Вы находитесь на странице: 1из 14

Author's personal copy

Nano Energy (2012) 1, 552565

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/nanoenergy

REVIEW

Nanostructured activated carbons from natural


precursors for electrical double layer capacitors
Lu Weia,b, Gleb Yushina,n
a

School of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0245, USA
School of Materials Science and Engineering, Northwestern Polytechnical University, Xian, Shaanxi 710072, PR China

Received 4 April 2012; received in revised form 4 May 2012; accepted 4 May 2012
Available online 22 May 2012

KEYWORDS

Abstract

Activated carbon;
Graphene;
Natural precursor;
Electrochemical
capacitor;
Double layer
capacitor

The development of energy-sustainable and energy-efcient economy depends on the ability to


produce low-cost high-performance renewable materials for electrical energy storage devices.
The electrical double layer capacitors (EDLCs) with nanostructured activated carbon (AC)
electrodes from natural precursors have attracted considerable attention due to their great
cycle stability, combined with moderate cost and attractive overall performance. Such ACs
offer high specic surface area, high electrical conductivity, relatively low price, easy and
environmental friendly production in large quantities. The recent developments in the synthesis
of such AC materials allow for the greatly enhanced specic capacitance in a wide range of
electrolytes. This review provides a summary of a recent research progress in synthesis and
understanding the critical structure-property relationships for nanostructured ACs and highlights the trends for the future developments of ACs for EDLC applications.
& 2012 Elsevier Ltd. All rights reserved.

Introduction
Electrical double layer capacitors (EDLCs) are one of the
promising electrochemical energy storage devices with high
power characteristics. The use of EDLCs range from consumer electronics to memory backup systems and uninterruptable power sources to smart grid systems to energy
efcient industrial equipment and hybrid electric vehicles
(HEVs) [1,2]. For HEV applications EDLCs could either be
used alone or in combination with fuel cells or rechargeable
batteries to deliver high power for vehicle acceleration and
n

quickly adsorb the energy converted from braking [1,2].


Unlike batteries, where energy storage is obtained via redox
reactions, EDLCs are based on the charge separation
occurring at an electrodeelectrolyte interface. In the
simplest conguration (Figure 1), an EDLC consists of two
high surface area porous carbon electrodes immersed into
electrolyte and separated by an ion-conducting but electron-insulating membrane. The use of porous carbons is
motivated by their high electrical conductivity, low cost and
high chemical stability. Upon the application of a potential
to one of the electrodes, the ions of the opposite sign

Corresponding author at: 771 Ferst Dr. NW, Room 371, Atlanta, GA 30332-0245, USA. Tel.: +1 404 385 3261; fax: +1 404 894 9140.
E-mail address: yushin@gatech.edu (G. Yushin).

2211-2855/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.nanoen.2012.05.002

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors

Figure 1 The schematic of an EDLC.

accumulate on its surface in quantity proportional to the


applied voltage, forming a so-called electrical double layer.
This double layer consists of an electrical space charge from
the electrode side and an ion space charge from the electrolyte side [3]. The charge storage mechanism in a pure
EDLC is non-Faradaic, which means that during charging and
discharging of this device no charge or mass transport takes
place across the electrodeelectrolyte interface and the
energy storage is purely electrostatic. Because electrostatic
interactions are not detrimental to the integrity and stability
of the electrodes, in contrast to Faradaic processes common
in battery electrodes, EDLCs can undergo 1000,000 cycles of
chargedischarge in organic electrolytes with lower than 10%
degradation. However, the energy density of these devices
is an order of magnitude lower than that of rechargeable
batteries [4].
In order to achieve high energy and power densities, EDLC
electrode materials must possess a large specic surface area
(SSA) available for ion electro-adsorption, pore size distribution (PSD) optimized for the desired combination of energy
and power characteristics, good electrical conductivity, and
wettability by the electrolyte utilized in the device construction. The SSA and PSD are often considered to be two
key factors controlling the energy and power density of
EDLCs [5]. To improve the state of the art EDLCs, various
porous carbon materials, such as activated carbons (ACs),
carbide derived carbons (CDCs), zeolite-templated carbons
(ZTCs), carbon nanotubes (CNTs), carbon onions, carbon
aerogels, and graphene, have been studied extensively as
electrodes in EDLCs. CDCs [615] and ZTCs [1619] offer tight
pore size control, exhibit high surface area, high conductivity
and other attractive properties, but these technologies have
not been yet commercialized to achieve a large scale
production required for the material adoption by EDLC
manufacturers. The requirements to employ multi-step purication procedures and the use of toxic chemicals slow down
the adoption of both technologies. For example, highly
corrosive Cl2 gas is commonly utilized for metal etching from

553

carbides during the CDC production and hydrouoric (HF)


acid etching is used for zeolite template etching during the
ZTC synthesis. CNTs [2026], carbon onions [20,27,28] and
carbon aerogels [2931] with large external surface area and
excessive porosity commonly offer high rate capability but
low capacitance per unit mass and, more importantly, per
unit volume. Vertically aligned CNTs [32] can deliver high
power density and high specic capacitance if functionalized
with surface groups during plasma-etching treatment
[33,34], but their preparation is expensive, and long-term
stability of such materials has not been reached so far. In
spite of the moderately high synthesis cost, graphene, a twodimensional monolayer of graphite, has recently received
rapidly growing attention in supercapacitors [3541] as it
possesses high electrical conductivity and a high theoretical
SSA of over 2600 m2 g1. However, due to the unavoidable
aggregation of graphene nanosheets, the surface area of
graphene is usually much lower than the theoretical one and
more importantly, due to very high pore volume and thus low
density, graphene electrodes commonly offer poor volumetric capacitance. Finally, graphene electrodes suffer from
poor control over the PSD with the lack of ne-tuned
micropores needed to achieve high specic capacitance
[4244] and the lack of macropores or large mesopores
commonly present between the individual AC or CDC or
ZTC particles and needed for rapid charging and discharging
of EDLCs. Activation of graphene [41] improves its performance characteristics but does not yet match the performance of most advanced ACs, CDCs and ZTCs [1,8,10,16,45].
Therefore, in the view of the authors, ACs will likely
dominate EDLC market, in spite of the development of
alternative routes for porous carbon synthesis.
AC is the oldest and most common type of porous carbons.
The use of AC in Egypt was described as early as in 1550 BC
[46]. Industrial production of ACs in the U.S. started in 1913
[47]. At present, nearly all commercial EDLCs utilize high
surface area AC powers due to their well-developed manufacturing technologies, easy production in large quantities,
relatively low cost and great cycle stability. Petroleum coke,
pitch and coal used to be the most common precursors for
commercial AC productions, but the decreasing availability of
fossil fuels, the growing global energy demand, and an
increased awareness of the environmental impacts of fossil
fuel combustion led to the AC productions from sustainable
and renewable resources, such as nutshells, wood, starch,
sucrose, cellulose, corn grain, banana ber, coffee grounds,
sugar cane bagasse, and others. The costs of some of the raw
materials and the carbon yields of pyrolysis from these
carbon sources were presented in Table 1.
In the view of the authors, although conventional carbon
sources such as petroleum coke and charcoal have high
carbon yields, lower-priced and broadly available natural
renewable precursors will likely dominate the future AC
production. Some of the low-cost materials, such as sucrose,
additionally offer high chemical purity and uniform structure,
and thus may allow one to achieve uniform and reproducible
properties of the produced ACs, which is critically important
property for many demanding EDLC applications. Furthermore, the recent progresses in AC synthesis allow the
achievement of well-controlled properties, suitable pore
size distribution, low pore tortuosity, and greatly enhanced
capacitive characteristics [5052].

Author's personal copy


554

L. Wei, G. Yushin

Table 1 Raw material costs and carbon yields from


pyrolysis of some of the natural precursors [48,49].
Raw material

Petroleum coke
Charcoal
Lignite
Coconut shell
Wood
Potato starch
Sucrose
Cellulose
Corn grain
Banana ber

Raw material
cost
($ kg1)

Carbon yield
from pyrolysis
(wt.%)

1.4
1.2
0.75
0.25
0.8
1.0
0.25
0.65
0.25
4

90
90
50
30
25
45
o45
o45
o45
o45

Synthesis of activated carbons


The classic AC production route is generally divided into two
categories: (i) physical activation and (ii) chemical activation. In physical activation, carbon precursor is pyrolyzed at
temperatures in the range 600900 1C, in absence of oxygen
(usually in inert atmosphere of N2) to remove non-carbon
species, and followed by exposing to oxidizing atmospheres
(CO2, steam, or a mixture of both) at temperatures above
450 1C, usually in the temperature range of 6001200 1C to
develop porosity by partial etching of carbon. In chemical
activation, prior to carbonization, the carbon precursor is
impregnated with certain chemicals (such as KOH, NaOH,
H3PO4, ZnCl, H2SO4, among a few), then the precursor is
carbonized at slightly lower temperatures of 450900 1C. It is
believed that the carbonization/activation step proceeds
simultaneously with the chemical activation. Chemical activation is preferred over physical activation owing to the
lower temperatures and shorter time needed for activating
materials. Moreover, chemical activation generally results in
more uniform pore size distribution and higher specic
capacitance in both aqueous and organic electrolytes. After
activation, a porous network in the bulk of the carbon
particles is produced, and micropores (pore size o2 nm),
mesopores (pore size between 250 nm) and macropores
(pore size 450 nm) can be created in carbon grains.
Accordingly, the porous structure of carbon is characterized
by a broad distribution of pore sizes. Longer activation time
or higher temperature leads to larger mean pore size [53].

Application of activated carbons for EDLCs in


different electrolytes
Different kinds of electrolytes have been used in EDLCs with
AC electrodes: (i) aqueous (solution of acids, bases and salts),
(ii) organic (most commonly solution of tetraethylammonium
tetrauoroborate (TEATFB) salt either in anhydrous acetonitrile (AN) or propylene carbonate (PC) solvents), and (iii) ionic
liquid (IL). Aqueous electrolytes are typically stable to 0.6
1.0 V in symmetric EDLCs, organic electrolytes to 2.22.9 V,

and ILs to 2.64.0 V. The energy density of an EDLC strongly


depends on the maximum voltage applied to the device
according to Eqs. (1) and (2) below:


C  C
EDLC 2
E EDLC
 Vmax

1
C C
where EEDLC is the energy density of an EDLC, C and C + are
the capacitances of negative and positive electrodes, and
EDLC
is the maximum voltage applied to the EDLC. In order to
Vmax
maximize the energy of an EDLC the capacitance and
resistance of both electrodes should be the same. In this case:
E EDLC

1
EDLC 2
C  Vmax

Since the choice of electrolyte determines the maximum


operating voltage, it strongly affects the energy density
of EDLCs. The high voltage stability of ILs, make IL-based
EDLCs of very high energy density. In addition, ILs are nonammable, non-volatile and non-toxic, which are important
attributes for many mobile applications, including their
potential use in hybrid engines [5458]. Serious shortcomings of ILs are their high (often prohibitively high) current
cost and commonly low ionic mobility at room temperature.
The advantages of using aqueous electrolytes include
their very low cost and high ionic conductivity. Serious
disadvantages, however, include their low maximum applied
voltage and the corrosion of EDLC electrodes observed at
higher temperatures and voltages (particularly for acidbased electrolytes, such as H2SO4 solutions), which limits
their lifetime and cycle life of the EDLCs.
Organic electrolyte-based EDLCs offer cycle life in excess
of 500,000 and a moderately high operational voltage. They
can operate in the temperature window from 30 1C to
+ 50 1C (in case of an AN solvent) and are used in the
majority of commercial EDLCs with AC electrodes. However,
they suffer from high ammability and potential explosion
risk because organic solvents exhibit high vapor pressure.
Selected characteristics of ACs produced from natural
precursors (including fossil fuels and natural renewable
materials) for EDLC use in different electrolytes are shown
in Tables 2 and 3. Conventional ACs from fossil fuels were
generally produced by chemical activation (Table 2), which
allows the SBET of ACs to range from 8703100 m2 g1, and
the highest specic capacitance of 330 F g1 in aqueous
electrolytes, 220 F g1 in organic electrolytes and 136 F g1
in ionic liquids to be achieved. The double layer capacitances of ACs from natural renewable materials reach
35250 F g1 in organic electrolytes, and 60150 F g1 in
ionic liquids (at high temperatures). This value can even
exceed 300 F g1 in aqueous electrolytes (Table 3), but at a
lower cell voltage because the electrolyte voltage window
is limited by water decomposition. The highest specic
capacitances of ACs from natural renewable materials can
clearly reach and even exceed the values for ACs from fossil
fuels in all of the three types of electrolytes. It indicates
that ACs from natural, renewable, low-cost and environmental-friendly materials will likely replace the more
conventional ACs from fossil fuels and dominate the future
AC market.
Among the ACs applied in aqueous electrolytes, waste
coffee ground-derived carbon (CGC) treated with ZnCl2
(Table 3) showed the highest specic capacitance of 368 F g1

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors

Table 2

Conventional activated carbons from fossil fuels for EDLC electrodes.

Carbon source Activation SBET a


method
(m2 g1)

Petroleum
coke
Petroleum
coke
Coal

Highest capacitance
(F g1)

Electrolyte

Ref.

Aqueous

Organic

Ionic liquid

1M
H2SO4
6 M KOH

[59]

[60]

[61]

1 M LiClO4/PC

[62]
[61]

[63]

EMITFMS
PVdF-HFP/EMITf/
MgTf2
1Me3BuImBF4

[64]
[65]

KOH

2280

213 (two-electrode cell)

KOH

1590

330 (two-electrode cell)

KOH

3150

317 (two-electrode cell)

Coal
Pitch

KOH
KOH

2650
2660

Pitch

KOH

2170

1M
H2SO4
220 (three-electrode cell)
299 (two-electrode cell) 1 M
H2SO4
140 (two-electrode cell)

Pitch
Charcoal

KOH
NaOH

2200(SDFT)b
875

130 (two-electrode cell)


136 (two-electrode cell)

1 M Et4NBF4/
PC

Peat

K2S

1090

67 (two-electrode cell)

a
b

[66]

BET specic surface area.


DFT specic surface area.

Table 3

Activated carbons from natural renewable precursors for EDLC electrodes.

Carbon source

Activation
method

SBET a
(m2 g1)

Highest capacitance
(F g1)

Electrolyte

Ref.

Aqueous

Organic

Ionic liquid

6 M KOH

1 M HNO3
30 wt.% H2SO4

1Me3BuImBF4

EMImBF4

EMImBF4

[66]
[67]
[1]
[45]
[45]
[68]
[69]
[69]
[45]
[45]
[70]
[71]
[49]
[72]
[73]
[74]
[75]
[76]

[77]
[78]

1 M TEABF4/AN
1.2 M MeEt3NBF4/
AN

1 M (C2H5)4NBF4/
AN

[79]
[80]
[81]

[82]
[83]
[83]

[84]
[84]

Furfurol
Coconut shell
Coconut shell
Coconut shell
Eucalyptus wood
Firwood
Bamboo
Bamboo
Cellulose
Potato starch
Potato starch
Starch
Sucrose
Sucrose
Beer lees
Banana ber
Corn grain
Sugar cane
bagasse
Apricot shell
Sunower seed
shell
Coffee ground
Coffee ground
Wheat straw

Steam
KOH
KOH
KOH
KOH
Steam
KOH
KOH
KOH
KOH
KOH
KOH
CO2
CO2
KOH
ZnCl2
KOH
ZnCl2

1040
1660
1515
1515
2970
1130
1250
1290
2460
2270
2340
1510
2100
1940
3560
1100
3200
1790

111 (two-electrode cell)


79 (two-electrode cell)
127 (two-electrode cell)
118 (two-electrode cell)
236 (two-electrode cell)
142(three-electrode cell)
65 (two-electrode cell)
35 (two-electrode cell)
185 (two-electrode cell)
180 (two-electrode cell)
335 (two-electrode cell)
194 (two-electrode cell)
163 (two-electrode cell)
148 (two-electrode cell)
188(three-electrode cell)
74 (three-electrode cell)
257 (two-electrode cell)
300 (two-electrode cell)

6 M KOH
30 wt.% KOH
1 M H2SO4

0.1 M H2SO4
1 M NaSO4
6 M KOH
1 M H2SO4

1M
1M

1M
1M
1M

NaOH
KOH

2335
2510

339 (two-electrode cell)


311 (two-electrode cell)

6 M KOH
30 wt.% KOH

ZnCl2
ZnCl2
KOH

1020
1020
2316

368 (two-electrode cell)


134 (two-electrode cell)
251 (two-electrode cell)

1 M H2SO4

Fish scale
Cherry stone
Cherry stone

Not shown
KOH
KOH

2270
1300
1300

168 (two-electrode cell)


230 (two-electrode cell)
120 (two-electrode cell)

7 M KOH
2 M H2SO4

Rice husk
Rice husk

NaOH
KOH

1890
1390

210 (two-electrode cell)


180 (two-electrode cell)

3 M KCl
3 M KCl

555

BET specic surface area.

TEABF4/AN
TEABF4/AN

Et4NBF4/PC
TEABF4/AN
TEABF4/AN

Author's personal copy


556

L. Wei, G. Yushin

in 1 M H2SO4 [79]. Interestingly, compared with commercial


AC Maxsorb (derived from petroleum pitch and activated by
KOH), coffee ground-based activated carbon showed lower
BET specic surface area and total pore volume, but a
higher specic capacitance per unit surface area (Table 4).
Both carbons contained a small number of mesopores in 2
4 nm width, however, the CGC had a greater ratio of narrow
microporous (o1 nm) to total pore volume, which could be
more important for achieving high capacitance. Moreover,
high-resolution X-ray photoelectric spectroscopy (XPS) spectra identied 6.8 at.% oxygen and 1.5 at.% nitrogen on the
surface of CGC. In contrast, the Maxsorb surface contained
less than 2 at.% oxygen and no nitrogen. The CGC electrodes
exhibited high energy density of up to 20 W h kg1 and
stable performance over 10,000 cycles at a cell potential of
1.2 V and current load of 5 A g1 (Figure 2b). The good
performance of CGC was attributed to a well-developed
porosity with a distribution of mainly micropores and small
mesopores (24 nm wide), and the presence of electrochemically active quinone oxygen groups and nitrogen functional groups that contributed to pseudocapacitance due to
stable Faradaic-based reactions. Such pseudocapacitance
contributions, however, evidently did not reveal themselves
in the rectangularly-shaped cyclic voltammetry curves
(Figure 2a).
The highest capacitance of 251 F g1 in organic electrolyte (1.2 M Methyltriethylammonium Tetrauoroborate
(MeEt3NBF4)/AN) was obtained by using AC made from
carbonization of wheat straw and followed by KOH activation [81]. The wheat straw-based activated carbon (WSC)
with honeycomb-like morphology and lots of nano holes

(Figure 3) presented micro- and mesoporous structure with


the peak pore size of 2.1 nm (Figure 4a), and the sizes of
pores covering a broad range of 0.410 nm. The BET surface
area and total pore volume from the nitrogen adsorption
isotherm reached 2316 m2 g1 (Table 2) and 1.496 cm3 g1,
respectively. By studying the mass loss from the thermal
gravimetric analysis (TGA) of AN impregnated activated
carbon, the author concluded that the pores larger than
0.85 nm could be most efciently impregnated by electrolytes. It is important to note that if the pore size distribution is correct, the high specic capacitance was achieved
in this largely mesoporous sample. Since the surface chemistry analysis was not conducted on these samples, the
potential contribution of functional groups to the pseudocapacitance remains unclear.
Figure 4b showed the relationship between specic capacitance of WSC and scan rate in the organic electrolytes.
Although WSC has relatively large pore size, the specic
capacitance drops quickly with the increase of scan rate from
2 mV s1 to 300 mV s1. This may be induced by the bottleneck pores existing within the WSC, which can slow down the
ion transport at fast scan rates.
The capacitive behavior of ACs was also investigated in
several ILs at both room and elevated temperatures. Among
them, AC prepared via a pyrolysis of sucrose followed by
activation in the stream of CO2 gas at 900 1C showed
moderately high specic capacitance of 148 F g1 in
1-ethyl-3-methylimidazolium tetrauoroborate (EMImBF4)
[72]. The sucrose-derived carbon powder showed irregular
morphology and the particle size of 115 mm with most of
the particles in the range of 37 mm (Figure 5a). Particles

Table 4 Surface texture properties and specic capacitance of coffee ground-based activated carbon (CGC) compared to
Maxsorb, from Ref. [79].
Sample

SBET
(m2 g1)

Total pore volume


(cm3 g1)

Small micropore volume


(o1 nm) (cm3 g1)

CSAa
(mF cm2)

CGC
Maxsorb

1019
1840

0.48
0.84

0.21
0.34

24
14

Specic capacitance per unit surface area at a current load of 5 A g1.

Figure 2 (a) Cyclic voltammograms of coffee ground-based activated carbon (CGC) in a two-electrode cell with 1 M H2SO4
electrolyte, and (b) electrochemical stability of CGC and Maxsorb over 5000 cycles at a cell potential of 01 V (closed symbols)
followed by 5000 cycles with a cell potential of 01.2 V (open symbols), from Ref. [79].

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors

557

Figure 3 SEM images of the wheat straw-based activated carbon prepared at 700 1C for 120 min in different magnication, from
Ref. [81].

Figure 4 (a) Incremental pore volume vs. pore width for the wheat straw-based activated carbon, and (b) the relationship between
specic capacitance of wheat straw-based activated carbon and scan rate in the organic electrolytes of MeEt3NBF4/PC and
MeEt3NBF4/AN, from Ref. [81].

Figure 5

SEM images of the sucrose-derived carbon powder before (a) and after (b) CO2 activation, from Ref. [72].

smaller than 1 mm were also present but their total volume


was less than 5%, according to the SEM image analysis. The
small particle size can promote the electrolyte diffusion by
providing a short ion-transport pathway. The activation with
CO2 has not changed the particle shape and the particle size
distribution to any signicant extent (Figure 5b). By controlling the activation time (from 0 h to 6 h), the sucrose-derived
carbon after CO2 activation for 6 h (ASC-6h) exhibited a
moderately high SSA of 1940 m2 g1 and a micropore volume

of 0.87 cm3 g1. Although ASC-6h has only micropores (pore


size o2 nm), it showed good rate performance and capacitance retention with scan rate increasing from 1 mV s1 to
500 mV s1 (Figure 6). Interestingly, at slow scan rates the
specic capacitance normalized by the surface area is slightly
higher for ASC-4h than for ASC-6h samples in EMImBF4
(Figure 6b). This may be explained by the hypothesis that
larger normalized capacitance in IL can be achieved in pores
with dimensions approaching (slightly larger) the size of the

Author's personal copy


558
cations [55]. The high specic capacitance in IL and good rate
capability of the ASC electrodes have likely resulted from the
large SSA, high percentage of micropores and small content
of bottle-neck pores obtained by a prolonged (6 h) activation
process.
From Table 3, most of the carbons were activated by
chemical activation. By controlling the mass ratio of carbon
samples and chemicals (such as KOH or ZnCl2), high SSA,
large pore volume and pores in the range of 0.44 nm were
achieved. The highest SSA of 3560 m2 g1 was obtained by
KOH activation on the beer lees-based carbons (Table 3).
The textual properties of the beer lees-based ACs such as
pore volume and pore size were gradually increasing with
the increase of mass ratio from 1:2 to 1:4 between beer lees
char and KOH (Figure 7). The obtained ACs were noted as
BL-1/2, BL-1/3 and BL-1/4. The pore size distribution of the
.
BL-1/2 is mostly located in the microporocity below 10 A
increases as the following
The fractions of porosity over 10 A
order: BL-1/2oBL-1/3oBL-1/4. Especially, in the case of
the BL-1/4 sample, the fraction of porosity in the range of
opore width (W)o30 A
is highly developed.
20 A

L. Wei, G. Yushin
Furthermore, the KOH activation also introduced oxygen
functional groups on the BL-ACs (Table 5). XPS results showed
that the CO type groups increase consistently with the
mass ratio between the carbonized beer lees char and
KOH whereas the graphitic carbon and C=O groups slightly
decrease. The oxygen functional groups on the surface of ACs
may contribute to the capacitance of EDLCs owing to both the
enhanced wetting of ACs and the pseudocapacitance contribution in aqueous electrolyte solutions [59,60,85]. In this
work, the specic capacitances of the BL-ACs decreased with
the increasing of graphic carbon and C=O functional groups
contents, whereas it was enhanced by the increasing of CO
groups. In addition to the EDLC capacitance enhancement,
the positive property of the CO functional groups of BL-ACs is
their high stability during EDLC operation [60,85].
In summary, chemical activation with a well-controlled
mass ratio between the carbon sample and the activating
agent, can produce ACs with higher specic surface area,
larger pore volume, and uniform pore size distribution.
When the activation process introduces surface functional
groups, higher specic capacitance in both aqueous and

Figure 6 Inuence of scan rate on (a) gravimetric specic capacitance and (b) specic capacitance normalized by the surface area
of activated sucrose derived carbons at 60 1C in EMImBF4, from Ref. [72].

Figure 7 (a) Pore size distributions of beer lee-based activated carbons calculated by DFT equation from N2 adsorption/desorption
isotherms, and (b) pore volume distributions, from Ref. [73].

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors

Table 5
[73].

Oxygen functionalities of BL-ACs, from Ref.

Sample C1s
Graphitic
(%)
BL-1/2 71.6
BL-1/3 66.8
BL-1/4 68.0

CO
(%)

C= O
(%)

OC = O
(%)

Carbonate
(%)

13.3
17.8
18.4

7.8
6.9
6.7

4.0
5.1
4.8

3.0
3.1
1.9

organic electrolytes is commonly observed. Physical activation needs higher temperature and longer activation time to
develop suitable pore size distribution and help to eliminate
the small bottle-neck pores in the ACs, which may hamper
the rapid electrolyte diffusion and greatly reduce the
capacitance retention of EDLCs at high scan rates and
current densities [49,68,72]. We should also note that no
linear correlation between the specic capacitance and the
specic surface area of ACs was observed. In the next
section, we review other properties of AC electrodes that
impact the overall performance of EDLCs.

Structure-property relationships
In the last decade signicant effort was put out into studies
that reveal critical properties of activated carbon electrodes that may affect the energy and power density as well as
specic energy and power of EDLCs, including the pore size
effect, particle size and electrode thickness effect, carbon
surface structure and wetting, elimination of small bottleneck pores, and surface functional groups.

Pore size effect


Traditionally, a Helmholtz electric double layer (EDL) consisting of solvated ions was used to understand the energy
storage mechanism. Prior to 2006 solvated ions were often
considered not being capable to enter the micropores if their
size exceeded the pore dimensions. Both the removal of the
solvation shells and de-solvated ion insertion into the smallest pores were commonly believed to be negligible in EDLCs.
Thus, initial research on activated carbons was directed
towards increasing the pore volume by developing high SSA
and dening the activation process. However, the capacitance increase was limited even for the most porous samples
[53]. By studying over 30 different activated carbons in KOH
electrolyte, it was reported in 1996 that there was no linear
relationship between the SSA (either BET or DFT) and the
capacitance [86]. Furthermore, increasing SSA with a simultaneous increase in the average pore size was shown not to
increase the specic capacitance either. Another systematic
study investigated the correlation of the pore size and SSA
accessible to N2 gas to that accessible by positive and
negative ions in both aqueous and organic electrolytes [44].
In 2000 the authors concluded that a distortion of solvation
shells is possible and dependent on both the electrolyte ion
and the solvent. In the investigated electrolytes the effective
relative size of ions (with or without full solvation shells) that

559

reached the internal surface area of ACs followed: size of N2


gas moleculeENa + (aq)ECl- (aq)oBF-4oTEA + (tetraethylammonium chloride) (PC)oLi + (PC) and that K + (aq)oNa +
(aq)oLi + (aq) [44]. In spite of this important nding, this
publication did not initially receive a proper attention in the
scientic community. Yet, this preliminary work was instrumental for developing better understanding of the electrosorption of ions in activated carbons and demonstrated that
improvements in EDLC energy storage characteristic depends
critically on the relative dimensions of pores, electrolyte ions
and the solvent molecules as well as on the solvation energy.
In 2006, systematic studies performed by two research
groups in common organic and aqueous electrolytes [7,87]
showed signicant enhancement of the specic capacitance in
small micropores, where the ion solvation shell becomes
highly distorted and partially removed (Figure 8) [7]. The
resulting smaller charge separation distance between the ion
centers and the pore walls leads to greatly increased capacitance [7,8789]. In this case ions do not create layers on the
surface of both walls of the slit-shape pores or the inner
cylinder coating on the surface of the cylinder-shape pores.
Instead, more appropriate models of ion electro-sorption
would constitute formation of an ion monolayer sandwiched
between slit pore walls or the formation of an ion wire inside a
cylindrical pore [89,90]. Yet, when the pore size becomes
smaller than the critical (often unsolvated) ion size, the access
of the ions into the pores becomes limited [7,12] and,
additionally, the transport of ions inside such pores becomes
slow. Slightly larger well-aligned pores are needed for rapid
chargingdischarging [16]. Therefore, a tight control over the
carbon pore size distribution in the micropore range (primarily
in the 0.51 nm range) was suggested to be critically important
for maximizing the energy storage characteristics of EDLCs,
while the co-existence of straight interconnected pores within
individual particles (larger than about 1 nm) was found needed
for improvements in power characteristics of EDLCs.
The generality of the simple model predicting enhanced
ion capacitance in the smallest pores which are still larger
than de-solvated ion size for a broad range of various porous
carbons and electrolytes remains a topic of debates in the
scientic community. Such an enhancement has not been
always observed (Figure 4). We suggest that a more comprehensive physics-based model that would additionally consider
the details of the carbon pore wallssolvent molecules, pore
wallsions and solvention interactions as well as the realistic
shape of carbon pores, the presence functional groups,
defects and impurities (dopants) within the carbon walls
should be developed in order to offer prognostic abilities and
better agreement with various experiments. Better experimental techniques capable of providing direct information
about the ion adsorption sites could be vital for the future
realization of such a model.

Particle size and electrode thickness effect


The power characteristics of EDLCs are commonly limited not
by the electrical resistance of the electrodes but by the
resistance faced by the ions traveling to or from the electroadsorption sites and, to a less extent, by the resistance of the
interface between the current collector foils and the electrodes. Once the interface resistance is minimized [9193],

Author's personal copy


560

L. Wei, G. Yushin

Figure 8 Effect of pore size on the specic capacitance normalized by BET SSA for the carbons in TEATFB-based electrolyte, from
Ref. [7]: (A) The normalized capacitance increases in sub-nm pores; (B to D) schematic illustration of solvated ions in pores with
distance between adjacent pore walls (B) greater than 2 nm, (C) between 1 nm and 2 nm, and (D) less than 1 nm.

improving EDLC power can be achieved by improving the rate


of ion transport within porous carbon [8,10,16,20,9496],
minimizing the ion diffusion distance [97] or by utilizing both
approaches [27,40,98].
Commercial EDLC electrodes consist of micro-sized activated carbon powders with large mesopores or macropores
between the particles. Therefore, while both the electrode
thickness and size of microporous carbon particles are
important for considering the overall ion diffusion time,
increasing the electrode thickness from 20 mm to 200 mm
commonly leads to a moderate decrease in the EDLCs power.
Larger electrode thickness is highly desirable for minimizing
the relative weight, volume and cost of the inactive device
components, such as metal foil current collectors and
separator membrane. Therefore, the majority of commercial
EDLCs have electrode thickness in excess of 100 mm. In order
to minimize the ion diffusion distance within individual
carbon particles, the particle size can be reduced to submicron and even 1030 nm range [97]. However, the use of
porous nanoparticles reduces both the electrode density (and
thus the energy density of the fabricated device) and the size
of pores between the individual particles, which may ultimately lead to high ionic resistance in thick electrodes and
reduced power density.
Compared with AC electrodes, when electrodes are produced not from microporous particles, but from the individual
non-porous nanoparticles of graphene, carbon onions and
CNTs, the electrodes do not normally contain large mesopores
or macropores, and most of the electrode pores are in the
range of 15 nm. In this case the time for the ion diffusion
through the full electrode becomes roughly proportional to
the square of the electrode thickness. As such, very high EDLC
power could only be achieved in relatively thin electrodes

[40,41]. Indeed, most of the relevant studies either do not


report the electrode thickness or minimize it to 25 mm or
below, which produced unfair comparisons to state-of-the art
EDLCs. Unfortunately, the high power density characteristics
of graphene, CNTs and carbon onion electrodes is extremely
challenging to exploit in electrodes with 100300 mm thickness
or mass loading on the order of 1020 mg cm2, which is
desired for the majority of commercial devices. Furthermore,
by forming large interconnected mesopores and eliminating
the micropores within carbon electrodes to achieve higher
power characteristics, the EDLCs volumetric capacitance
becomes greatly reduced. Finally, strictly mesoporous electrodes may require a signicantly increased amount of electrolyte needed to ll the enlarged electrode pore volume. As the
weight of electrolyte becomes higher than the weight of
carbon, not only volumetric but also gravimetric device energy
density may be reduced. As a result, for many applications
currently utilizing commercial EDLC with AC electrodes which
do not require ultra-fast (0.10.001 s) charging time or on-chip
integration with electronic devices, the use of non-porous
carbon nanoparticles is unlikely to offer any advantages.
Similar arguments are also valid for many bulk or thin lm
electrodes having no mesopores [99101]. Such technology
may likely target only specialized applications and be considerably more expensive (on a cost per energy basis) than
commercial EDLC devices.

Carbon surface structure and wetting


The positive effect of electrolyte wetting on the carbon surface has been discussed in multiple publications [102105].
Improved electrolyte wetting behavior commonly improves

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors
the chargedischarge rate and allows one to achieve higher
specic capacitance value. In one study [102], the author
pointed out that without post heat-treating, the prepared
carbon electrodes had not only a hydrophobic nature but also
a weak afnity with KOH electrolytes. However, the post
heat-treating in air atmosphere increased the wettability of
carbon with the electrolyte of KOH aqueous solution, and the
capacitances of electrodes were also enhanced. Furthermore,
defects on surface of carbon were proposed to interact with
the solvation shells of the ions, decreasing the average carbon
surface-ion separation distance, and thereby increasing specic capacitance [20,106]. In Ref. [20], the author observed
the capacitance of carbon electrodes (carbon onions and
carbon nanotubes) steadily decreasing with the increasing of
annealing temperature when normalized per unit area. As the
average pore size and pore size distribution did not change
noticeably, the agglomeration of carbon onions at high
temperatures should not hinder the electrolyte accessibility.
As the electrode electrical conductivity increased, the
observed dependence suggested that the surface graphitization associated to a defect density decrease had a negative
effect on the ion adsorption on the carbon surface under
applied potential. This demonstrates a pronounced effect of
surface defects on capacitance.

Elimination of small bottle-neck pores


Tortuous and bottle-neck pores may dramatically slow down
the ion transport in microporous carbon [16]. By improving
the degree of pore alignment in ZTC the characteristics
charge/discharge time in EDLC was found to increase by
3 orders of magnitude [16].
In order to achieve higher surface areas and eliminate
bottle-neck pores (which may exist in ACs and slow down
transport of ions) while uniformly enlarging the smallest
micropores produced in the course of carbonization of
organic precursors, several promising routes were proposed.
According to one method an equilibrium content of oxygencontaining functional groups are uniformly formed on the
porous carbon surface during room temperature treatment
in acids [107]. These groups together with the adjusted
carbon atoms are later removed via heat treatment at
900 1C. Repetition of the process of uniform formation of

561

chemisorbed oxygen functional groups and subsequently


removing them, allows for the uniform pore broadening
needed to achieve the optimum PSD [107].
In another study, an environmentally friendly low-temperature hydrothermal carbonization was utilized in order to
introduce a network of uniformly distributed oxygen within
the carbon structure in one step [45]. This material, produced from natural precursors (such as wood dust and potato
starch), was then transformed into microporous carbons
(pore size almost o2 nm) (Figure 9a) with high SSA of
21002960 m2 g1 via simultaneous heat-treatment (and thus
uniform removal of the oxygen-containing functional groups
from the internal material surface) and opening of closed and
bottle-neck pores by activation. A very high specic capacitance of 140240 F g1 was demonstrated in a TEATFB-based
organic electrolyte with very good capacity retention at fast
sweep rates (Figure 9b). Data comparison with the performance of commercial activated carbon YP-17D (produced
from coconut shell and KOH activation) showed doubling the
specic capacitance and, thus, doubling the energy density
achieved via utilization of the hydrothermal carbonization
route. The elimination of small bottle-neck pores and a
decrease in the concentration of obstacles for ion diffusion
resulted fast frequency response and ion transport [45].
These ndings not only prove that mesopores are not
required for rapid ion diffusion but also provide guidance
for the optimal design of activated carbons with improved
power storage characteristics.

Surface functional groups


Activated carbons with oxygen-containing functional groups
(such as COOH, =CO, and others) can be formed by physical
and chemical activation of porous carbons in O2, HNO3, KOH,
by plasma treatment or by using electrochemical oxidation
techniques [33,52,59,60,85,108110]. Some of these functional groups were understood to be acidic and electrochemically active, and contribute to the capacitance of EDLCs,
which is called pseudocapacitance generally credited to
Faradaic reaction of these groups with electrolyte ions
[111]. The oxygen-containing surface groups on activated
carbons can be divided into three types, depending on the
nature of the CO bonds: (i) chemically xed groups (e.g.,

Figure 9 (a) The pore size distribution for hydrothermally synthesized carbons after activation, and (b) specic capacitance of the
carbon samples in 1 M TEABF4/AN electrolyte at different CV scan rates in comparison with that of commercially available YP-17D
activated carbon, from Ref. [45].

Author's personal copy


562
carbonyl), which are degassed as CO, only upon heating the
carbon to above 800 1C. These groups are believed to be
electrochemically inactive, and their main effect is a shift of
the point of zero charge (PZC). (ii) Surface groups (e.g.,
carboxylic), which provide surface acidity, and are degassed
at above 400 1C, mostly as CO2. (iii) Surface groups with
electrochemical redox activity such as quinine/hydroquinone
moieties, which can also be degassed between 400800 1C.
Such surface groups are easily recognized by the electrochemical response of carbon electrodes (e.g., a couple of
reversible peaks in cyclic voltammetric measurements). It
should be noted that these redox sites are active commonly
only under acidic conditions [112].
Oxygen species on the surface are commonly disliked for
the carbon materials used in non-aqueous electrolyte solutions, because they were found to be unstable and have
adverse effects on the reliability of capacitors, causing selfdischarge, leakage current, degradation and other negative
contributions [52]. Substantial costs are consumed to remove
oxygen in production of activated carbons for commercial
EDLCs. The formation of stable (yet redox-active in organic
electrolytes) functional groups on carbon has been suggested
by multiple research groups to enhance the energy density of
EDLCs [33,34,60,85]. However, such research efforts have not
yet been very successful and functionalized carbons are not
utilized in commercial devices. The challenge comes from
the difculties involved in the selective functionalization of
carbons, where all electrochemically unstable groups could
be removed from the carbon surface while all the stable ones
retained. Unfortunately, most publications in research journals describing the development of ACs for EDLC use in
organic electrolyte only focus on the maximum specic
capacitance but do not report on their surface chemistry
and its effect on the cycle stability or self-discharge characteristics of these novel materials. Therefore, no comprehensive understanding exists on the specic contributions of
various functional groups to both the desirable and undesirable performance characteristics of EDLCs.
In contrast to the generally accepted need to eliminate
functional groups from carbons used in IL- or organic electrolyte-based EDLCs, many oxygen-containing redox active functional groups increase both the capacitance and high rate
performance in aqueous electrolyte solutions without signicant penalties. The wetting of the hydrophobic carbon surface in aqueous electrolyte solution can be poor. Therefore,
the presence of oxygen-containing polar functional groups
(hydroxyl, carboxyl, carbonyl, quinine and others) in ACs
greatly increase their hydrophilicity and surface area accessible to aqueous electrolytes [59,73,108,113]. In addition to
improved wetting, pseudocapacitance from selected functional groups may contribute to over 22% of the total capacitance [114]. We shall note however, that certain functional
groups may also lead to the leakage and degradation of EDLCs
in strong acidic and basic aqueous electrolytes, similar to
their negative effects in organic electrolytes. In another study
[109], the effect of specic functional groups on electric
double layer characteristics in both the aqueous electrolyte
and organic electrolyte was compared. The author pointed
out in the case of the commonly used acidic aqueous
electrolyte (1 M H2SO4 solution), the capacitance depended
more on the presence of oxygen functional groups than on
BET surface area. Importantly, while both phenolic hydroxyl

L. Wei, G. Yushin
and carboxyl functional groups led to the capacitance increase, the more polar carboxylic groups caused a signicant
leak current due to their catalytic effect on the following
reaction:
2OH 2h -H2 O0:5O2 m

Furthermore, the EDLC with functionalized electrodes


experienced signicant (25%) fading after prolonged (30,000)
chargedischarge cycling (00.8 V). Such experiment also
revealed decrease in the micropore volume in the cycled
electrodes and increase in the number of functional groups
relative to the CC and CH bonds (1.6 times for the negative
electrode and 2.5 for the positive electrode) due to electrode
oxidation. Employment of pH-neutral electrolytes could minimize the effects of some of the degradation mechanisms
involved [115].
In the case of the organic electrolyte (0.5 M LiClO4/PC
solution), pore structure was found to be a more dominant
factor than functional groups present. While in some electrodes, increase in the acidity (number of functional groups)
caused slight increase in the specic capacitance, other electrodes experienced up to 50% decrease. This study further
emphasizes the importance of gaining more comprehensive
understanding of the complex interactions between the ions
and solvent molecules of electrolyte with carbon surface
functionalities and defects and how the pore size and shapes
may affect these interactions. Such knowledge will assist in
the development of stable high-energy, high-power, lowleakage EDLCs based on electrodes produced from low-cost
renewable precursors.

Summary and outlook


Activated carbons produced by different activation processes
from various precursors are the most widely used electrodes
for commercial EDLCs. With the development of sustainable
and renewable energy sources and the advancements of AC
synthesis techniques, AC materials have progressed tremendously. ACs with better controlled PSD and surface chemistry
have demonstrated greatly enhanced capacitive characteristics in different electrolytes. Therefore, in spite of the
development of other porous carbon synthesis technologies,
ACs will likely remain the material of choice for commercial
EDLC applications in the near future.
The low cost and high volumetric energy density are
currently the most critical parameters describing the value
proposition of EDLCs to their customers and users. Therefore,
further development of dense and inexpensive carbon materials with high capacitance per unit volume is critically
needed [51]. The continuing study of microporous activated
carbon materials from low-cost natural precursors should be
aimed to increase volumetric capacitance without sacricing
stability and negatively affecting the leakage currents in
EDLCs based on high voltage electrolytes. Furthermore, the
ability to tailor pore sizes and minimize the pore tortuosity in
the micropores to facilitate fast ion-transport kinetics will be
important for fabrication of EDLCs with improved power
storage characteristics. Gaining better fundamental understanding of the ion transport and adsorption phenomena and
the effects of various carbon defects and functional groups
on both the capacitance and undesirable electrochemical

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors
side reactions will likely be critical for the successful
development of improved AC technology. The advancement
of these research topics will be instrumental for the extensive applications of EDLCs in the smart grid and transportation elds.

Acknowledgement
This work was partially supported by a gift from a Semiconductor Research Corporation (SRC) and ACS Petroleum
Research Fund (grant # 49045-DNI10).

References
[1] L. Wei, M. Sevilla, A.B. Fuertes, R. Mokaya, G. Yushin,
Advanced Functional Materials 22 (2011) 827834.
[2] B. Daffos, P.L. Taberna, Y. Gogotsi, P. Simon, Fuel Cells 10
(2010) 819824.
[3] B.E. Conway, Electrochemical Supercapacitors: Scientic
Fundamentals and Technological Applications, Kluwer Academic/Plenum Publishers, New York, USA, 1999.
[4] E. Pollak, N. Levy, L. Eliad, G. Salitra, A. Soffer, D. Aurbach,
Israel, Journal of Chemistry 48 (2008) 287303.
[5] H. Yang, M. Yoshio, K. Isono, R. Kuramoto, Electrochemical
and Solid-State Letters 5 (2002) A141A144.
[6] G. Yushin, R. Dash, J. Jagiello, J.E. Fischer, Y. Gogotsi,
Advanced Functional Materials 16 (2006) 22882293.
[7] J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon,
P.L. Taberna, Science 313 (2006) 17601763.
[8] Y. Korenblit, M. Rose, E. Kockrick, L. Borchardt, A. Kvit,
S. Kaskel, G. Yushin, ACS Nano 4 (2010) 13371344.
[9] J. Eskusson, A. Janes, A. Kikas, L. Matisen, E. Lust, Journal of
Power Sources 196 (2011) 41094116.
[10] M. Rose, Y. Korenblit, E. Kockrick, L. Borchardt, M. Oschatz,
S. Kaskel, G. Yushin, Small 7 (2011) 11081117.
[11] R. Dash, J. Chmiola, G. Yushin, Y. Gogotsi, G. Laudisio, J. Singer,
J. Fischer, S. Kucheyev, Carbon 44 (2006) 24892497.
[12] J. Chmiola, G. Yushin, R. Dash, Y. Gogotsi, Journal of Power
Sources 158 (2006) 765772.
[13] E.N. Hoffman, G. Yushin, B.G. Wendler, M.W. Barsoum,
Y. Gogotsi, Materials Chemistry and Physics 112 (2008) 587591.
[14] G. Yushin, E.N. Hoffman, M.W. Barsoum, Y. Gogotsi,
C.A. Howell, S.R. Sandeman, G.J. Phlllips, A.W. Lloyd,
S.V. Mikhalovsky, Biomaterials 27 (2006) 57555762.
[15] M. Oschatz, E. Kockrick, M. Rose, L. Borchardt, N. Klein,
I. Senkovska, T. Freudenberg, Y. Korenblit, G. Yushin,
S. Kaskel, Carbon 48 (2010) 39873992.
[16] A. Kajdos, A. Kvit, F. Jones, J. Jagiello, G. Yushin, Journal of
the American Chemical Society 132 (2010) 32523253.
[17] C. Portet, Z. Yang, Y. Korenblit, Y. Gogotsi, R. Mokaya, G. Yushin,
Journal of the Electrochemical Society 156 (2009) A1A6.
[18] C.O. Ania, V. Khomenko, E. Raymundo-Pin
ero, J.B. Parra,

F. Beguin,
Advanced Functional Materials 17 (2007) 18281836.
[19] H. Nishihara, H. Itoi, T. Kogure, P.-X. Hou, H. Touhara,
F. Okino, T. Kyotani, Chemistry A European Journal 15
(2009) 53555363.
[20] C. Portet, G. Yushin, Y. Gogotsi, Carbon 45 (2007) 25112518.
[21] E. Frackowiak, F. Be guin, Carbon 39 (2011) 937950.
[22] G. Lota, K. Fic, E. Frackowiak, Energy & Environmental
Science 4 (2011) 15921605.
[23] B.Q. Wei, Abstracts of PapersAmerican Chemical Society
241 (2011).
[24] A. Izadi-Najafabadi, T. Yamada, D.N. Futaba, H. Hatori, S. Iijima,
K. Hata, Electrochemistry Communications 12 (2010) 16781681.
[25] H. Zhang, G. Cao, Y. Yang, Energy & Environmental Science
2 (2009) 932943.

563

[26] V.V.N. Obreja, Physics E (Low-Dimensional Systems & Nanostructures) 40 (2008) 25962605.
[27] D. Pech, M. Brunet, H. Durou, P.H. Huang, V. Mochalin,
Y. Gogotsi, P.L. Taberna, P. Simon, Nature Nanotechnology
5 (2010) 651654.
[28] E.G. Bushueva, P.S. Galkin, A.V. Okotrub, L.G. Bulusheva,
N.N. Gavrilov, V.L. Kuznetsov, S.I. Moiseekov, Physica Status
Solidi B Basic Solid State Physics 245 (2008) 22962299.
[29] X. Wang, L. Liu, X. Wang, L. Bai, H. Wu, X. Zhang, L. Yi,
Q. Chen, Journal of Solid State Electrochemistry 15 (2011)
643648.
[30] B.B. Garcia, S.L. Candelaria, D.W. Liu, S. Sepheri, J.A. Cruz,
G.Z. Cao, Renewable Energy 36 (2011) 17881794.
[31] A. Halama, B. Szubzda, G. Pasciak, Electrochimica Acta 55
(2010) 75017505.
[32] K. Evanoff, J. Khan, A.A. Balandin, A. Magasinski, W.J. Ready,
T.F. Fuller, G. Yushin, Advanced Materials 24 (2011) 533537.
[33] G. Lota, J. Tyczkowski, R. Kapica, K. Lota, E. Frackowiak,
Journal of Power Sources 195 (2010) 75357539.
[34] W. Lu, L. Qu, K. Henry, L. Dai, Journal of Power Sources 189
(2009) 12701277.
[35] B. Xu, S.F. Yue, Z.Y. Sui, X.T. Zhang, S.S. Hou, G.P. Cao,
Y.S. Yang, Energy & Environmental Science 4 (2011)
28262830.
[36] Y. Sun, Q. Wu, G. Shi, Energy & Environmental Science 4
(2011) 11131132.
[37] M. Pumera, Energy & Environmental Science 4 (2011) 668674.
[38] L.L. Zhang, R. Zhou, X.S. Zhao, Journal of Materials Chemistry 20 (2010) 59835992.
[39] J.J. Yoo, K. Balakrishnan, J.S. Huang, V. Meunier, B.G. Sumpter,
A. Srivastava, M. Conway, A.L.M. Reddy, J. Yu, R. Vajtai,
P.M. Ajayan, Nano Letters 11 (2011) 14231427.
[40] J.R. Miller, R.A. Outlaw, B.C. Holloway, Science 329 (2010)
16371639.
[41] Y.W. Zhu, S. Murali, M.D. Stoller, K.J. Ganesh, W.W. Cai,
P.J. Ferreira, A. Pirkle, R.M. Wallace, K.A. Cychosz,
M. Thommes, D. Su, E.A. Stach, R.S. Ruoff, Science 332
(2011) 15371541.
[42] E. Raymundo-Pinero, K. Kierzek, J. Machnikowski, F. Beguin,
Carbon 44 (2006) 24982507.
[43] J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon, Science
313 (2006) 17601763.
[44] G. Salitra, A. Soffer, L. Eliad, Y. Cohen, D. Aurbach, Journal
of the Electrochemical Society 147 (2000) 24862493.
[45] L. Wei, M. Sevilla, A.B. Fuertes, R. Mokaya, G. Yushin,
Advanced Energy Materials 1 (2011) 356361.
[46] D.O. Cooney, Activated Charcoal: Antidotal and Other Medical Uses, M. Dekker, New York, USA, 1980.
[47] F.S. Baker, C.E. Miller, A.J. Repik, Kirk-Othmer Encyclopedia
of Chemical Technology, John Wiley & Sons, New York, USA,
1992.
[48] G.G. Stavropoulos, A.A. Zabaniotou, Fuel Processing Technology 90 (2009) 952957.
[49] L. Wei, G. Yushin, Carbon 49 (2011) 48304838.
[50] L.L. Zhang, X.S. Zhao, Chemical Society Reviews 38 (2009)
25202531.
[51] Y.P. Zhai, Y.Q. Dou, D.Y. Zhao, P.F. Fulvio, R.T. Mayes, S. Dai,
Advanced Materials 23 (2011) 48284850.
[52] M. Inagaki, H. Konno, O. Tanaike, Journal of Power Sources
195 (2010) 78807903.
[53] P. Simon, Y. Gogotsi, Nature Materials 7 (2008) 845854.
[54] M. Lazzari, F. Soavi, M. Mastragostino, Journal of the
Electrochemical Society 156 (2009) A661A666.
[55] C. Largeot, C. Portet, J. Chmiola, P.-L. Taberna, Y. Gogotsi,
P. Simon, Journal of the American Chemical Society 130
(2008) 27302731.
[56] T. Sato, G. Masuda, K. Takagi, Electrochimica Acta 49 (2004)
36033611.

Author's personal copy


564
[57] A. Balducci, R. Dugasa, P.L. Tabernaa, P. Simona, D. Ple eb,
M. Mastragostinoc, S. Passerinid, Journal of Power Sources
165 (2007) 922927.
[58] A. Balducci, W.A. Henderson, M. Mastragostino, S. Passerini,
P. Simon, F. Soavi, Electrochimica Acta 50 (2005) 22332237.
[59] M.J. Bleda-Martinez, J.A. Macia-Agullo, D. Lozano-Castello,
E. Morallon, D. Cazorla-Amoros, A. Linares-Solano, Carbon 43
(2005) 26772684.
[60] C.T. Hsieh, H. Teng, Carbon 40 (2002) 667674.
[61] K. Kierzek, E. Frackowiak, G. Lota, G. Gryglewicz,
J. Machnikowski, Electrochimica Acta 49 (2004) 11691170.
[62] D. Lozano-Castello, D. Cazorla-Amoros, A. Linares-Solano,
S. Shiraishi, H. Kurihara, A. Oya, Carbon 41 (2003) 17651775.
[63] D. Zhai, B. Li, H. Du, G. Wang, F. Kang, Journal of Solid State
Electrochemistry 15 (2011) 787794.
[64] R. Mysyk, V. Ruiz, E. Raymundo-Pinero, R. Santamaria,
F. Beguin, Fuel Cells 10 (2010) 834839.
[65] G.P. Pandey, S.A. Hashmi, Y. Kumar, Energy Fuels 24 (2010)
66446652.
[66] K.K. Denshchikov, M.Y. Izmaylova, A.Z. Zhuk, Y.S. Vygodskii,
V.T. Novikov, A.F. Gerasimov, Electrochimica Acta 55 (2010)
75067510.
[67] P.W. Zhou, B.H. Li, F.Y. Kang, Y.Q. Zeng, New Carbon
Materials 21 (2006) 125131.
[68] F.C. Wu, R.L. Tseng, C.C. Hu, C.C. Wang, Journal of Power
Sources 138 (2004) 351359.
[69] Y.J. Kim, B.J. Lee, H. Suezaki, T. Chino, Y. Abe, T. Yanagiura,
K.C. Park, M. Endo, Carbon 44 (2006) 15921595.
[70] S. Zhao, C.Y. Wang, M.M. Chen, J. Wang, Z.Q. Shi, Journal of
Physics and Chemistry of Solids 70 (2009) 12561260.
[71] Q.Y. Li, H.Q. Wang, Q.F. Dai, J.H. Yang, Y.L. Zhong, Solid
State Ionics 179 (2008) 269273.
[72] L. Wei, G. Yushin, Journal of Power Sources 196 (2011)
40724079.
[73] S.G. Lee, K.H. Park, W.G. Shim, M.S. Balathanigaimani,
H. Moon, Journal of Industrial and Engineering Chemistry 17
(2011) 450454.
[74] V. Subramanian, C. Luo, A.M. Stephan, K.S. Nahm,
S. Thomas, B. Wei, Journal of Physical Chemistry C 111
(2007) 75277531.
[75] M.S. Balathanigaimani, W.G. Shim, M.J. Lee, C. Kim,
J.W. Lee, H. Moon, Electrochemistry Communications 10
(2008) 868871.
[76] T.E. Rufford, D. Hulicova-Jurcakova, K. Khosla, Z.H. Zhu,
G.Q. Lu, Journal of Power Sources 195 (2010) 912918.
[77] B. Xu, Y.F. Chen, G. Wei, G.P. Cao, H. Zhang, Y.S. Yang,
Materials Chemistry and Physics 124 (2010) 504509.
[78] X.A. Li, W. Xing, S.P. Zhuo, J. Zhou, F. Li, S.Z. Qiao, G.Q. Lu,
Bioresource Technology 102 (2011) 11181123.
[79] T.E. Rufford, D. Hulicova-Jurcakova, Z.H. Zhu, G.Q. Lu,
Electrochemistry Communications 10 (2008) 15941597.
[80] T.E. Rufford, D. Hulicova-Jurcakova, E. Fiset, Z.H. Zhu, G.Q. Lu,
Electrochemistry Communications 11 (2009) 974977.
[81] X. Li, C. Han, X. Chen, C. Shi, Microporous and Mesoporous
Materials 131 (2010) 303309.
[82] W.X. Chen, H. Zhang, Y.Q. Huang, W.K. Wang, Journal of
Materials Chemistry 20 (2010) 47734775.
[83] M. Olivares-Marin, J.A. Fernandez, M.J. Lazaro, C. FernandezGonzalez, A. Macias-Garcia, V. Gomez-Serrano, F. Stoeckli,
T.A. Centeno, Materials Chemistry and Physics 114 (2009)
323327.
[84] Y.P. Guo, J.R. Qi, Y.Q. Jiang, S.F. Yang, Z.C. Wang, H.D. Xu,
Materials Chemistry and Physics 80 (2003) 704709.

L. Wei, G. Yushin
[85] K. Okajima, K. Ohta, M. Sudoh, Electrochimica Acta 50
(2005) 22272231.
[86] H. Shi, Electrochimica Acta 41 (1996) 16331639.
[87] E. Raymundo-Pin
ero, K. Kierzek, J. Machnikowski, F. Be guin,
Carbon 44 (2006) 24982507.
[88] G. Feng, R. Qiao, J.S. Huang, B.G. Sumpter, V. Meunier, ACS
Nano 4 (2010) 23822390.
[89] J.S. Huang, B.G. Sumpter, V. Meunier, Angewandte Chemie
International Edition 47 (2008) 520524.
[90] J.S. Huang, B.G. Sumpter, V. Meunier, Chemistry A European
Journal 14 (2008) 66146626.
[91] P.L. Taberna, P. Simon, J.F. Fauvarque, Journal of the
Electrochemical Society 150 (2003) A292A300.
[92] C. Portet, P.L. Taberna, P. Simon, E. Flahaut, Journal of the
Electrochemical Society 153 (2006) A649A653.
[93] C. Portet, P.L. Taberna, P. Simon, C. Laberty-Robert, Electrochimica Acta 49 (2004) 905912.
[94] C.G. Liu, Z.N. Yu, D. Neff, A. Zhamu, B.Z. Jang, Nano Letters
10 (2010) 48634868.
[95] J. Li, X.Y. Wang, Q.H. Huang, S. Gamboa, P.J. Sebastian,
Journal of Power Sources 158 (2006) 784788.
[96] B.Z. Fang, L. Binder, Journal of Power Sources 163 (2006)
616622.
[97] C. Portet, G. Yushin, Y. Gogotsi, Journal of the Electrochemical Society 155 (2008) A531A536.
[98] C.S. Du, J. Yeh, N. Pan, Nanotechnology 16 (2005) 350353.
[99] M. Heon, S. Loand, J. Applegate, R. Nolte, E. Cortes,
J.D. Hettinger, P.L. Taberna, P. Simon, P.H. Huang, M. Brunet,
Y. Gogotsi, Energy & Environmental Science 4 (2011)
135138.
[100] J. Chmiola, C. Largeot, P.L. Taberna, P. Simon, Y. Gogotsi,
Science 328 (2010) 480483.
[101] V. Presser, L.F. Zhang, J.J. Niu, J. McDonough, C. Perez, H. Fong,
Y. Gogotsi, Advanced Energy Materials (2011) 423430.
[102] S.J. Kim, S.W. Hwang, S.H. Hyun, Journal of Materials Science
40 (2005) 725731.
[103] S.W. Hwang, S.H. Hyun, Journal of Non-Crystalline Solids 347
(2004) 238245.
[104] A.G. Pandolfo, A.F. Hollenkamp, Journal of Power Sources 157
(2006) 1127.
[105] I. Bispo-Fonseca, J. Aggar, C. Sarrazin, P. Simon,
J.F. Fauvarque, Journal of Power Sources 79 (1999) 238241.
[106] K.G. Gallagher, G. Yushin, T.F. Fuller, Journal of the Electrochemical Society 157 (2010) B820B830.
[107] L. Eliad, E. Pollak, N. Levy, G. Salitra, A. Soffer, D. Aurbach,
Applied Physics A (Materials Science & Processing) 82 (2006)
607613.
[108] V. Ruiz, C. Blanco, E. Raymundo-Pinero, V. Khomenko,
F. Beguin, R. Santamaria, Electrochimica Acta 52 (2007)
49694973.
[109] H. Oda, A. Yamashita, S. Minoura, M. Okamoto, T. Morimoto,
Journal of Power Sources 158 (2006) 15101516.
[110] C.C. Hu, C.C. Wang, Journal of Power Sources 125 (2004)
299308.
[111] B.E. Conway, V. Birss, J. Wojtowicz, Journal of Power Sources
66 (1997) 114.
[112] M. Noked, A. Soffer, D. Aurbach, Journal of Solid State
Electrochemistry 15 (2011) 15631578.
[113] M.J. Bleda-Martinez, D. Lozano-Castello, E. Morallon, D. CazorlaAmoros, A. Linares-Solano, Carbon 44 (2006) 26422651.
[114] D.Y. Qu, Journal of Power Sources 109 (2002) 403411.
[115] K. Fic, G. Lota, M. Meller, E. Frackowiak, Energy & Environmental Science 5 (2012) 58425850.

Author's personal copy


Nanostructured activated carbons from natural precursors for electrical double layer capacitors
Lu Wei is a visiting Ph.D. student of the
School of Materials Science and Engineering
(MSE) at Georgia Institute of Technology
(USA) with Prof. Gleb Yushin as her advisor,
effective Sept. 2008. She received a B.S. in
MSE from Zhengzhou University (China) in
Jun. 2005. Upon graduation she joined
Northwestern
Polytechnical
University
(NWPU, China) as a graduate student in
the National Key Laboratory of Thermostructural Composite Materials. At NWPU,
Lu worked on the structural design and preparation techniques of
high temperature oxidation protective coatings for carbon/carbon
composites. At Georgia Institute of Technology, she has been
engaged in the synthesis of innovative nanostructured materials
for energy storage applications, such as supercapacitors and Libatteries. Her research interests are focused on the investigation of
low cost nanoporous carbon electrodes from environmentally
friendly precursors and the manufacturing of on-chip micro-supercapacitors with precisely dened dimensions. Lus recent work has
been published in Advanced Functional Materials, Advanced Energy
Materials, Carbon, and the Journal of Power Sources, etc.

565

Gleb Yushin is an Associate Professor in the


School of Materials Science & Engineering
and Director of the Center for Nanostructured Materials for Energy Storage at
Georgia Institute of Technology. The current
research activities of his laboratory are
focused on electrochemical energy storage
technologies with emphasize on nanostructured and nanocomposite materials for use
in advanced lithium-ion batteries and
supercapacitors. For his contributions to
these and other areas, he has received numerous awards and
recognition, including the R&D 100 Award, NSF Faculty Early Career
Development Award, US AFOSR Young Investigator Award, Honda
Initiation Award, Petroleum Research Fund Doctoral New
Investigator Award, NASA Nano 50 Award and Roland B. Snow
Award from the American Ceramic Society, Micromeritics Instrument Award, Creating and Energy Options Award from the Strategic
Energy Institute. Prof. Yushin received a Ph.D. degree in Materials
Science from North Carolina State University, USA, and an M.S.
degree in Physics from Saint-Petersburg Technical University,
Russia.

Вам также может понравиться