You are on page 1of 24

www.sciencemag.

org/content/350/6267/1508/suppl/DC1

Supplementary Materials for


Nitrogen-doped mesoporous carbon of extraordinary capacitance for
electrochemical energy storage
Tianquan Lin, I-Wei Chen, Fengxin Liu, Chongyin Yang, Hui Bi, Fangfang Xu,
Fuqiang Huang*
*Corresponding author. E-mail: huangfq@mail.sic.ac.cn

Published 18 December 2015, Science 350, 1508 (2015)


DOI: 10.1126/science.aab3798

This PDF file includes:


Materials and Methods
Figs. S1 to S14
Tables S1 and S2
References

Methods
Synthesis and N-doping of ordered mesoporous carbons and few-layer carbon
The materials below were fabricated either with a silica template or without. In both cases,
few-layer carbon was CVD grown with the aid of carbon source and nickel catalyst. If a template
was used, it was later removed by HF etching. In both cases, a self-supported ordered
mesoporous few-layer carbon (OMFLC) structure in Fig. 1A was obtained. Nitrogen doping was
implemented by including nitrogen source in the precursor and in the CVD gas of the above
process to obtain N-doped OMFLC (OMFLC-N) having N incorporated at various carbon
locations in Fig. 1B. The N content was adjustable by the precursor composition and by
post-CVD acid-oxidation. A more detailed description follows.
Ordered mesoporous carbon (OMC) was synthesized using ordered mesoscopic silica as the
template and polyfurfuryl alcohol (PFA) as the carbon source. The template (SBA-15) was
prepared following the procedure in the literature (11) using tetraethyl orthosilicate (TEOS) as
the silica source and PEO20PPO70PEO20 (Pluronic P123, Sigma-Aldrich), a triblock copolymer,
as the surfactant. The OMC was next prepared by adding 1.0 g of SBA-15 to a solution
containing 1.8 g PFA and 20 mL ethanol, then held at 150 C before carbonization under a
nitrogen atmosphere at 800 C. Finally, the template (SBA-15) was removed by HF (30%)
solution etching to obtain the remaining OMC. Nickel impregnated templates (SBA-15/Ni) were
also prepared from SBA-15 by adding 5 mL g1 of Ni(NO3)26H2O solution (0.5 mol L1). After
drying and calcination under a H2/Ar (10%) flow at 450 C, a product (SBA-15/Ni) of ordered
mesoporous silica with nickel nanocrystals was obtained.
Nitrogen-doped ordered mesoporous few-layer carbon (OMFLC-N) using SBA-15/Ni
template was prepared by CVD. Filled with PFA and dicyandiamide (DCDA), the template after
drying was heated under a hydrogen and argon flow to 1000 C. CVD was initiated by
introducing CH4 and NH3 into the gas ow with Ar:CH4:H2:NH3=300:10:20:100 sccm
(standard-state cubic centimeter per minute). After that the sample was cooled under hydrogen
flow and the template was similarly removed by HF etching to recover the product OMFLC-N.
The N content was adjusted by changing the amount of DCDA and further modified by a HNO3
treatment that partially oxidized N into NO by immersing the samples into concentrated HNO3
for 12 h at room temperature. For example, the mass ratio of PFA:DCDA=1:1 enabled us to

obtain OMFLC-N S1 containing 8.2% N and, after HNO3 treatment, 7.5%; if the ratio increases
to 1:2, the content of nitrogen increases to 11.9% and, after HNO3 treatment, 10.6%. Undoped
OMFLC was similarly synthesized without using DCDA and NH3.
In addition to the above method that was used to obtain OMFLC-N S1, S2, S3 and S4, as
well as their mixture SM (at the ratio of S1:S2:S3=0.3:0.3:0.4), a simplified method was used to
obtain N-doped mesoporous few-layer carbon with properties comparable to OMFLC-N SM.
This is by combining CVD with a sol-gel process containing a sol made of polyethylene glycol
(PEG, as an inexpensive pore-forming agent and C source), urea (as an inexpensive N source)
and nickel nitrate (as before). The gel precursor was first prepared by dissolving Ni(NO3)26H2O
(1.0 g), urea (2.0 g) and PEG (2.5 g) in 10 mL ethanol under vigorous stirring, followed by
gelation at 30 C for 8 h and heating at 80 C for 4 h. Few-layer graphene-like carbon was grown
on the resulting xerogel at 1,000 oC under a gas mixture (Ar:CH4:H2:NH3=300:10:20:100 sccm)
flow. The N content was also adjusted by changing the amount of urea, and further modified by a
HNO3 treatment that partially oxidized N into NO by immersing the samples into concentrated
HNO3 for 12 h at room temperature.
Characterization of materials
Low angle X-ray diffraction (-2 scan) to determine the superstructure of OMC, OMFLC
and OMFLC-N was performed on a Brucker D8 powder X-ray diffractometer using Cu K
radiation. Nitrogen adsorption-desorption isotherms at 77 K were measured on a Micromeritics
Tristar 3000 system using vacuum-degassed samples (180 C for at least 6 h). The isotherms
were used to calculate (a) the specific surface area by the Brunauer-Emmett-Teller (BET)
method and (b) the pore volume and pore size by the BarrettJoynerHalenda (BJH) method.
Scanning electron microscopy (SEM) images were obtained in a field emission Magellan 400
microscope (FEI Company). Transmission electron microscopy (TEM), high angle annular dark
field (HAADF), electron energy loss spectroscopy (EELS, Gatan) examinations were conducted
at 200 kV in a JEOL 2011 microscope operated. X-ray photoelectron spectroscopy (XPS) was
collected in a RBD upgraded PHI-5000C ESCA system (Perkin Elmer) with Mg K radiation
(h =1253.6 eV). Binding energies were calibrated using containment carbon (C 1s =284.6 eV).
Because of the thin thickness (<2 nm) of the few layer carbon OMFLC-N used in this work and
the long mean free path of photoelectron (about 3.2 nm), XPS N/O/C 1s signal is representative
2

of the bulk composition when normalized by the C signal.(29) Raman spectra were collected in a
Thermal Dispersive Spectrometer using a 10 mW laser with an excitation wavelength of 532 nm.
Zeta potential of colloids in suspension was measured by a Zeta Nanoseries (Nano ZS90,
Malvern Instrument Ltd.)

Electrochemical characterization
To fabricate electrodes, active material (OMC, OMFLC, or OMFLC-N) powders were mixed
in N-methyl-2-pyrrolidone (NMP) to form a 10 mg mL1 of homogeneous slurry. In the above,
polyvinylidene fluoride (PVDF) of 5 wt.% was added as binder when more than 2.0 mg cm2
active material was loaded in the electrode; at lower mass loading, no PVDF was used. The
slurry was coated onto a porous 3D-graphene foam (1.2 mm thick, as current collector), which is
highly compressible. The preform (electrode) structure was dried under vacuum at 100 C for 10
h to remove NMP, before being compressed into a disk to form the electrode. The final electrode
was ~10 thinner than the original foam thickness, according to the SEM viewing of the
cross-section (Fig. S14 shows four samples with an average thickness of 102 m), but despite
including the 3D graphene it still maintained a BET surface area of 1,280 m2 g-1 indicating nearly
all powder surfaces remained accessible. The electrode thickness increased with the mass
loading in the electrode. The measured specific capacitance of pristine 3D-graphene foam with
density of 1.1 mg cm2 is about 30 F g1.
Electrochemical experiments were performed in several configurations and electrolytes. For
aqueous electrolytes, 0.5 M H2SO4 (pH 0), 2 M Li2SO4 (pH 0.5, 1.8, 9.2), and 1 M KOH (pH 14)
solutions were used. In three-electrode cells, in addition to the working electrode of active
material, platinum as the counter electrode and Ag/AgCl as the reference electrode were used. In
symmetric cells, two identical (by weight, size and composition) active-material electrodes were
used as cathode and anode. Cyclic voltammetry (CV) tests and galvanostatic chargedischarge
(CC) tests were performed using an electrochemical analyzer, CHI 660E, under ambient
conditions. Electric impedance spectroscopy (EIS) was performed with an amplitude of 10 mV,
from 10 mHz to 100 kHz.
Symmetric devices with two identically configured electrodes were also packaged. Since the
electrodes were highly conductive, they were directly connected to the electrochemical analyzer
by alligators without any current collector other than the 3D graphene foam that was an integral
3

part of the electrode construction. After inserting an ion-porous separator (Celgard 3501)
between the two electrodes, the electrode/electrolyte assembly was wrapped within a Kapton
tape seal to complete the package.

Electrochemical Data
For a nonlinear Faradaic capacitor, the capacitance is obtained from the integrated form

dQ
C= =
dV

I dt
=
dV

it
V

(1)

In the above, i is the average current during the CV/CC charging or discharging cycle, V is the
potential window, and t is the charging or discharging time.

(a) CV test
Since dV/dt=v is a set constant (i.e., the scan rate), integration over the entire loop gives the
average capacitance

C=

S Area
2vV

(2)

Here S Area = I d V is the loop area. The specific capacitance is the capacitance divided by the
mass of the electrode.

C=

S Area
2mvV

(3)

The volumetric capacitance is likewise obtained by dividing capacitance by the volume of the
electrode.
The energy density is obtained from the integral E = V d Q . To average over the
charging/discharging cycle, absolute values are used in integration
4

E=

V d Q = VI d V
2

(4)

2v

The specific energy is the energy divided by the mass of the electrode

E=

VI d V

(5)

2mv

Similarly, the energy density is obtained by dividing the energy by the volume of the electrode.
(b) CC test
Since I is a set constant (i.e., the charging/discharging rate), i=I so Eq (1) reduces to
C=It/V

(6)

(For our Faradaic capacitance, the charging/discharging curves are rather linear, so V/t is
essentially the slope.) The specific capacitance is the capacitance divided by the mass of the
electrode
C=

I
m(V t )

(7)

The volumetric capacitance is likewise obtained by dividing capacitance by the volume of the
electrode.
The energy density is obtained from the integration

E = =
V dQ

VI d t I =
V dt
=

V dt

(8)

In the above, the last integral over both half cycles obtains an average.
The device energy and power are calculated considering the entire device weight and volume,
including the electrodes, the electrolyte, the separator and the packing material. Power (Pwt) is
obtained from Pwt=V2/4RESR and divided by the device weight and volume to obtain specific

power and power density, respectively. Here, V is the operating voltage and RESR is the
equivalent serial resistance (ESR) of the device. ESR is obtained from the CC test by dividing
the voltage drop (Vdrop) upon current reversal by twice the value of I, i.e., RESR=Vdrop/2I.
The results obtained from Eq (2, 4) from the CV test and Eq (7, 8) from the CC test are
compared in Table S2. Here, the specific capacitance, specific energy and specific power
consider the weight of active material in the electrode only. To self-consistently compare the
capacitance of a symmetric EC with the capacitance measured in a three-electrode cell, we
follow Stoller and Ruoff (23) and multiply the nominal specific capacitance of a symmetric EC
by 4 to account for the number (2) of electrodes and the serial connection of two identical
capacitors (having total capacitance =C/2).

Fig. S1. TEM and SEM images of OMFLC and OMFLC-N. (A) TEM image of OMFLC. (B,
C) TEM images of OMFLC-N at two magnifications, showing ordered channels of pores and
graphene-like structure. Few-layer graphene-like structure seen in (C). (D) SEM and
corresponding (E) dark field and (F) bright field STEM images of OMFLC-N.

Fig. S2. Properties of OMC, OMFLC and OMFLC-N (S1). (A) N2 adsorptiondesorption
isotherms. All exhibit typical Langmuir hysteresis indicating presence of well-defined
mesopores. (B) Raman spectra. Characteristic 2D band of few-layer graphene-like structure
found in OMFLC and OMFLC-N but not OMC. Few-layer graphene-like structure in OMFLC is
estimated to have fewer than five layers according to peak position (2,681 cm-1) and
full-width-at-half-maximum (55 cm1) of 2D band and intensity ratio (~0.65) of 2D band to G
band (30). This is consistent with HRTEM observation. Blue shifts of G-band (5 cm1) and the
2D band (13 cm1) of OMFLC-N indicate higher Fermi level according to the literature (31). (C)
Resistance of OMC, OMFLC and OMFLC-N (S1), indicating the improvement of structural
order of carbon (12).

Fig. S3. EELS and XPS. (A) Electron energy loss spectroscopy (EELS) spectra from OMC,
OMFLC and OMFLC-N (S1). (B) EELS features at 410 eV characteristic of N bonding. (CF)
Bonding of OMFLC-N. High-resolution XPS spectra for N 1s of OMFLC-N samples (S1 and
S3: without HNO3 treatment; S2 and S4: with HNO3 treatment). (C) OMFLC-N S1; inset:
locations of N-dopant in carbon sheet, showing pyridinic N (N-6), pyrrolic N (N-5), and
graphitic N (N-Q). (D) Sample S2. (E) Sample S3. (F) Sample S4. Here, pyridinic N refers to
any N with one p-electron on system, and pyrrolic N refers to any N with two p-electrons on
system; the latter is not limited to five-member ring environment as in pyrrole (14, 32).

Fig. S4. Performance of OMFLC-N (S1 and SM) in acidic electrolyte. Electrochemical
performance of the OMFLC-N with 8.2 at.% N (S1) and mixed OMFLC-N (SM) in 0.5 M
H2SO4 electrolyte. (A) CV curves of S1 at voltage scan rates of 2100 mV s1. Here, current is
normalized by voltage scan rate . (B) Specific capacitance calculated from CV curves vs.
voltage scan rate for S1. CC curves at current densities of 140 A g1 of (C) S1 and (E) SM.
Specific capacitance calculated from CC curves vs. current density for (D) S1 and (F) SM.

10

Fig. S5. YP-50 vs OMFLC-N (S1). Electrochemical performance of YP-50 and our OMFLC-N
(S1) in 0.5 M H2SO4 (pH 0) and 2.0 M Li2SO4 (pH 1.8) electrolytes. CC curves of YP-50 and S1
at sweep rates of 1 A g1 in (A) 0.5 M H2SO4 and (C) 2.0 M Li2SO4 electrolyte. CV curves at
current densities of in 2 mV s1 (B) 0.5 M H2SO4 and (D) 2.0 M Li2SO4 electrolyte. (E)
Percentage capacitance retantion of YP-50 in 0.5 M H2SO4 () and 2.0 M Li2SO4 () in cycling
is less than that of OMFLC-N in 0.5 M H2SO4 () and 2.0 M Li2SO4 (). (F) Complex-plane
plots of AC impedance, inset: enlarged view.

11

Fig. S6. Comparison of impedance and kinetics. (A) Enlarged view of Cole-Cole plots of AC
impedance (Fig. 2C) of OMC, OMFLC, OMFLC-N (S1 and SM). (B) Specific capacitance of
OMC, OMFLC and OMFLC-N (S1 and SM) electrodes vs. root-inverse sweep rate, v1/2, with v
from 2 to 500 mV s1, in 0.5 M H2SO4.

12

Fig. S7. Performance of OMFLC-N (S1) in basic electrolyte. (AD) Electrochemical


performance of the OMFLC-N (S1) in 1.0 M KOH electrolyte. (A) CV curves at voltage scan
rates of 250 mV s1. Rectangular CV shape observed at all rates indicating efficient
double-layer formation. Here, current is normalized by voltage scan rate . (B) Specific
capacitance calculated from CV curves vs. voltage scan rate. (C) CC curves at current density of
120 A g1. (D) Specific capacitance calculated from CC curves vs. current density. (E) CV
curves at 2 mV s1 of OMFLC in 1.0 M KOH (alkaline, pH 14) and 0.5 M H2SO4 (acidic, pH 0)
electrolytes, showing nearly identical capacitance.

13

Fig. S8. Threshold voltage of water splitting. Determined by H2 and O2 accumulation


(measured by gas chromatography) in sealed symmetric cell with two OMFLC-N (SM)
electrodes of same capacitance in (A) 0.5 M H2SO4 (pH 0) and (B) 2 M Li2SO4 (pH 1.8) over 24
h under CV sweeping at 2 mV s1.

Fig. S9. Behavior in Li2SO4 electrolytes. (A) CV curves of OMFLC-N (S1) in Li2SO4
electrolytes with different pH at sweep rates of 2 mV s1. Li2SO4 (pH 1.8) electrolyte contains
sufficient H+ for the redox process, as evidenced by the redox peak at a potential similar to that
0.5 M H2SO4. The same is true at pH 0.5 although it is so acidic that it also generates H2. In basic
electrolyte (pH 9.2), the redox peak is lost. (B) CV curves at 2 mV s1 of OMFLC-N (S1) in 1.0
M KOH and Li2SO4 (pH 9.2) electrolytes, showing approximate capacitance.

14

Fig. S10. Microstructure of sol-gel/CVD N-doped mesoporous few-layer carbon. (A) TEM
image showing ordered channels of pores, (B) High resolution TEM image showing
few-layer-graphene like structure.

Fig. S11. Properties of as-prepared sol-gel/CVD N-doped mesoporous few-layer carbon. (A)
N2 adsorptiondesorption isotherm exhibits a typical Langmuir hysteresis indicating presence of
well-defined mesopores. (B) Pore-size distributions. (C) Low-angle X-ray diffraction patterns
showing characteristic (100), (110) and (200) peaks of hexagonal packing. (D) Raman spectrum
showing characteristic 2D band of few-layer-graphene like structure, estimated to have fewer
than five layers according to peak position (2,691 cm-1) and full-width-at-half-maximum (65
cm1) of 2D band and intensity ratio (~0.52) of 2D band to G band. This is consistent with high
resolution TEM observation. (E) High-resolution XPS spectrum for N 1s corresponding to (F)
locations of N-dopant in few-layer carbon, showing pyridinic N (N-6), pyrrolic N (N-5), and
graphitic N (N-Q). The N content is 8.9 at.%.
15

Fig. S12. Electrochemical performances of the sol-gel/CVD N-doped mesoporous few-layer


carbon in 0.5 M H2SO4 and 2 M Li2SO4 (pH 1.8) electrolytes obtained by three-electrode
cells. (A) CV curves at voltage scan rate of 2 mV s1. (B) Specific capacitance calculated from
CV curves vs. voltage scan rate. (C) CC curves at a current density of 1 A g1. (D) Specific
capacitance calculated from CC curves vs. current density.

Fig. S13. Electrochemical performance of symmetric cells. With sol-gel/CVD N-doped


mesoporous few-layer carbon cathode and anode in 0.5 M H2SO4 electrolyte and 2 M Li2SO4
(pH 1.8) electrolytes. (A) Cyclic voltammetry at 2 mV s1 scan rate. (B) Galvanostatic
charge/discharge curves at 1.0 A g1.

16

Fig. S14. SEM images of OMFLC-N electrode. (A) Top-view and (BF) cross-section images
of OMFLC-N electrode after compressing the powder/3D-graphene assembly from the original
thickness 1.2 mm.

17

Table S1. N and O content correlated to redox potential in OMFLC-N samples.


Sample

N
O
(at.%) (at.%)

%N-5 %N-6 %N-Q %N-O %N-5 +


%N-6

%N-5* (V) #

S1

8.2

0.6

1.4

6.1

0.7

7.5

0.40

S2

7.5

6.2

0.9

5.1

0.6

0.9

6.0

0.5

0.50

S3

11.9

0.8

2.2

5.8

3.9

8.0

0.29

S4

10.6

7.7

1.2

4.0

3.7

1.7

5.2

1.0

0.53

%N-6, %N-5, and %N-Q are the percentage of pyridinic N (N-6), pyrrolic N (N-5), and graphitic N (N-Q)
in few-layer carbon, respectively; see a schematic of these nitrogen locations in Fig. 1B, Fig. S3C, and Fig. S11F.
Likewise, %N-O is the percentage of N-O, i.e., N associated with one O. * %N-5 is the decrement of %N-5 after
HNO3 treatment; # Redox potential at peak charging current in the CV curve. Oxidative HNO3 treatment caused the
least stable N-5 to substantially convert to NO without affecting the most stable N-Q, as suggested by the
correlation between %N-O and %N-5. Note that %N-5+%N-6 represents the total non-N-Q fraction of nitrogen,
and it decreases in the order of Samples S3, S1, S2 and S4 whose redox potentials () increase in the same order.
Thus, non-N-Q nitrogen is more responsible for redox reactions than N-Q nitrogen.

18

Table S2. Properties of YP-50, OMC, OMFLC and OMFLC-N in aqueous electrochemical cells.*
CC test at 1 A g1
Samples

CV test at 2 mV s1

3-electrode
capacitance
(F g1)

Symmetric-cell
capacitance
(F g1)

Specific
energy
(Wh kg1)

Specific
power
(kW kg1)

3-electrode
capacitance
(F g1)

Symmetric-cell
capacitance
(F g1)

Specific
energy
(Wh kg1)

H2SO4

175

155

6.0

17.5

180

165

8.0

Li2SO4

160

150

12.5

25.5

170

160

14.0

H2SO4

135

130

5.5

8.5

165

155

7.5

Li2SO4

115

105

8.0

10.0

145

135

11.5

H2SO4

325

315

14.0

19.5

330

320

15.5

Li2SO4

300

290

22.5

30.5

300

285

25.0

OMFLC-N

H2SO4

715

625

25.5

34.5

665

575

28.5

(S1)

Li2SO4

690

590

40.0

40.5

630

540

47.5

OMFLC-N

H2SO4

730

640

27.0

38.0

675

590

29.0

(S2)

Li2SO4

690

590

40.5

45.0

635

550

48.0

OMFLC-N

H2SO4

665

595

24.5

32.5

615

545

27.0

(S3)

Li2SO4

600

530

38.0

38.0

565

495

43.0

OMFLC-N

H2SO4

855

840

36.5

42.5

820

790

39.5

(SM)

Li2SO4

780

740

54.5

44.0

725

715

63.0

YP-50

OMC

OMFLC

Electrolyte

*Mass loading of active material per electrode is 0.5 mg cm2 in three-electrode-cell and symmetric-cell. The gravimetric basis is that of active
material only. #0.5 M H2SO4 (pH 0) and 2 M Li2SO4 (pH 1.8).
19

References
1. J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon, P. L. Taberna, Anomalous increase in
carbon capacitance at pore sizes less than 1 nanometer. Science 313, 17601763 (2006).
Medline doi:10.1126/science.1132195
2. Y. Zhu, S. Murali, M. D. Stoller, K. J. Ganesh, W. Cai, P. J. Ferreira, A. Pirkle, R. M.
Wallace, K. A. Cychosz, M. Thommes, D. Su, E. A. Stach, R. S. Ruoff, Carbon-based
supercapacitors produced by activation of graphene. Science 332, 15371541 (2011).
Medline doi:10.1126/science.1200770
3. P. Simon, Y. Gogotsi, Materials for electrochemical capacitors. Nat. Mater. 7, 845854
(2008). Medline doi:10.1038/nmat2297
4. X. Yang, C. Cheng, Y. Wang, L. Qiu, D. Li, Liquid-mediated dense integration of graphene
materials for compact capacitive energy storage. Science 341, 534537 (2013). Medline
doi:10.1126/science.1239089
5. J. Xia, F. Chen, J. Li, N. Tao, Measurement of the quantum capacitance of graphene. Nat.
Nanotechnol. 4, 505509 (2009). Medline doi:10.1038/nnano.2009.177
6. M. F. El-Kady, V. Strong, S. Dubin, R. B. Kaner, Laser scribing of high-performance and
flexible graphene-based electrochemical capacitors. Science 335, 13261330 (2012).
Medline doi:10.1126/science.1216744
7. J. Huang, B. G. Sumpter, V. Meunier, Theoretical model for nanoporous carbon
supercapacitors. Angew. Chem. Int. Ed. 47, 520524 (2008). Medline
doi:10.1002/anie.200703864
8. L. L. Zhang, X. Zhao, H. Ji, M. D. Stoller, L. Lai, S. Murali, S. McDonnell, B. Cleveger, R.
M. Wallace, R. S. Ruoff, Nitrogen doping of graphene and its effect on quantum
capacitance, and a new insight on the enhanced capacitance of N-doped carbon. Energy
Environ. Sci. 5, 96189625 (2012). doi:10.1039/c2ee23442d
9. G. Lota, B. Grzyb, H. Machnikowska, J. Machnikowski, E. Frackowiak, Effect of nitrogen in
carbon electrode on the supercapacitor performance. Chem. Phys. Lett. 404, 5358
(2005). doi:10.1016/j.cplett.2005.01.074

10. H. Li, J. Wang, Q. Chu, Z. Wang, F. Zhang, S. Wang, Theoretical and experimental specific
capacitance of polyaniline in sulfuric acid. J. Power Sources 190, 578586 (2009).
doi:10.1016/j.jpowsour.2009.01.052
11. D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka, G. D. Stucky,
Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 angstrom
pores. Science 279, 548552 (1998). Medline doi:10.1126/science.279.5350.548
12. P. M. Vora, P. Gopu, M. Rosario-Canales, C. R. Prez, Y. Gogotsi, J. J. Santiago-Avils, J.
M. Kikkawa, Correlating magnetotransport and diamagnetism of sp2-bonded carbon
networks through the metal-insulator transition. Phys. Rev. B 84, 155114 (2011).
doi:10.1103/PhysRevB.84.155114
13. J. Casanovas, J. M. Ricart, J. Rubio, F. Illas, J. M. Jimnez-Mateos, Origin of the large N 1s
binding energy in X-ray photoelectron spectra of calcined carbonaceous materials. J. Am.
Chem. Soc. 118, 80718076 (1996). doi:10.1021/ja960338m
14. D. Wei, Y. Liu, Y. Wang, H. Zhang, L. Huang, G. Yu, Synthesis of N-doped graphene by
chemical vapor deposition and its electrical properties. Nano Lett. 9, 17521758 (2009).
Medline doi:10.1021/nl803279t
15. D. W. Wang, F. Li, L. C. Yin, X. Lu, Z.-G. Chen, I. R. Gentle, G. Q. Lu, H. M. Cheng,
Nitrogen-doped carbon monolith for alkaline supercapacitors and understanding nitrogeninduced redox transitions. Chem. Eur. J. 18, 53455351 (2012). Medline
doi:10.1002/chem.201102806
16. T. Brezesinski, J. Wang, S. H. Tolbert, B. Dunn, Ordered mesoporous -MoO3 with isooriented nanocrystalline walls for thin-film pseudocapacitors. Nat. Mater. 9, 146151
(2010). Medline doi:10.1038/nmat2612
17. K. Brezesinski, J. Wang, J. Haetge, C. Reitz, S. O. Steinmueller, S. H. Tolbert, B. M.
Smarsly, B. Dunn, T. Brezesinski, Pseudocapacitive contributions to charge storage in
highly ordered mesoporous group V transition metal oxides with iso-oriented layered
nanocrystalline domains. J. Am. Chem. Soc. 132, 69826990 (2010). Medline
doi:10.1021/ja9106385

18. X. Lang, A. Hirata, T. Fujita, M. Chen, Nanoporous metal/oxide hybrid electrodes for
electrochemical supercapacitors. Nat. Nanotechnol. 6, 232236 (2011). Medline
doi:10.1038/nnano.2011.13
19. V. Augustyn, J. Come, M. A. Lowe, J. W. Kim, P. L. Taberna, S. H. Tolbert, H. D. Abrua,
P. Simon, B. Dunn, High-rate electrochemical energy storage through Li+ intercalation
pseudocapacitance. Nat. Mater. 12, 518522 (2013). Medline doi:10.1038/nmat3601
20. S. Ardizzone, G. Fregonara, S. Trasatti, Inner and outer active surface of RuO2
electrodes. Electrochim. Acta 35, 263267 (1990). doi:10.1016/0013-4686(90)85068-X
21. Y. Surendranath, M. W. Kanan, D. G. Nocera, Mechanistic studies of the oxygen evolution
reaction by a cobalt-phosphate catalyst at neutral pH. J. Am. Chem. Soc. 132, 16501
16509 (2010). Medline doi:10.1021/ja106102b
22. G. A. Snook, P. Kao, A. S. Best, Conducting-polymer-based supercapacitor devices and
electrodes. J. Power Sources 196, 112 (2011). doi:10.1016/j.jpowsour.2010.06.084
23. M. D. Stoller, R. S. Ruoff, Best practice methods for determining an electrode material's
performance for ultracapacitors. Energy Environ. Sci. 3, 12941301 (2010).
doi:10.1039/c0ee00074d
24. V. Khomenko, E. Frackowiak, F. Bguin, Determination of the specific capacitance of
conducting polymer/nanotubes composite electrodes using different cell configurations.
Electrochim. Acta 50, 24992506 (2005). doi:10.1016/j.electacta.2004.10.078
25. Q. Gao, L. Demarconnay, E. Raymundo-Piero, F. Bguin, Exploring the large voltage range
of carbon/carbon supercapacitors in aqueous lithium sulfate electrolyte. Energy Environ.
Sci. 5, 96119617 (2012). doi:10.1039/c2ee22284a
26. A. F. Burke, Advanced batteries for vehicle applications. In Encyclopedia of Automotive
Engineering, D. Crolla, D. E. Foster, T. Kobayashi, N. Vaughan, Eds. (Wiley, 2014), pp.
120.
27. A. Burke, M. Miller, The power capability of ultracapacitors and lithium batteries for electric
and hybrid vehicle applications. J. Power Sources 196, 514522 (2011).
doi:10.1016/j.jpowsour.2010.06.092

28. D. Linden, T. B. Reddy, Handbook of Batteries (McGraw-Hill, New York, ed. 3, 2001).
29. S. Tanuma, C. J. Powell, D. R. Penn, Calculations of electron inelastic mean free paths II.
Data for 27 elements over the 502000 eV range. Surf. Interface Anal. 17, 911926
(1991). doi:10.1002/sia.740171304
30. A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D.
Jiang, K. S. Novoselov, S. Roth, A. K. Geim, Raman spectrum of graphene and graphene
layers. Phys. Rev. Lett. 97, 187401 (2006). Medline doi:10.1103/PhysRevLett.97.187401
31. A. Das, S. Pisana, B. Chakraborty, S. Piscanec, S. K. Saha, U. V. Waghmare, K. S.
Novoselov, H. R. Krishnamurthy, A. K. Geim, A. C. Ferrari, A. K. Sood, Monitoring
dopants by Raman scattering in an electrochemically top-gated graphene transistor. Nat.
Nanotechnol. 3, 210215 (2008). Medline doi:10.1038/nnano.2008.67
32. Y. F. Lu, S. T. Lo, J. C. Lin, W. Zhang, J. Y. Lu, F. H. Liu, C. M. Tseng, Y. H. Lee, C. T.
Liang, L. J. Li, Nitrogen-doped graphene sheets grown by chemical vapor deposition:
Synthesis and influence of nitrogen impurities on carrier transport. ACS Nano 7, 6522
6532 (2013). Medline doi:10.1021/nn402102y

You might also like