22.02.2015 Views

Downloaded

Downloaded

Downloaded

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

www.rsc.org/materials Volume 22 | Number 41 | 7 November 2012 | Pages 21781–22310<br />

ISSN 0959-9428<br />

COMMUNICATION<br />

Jieshan Qiu et al.<br />

Hierarchical activated carbon nanofiber webs with tuned structure fabricated by<br />

electrospinning for capacitive deionization<br />

0959-9428(2012)22:41;1-C


Journal of<br />

Materials Chemistry<br />

Dynamic Article Links C <<br />

Cite this: J. Mater. Chem., 2012, 22, 21819<br />

www.rsc.org/materials<br />

COMMUNICATION<br />

Hierarchical activated carbon nanofiber webs with tuned structure fabricated<br />

by electrospinning for capacitive deionization<br />

<strong>Downloaded</strong> on 06 October 2012<br />

Published on 10 September 2012 on http://pubs.rsc.org | doi:10.1039/C2JM34890J<br />

Gang Wang, a Qiang Dong, a Zheng Ling, a Chao Pan, ab Chang Yu a and Jieshan Qiu* a<br />

Received 24th July 2012, Accepted 28th August 2012<br />

DOI: 10.1039/c2jm34890j<br />

Novel hierarchical activated carbon nanofiber (ACF) webs with<br />

tuned structure have been fabricated by incorporating carbon black<br />

(CB) into an electrospun polymer solution, followed by heat treatment.<br />

The as-made electrospun ACF webs show superior capacities<br />

as electrode materials in capacitive deionization (CDI) for desalination<br />

due to their advantageous hierarchical structures.<br />

The consumption and sustainable supply of fresh water have become<br />

one of the top priority issues faced by the global community. 1 In the<br />

past few years, the capacitive deionization (CDI) technology for<br />

desalination of salt water has drawn much attention because of its<br />

potential as an energy-efficient alternative to membrane desalination<br />

and thermal processes currently available for producing fresh water<br />

from salted water sources. 2–10 For the CDI technology, electrode<br />

materials with tuned pore structure and functions, such as conductivity,<br />

are the key to an efficient desalination process. 11 Up to now, a<br />

number of porous carbons with different forms and textures have<br />

been tested as electrosorptive electrodes. 12–15 Unfortunately, the<br />

desalination performance of the electrodes made of conventional<br />

porous carbons such as microporous carbon particles or powders is<br />

far from satisfactory because some micropores in the electrodes are<br />

not accessible to ions. 16 As such, hierarchical porous carbons with<br />

well-defined pore structures are highly desirable, 17,18 in which<br />

micropores, mesopores and macropores are connected well in a<br />

balanced way. The micropores would help lead to an enhanced<br />

electrical double-layer capacitance, the mesopores would provide<br />

low-resistance pathways and channels for the ions through the<br />

porous fibers, while the macropores would function as ion-buffering<br />

reservoirs to minimize the diffusion distances to the interior surfaces<br />

of the porous carbons. 19,20 With this in mind, several kinds of new<br />

carbon materials such as carbon aerogel, 21 carbon nanotubes, 22 graphene<br />

23 and ordered mesoporous carbon 24 have been tested as the<br />

electrode materials for the CDI process. Of the tested carbon materials,<br />

carbon aerogels with a monolithic structure are found to be one<br />

of the promising electrode materials for CDI systems because they<br />

have controllable mesopore size distribution and better solid phase<br />

a<br />

Carbon Research Laboratory, Liaoning Key Lab for Energy Materials<br />

and Chemical Engineering, State Key Lab of Fine Chemicals, Dalian<br />

University of Technology, Dalian 116024, China. E-mail: jqiu@dlut.edu.<br />

cn; Fax: +86 411 84986080; Tel: +86 411 84986024<br />

b<br />

College of Science, Dalian Ocean University, Dalian 116023, China<br />

conductivity, but some bottle-neck problems such as high production<br />

cost and poor mechanical properties limit the carbon aerogels for<br />

practical use as an ideal electrode material in CDI. 25<br />

Electrospinning is a simple yet effective technique that is capable of<br />

fabricating continuous nanofiber webs. 26 In the literature, there are<br />

several reports on the synthesis of self-standing thin webs consisting<br />

of porous carbon nanofibers for energy storage applications, 27,28 but<br />

to our best knowledge, little has been done about the potential of<br />

activated carbon nanofiber (ACF) webs made by electrospinning for<br />

CDI. The concerned ACF webs featuring high porosity and freestanding<br />

nanofibers are expected to be an excellent alternative as<br />

electrode materials in CDI. This has been partly demonstrated<br />

recently by our work that electrospun ACF webs with relatively<br />

developed microporous structure show a high desalination capacity<br />

as CDI electrodes. 29 Nevertheless, there is still room for further<br />

enhancing their desalination capacity, which can be done by further<br />

optimizing the porous structure. Here we report a new strategy for<br />

fabricating hierarchical ACF webs with tuned structure for CDI by<br />

electrospinning, in which carbon black (CB) was incorporated or<br />

embedded in situ into the electrospun ACF webs that were subsequently<br />

activated in flowing CO 2 to tailor the hierarchical pore size<br />

distribution. This helps greatly improve the conductivity and the saltremoval<br />

capacity of the ACF webs that are free-standing and have a<br />

monolithic hierarchical structure with a well-developed yet balanced<br />

micro-, meso- and macro-porosity. In this novel strategy reported<br />

here, no binders are used in contrast to the traditional methods. This<br />

makes it possible to fabricate electrodes with an improved saltremoval<br />

capacity that is much better than other carbon electrodes for<br />

CDI reported in literature. 30,31<br />

The as-made ACF/CB900 webs have a smooth surface (see<br />

Fig. 1a), and are robust and flexible yet free-standing (see Fig. 1b).<br />

Because of this, they can be folded like non-woven cloth, and can<br />

be rolled up easily into a scroll. The continuous skeletons of the<br />

ACF/CB900 webs help reduce the internal resistance of the electrodes,<br />

11 and result in a large electrosorption capacity. The SEM<br />

image in Fig. 1c shows a composite web with 5 wt% CB loading, in<br />

which some small nanoparticles of agglomerated CB can be seen, as<br />

marked in Fig. 1c in the white circles. It seems that the fibers in the<br />

ACF/CB900 webs (Fig. 1d) have a morphology differing from those<br />

without any CB (Fig. 1e and f), and the agglomerated CB would<br />

cause some changes of the smooth cylindrical shape of the fibers to<br />

some degree. In the electrospinning step, the CB particles are oriented<br />

along the electrospinning needle created ‘streamlines’ in the PAN<br />

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem., 2012, 22, 21819–21823 | 21819


<strong>Downloaded</strong> on 06 October 2012<br />

Published on 10 September 2012 on http://pubs.rsc.org | doi:10.1039/C2JM34890J<br />

Fig. 1 (a) and (b) optical images of ACF/CB900 webs; (c) and (d) SEM images of ACF/CB900 surface; (e) and (f) SEM images of ACF900 surface; (g)<br />

TEM image of ACF/CB900 webs, where small nanoparticles of CB can be seen, as shown in white circles in (c).<br />

solution due to the elongation effect of the fluid jet, and the<br />

agglomerated CB particles are visually parallel to the long-axis of<br />

individual fibers, as evidenced in Fig. 1d. A typical TEM image<br />

(Fig. 1g) shows clearly that the CB particles are embedded in the<br />

PAN-based nanofiber matrix and some pores are formed both in the<br />

nanofibers and in the junction between the nanofiber and CB that<br />

possibly results from the CO 2 activation.<br />

It has been found that the electrical conductivity of the composite<br />

webs can be significantly improved by the embedded CB. The electrical<br />

resistance for the ACF900 webs without CB is 4.61 U cm on<br />

average, while for ACF/CB900 webs with CB, it is only 1.65 U cm.<br />

This significant reduction in the electrical resistance of ACF/CB900<br />

webs is obviously due to the embedded CB in the ACF. Because of<br />

this, the electrical conductivity of ACF/CB900 webs increases by 3<br />

times in comparison to the pristine ACF900 webs without any CB.<br />

This is beneficial for the formation of an electrical double-layer for<br />

charge storage that finally leads to a good capacitive deionization<br />

performance. 22,32<br />

Fig. 2a shows the N 2 adsorption–desorption isotherms of ACF900<br />

and ACF/CB900 webs. According to the IUPAC classification, the<br />

adsorption isotherms of the ACF900 webs are typical type I, indicative<br />

of the dominance of micropores in the porous structure. For the<br />

ACF900 web, its BET surface area is 656 m 2 g 1 (Fig. 2a inset) and its<br />

average pore diameter is 2.3 nm. While for the ACF/CB900 web with<br />

a hysteresis loop at a relative pressure P/P 0 ¼ 0.4–1.0, its surface area<br />

is 428 m 2 g 1 and its average pore diameter is 3.3 nm. The nitrogen<br />

adsorption curves feature typical combined characteristics of type I/II<br />

behaviours due to the capillary condensation in the mesopores and<br />

macropores. 33 The PSDs of ACF900 and ACF/CB900 webs calculated<br />

by dislocalized density function theory (DLDFT) method are<br />

shown in Fig. 2b. Obviously, in comparison with the pristine ACF<br />

with a mesopore ratio of 8%, the ACF/CB900 has a higher mesopore<br />

ratio of 40%, implying that some micropores have merged and<br />

transformed into mesopores with a pore size smaller than 5 nm. For<br />

this phenomenon, the possible reason is below. 34 Firstly, CB used in<br />

the present work has a higher mesopore and macropore volume.<br />

Secondly, more new mesopores are created between the CB and the<br />

PAN matrix due to phase separation and poor contacts, and thirdly,<br />

the incorporation of CB within the ACF leads to some larger pores<br />

due to the difference in shrinkage between the PAN and CB in the<br />

CO 2 burn-off step. The macropores in the ACF/CB900 sample are<br />

actually a combined contribution both from the pores in CB and<br />

from the interconnected open channels between the nanofibers<br />

interconnected in the 3D architecture. Because of these combined<br />

effects, both the mesopore–macropore ratio and the average pore size<br />

of ACF/CB900 webs are higher than the pristine ACF webs, as can<br />

be seen in Fig. 2b inset. The results have evidenced that the<br />

Fig. 2 (a) Nitrogen adsorption–desorption isotherms, inset is the BET,<br />

mesopore volume and mesopore ratio of the corresponding samples; (b)<br />

the micropore and meso-macro-pore size distributions (inset) of ACF900<br />

and ACF/CB900 webs.<br />

21820 | J. Mater. Chem., 2012, 22, 21819–21823 This journal is ª The Royal Society of Chemistry 2012


<strong>Downloaded</strong> on 06 October 2012<br />

Published on 10 September 2012 on http://pubs.rsc.org | doi:10.1039/C2JM34890J<br />

incorporation of CB in the electrospinning process can result in a<br />

well-developed hierarchical pore structure that leads to lower resistance<br />

and shorter diffusion pathways of species in the CDI process.<br />

The as-made webs were tested as the CDI electrode in a batch-type<br />

electrosorptive setup operated at 1.6 V. The CDI cell was constructed<br />

by directly attaching the webs to a graphite current collector. The<br />

CDI cell was first washed with deionized water until the solution<br />

conductivity reached the pure water level. Fig. 3a shows the<br />

adsorptive capacity of electrodes made of ACF/CB900 web and<br />

activated carbon powder (AC). It can be seen that for the ACF/<br />

CB900 electrode, the desalination capacity is much higher than the<br />

AC electrode, and increases more drastically in the initial stage, and<br />

after 60 min, it gradually increases until equilibrium is reached. The<br />

web-like ACF electrode has more benefits than the electrode made of<br />

traditional powder-like AC that must be mixed and combined with a<br />

polymer binder, and the binder would block some of the pores in the<br />

activated carbon material thus increasing the internal resistance,<br />

which subsequently results in a lower adsorption capacity. 12,35,36<br />

To clarify the effect of CB on the desalination of the as-made ACF<br />

electrodes, the desalination capacities of the ACF and the ACF/CB<br />

web electrodes activated at different temperatures are compared, as<br />

shown in Fig. 3b. The ACF/CB electrode shows a much higher<br />

electrosorption capacity than the ACF electrode without CB. For the<br />

ACFs activated at 700 C, the electrosorption capacity of the ACF/<br />

CB electrode is 1.99 mg g 1 , while in the case of the ACF electrode, it<br />

is only 0.92 mg g 1 . The electrosorption capacity of the electrode with<br />

CB activated at 700 C is 116% higher than the ACF electrode<br />

without CB. As the activation temperature increases from 700 Cto<br />

Fig. 4 Electrosorption and regeneration cycles of the ACF/CB900<br />

electrode (left), and applied voltage curves vs. time (right), for which the<br />

test conditions are V NaCl ¼ 50 mL, C 0 ¼ 90 mg L 1 , and M ¼ 0.224 g.<br />

900 C, the electrosorption capacity for all of the electrodes increases<br />

continuously. An electrosorption capacity of 9.13 mg g 1 is achieved<br />

for the ACF/CB900 electrode, which is 90% higher than the electrosorption<br />

capacity of the ACF900 electrode, which is only 4.8 mg<br />

g 1 . In general, the electrosorption capacity of the ACF/CB electrodes<br />

increases by 84–116% in comparison to the ACF electrode<br />

without CB, depending on the activation temperature. The excellent<br />

CDI performance of the ACF/CB electrodes can be attributed to the<br />

hierarchical pore size distribution resulting from the incorporation<br />

and embedment of CB in the electrospinning process.<br />

One of the parameters of key concern for CDI cells is their lifelong<br />

performance. 37 To evaluate the reversibility of the electrosorption<br />

capacity of the CDI cells, tests for cyclic operation and regeneration<br />

were performed, in which the solution with an initial conductivity of<br />

197 mS cm 1 was used. It should be noted that the reversed voltage<br />

of 1.6 V was only applied for 5 min at the beginning of the<br />

regeneration cycle, and the flow rate was increased from 5 mL min 1<br />

to 60 mL min 1 , and this shortens the duration of the ion-release step<br />

from the electrode. However, according to ref. 9, the duration of the<br />

regeneration step is generally equal to the duration of the electrosorption<br />

step if the reversed applied voltage is kept at a constant value<br />

of 0 V. In our case, the adsorption time was 150 min while the<br />

regeneration time used was 100 min, and this is more advantageous in<br />

practice. Fig. 4 shows the variation of the solution conductivity over 4<br />

cycles of adsorption and regeneration, which has evidenced that the<br />

reversibility of the CDI cells is excellent, i.e. the high performance<br />

electrosorption process can be repeated satisfactorily for the CDI cells<br />

Fig. 3 (a) Electrosorption capacities of ACF/CB900 web and AC electrodes<br />

as a function of time in CDI; and (b) the desalination capacity of<br />

ACF and ACF/CB web electrodes activated at different temperatures, for<br />

which the test conditions are V cell ¼ 1.6 V, V NaCl ¼ 50 mL, and C 0 ¼<br />

90 mg L 1 .<br />

Fig. 5<br />

Schematic of the CDI set-up.<br />

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem., 2012, 22, 21819–21823 | 21821


<strong>Downloaded</strong> on 06 October 2012<br />

Published on 10 September 2012 on http://pubs.rsc.org | doi:10.1039/C2JM34890J<br />

in our work. This has been repeated over 20 times and no decline in<br />

the electrosorption capacity is observed.<br />

Conclusions<br />

In summary, we have developed a simple yet effective strategy to<br />

make free-standing ACF electrodes with well-developed hierarchical<br />

porous structure that show excellent performance in CDI water<br />

desalination. The micro-meso-macroporous structures of the asmade<br />

ACFs can be tuned by incorporating mesoporous carbon<br />

blacks with good electric conductivity in the electrospinning process<br />

followed by CO 2 activation. The ACF web electrodes with a balanced<br />

hierarchical structure help overcome the mass-transport limitation<br />

that is an intrinsic shortcoming for micropore dominated materials.<br />

The novel materials reported here exhibit a high capacity for salt<br />

removal, and are of potential use in CDI desalination as electrodes.<br />

The desalination performance of the ACF web electrodes can be<br />

further improved by optimizing the properties of CB and the process<br />

parameters such as the weight ratio of CB to ACF in the precursor.<br />

Experimental<br />

Materials<br />

The precursor polyacrylonitrile (PAN, M w ¼ 150 000) and solvent<br />

N,N-dimethyl formamide (DMF) were obtained from Aldrich<br />

Chemical Co (USA). Carbon black (CB, ketjen EC-600JD, Japan)<br />

with a mesopore surface area of 393 m 2 g 1 wasusedasafillerof<br />

electrospun activated carbon nanofiber (ACF). For comparison,<br />

powder-like coconut shell-based activated carbon (AC) was used as<br />

an electrode material, of which the BET surface area was 917 m 2 g 1<br />

and the average pore diameter was 2.2 nm.<br />

Preparation method<br />

The PAN-based fibers were prepared by electrospinning using a<br />

10 wt% PAN solution in DMF, and the CB-embedded PAN-based<br />

fibers were prepared by electrospinning using a composite solution<br />

made of 5 wt% CB in 10 wt% PAN solution in DMF. The solution<br />

was sonicated for 2 h to ensure that the CB was dispersed uniformly<br />

before being mixed with the PAN. The two kinds of polymer solutions<br />

with/without CB were electrospun from a syringe tip onto a<br />

rotating metal drum wrapped in aluminium foil. Below are the<br />

conditions for electrospinning: a feeding rate of 1.0 mL h 1 polymer<br />

solution, a supplied voltage of 22 kV, a tip to drum distance of 15 cm,<br />

and a drum rotation rate of 300 rpm. The electrospun fiber webs were<br />

heated at 1 Cmin 1 and stabilized at 280 Cinairfor2h.The<br />

stabilized fibers were activated in flowing CO 2 for 0.5 h at 700–900<br />

C, resulting in ACF webs denoted as ACFX and ACF/CBX (CB<br />

containing ACF), where X stands for the activation temperature.<br />

Measurements<br />

The morphology and structures of the samples were examined using<br />

scanning electron microscopy (SEM, Hitachi S-4700, Japan) and<br />

transmission electron microscopy (TEM, FEI Tecnai G20, USA).<br />

The specific surface area was measured by nitrogen adsorption<br />

(ASAP2020, Micromeritics, USA). The electrical resistance was<br />

measured 5 times with a four-probe method at room temperature, of<br />

which the values were averaged.<br />

Schematic of the CDI set-up<br />

Activated carbon electrode was fabricated using direct coating<br />

method. 38 A carbon slurry was prepared as a suspension of activated<br />

carbon powder, acetylene black and poly(vinylidene fluoride) in dimethylacetamide<br />

in a weight ratio of 75 : 10 : 15. The slurry mixture<br />

was stirred for 6 h before being cast onto a graphite sheet to make a<br />

300 mm thick coating. The coated electrode was dried at 80 Cunder<br />

vacuum for 4 h to get rid of the organic solvent.<br />

The webs were used as CDI electrodes and directly attached to the<br />

graphite current collectors to make a CDI cell. The adsorption<br />

removal efficiency of ions on the web electrodes with a weight of ca.<br />

0.2 g was measured at 25 C in a flow-through setup, as shown in<br />

Fig. 5. For each run, the solution was continuously pumped into the<br />

cell at a rate of 5 mL min 1 using a peristaltic pump (Loner BT100,<br />

China). During the measurement, the potential difference between<br />

the two electrodes was kept at a constant voltage of V cell ¼ 1.6 V<br />

using a programmable DC power supply (PST-3202, Gwinstek,<br />

Taiwan). All experiments were conducted with a C 0 ¼ 90 mg L 1<br />

(197 mScm 1 ) NaCl solution. The variation of NaCl concentration in<br />

the solution was continuously monitored using an ion conductivity<br />

meter (ET915, eDAQ TECH, Australia). In the regeneration process,<br />

the flow rate was kept at 60 mL min 1 . After applying a reverse<br />

voltage of 1.6Vfor5min,itwasswitchedto0Vuntiltheregeneration<br />

process ended.<br />

The electrosorptive capacity (Q t ) is defined as below:<br />

Q t ¼ ðC 0 C t ÞV NaCl<br />

(1)<br />

M<br />

where C 0 (mg L 1 ) is the initial concentration of NaCl, C t is the<br />

instantaneous concentration of NaCl measured at time t; V NaCl is<br />

the solution volume (mL); and M is the mass of the ACF web<br />

electrodes (g).<br />

Acknowledgements<br />

We thank Prof. Yury Gogotsi at Drexel University, USA, for useful<br />

discussion. This work was partly supported by the Dalian Science<br />

and Technology Bureau of China (no. 2010A17GX095), the<br />

Research Start-up Fund in DUT (no. DUT12RC(3)04), and the<br />

NSFC (nos. 20836002, 51102033).<br />

Notes and references<br />

1 U. Yermiyahu, A. Tal, A. Ben-Gal, A. Bar-Tal, J. Tarchitzky and<br />

O. Lahav, Science, 2007, 318, 920–921.<br />

2 M. E. Suss, T. F. Baumann, W. L. Bourcier, C. M. Spadaccini,<br />

K. A. Rose, J. G. Santiago and M. Stadermann, Energy Environ.<br />

Sci., 2012, DOI: 10.1039/c2ee21498a.<br />

3 R. Zhao, P. M. Biesheuvel and A. van der Wal, Energy Environ. Sci.,<br />

2012, DOI: 10.1039/c2ee21737f.<br />

4 M. W. Ryoo and G. Seo, Water Res., 2003, 37, 1527–1534.<br />

5 J. Y. Lee, S. J. Seo, S. H. Yun and S. H. Moon, Water Res., 2011, 45,<br />

5375–5380.<br />

6 M. A. Anderson, A. L. Cudero and J. Palma, Electrochim. Acta, 2010,<br />

55, 3845–3856.<br />

7 P. M. Biesheuvel, J. Colloid Interface Sci., 2009, 332, 258–264.<br />

8 J. Biener, M. Stadermann, M. Suss, M. A. Worsley, M. M. Biener,<br />

K. A. Rose and T. F. Baumann, Energy Environ. Sci., 2011, 4, 656–667.<br />

9 P. M. Biesheuvel, R. Zhao, S. Porada and A. van der Wal, J. Colloid<br />

Interface Sci., 2011, 360, 239–248.<br />

10 J. C. Farmer, S. M. Bahowick, J. E. Harrar, D. V. Fix,<br />

R. E. Martinelli, A. K. Vu and K. L. Carroll, Energy Fuels, 1997,<br />

11, 337–347.<br />

21822 | J. Mater. Chem., 2012, 22, 21819–21823 This journal is ª The Royal Society of Chemistry 2012


<strong>Downloaded</strong> on 06 October 2012<br />

Published on 10 September 2012 on http://pubs.rsc.org | doi:10.1039/C2JM34890J<br />

11 S. Nadakatti, M. Tendulkar and M. Kadam, Desalination, 2011, 268,<br />

182–188.<br />

12 S. Porada, L. Weinstein, R. Dash, A. van der Wal, M. Bryjak,<br />

Y. Gogotsi and P. M. Biesheuvel, ACS Appl. Mater. Interfaces,<br />

2012, 4, 1194–1199.<br />

13 L. Y. Lee, H. Y. Ng, S. L. Ong, G. Tao, K. Kekre, B. Viswanath,<br />

W. Lay and H. Seah, Water Res., 2009, 43, 4769–4777.<br />

14 P. Xu, J. E. Drewes, D. Heil and G. Wang, Water Res., 2008, 42,<br />

2605–2617.<br />

15 L. Li, L. Zou, H. Song and G. Morris, Carbon, 2009, 47, 775–781.<br />

16 K. L. Yang, T. Y. Ying, S. Yiacoumi, C. Tsouris and E. S. Vittoratos,<br />

Langmuir, 2001, 17, 1961–1969.<br />

17 R. Mayes, C. Tsouris, J. Kiggans, S. Mahurin, D. DePaoli and S. Dai,<br />

J. Mater. Chem., 2010, 20, 8674–8678.<br />

18 C. Tsouris, R. Mayes, J. Kiggans, K. Sharma, S. Yiacoumi,<br />

D. DePaoli and S. Dai, Environ. Sci. Technol., 2011, 45, 10243–10249.<br />

19 G. Hasegawa, M. Aoki, K. Kanamori, K. Nakanishi, T. Hanada and<br />

K. Tadanaga, J. Mater. Chem., 2011, 21, 2060–2063.<br />

20 P. Adelhelm, Y. S. Hu, L. Chuenchom, M. Antonietti, B. M. Smarsly<br />

and J. Maier, Adv. Mater., 2007, 19, 4012–4017.<br />

21 C. J. Gabelich, T. D. Tran and S. I. H. Mel, Environ. Sci. Technol.,<br />

2002, 36, 3010–3019.<br />

22 C. Nie, L. Pan, H. Li, T. Chen, T. Lu and Z. Sun, J. Electroanal.<br />

Chem., 2012, 666, 85–88.<br />

23 H. Li, T. Lu, L. Pan, Y. Zhang and Z. Sun, J. Mater. Chem., 2009, 19,<br />

6773–6779.<br />

24 Z. Peng, D. Zhang, L. Shi and T. Yan, J. Mater. Chem., 2012, 22,<br />

6603–6612.<br />

25 K. Y. Foo and B. H. Hameed, J. Hazard. Mater., 2009, 170, 552–<br />

559.<br />

26 R. Sahay, P. S. Kumar, R. Sridhar, J. Sundaramurthy, J. Venugopal,<br />

S. G. Mhaisalkar and S. Ramakrishna, J. Mater. Chem., 2012, 22,<br />

12953–12971.<br />

27 C. Kim, B. T. N. Ngoc, K. S. Yang, M. Kojima, Y. A. Kim,<br />

Y. J. Kim, M. Endo and S. C. Yang, Adv. Mater., 2007, 19, 2341–<br />

2346.<br />

28 V. Presser, L. Zhang, J. J. Niu, J. McDonough, C. Perez, H. Fong and<br />

Y. Gogotsi, Adv. Energy Mater., 2011, 1, 423–430.<br />

29 G. Wang, C. Pan, L. Wang, Q. Dong, C. Yu, Z. Zhao and J. Qiu,<br />

Electrochim. Acta, 2012, 69, 65–70.<br />

30 L. Wang, M. Wang, Z. H. Huang, T. Cui, X. Gui, F. Kang, K. Wang<br />

and D. Wu, J. Mater. Chem., 2011, 21, 18295–18299.<br />

31 M. Wang, Z. H. Huang, L. Wang, M. X. Wang, F. Kang and H. Hou,<br />

New J. Chem., 2010, 34, 1843–1845.<br />

32 Y. Zhan, C. Nie, H. Li, L. Pan and Z. Sun, Electrochim. Acta, 2011,<br />

56, 3164–3169.<br />

33 Z. Hu, M. P. Srinivasan and Y. Ni, Adv. Mater., 2000, 12, 62–65.<br />

34 Y. W. Ju, G. R. Choi, H. R. Jung and W. J. Lee, Electrochim. Acta,<br />

2008, 53, 5796–5803.<br />

35 B. H. Park and J. H. Choi, Electrochim. Acta, 2010, 55, 2888–<br />

2893.<br />

36 X. Wang, J. S. Lee, C. Tsouris, D. W. DePaoli and S. Dai, J. Mater.<br />

Chem., 2010, 20, 4602–4608.<br />

37 X. Z. Wang, M. G. Li, Y. W. Chen, R. M. Cheng, S. M. Huang,<br />

L. K. Pan and Z. Sun, Appl. Phys. Lett., 2006, 89, 053127.<br />

38 Y. J. Kim and J. H. Choi, Water Res., 2010, 44, 990–996.<br />

This journal is ª The Royal Society of Chemistry 2012 J. Mater. Chem., 2012, 22, 21819–21823 | 21823

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!