5777.

Cooper A. I., 2003. Porous materials and supercritical fluids. Advanced Materials 15, 1049-

1059.

Davis M. E., 2002. Ordered porous materials for emerging applications. Nature 417, 813-821.

Xiao et al., 2011. Hierarchically porous graphene as a lithium-air battery electrode. Nano Letter

11, 5071-5078.

Jin et al., 2014. Ionic liquid-assisted synthesis of microporous carbon nanosheets for use in

high rate and long cycle life supercapacitors. Advanced Materials. 26, 3700-3705.

Han et al., 2014. Porous craphene materials for advanced electrochemical energy storage and

conversion devices. Advanced Materials 26, 849-864.

Zheng et al., 2005. Two-dimensional porous carbon: synthesis and ion-transport properties.

Advanced Materials 27, 5388-5395.

Niu et al., Graphene-supported mesoporous carbons prepared with thermally removable

templates as efficient catalysts for oxygen electroreduction. Small 2016, 12, 1900-1908.

Hao et al., 2013. Sandwich-type microporous carbon nanosheets for enhanced supercapacitor

performance. Advanced Energy Materials 3, 1421-1427.

Page 20: Construction of porous N-doped graphene layer for ...

20

Xu et al., 2016. Nitrogen-doped porous carbon superstructures derived from hierarchical

assembly of polyimide nanosheets. Advanced Materials 28,1981-1988.

Wen et al., 2012. Crumpled nitrogen-doped graphene nanosheets with ultrahigh pore volume

for high-performance supercapacitor. Advanced Materials 24, 5610-5616.

Meng et al., General oriented formation of carbon nanotubes from metal-organic frameworks.

Journal of American Chemistry society 139, 24, 8212-8221.

Tang et al., 2009. Electrocatalytic activity of nitrogen-doped carbon nanotube cups. Journal

of American Chemistry Society 131, 13200-13201.

Wu et al., 2015. Graphitic carbon nitride nanosheet electrode-based high-performance ionic

actuator. Nature Communications 6, 7258.

Geng et al., 2011. High oxygen-reduction activity and durability of nitrogen-doped graphene.

Energy & Environmental Science 4, 760-764.

Zheng et al., 2011. Nanoporous graphitic-C3N4@carbon metal-free electrocatalysts for highly

efficient oxygen reduction. Journal of American Chemistry Society 2011, 133, 20116-20119.

Kurak K. A. and A. B. Anderson, 2009. Nitrogen-treated graphite and oxygen electroreduction

on pyridinic edge sites. Journal of Physical Chemistry C 113, 6730-6734.

Guo et al., 2016. Active sites of nitrogen-doped carbon materials for oxygen reduction reaction

clarified using model catalysts, Science 351, 361-365.

Hao et al., 2015. Graphene-based nitrogen self-doped hierarchical porous carbon aerogels

derived from chitosan for high performance supercapacitors. Nano Energy 15, 9-23.

Pandiselvi K., and S. Thambidurai, 2014. Chitosan-ZnO/polyaniline ternary nanocomposite for

high-performance supercapacitor. Ionics 20, 551-561.

Primo et al., 2012. From biomass wastes to large-area, high-quality, N-doped graphene:

catalyst-free carbonization of chitosan coatings on arbitrary substrates. Chemical

Communications 48, 9254-9256.

Page 21: Construction of porous N-doped graphene layer for ...

21

Mateo et al., 2016. 111 oriented gold nanoplatelets on multilayer graphene as visible light

photocatalyst for overall water splitting. Natature Communications 7, 111819.

Li et al., 2012. Synthesis of monolayer-patched graphene from glucose. Angewandte Chemie

International Edition 5, 9689-9692.

Chen et al., 2016. Growth of g-C3N4 on mesoporous TiO2 spheres with high photocatalytic

activity under visible light irradiation. Applied Catalysis B-Environmental 188, 342-350.

Chen et al., 2011. Three-dimensional flexible and conductive interconnected graphene

networks grown by chemical vapour deposition. Nature Materials 10, 424-428.

Ferrari A. C., 2007. Raman spectroscopy of graphene and graphite: disorder, electron–phonon

coupling, doping and nonadiabatic effects. Solid State Communications 143, 47-57.

Wu et al., 2014. Raman Spectroscopy of Graphene. Acta Chimica Sinica 72, 301-318.

Guo et al., 2010. Controllable N-doping of graphene. Nano Letter 10, 4975-4980.

Ajlani et al., 2016. Raman spectroscopy study of annealed incommensurate graphene bilayer

on SiO2 substrate. Superlattices Microstructures 90, 96-106.

Jorio A., and L. Cançado, 2013. Raman spectroscopy of twisted bilayer graphene. Solid State

Communications 175, 3-12.

Zhang et al., 2011. Non-covalent doping of graphitic carbon nitride polymer with graphene:

controlled electronic structure and enhanced optoelectronic conversion. Energy Environmental

Science 4, 4517-4521.

Nicholls et al., 2013. Probing the bonding in nitrogen-doped graphene using electron energy

loss spectroscopy. ACS Nano 7, 7145-7150.

Suenaga K., and M. Koshino, 2010. Atom-by-atom spectroscopy at graphene edge. Nature 468,

1088-1090.

Hellgren et al., 2016. Interpretation of X-ray photoelectron spectra of carbon-nitride thin films:

new insights from in Situ XPS. Carbon 108, 242-252.

Page 22: Construction of porous N-doped graphene layer for ...

22

Jansen et al., 1995. XPS of nitrogen-containing functional groups on activated carbon. Carbon

33, 1021-1027.

Gammon et al., 2003. Experimental comparison of N(1s) X-ray photoelectron spectroscopy

binding energies of hard and elastic amorphous carbon nitride flms with reference organic

compounds. Carbon 41, 1917-1923.

Scardamaglia et al., 2014. Nitrogen implantation of suspended graphene flakes: annealing

effects and selectivity of sp2 nitrogen species. Carbon 73, 371-381.

Lai et al., 2012. Exploration of the active center structure of nitrogen-doped graphene-based

catalysts for oxygen reduction reaction. Energy Environmental Science 5, 7936-7942.

Parvez et al., 2012. Nitrogen-doped graphene and its iron-based composite as efficient

electrocatalysts for oxygen reduction reaction. ACS Nano 2026, 9541-9550.

Jeon et al., 2011. Formation of large-area nitrogen-doped graphene film prepared from simple

solution casting of edge-selectively functionalized graphite and its electrocatalytic activity.

Chemistry Materials 23, 3987-3992.

Lee et al., 2010. Electrochemical oxygen reduction on nitrogen doped graphene sheets in acid

media. Electrochemistry Communications 12, 1052-1055.

Cong et al., 2014. Facile synthesis of mesoporous nitrogen-doped graphene: an efficient

methanol-tolerant cathodic catalyst for oxygen reduction reaction, Nano Energy 3, 55-63.

Sheng et al., 2011. Catalyst-free synthesis of nitrogen-doped graphene via thermal annealing

graphite oxide with melamine and its excellent electrocatalysis. ACS Nano 5, 4350-4358.

Liu et al., 2014. Nitrogen-doped graphene nanoribbons as efficient metal-free electrocatalysts

for oxygen reduction, ACS Applied Material & Interfaces 6, 4214-4222.

Page 23: Construction of porous N-doped graphene layer for ...

23

Tables

Table 1. BET surface area, micropore volume and single point adsorption total volume at 0.98

of P/P0 of g-C3N4 and NGS samples synthesized at varying temperatures from 800 to 1100 °C.

Sample

BET surface area

(m2 g-1)

Micropore

volume

(cm3 g-1)

Single point total

volume (cm3 g-1)

(at 0.98 (P/P0))

g-C3N4 99.5 0.008

0.36

NGS-800 443.4 0.057

1.94

NGS-900 484.1 0.061

2.59

NGS-1000 1118.3 0.134

5.91

NGS-1100 509.8 0.095

2.46

Page 24: Construction of porous N-doped graphene layer for ...

24

Figure Captions

Figure 1. Schematic representation of the synthesis of porous NGS: adsorption of chitosan on

both surfaces of a g-C3N4 nanosheet forming two large planar layers (chitosan film); during the

pyrolysis process, the chitosan film converting into porous N-doped graphene layer

accompanied by decomposition of g-C3N4 nanosheets during the high-temperature treatment.

Figure 2. SEM images of (a) g-C3N4 nanosheets and (b) sandwiched GCN-CS, (c) SEM

image of NGS-1000, and (d) TEM image of NGS-1000.

Figure 3. (a) HRTEM image of NGS-1000 (inset: the corresponding SAED pattern),

and (b) HRTEM image of NGS-1000, partial NGS with pores (~ 2.5 nm) circled in

yellow.

Figure 4. (a) Nitrogen adsorption/desorption isotherms and (b) DFT pore size

distribution of NGS-800, NGS-900, NGS-1000, NGS-1100, and g-C3N4, respectively.

Figure 5. (a) Raman spectra of NGS-800, NGS-900, NGS-1000, and NGS-1100,

respectively, (b) EELS spectrum of the region labelled in bright-field STEM image

(inset) of NGS-1000, (c) N1s XPS spectra of NGS-1000 (inset: schematic illustration of

nitrogen-doped graphene layer), and (d) Relative fraction of N (left axis) and total N

content (right axis) in NGS samples synthesized at different temperatures (800, 900,

1000, and 1100 °C).

Figure 6. (a) CV curves of NG-800, 900, 1000, and 1100 in argon- and O2- saturated

0.1 M KOH aqueous solution with a scan rate of 50 mV s-1; (b) LSV curves of NGS

synthesis at varying temperatures from 800, 900, 1000 to 1100 °C; (c) LSV curves of

NGS-derived from bulk g-C3N4, Pt/C, and NGS-1000; and (d) Current-time

chronoamperometric response of NGS-1000 and Pt/C at 0.80 V with a rotating speed of

1600 rpm in O2-saturated 0.1 M KOH solution.

Page 25: Construction of porous N-doped graphene layer for ...

25

Figure 1.

Page 26: Construction of porous N-doped graphene layer for ...

26

Figure 2.

Page 27: Construction of porous N-doped graphene layer for ...

27

Figure 3.

Page 28: Construction of porous N-doped graphene layer for ...

28

Figure 4.

Page 29: Construction of porous N-doped graphene layer for ...

29

Figure 5.

Page 30: Construction of porous N-doped graphene layer for ...

30

Figure 6.

Page 31: Construction of porous N-doped graphene layer for ...

1

Supporting Information

Construction of porous N-doped graphene layer for efficient oxygen

reduction reaction

Xiaofang Chen1, 2, Yan Liang1, Li Wan1, Zongli Xie2, Christopher D. Easton2, Laure Bourgeois3,

Ziyu Wang4, Qiaoliang Bao4,Yonggang Zhu5, Shanwen Tao1, Huanting Wang1*

1Department of Chemical Engineering, Monash University, Victoria 3800, Australia

2CSIRO Manufacturing, Private Bag 10, Clayton South, Victoria 3169, Australia

3Monash Centre for Electron Microscopy – Department of Materials Science and Engineering,

Monash University, Victoria 3800, Australia

4Department of Materials Science and Engineering, Monash University, Victoria 3800, Australia

5School of Science, RMIT University Melbourne, VIC 3001, Australia

*Corresponding author. E-mail: [email protected]

Page 32: Construction of porous N-doped graphene layer for ...

2

Figure S1. TEM image of g-C3N4 nanosheets.

Page 33: Construction of porous N-doped graphene layer for ...

3

Figure S2. FTIR spectra of g-C3N4, chitosan, and GCN-CS.

Page 34: Construction of porous N-doped graphene layer for ...

4

Figure S3. (a) Carbonized chitosan (synthesis temperature: 1000 °C) and (b) NGS derived from

unexfoliated g-C3N4.

Page 35: Construction of porous N-doped graphene layer for ...

5

Figure S4. Thermogravimetric (TG) analysis of GCN-CS and chitosan.

Page 36: Construction of porous N-doped graphene layer for ...

6

Figure S5. SEM images of NGS synthesized at varying temperatures: (a) 580, (b) 700, (c) 800,

(d) 900, and (e) 1100 °C.

Page 37: Construction of porous N-doped graphene layer for ...

7

Figure S6. TEM images of NGS synthesized at varying temperatures: (a) 580, (b) 700, (c) 800,

(d) 900, and (e) 1100 °C.

Page 38: Construction of porous N-doped graphene layer for ...

8

Figure S7. HRTEM images of NGS synthesized at 580, 700, 800, 900, 1100 °C from (a) to (e),

respectively.

Page 39: Construction of porous N-doped graphene layer for ...

9

Table S1. Surface species contents of samples NGS-580, NGS-700, NGS-800, NGS-900, NGS-

1000 and NGS-1100 determined by XPS.

Sample

C content

(atomic %)

N content

(atomic %)

O content

(atomic %)

Graphitic carbon nitride 42.28 56.97 0.00

NGS-580 75.77 19.21 4.00

NGS-700 72.43 21.68 4.72

NGS-800 82.87 11.70 4.34

NGS-900 87.50 8.66 3.09

NGS-1000 90.71 4.12 4.47

NGS-1100 89.74 2.29 5.81

Page 40: Construction of porous N-doped graphene layer for ...

10

4

Figure S8. Atomic concentration of N species in NGS samples synthesized at different heating

temperatures.

Page 41: Construction of porous N-doped graphene layer for ...

11

Figure S9. (a) LSV curves of NGS-1000 with different rotating speeds (400, 900, 1600 and 2500

rpm), (b) K-L plots for NGS-1000 calculated from LSV curves.

Page 42: Construction of porous N-doped graphene layer for ...

12

Figure S10. LSV curves of (a) NGS-800, (b) NGS-900, and (c) NGS-1100, respectively; K-L

plots (d) NGS-800, e) NGS-900, and (f) NGS-1100, calculated from LSV curves, respectively.

Page 43: Construction of porous N-doped graphene layer for ...

13

Table S2. The catalytic performance of non-metal-doped carbon materials for ORR in recent

literature.

O2-saturated 0.1 M KOH, 1600 rpm

Sample

highest

limiting

current density

(mA cm-2)

E onset

potential

(vs RHE)

Number

of

electron

transfer

Catalyst

loading

(μg cm-

2)

SBET

(m2 g-1) Ref

N-doped graphene nanoribbon 3.2 0.81 3.9 141

-

Liu et

al., 2014

Co2O3/N-doped mesoporous

graphene 5.0 0.88 4.0 170

-

Liang et

al., 2011

S, N doped graphene/carbon

nanotube composite 4.51 0.81 4.0 408

-

Higgins

et al.,

2014

N-doped carbon sheets 5.78 0.95 3.70 600

589

Wei et

al., 2014

Monolayer-patched graphene - 0.79 3.9 286

~ 800

Li et al.,

2012

N-doped porous carbon

nanosheets template from g-

C3N4 5.79 0.89 3.9 161

1077

Yu et

al., 2016

S, N dual-doped mesoporous

graphene 5.79 0.9 3.6 -

-

Liang et

al., 2012

N-doped carbon nanotubes ~ 5.1 0.87 0.393 -

1389

Meng et

al., 2017

Graphene/mesporous carbon 5.86 0.99 > 3.95 204

821

Niu et

al., 2016

Page 44: Construction of porous N-doped graphene layer for ...

14

N-doped graphene 5.8 0.98 3.3 200 1183

This

work

Note: “-” not applicable

Figure S11. (a) CV curves of NGS-1000 in 0.5 M H2SO4 solution with a scan speed of 50 mV/s,

(b) LSV curves of NGS-1000 in O2-saturated 0.5 M H2SO4 solution with different rotating speeds

(400, 900, 1600, 2500 rpm), (c) LSV curves of NGS-100 and Pt/C in O2-saturated solution with a

Page 45: Construction of porous N-doped graphene layer for ...

15

rotating speed 1600 rpm, and (d) K-L plots for NGS-1000 calculated from LSV curves (inset) and

electron transfer numbers at different potentials.

References

Liu et al., 2014. Nitrogen-doped graphene nanoribbons as efficient metal-free electrocatalysts for

oxygen reduction. ACS Applied Material & Interfaces 6, 4214-4222.

Liang et al., 2011. Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction

reaction. Nature Mateials. 10, 780-786.

Higgins et al., 2014. Oxygen reduction on graphene-carbon nanotube composites doped

sequentially with nitrogen and sulfur. ACS Catalysis 4, 2734-2740.

Wei et a., 2014. Nitrogen-doped carbon nanosheets with dize-fefined mesopores as highly efficient

metal-free catalyst for the oxygen reduction reaction. Angewandte Chemie International Edition

53, 1570-1574.

Li et al., 2012. Synthesis of monolayer-patched graphene from glucose. Angewandte Chemie

International Edition 51, 9689-9692.

Yu et al., 2016. Nitrogen-doped porous carbon nanosheets templated from g-C3N4 as metal-free

electrocatalysts for efficient oxygen reduction reaction. Advanced Materials 28, 5080-5086.

Liang et al., 2012. Sulfur and nitrogen dual-doped mesoporous graphene electrocatalyst for oxygen

reduction with synergistically enhanced performance. Angewandte Chemie International Edition

51, 11496-11500.

Page 46: Construction of porous N-doped graphene layer for ...

16

Niu et al., Graphene-supported mesoporous carbons prepared with thermally removable templates

as efficient catalysts for oxygen electroreduction. Small 2016, 12, 1900-1908.

Meng et al., General oriented formation of carbon nanotubes from metal-organic frameworks.

Journal of American Chemistry society 139, 24, 8212-8221.