You are on page 1of 9

Article

pubs.acs.org/JPCC

Formation Mechanisms of Graphitic-N: Oxygen Reduction and


Nitrogen Doping of Graphene Oxides
Wei-Wei Wang,, Jing-Shuang Dang, Xiang Zhao,*, and Shigeru Nagase*,

Institute for Chemical Physics & Department of Chemistry, School of Science, Xian Jiaotong University, Xian 710049, China
Fukui Institute for Fundamental Chemistry, Kyoto University, Kyoto 606-8103, Japan

S Supporting Information
*

ABSTRACT: Deoxygenation and nitrogen-doping mechanisms of graphene oxides with participation of foreign NH3 molecules were
investigated by density functional theory calculations. First, reduction on
perfect graphene oxide without any structural defect is proved to be
facilitated at high temperature, but the following doping process is
impracticable because of the huge energy requirement for CC cleavage.
To elucidate the formation of hexagonal graphitic-N, we explored oxygen
reduction and subsequent nitrogen-doping processes on defective graphene
oxides with single vacancies for the rst time. All possible reaction
pathways were taken into account, and the results demonstrate that the
formation of graphitic-N from NH3 and defective graphene oxides with one
carbonyl or two hydroxyl groups is feasible in energy. The dominant
reaction route is found to be exothermic with a practical reaction rate of
2.26 106 s1 at 900 C, which is in a good agreement with experimental
observations.

INTRODUCTION
Ever since its isolation in 2004, the 2D graphene has attracted
tremendous attention owing to its exotic properties.1 To date,
great achievements have been made on graphene-based
nanomaterials in the elds of chemistry, physics, and life
science.219 Besides the well-recognized pristine graphene
structure, both experimental and theoretical eorts were
made to design and synthesize heteroatom-doped graphene
materials to tailor the physicochemical properties for extending
its applications. For example, owing to the unique electronic
properties, nitrogen-doped graphene (NG, in which one or
more carbon atoms are substituted by nitrogen atoms) is
expected to be utilized in fuel cells and other electrochemical
devices.2026
In general, there are two approaches to obtain nitrogendoped graphene in experiments. First, NG can be yielded by
directly mixing the carbon source with nitrogen-containing
compounds (such as ammonia and hydrazine) as precursors.27,28 Beside such direct synthesis, another way is
postsynthesis treatment. By thermal or plasma treatment on
existing graphene or graphene oxide (GO), nitrogen atoms can
be inserted into the network of the carbon skeleton (edges,
defective sites, and also the interior of the carbon
sheet).2325,2936Structurally, there are three types of bonding
congurations for a doped atomic nitrogen in graphene,
designated as graphitic-N, pyridinic-N, and pyrrolic-N (see
Figure 1). Pyridinic-N and pyrrolic-N are formed at the edge or
defective sites of graphene. Graphitic-N, which bonds with
three carbon atoms, is formed inside the network of graphene.
2016 American Chemical Society

Figure 1. Schematic representation of nitrogen-doped graphene.

In 2009, Li et al. reported a simple chemical method to


obtain NG in bulk quantities.32 By annealing GO in NH3, the
N-doped graphene sheets were synthesized at elevated
temperatures. XPS spectra revealed that nitrogen atoms were
embedded in the carbon skeleton, and the graphitic-N exhibited
a signicant role over 900 C.32 Although such an eective
method has been proposed in experiments, the oxygen
reduction and following nitrogen-doping processes of GO are
still less known. Therefore, discussions on reaction pathways,
and corresponding inuential factors in theory are necessary to
understand the mechanisms in detail. Moreover, to the best of
our best knowledge, almost all theoretical studies on the
reactivity of GO focus only on the oxygen reductions, and there
Received: October 29, 2015
Revised: February 23, 2016
Published: February 24, 2016
5673

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

Figure 2. Computational models of graphene oxides. Atom representations are O (red), C (white), and H (pink). GO-1: perfect graphene with an
epoxide group. GO-2: defective graphene with a carbonyl group in the center. GO-3 and GO-4: defective graphene with two hydroxyl groups.

Figure 3. Fragmental structures for stationary points involved in route-a and route-b. Atoms representations are O (red), N (blue), C (white), and H
(pink). Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are given
in kcal mol1.

study is constrained on the formation of graphitic-N in basal


plane of GO.
All calculations were performed with the Gaussian 09
program.53 The M06-2X density functional in conjunction
with the basis set of 6-31G(d) was employed for all structural
optimizations and vibrational frequency calculations. The
correction of zero-point energy (ZPE) to the total energy of
each molecule was taken into account. The electronic ground
states for all GO models in Figure 2 were conrmed as closedshell singlet, with the singlettriplet splitting energies more
than 28.0 kcal mol1. In the present study, the structures of
transition states were located by using the Berny algorithm.
Vibrational analyses were conducted to clarify the nature of
stationary points as global minima or transition states with one
imaginary frequency. On the basis of transition-state theory
(TST), the reaction rates were evaluated by the Arrhenius
formula: exp(Eb/kBT),54 where is the attempt frequency,
T is the reaction temperature, Eb is the computed energy
barrier, and kB is the Boltzmann constant.

is no report on the nitrogen-doping mechanisms so far. In this


study, comprehensive density functional theory (DFT)
computations were performed to explore the reduction and
doping mechanisms of GO by using ammonia as the reducing
agent and meanwhile the nitrogen source for doping. Our
objectives are to elucidate the interactions between ammonia
with pristine and defective graphene oxides with various
oxygen-containing functional groups and to uncover the
formation mechanisms of graphitic-N from foreign NH3 and
parental GO sheet.

COMPUTATIONAL DETAILS

Experimental evidence has shown that the oxygen functionalities distribute on both the edge and basal plane of the
defective GO sheet, and those oxygen groups lead to a
separation of GO sheet into small in-plane aromatic
domains.3741 Accordingly, we modeled the structures of GO
as nite fragmental carbon sheets with various oxygen groups
(Figure 2), which have been extensively used for theoretical
calculations of graphene oxides.39,4245 The molecular radius of
graphene ake is 6.17 . As for the oxygen-containing
functionalities, epoxide (O), hydroxyl (OH), and carbonyl
(CO) groups were established as the major components
because their existence has been well identied.4652 In
addition, the minor component carboxyl (COOH) group
was excluded in the present work because carboxyl is
distributed only at the edge of GO, whereas our present

RESULTS AND DISCUSSION


We initially focused on the reaction of NH3 to the perfect
graphene sheet with single epoxide group (GO-1 in Figure 2).
Epoxide was chosen as the major oxygen-containing
component because this type of functional group exists in
abundance at both the edge and interior of an aromatic domain
of GO. A previous report by Tang et al. indicated that attack of
NH3 on the inner epoxide will cause the reduction of graphene
5674

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C


oxides, and the nitrogen-containing groups (NH2 and NH)
will externally attach onto the carbon surface.55 Therefore, the
adsorption of NH3 groups can be considered as the prelude of
nitrogen doping. Kinetically, the reaction of NH3 to GO with
single epoxide group was conrmed as a stepwise process. In
the rst instance, the foreign NH3 molecule is adsorbed on GO
with a binding energy of 4.6 kcal mol1. After that, a hydrogen
atom in NH3 transfers from nitrogen to oxygen, resulting in a
new CNH2 bond and a OH group on carbon surface. In the
third step, the second NH dissociation takes place and the H
atom attacks on oxygen to yield a H2O molecule. According to
the dierent binding sites of NH2 in the second step, the
reduction is divided into two reaction routes, as shown in
Figures 3 and 4. Route-a in which NH2 is located at the para

participated GO will generate a water molecule as the reduction


product and a NH2 group, which is adsorbed on carbon
surface;55 however, similar to the case of epoxide, because
dissociation of CC bonds on perfect graphene skeleton needs
huge energy requirements, the nitrogen substitution cannot
take place in practice. On the basis of this reason, the reduction
and doping reactions of GO with single hydroxyl group are not
considered in this work.
Dierent from the hypothetical GO model with structurally
perfect graphene layer, experimental observations suggested
that carbon surface is always interspersed with vacancies, and
the most reported defects are single vacancies (SVs) and
double vacancies (DVs), which are thermodynamically
stable.5662 In the case of SV, the missing carbon atom from
hexagonal lattice leads to a defective segment, which includes a
pentagonnonagon pair and an sp-hybridized carbon with
dangling bonds. Herein, the reactions of NH3 to defective GO
with SV (GO-2, GO-3, and GO-4 in Figure 2) were explored to
determine whether it is feasible to dope nitrogen into the basal
plane of carbon sheet to generate the experimentally observed
graphitic-N. The carbonyls and hydroxyls were selected as the
oxygen-containing groups because of the identied large
amounts of sp3 carbon with CO and CO in experiments.32
As for the double vacancies, because it is impossible to generate
the graphitic-N after injecting only one nitrogen atom,
investigations on GO with DV are beyond our present
research. Therefore, in the following discussions, our main
goal is to study the reduction and doping processes of GO with
SV to uncover the formation of graphitic-N and to clarify the
role of defects in carbon skeleton on chemical doping.
Herein, the reactions between ammonia and defective GO
with a carbonyl group were studied kinetically. First,
optimizations of two dierent models with single carbonyl
group were performed, and the results show that the structure
GO-2 shown in Figure 2 is 17.6 kcal/mol lower in energy than
the other model (labeled as GO-2b, shown in Figure S1 in
Supporting Information). Therefore, GO-2 was employed as
the precursor for further kinetic computations. The optimized
structures and calculated reaction channels are shown in
Figures 710. After adsorption of NH3 to generate the identical
intermediate (INT1-ghi in Figure 710) in the rst step, three
distinct pathways based on dierent addition sites of NH3 were
discussed. The calculation results indicate that NH3 prefers to
attack from the pentagonal site (C2, in route-g) to generate a
stable doping product with graphitic-N (NG-H-1 in Figure 7)
after NH dissociations and water formation. Route-g is the
only exothermic reaction with a negative reaction energy of
55.5 kcal mol1 and exhibits the lowest energy barrier of 55.3
kcal mol1 among all possible reaction pathways. Such an
activation barrier is much lower than that of the previously
mentioned nitrogen-doping process from a perfect graphene
layer (ca. 110.0 kcal mol1), implying that structural defect is
essential for nitrogen doping to form graphitic-N. After the
reduction of carbonyl by transferring two hydrogen atoms from
ammonia (NG-H-1), the single carbon vacancy on carbon
surface is healed by the external nitrogen.
As depicted in Figure 10, in comparison with route-g, the
energy barriers of two other channels are much higher. In the
case of route-h, NH3 rst attacks the carbon next to the
pentagon (C3), which exhibits the shortest distance to foreign
nitrogen in INT1-ghi of 3.30 . Because the defective vemembered ring in GO-2 is not destroyed during nitrogen
doping, the nal doping network contains a defective

Figure 4. Reaction energy proles for ammonia participated reduction


of GO-1 (route-a in red and route-b in green). Relative energies
referred to the total energy of reactants (GO-1+NH3) are calculated at
the M06-2X/6-31G(d) level of theory, units in kcal mol1.

position of oxygen, is considered as a more favorable pathway


because of a lower energy barrier for hydrogen transfer to form
the H2O molecule. On the basis of our DFT calculations, the
second NH dissociation (TS3-a) acts as the rate-determining
step with an energy barrier of 42.0 kcal mol1 (route-a),
suggesting that the reduction of epoxide on graphene oxide is
facilitated at elevated temperature. After elimination of H2O,
the external nitrogen atom is attached onto the basal plane of
the graphene to form a NH-containing reduced product (GNH-1), as shown in Figure 3.
Furthermore, the subsequent nitrogen-doping process on
pristine graphene (G-NH-1 in Figures 3 and 5) was explored
here. In general, nitrogen doping includes two major steps: N
H dissociation and CN formation. According to a dierent
sequence of the two steps and distinct positions of newly
formed CN bonds, four reaction pathways were found to
elucidate the doping process. As shown in Figures 5 and 6, the
reaction is preferred to generate a sp3 carbon in the rst step
(INT1-cd in route-c(d)) and then to form new CN bonds
(C4N bond of NG-CH-1 in route-c and C3N bond of NGCH-2 in route-d, respectively); however, energetically, even
assisted by the migrated hydrogen atom (route-d, in Figures 5
and 6), this substitution process is still endothermic, with a
huge activation barrier of 107.2 kcal mol1, which indicates that
the doping is unlikely to take place even at elevated
temperatures.
Accordingly, direct doping from awless graphene network
seems to be unpractical. It should be mentioned that such a
conclusion can also apply to the case of GO with hydroxyl
groups located on the interior of the aromatic domain. It has
been demonstrated that the reaction of NH3 to hydroxyl5675

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

Figure 5. Fragmental structures for stationary points involved in route-c, -d, -e, and -f. Atoms representations are O (red), N (blue), C (white), and
H (pink).Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are
given in kcal mol1.

Figures 9 and 10 that the reaction cannot embed the nitrogen


inside the network of graphene to generate the graphitic-N.
Moreover, the reaction needs to overcome a considerably huge
barrier of 109.7 kcal mol1, implying that route-i is kinetically
unfavorable. Therefore, according to the previously mentioned
three reaction routes in Figures 79, route-g is considered to
be the dominant mechanism for CO reduction and nitrogen
doping, both thermodynamically and kinetically.
Besides carbonyl, hydroxyl is another important oxygencontaining group in GO. It should be mentioned that if only
one OH is participating, the reduction products must be a
water molecule and an adsorbed NH2 group. The two NH
bonds in NH2 are stable, and obviously NH dissociation
cannot take place for further nitrogen doping. Therefore, the
nitrogen doping with the presence of OH should be
considered as a multihydroxyl participated proceed. In the
present work, the reactions between NH3 and defective GO
with two hydroxyl groups were investigated. As shown in Figure
2 and Figure S2 in the Supporting Information, the two models
GO-3 and GO-4, which are proved to be the two most stable
GOs with two hydroxyl groups among all possible congurations, were established as the precursors. GO-3 is 0.2 kcal/
mol lower than GO-4 in energy and at least 1.2 kcal/mol lower
than other species. In GO-3 and GO-4, one of the two

Figure 6. Reaction energy proles for nitrogen doping of GO-1


(route-c in green, route-d in red, route-e in pink, and route-f in blue).
Relative energies referred to the energy of reactant (G-NH-1) are
calculated at the M06-2X/6-31G(d) level of theory, units in kcal
mol1.

pentagonheptagonpentagon adjacency (NG-H-2 in Figure


8). As seen in Figures 8 and 10, the much higher energy barrier
(77.5 kcal mol1) as well as the positive reaction energy (2.1
kcal mol1) both suggest that the formation of this pentagon
and heptagon-containing heteroconguration is energetically
unfavorable in practice. In the case of route-i, it is evident from
5676

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

Figure 7. Fragmental structures for stationary points involved in route-g. Atoms representations are O (red), N (blue), C (white), and H (pink).
Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are given in kcal
mol1.

Figure 8. Fragmental structures for stationary points involved in route-h. Atoms representations are O (red), N (blue), C (white), and H (pink).
Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are given in kcal
mol1.

Figure 9. Fragmental structures for stationary points involved in route-i. Atoms representations are O (red), N (blue), C (white) and H (pink).
Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are given in kcal
mol1.

NH3 on the OH group by hydrogen bonding interaction


(INT1-j and INT1-k) is exothermic by 14.9 kcal mol1 in
GO-3 (12.5 kcal mol1 in GO-4). In contrast with the
barrierless adsorption of NH3 to hydroxyl, the adsorption of
NH3 to carbon surface in Figures 13 and 14 (INT1-l and INT1m) is demonstrated as an endothermic reaction with an

hydroxyls is located on the dangling carbon, and the other one


is adsorbed around the vacancy. A hydrogen bond is formed
between the two OH groups for stabilization.
In the rst step, the foreign NH3 is adsorbed on GO. The
NH3 molecule can locate on both the carbon surface and the
OH group. As shown in Figure 11 and 12, the location of
5677

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

Figure 10. Reaction energy proles for reactions of ammonia with GO-2 (route-g in red, route-h in blue, and route-i in green). Relative energies
referred to the total energy of reactants (GO-2+NH3) are calculated at the M06-2X/6-31G(d) level of theory, units in kcal mol1.

Figure 11. Fragmental structures for stationary points involved in route-j and route-k. Atoms representations are O (red), N (blue), C (white), and
H (pink).Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are
given in kcal mol1.

Interestingly, we also calculated the proton transfer between


OH groups without the participation of NH3 for comparison.
As shown in Figures 15 and 16, the energy barrier is found to
be slightly increased to 12.7 kcal mol1 for GO-3 and 15.0 kcal
mol1 for GO-4, which indicates that the adsorbed NH3 can
promote the hydrogen transfer to form an eliminated NH3
H2O complex. After the removal of NH3H2O (GO-2 in
Figure 11), the OH group that locates on the dangling C6 is
reduced as a carbonyl group. In another word, GO-2 is
obtained from GO-3 and GO-4 by intramolecular dehydroxylation. As previously mentioned, the CO group in GO-2 can
react with NH3 to dissociate the NH bonds to generate a
water molecule and NH species (route-g in Figure 8). The
nitrogen atom in NH3 is successfully inserted into the graphene
surface to form a graphitic-N. Throughout the whole reaction
processes (route-j for dehydroxylation and route-g for
decarbonylation), dehydroxylation from two OH is much
easier, and the rate-determining step is the reduction of CO
and nitrogen doping in subsequent steps.
Kinetically, on the basis of the Arrhenius formula, the
calculated reaction rate of route-g at 900 C is 2.26 106 s1
(with a barrier of 55.3 kcal/mol and an attempt frequency of
429.5 cm1, which is originated from the frequency analysis of
rate-determining TS3-g), implying that the formation of
graphitic-N from NH3 and defective GO with a carbonyl
group at elevated temperature is available. (The calculated

Figure 12. Reaction energy proles for ammonia participated


dehydroxylation of GO-3 (route-j in red) and GO-4 (route-k in
green). Relative energies referred to the total energy of reactants (GO3+NH3) are calculated at the M06-2X/6-31G(d) level of theory, units
in kcal mol1.

activation barrier of 27.6 kcal mol1 in GO-3 (30.1 kcal mol1


in GO-4), suggesting that the NH3 molecule is preferred to
adsorb to OH rather than the carbon surface. Therefore, we
further investigated the following reduction and doping
reactions based on the structures of INT1-j and INT-k, as
shown in Figure 11. In the presence of adsorbed NH3, the
dehydroxylation from two OH groups was subsequently
calculated. The reaction barrier of hydrogen transfer is 10.9 kcal
mol1 in GO-3 and 12.8 kcal mol1 in GO-4, respectively.
5678

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

Figure 13. Fragmental structures for stationary points involved in route-l and route-m. Atoms representations are O (red), N (blue), C (white) and
H (pink).Other carbon atoms are omitted for clarity. The distances are represented in units of angstroms. Relative energies (in parentheses) are
given in kcal mol1.

Figure 14. Reaction energy proles for ammonia participated


dehydroxylation of GO-3 (route-l in red) and GO-4 (route-m in
green). Relative energies referred to the energy of reactant GO-3 are
calculated at the M06-2X/6-31G(d) level of theory, units in kcal
mol1.

Figure 16. Reaction energy proles for intramolecular dehydroxylation


of GO-3 (route-n in red) and GO-4 (route-o in green). Relative
energies referred to the energy of reactant GO-3 are calculated at the
M06-2X/6-31G(d) level of theory, units in kcal mol1.

ammonia molecules. In the case of perfect graphene oxide


without any defects (GO-1), de-epoxidation can be achieved by
NH dissociations, but the following nitrogen doping is
energetically unfavorable because of the endothermic character
and the huge activation barrier of 107.2 kcal mol1 for CC
cleavage. Moreover, defective graphene oxides with single
vacancies were modeled to expose the formation mechanism of
graphitic-N. Three distinct reaction pathways (route-g, -h, (i))
were discovered for decarbonylation and nitrogen doping.
Owing to the negative reaction energy of 55.5 kcal mol1 and
the lowest energy barrier of 55.3 kcal mol1, route-g is
predicted to be the most favorable pathway for doping. The
calculated reaction rate of route-g at 900 C (2.26 106 s1)
indicates that the formation of graphitic-N from NH3 and
defective GO with a carbonyl group (GO-2) is feasible at
elevated temperature. Furthermore, we discussed the reaction
of NH3 to defective GO with two hydroxyl groups (GO-3 and
GO-4). Our calculation results suggest that the NH3 molecule
is preferred to adsorbed with the hydroxyl group instead of the
carbon surface, and the reactivity of dehydroxylation between
two hydroxyls is improved in the presence of adsorbed NH3 to
generate the previously mentioned defective GO-2 with a single
carbonyl group. Similarly, subsequent nitrogen doping from
NH3 can take place, and the graphitic-N can be generated by
route-g.

Figure 15. Fragmental structures for stationary points involved in


route-n and route-o. Atoms representations are O (red), C (white),
and H (pink). Other carbon atoms are omitted for clarity. The
distances are represented in units of angstroms. Relative energies (in
parentheses) are given in kcal mol1.

reaction rates at other temperatures are listed in Table S1 in the


Supporting Information for comparison.) Such a computational
result is consistent with experimental evidence 32 and
successfully uncovers the nitrogen-doping mechanisms of GO
from foreign nitrogen sources for the rst time.

CONCLUSIONS
In the present work, we studied the oxygen reduction and
nitrogen-doping mechanisms of GO in the presence of
5679

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C

(9) Yin, Z.; Wu, S.; Zhou, X.; Huang, X.; Zhang, Q.; Boey, F.; Zhang,
H. Electrochemical Deposition of ZnO Nanorods on Transparent
Reduced Graphene Oxide Electrodes for Hybrid Solar Cells. Small
2010, 6, 307312.
(10) Ponomarenko, L. A.; Schedin, F.; Katsnelson, M. I.; Yang, R.;
Hill, E. W.; Novoselov, K. S.; Geim, A. K. Chaotic Dirac Billiard in
Graphene Quantum Dots. Science 2008, 320, 356358.
(11) Wang, S.; Wang, J.; Miraldo, P.; Zhu, M.; Outlaw, R.; Hou, K.;
Zhao, X.; Holloway, B. C.; Manos, D.; Tyler, T.; et al. High Field
Emission Reproducibility and Stability of Carbon Nanosheets and
Nanosheet-based Backgated Triode Emission Devices. Appl. Phys. Lett.
2006, 89, 183103.
(12) Chen, H.; Muller, M. B.; Gilmore, K. J.; Wallace, G. G.; Li, D.
Mechanically Strong, Electrically Conductive, and Biocompatible
Graphene Paper. Adv. Mater. 2008, 20, 35573561.
(13) Wang, L.; Lee, K.; Sun, Y. Y.; Lucking, M.; Chen, Z.; Zhao, J. J.;
Zhang, S. B. Graphene Oxide as an Ideal Substrate for Hydrogen
Storage. ACS Nano 2009, 3, 29953000.
(14) Liu, C.; Yu, Z.; Neff, D.; Zhamu, A.; Jang, B. Z. Graphene-Based
Supercapacitor with an Ultrahigh Energy Density. Nano Lett. 2010, 10,
48634868.
(15) Mohanty, N.; Berry, V. Graphene-Based Single-Bacterium
Resolution Biodevice and DNA Transistor: Interfacing Graphene
Derivatives with Nanoscale and Microscale Biocomponents. Nano Lett.
2008, 8, 44694476.
(16) Jiang, Z.; Wang, J.; Meng, L.; Huang, Y.; Liu, L. A Highly
Efficient Chemical Sensor Material for Ethanol: Al2O3/Graphene
Nanocomposites Fabricated from Graphene Oxide. Chem. Commun.
2011, 47, 63506352.
(17) Mao, S.; Lu, G.; Yu, K.; Bo, Z.; Chen, J. Specific Protein
Detection Using Thermally Reduced Graphene Oxide Sheet
Decorated with Gold Nanoparticle-Antibody Conjugates. Adv. Mater.
2010, 22, 35213526.
(18) Dong, X.; Shi, Y.; Huang, W.; Chen, P.; Li, L. J. Electrical
Detection of DNA Hybridization with Single-Base Specificity Using
Transistors Based on CVD-Grown Graphene Sheets. Adv. Mater.
2010, 22, 1649.
(19) Nguyen, P.; Berry, V. Graphene Interfaced with Biological Cells:
Opportunities and Challenges. J. Phys. Chem. Lett. 2012, 3, 1024
1029.
(20) Wang, X.; Li, X.; Zhang, L.; Yoon, Y.; Weber, P. K.; Wang, H.;
Guo, J.; Dai, H. N-Doping of Graphene Through Electrothermal
Reactions with Ammonia. Science 2009, 324, 768771.
(21) Reddy, A. L. M.; Srivastava, A.; Gowda, S. R.; Gullapalli, H.;
Dubey, M.; Ajayan, P. M. Synthesis of Nitrogen-doped Graphene
Films for Lithium Battery Application. ACS Nano 2010, 4, 63376342.
(22) Jin, Z.; Yao, J.; Kittrell, C.; Tour, J. M. Large-Scale Growth and
Characterizations of Nitrogen-Doped Monolayer Graphene Sheets.
ACS Nano 2011, 5, 41124117.
(23) Guo, B.; Liu, Q.; Chen, E.; Zhu, H.; Fang, L.; Gong, J. R.
Controllable N-Doping of Graphene. Nano Lett. 2010, 10, 4975
4980.
(24) Wang, Y.; Shao, Y.; Matson, D. W.; Li, J.; Lin, Y. NitrogenDoped Graphene and Its Application in Electrochemical Biosensing.
ACS Nano 2010, 4, 17901798.
(25) Jeong, H. M.; Lee, J. W.; Shin, W. H.; Choi, Y. J.; Shin, H. J.;
Kang, J. K.; Choi, J. W. Nitrogen-Doped Graphene for HighPerformance Ultracapacitors and the Importance of Nitrogen-Doped
Sites at Basal Planes. Nano Lett. 2011, 11, 24722477.
(26) Wu, Z. S.; Ren, W.; Xu, L.; Li, F.; Cheng, H. M. Doped
Graphene Sheets As Anode Materials with Superhigh Rate and Large
Capacity for Lithium Ion Batteries. ACS Nano 2011, 5, 54635471.
(27) Wei, D.; Liu, Y.; Wang, Y.; Zhang, H.; Huang, L.; Yu, G.
Synthesis of N-Doped Graphene by Chemical Vapor Deposition and
Its Electrical Properties. Nano Lett. 2009, 9, 17521758.
(28) Qu, L.; Liu, Y.; Baek, J. B.; Dai, L. Nitrogen-Doped Graphene as
Efficient Metal-Free Electrocatalyst for Oxygen Reduction in Fuel
Cells. ACS Nano 2010, 4, 13211326.

Overall, we introduced the detailed formation mechanisms of


nitrogen-doped graphene materials from various GO and
foreign nitrogen sources for the rst time. The present work
suggests that not only the reducing agent but also the defects in
carbon skeleton are essential for reduction and doping, which is
useful to understand the chemical-doping process of graphenebased materials for the design of hyperne functionalized
materials in the future.

ASSOCIATED CONTENT

S Supporting Information
*

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acs.jpcc.5b10607.
Structures and relative energies of single carbonyl and
two-hydroxylGO congurations, calculated reaction rates
at dierent temperatures, and full citations of refs 4, 11,
48, and 53. (PDF)

AUTHOR INFORMATION

Corresponding Authors

*X.Z.: Fax: +86 29 82668559. Tel: +86 29 82665671. E-mail:


xzhao@mail.xjtu.edu.cn.
*S.N.: E-mail: nagase@ims.ac.jp.
Author Contributions

The manuscript was written through contributions of all


authors. All authors have given approval to the nal version of
the manuscript.
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
This work has been nancially supported by the National
Natural Science Foundation of China (21171138, 21573172)
and the Specially Promoted Research Grant (22000009) from
the Ministry of Education, Culture, Sports, Science, and
Technology of Japan.

REFERENCES

(1) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang,
Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Electric Field Effect
in Atomically Thin Carbon Films. Science 2004, 306, 666669.
(2) Stankovich, S.; Dikin, D. A.; Dommett, G. H. B.; Kohlhaas, K. M.;
Zimney, E. J.; Stach, E. A.; Piner, R. D.; Nguyen, S. T.; Ruoff, R. S.
Graphene-based Composite Materials. Nature 2006, 442, 282286.
(3) Yoo, E.; Kim, J.; Hosono, E.; Zhou, H.; Kudo, T.; Honma, I.
Large Reversible Li Storage of Graphene Nanosheet Families for Use
in Rechargeable Lithium Ion Batteries. Nano Lett. 2008, 8, 2277
2282.
(4) Wang, D.; Choi, D.; Li, J.; Yang, Z.; Nie, Z.; Kou, R.; Hu, D.;
Wang, C.; Saraf, L. V.; Zhang, J.; et al. Self-Assembled TiO2
Graphene Hybrid Nanostructures for Enhanced Li-Ion Insertion. ACS
Nano 2009, 3, 907914.
(5) Wang, G.; Shen, X.; Yao, J.; Park, J. Graphene Nanosheets for
Enhanced Lithium Storage in Lithium Ion Batteries. Carbon 2009, 47,
20492053.
(6) Eda, G.; Lin, Y. Y.; Miller, S.; Chen, C. W.; Su, W. F.; Chhowalla,
M. Transparent and Conducting Electrodes for Organic Electronics
from Reduced Graphene Oxide. Appl. Phys. Lett. 2008, 92, 233305.
(7) Wang, X.; Zhi, L.; Mullen, K. Transparent, Conductive Graphene
Electrodes for Dye-Sensitized Solar Cells. Nano Lett. 2008, 8, 323
327.
(8) Wu, J.; Becerril, H. A.; Bao, Z.; Liu, Z.; Chen, Y.; Peumans, P.
Organic Solar Cells with Solution-processed Graphene Transparent
Electrodes. Appl. Phys. Lett. 2008, 92, 263302.
5680

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

Article

The Journal of Physical Chemistry C


(29) Shao, Y.; Zhang, S.; Engelhard, M. H.; Li, G.; Shao, G.; Wang,
Y.; Liu, J.; Aksay, I. A.; Lin, Y. Nitrogen-doped Graphene and Its
Electrochemical Applications. J. Mater. Chem. 2010, 20, 74917496.
(30) Geng, D.; Chen, Y.; Chen, Y.; Li, Y.; Li, R.; Sun, X.; Ye, S.;
Knights, S. High Oxygen-reduction Activity and Durability of
Nitrogen-doped Graphene. Energy Environ. Sci. 2011, 4, 760764.
(31) Sheng, Z. H.; Shao, L.; Chen, J. J.; Bao, W. J.; Wang, F. B.; Xia,
X. H. Catalyst-free Synthesis of Nitrogen-doped Graphene via
Thermal Annealing Graphite Oxide with Melamine and Its Excellent
Electrocatalysis. ACS Nano 2011, 5, 43504358.
(32) Li, X.; Wang, H.; Robinson, J. T.; Sanchez, H.; Diankov, G.; Dai,
H. Simultaneous Nitrogen Doping and Reduction of Graphene Oxide.
J. Am. Chem. Soc. 2009, 131, 1593915944.
(33) Imran Jafri, R.; Rajalakshmi, N.; Ramaprabhu, S. Nitrogen
Doped Graphene Nanoplateletsas Catalyst Support for Oxygen
Reduction Reaction in Proton Exchange Membrane Fuel Cell. J.
Mater. Chem. 2010, 20, 71147117.
(34) Lin, Y. C.; Lin, C. Y.; Chiu, P. W. Controllable Graphene Ndoping with Ammonia Plasma. Appl. Phys. Lett. 2010, 96, 133110.
(35) Zhang, L. S.; Liang, X. Q.; Song, W. G.; Wu, Z. Y. Identification
of the Nitrogen Species on N-doped Graphene Layers and Pt/NG
Composite Catalyst for Direct Methanol Fuel Cell. Phys. Chem. Chem.
Phys. 2010, 12, 1205512059.
(36) Parvez, K.; Yang, S.; Hernandez, Y.; Winter, A.; Turchanin, A.;
Feng, X.; Mullen, K. Nitrogen-Doped Graphene and Its Iron-Based
Composite As Efficient Electrocatalysts for Oxygen Reduction
Reaction. ACS Nano 2012, 6, 95419550.
(37) Liang, Y.; Wu, D.; Feng, X.; Mullen, K. Dispersion of Graphene
Sheets in Organic Solvent Supported by Ionic Interactions. Adv. Mater.
2009, 21, 16791683.
(38) Paredes, J. I.; Villar-Rodil, S.; Sols-Fernandez, P.; MartnezAlonso, A.; Tascon, J. M. D. Atomic Force and Scanning Tunneling
Microscopy Imaging of Graphene Nanosheets Derived from Graphite
Oxide. Langmuir 2009, 25, 59575968.
(39) Li, J. L.; Kudin, K. N.; McAllister, M. J.; Prudhomme, R. K.;
Aksay, I. A.; Car, R. Oxygen-Driven Unzipping of Graphitic Materials.
Phys. Rev. Lett. 2006, 96, 176101.
(40) Gomez-Navarro, C.; Weitz, R. T.; Bittner, A. M.; Scolari, M.;
Mews, A.; Burghard, M.; Kern, K. Electronic Transport Properties of
Individual Chemically Reduced Graphene Oxide Sheets. Nano Lett.
2007, 7, 34993503.
(41) Mkhoyan, K. A.; Contryman, A. W.; Silcox, J.; Stewart, D. A.;
Eda, G.; Mattevi, C.; Miller, S.; Chhowalla, M. Atomic and Electronic
Structure of Graphene-Oxide. Nano Lett. 2009, 9, 10581063.
(42) Frankcombe, T. J.; Bhatia, S. K.; Smith, S. C. Ab initio
Modelling of Basal Plane Oxidation of Graphenes and Implications for
Modelling Char Combustion. Carbon 2002, 40, 23412349.
(43) Sendt, K.; Haynes, B. S. Density Functional Study of the
Chemisorption of O2 Across Two Rings of the Armchair Surface of
Graphite. J. Phys. Chem. C 2007, 111, 54655473.
(44) Tachikawa, H.; Kawabata, H. Electronic States of Defect Sites of
Graphene Model Compounds: A DFT and Direct Molecular Orbital
Molecular Dynamics Study. J. Phys. Chem. C 2009, 113, 76037609.
(45) Gao, X.; Jang, J.; Nagase, S. Hydrazine and Thermal Reduction
of Graphene Oxide: Reaction Mechanisms, Product Structures, and
Reaction Design. J. Phys. Chem. C 2010, 114, 832842.
(46) Boukhvalov, D. W.; Katsnelson, M. I. Modeling of Graphite
Oxide. J. Am. Chem. Soc. 2008, 130, 1069710701.
(47) Li, Z.; Zhang, W.; Luo, Y.; Yang, J.; Hou, J. How Graphene is
Cut upon Oxidation? J. Am. Chem. Soc. 2009, 131, 63206321.
(48) Cai, W.; Piner, R. D.; Stadermann, F. J.; Park, S.; Shaibat, M. A.;
Ishii, Y.; Yang, D.; Velamakanni, A.; An, S. J.; Stoller, M.; et al.
Synthesis and Solid-State NMR Structural Characterization of 13CLabeled Graphite Oxide. Science 2008, 321, 18151817.
(49) Szabo, T.; Tombacz, E.; Illes, E.; Dekany, I. Enhanced Acidity
and pH-dependent Surface Charge Characterization of Successively
Oxidized Graphite Oxides. Carbon 2006, 44, 537545.
(50) He, H.; Klinowski, J.; Forster, M.; Lerf, A. A New Structural
Model for Graphite Oxide. Chem. Phys. Lett. 1998, 287, 5356.

(51) He, H.; Riedl, T.; Lerf, A.; Klinowski, J. Solid-State NMR
Studies of the Structure of Graphite Oxide. J. Phys. Chem. 1996, 100,
1995419958.
(52) Szabo, T.; Berkesi, O.; Forgo, P.; Josepovits, K.; Sanakis, Y.;
Petridis, D.; Dekany, I. Evolution of Surface Functional Groups in a
Series of Progressively Oxidized Graphite Oxides. Chem. Mater. 2006,
18, 27402749.
(53) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A.; et al. Gaussian 09, revision A.02; Gaussian, Inc.:
Wallingford, CT, 2009.
(54) Han, S.; Yoon, M.; Berber, S.; Park, N.; Osawa, E.; Ihm, J.;
Tomanek, D. Microscopic Mechanism of Fullerene Fusion. Phys. Rev.
B: Condens. Matter Mater. Phys. 2004, 70, 113402.
(55) Tang, S.; Cao, Z. Adsorption and Dissociation of Ammonia on
Graphene Oxides: A First-Principles Study. J. Phys. Chem. C 2012, 116,
87788791.
(56) Meyer, J. C.; Kisielowski, C.; Erni, R.; Rossell, M. D.; Crommie,
M. F.; Zettl, A. Direct Imaging of Lattice Atoms and Topological
Defects in Graphene Membranes. Nano Lett. 2008, 8, 35823586.
(57) Warner, J. H.; Rummeli, M. H.; Ge, L.; Gemming, T.;
Montanari, B.; Harrison, N. M.; Buchner, B.; Briggs, G. A. D.
Structural Transformations in Graphene Studied with High Spatial and
Temporal Resolution. Nat. Nanotechnol. 2009, 4, 500504.
(58) Ugeda, M. M.; Brihuega, I.; Guinea, F.; Gomez-Rodrguez, J. M.
Missing Atom as a Source of Carbon Magnetism. Phys. Rev. Lett. 2010,
104, 096804.
(59) Krasheninnikov, A. V.; Lehtinen, P. O.; Foster, A. S.; Nieminen,
R. M. Bending the Rules: Contrasting Vacancy Energetics and
Migration in Graphite and Carbon Nanotubes. Chem. Phys. Lett. 2006,
418, 132136.
(60) El-Barbary, A. A.; Telling, R. H.; Ewels, C. P.; Heggie, M. I.;
Briddon, P. R. Structure and Energetics of the Vacancy in Graphite.
Phys. Rev. B: Condens. Matter Mater. Phys. 2003, 68, 144107.
(61) Lee, G.-D.; Wang, C. Z.; Yoon, E.; Hwang, N.-M.; Kim, D.-Y.;
Ho, K. M. Diffusion, Coalescence, and Reconstruction of Vacancy
Defects in Graphene Layers. Phys. Rev. Lett. 2005, 95, 205501.
(62) Banhart, F.; Kotakoski, J.; Krasheninnikov, A. V. Structural
Defects in Graphene. ACS Nano 2011, 5, 2641.

5681

DOI: 10.1021/acs.jpcc.5b10607
J. Phys. Chem. C 2016, 120, 56735681

You might also like