+ All Categories
Home > Documents > Journal of Materials Chemistry A - Yao Research Group |...

Journal of Materials Chemistry A - Yao Research Group |...

Date post: 23-May-2018
Category:
Upload: truongkhuong
View: 213 times
Download: 0 times
Share this document with a friend
4
Flexible electrode for long-life rechargeable sodium-ion batteries: eect of oxygen vacancy in MoO 3x Yifei Li,a Dandan Wang,a Qinyou An, a Bush Ren, a Yaoguang Rong a and Yan Yao * ab Layered molybdenum trioxide (MoO 3 ) has rarely been studied as an electrode material for sodium ion batteries due to its low electronic conductivity and irreversible phase transitions. Here we demonstrate that MoO 3x , with a well-controlled oxygen vacancy, exhibits near theoretical capacity, excellent rate capability, and 2000 stable cycles with minimal capacity loss. The oxygen vacancy in MoO 3x is responsible for the two-plateau voltage prole, in contrast to the sloping feature observed in a-MoO 3 . This work highlights the impor- tance of oxygen vacancies in enabling long-life rechargeable sodium- ion batteries. The cost of lithium is expected to soar as it powers the next generation of electric vehicles due to limited resources, rendering the development of low-cost alternatives to lithium ion batteries of great importance. 1 The earthly abundance and low cost of raw materials has become of unparalleled superi- ority. 2 In this respect, ubiquitous distribution of sodium in the earth's crust has stimulated extensive research interest towards sodium-ion batteries. 36 Molybdenum oxides have enabled a wide range of applica- tions in gas sensing, catalysts, electronic devices and lubricants due to their low cost, resource abundance and nontoxicity. In particular, a layered molybdenum trioxide (a-MoO 3 ) has been studied as an attractive electrode in rechargeable lithium-ion batteries due to its high theoretical specic capacity of 372 mA h g 1 , corresponding to two electrons reaction per formula. 712 However, the poor cycling stability so far has limited its appli- cation as a battery electrode for two reasons. One is the low electronic and ionic conductivity, and the other is the irrevers- ible phase transition related to intralayer Li insertion. 7 Strate- gies were developed in the past to overcome such challenges: (i) nanosizing particle size to reduce the Li ion diusion distance, (ii) carbon coating to increase the electronic conductivity, 913 and (iii) pre-intercalating Li, Na or K ions in layered MoO 3 to suppress intralayer Li insertion. 1417 Oxygen decient MoO 3x was reported to have fast interca- lation and excellent electrochemical performance in lithium- ion batteries. 18 According to Law et al., 19 oxygen vacancies in MoO 3x led to increased electronic conductivity. Goodenough et al. showed that reduced MoO 2 had metallic conductivity as the Mo 4d orbitals split into t 2g and e g orbitals and one extra electron per metal cation in MoO 2 partially populated the higher t bands. 20 Therefore, partially reduced MoO 3x is bene- cial for electrochemical applications. 2124 To the best of our knowledge, MoO 3x so far has not been investigated as an electrode material in sodium-ion batteries. In this work, we demonstrate partially reduced MoO 3x grown on exible carbon cloth and investigate its electrochemical performance as a sodium-ion electrode material (Fig. 1), exploiting the reducing power of carbon cloth when heated with MoO 3 . The electrochemical characterizations reveal improved electrochemical performance with a reversible capacity close to the theoretical specic capacity of MoO 3 (156.3 mA h g 1 ) as well as a signicantly improved cycling stability (92% capacity retention aer 2000 cycles at 1000 mA g 1 ). Fig. 1 Schematic diagram of partially-reduced MoO 3x deposited on exible carbon cloth to enable facile ion diusion and electron conduction. a Department of Electrical and Computer Engineering and Materials Science and Engineering Program, University of Houston, Houston, Texas, 77204, USA. E-mail: [email protected] b Texas Center for Superconductivity (TcSUH), University of Houston, Houston, Texas, 77204, USA These authors contributed equally to this work. Cite this: J. Mater. Chem. A, 2016, 4, 5402 Received 14th February 2016 Accepted 11th March 2016 DOI: 10.1039/c6ta01342b www.rsc.org/MaterialsA 5402 | J. Mater. Chem. A, 2016, 4, 54025405 This journal is © The Royal Society of Chemistry 2016 Journal of Materials Chemistry A COMMUNICATION Published on 11 March 2016. Downloaded by University of Houston on 05/06/2016 17:28:27. View Article Online View Journal | View Issue
Transcript

Journal ofMaterials Chemistry A

COMMUNICATION

Publ

ishe

d on

11

Mar

ch 2

016.

Dow

nloa

ded

by U

nive

rsity

of

Hou

ston

on

05/0

6/20

16 1

7:28

:27.

View Article OnlineView Journal | View Issue

Flexible electrod

aDepartment of Electrical and Computer

Engineering Program, University of Housto

[email protected] Center for Superconductivity (TcSUH

77204, USA

† These authors contributed equally to th

Cite this: J. Mater. Chem. A, 2016, 4,5402

Received 14th February 2016Accepted 11th March 2016

DOI: 10.1039/c6ta01342b

www.rsc.org/MaterialsA

5402 | J. Mater. Chem. A, 2016, 4, 540

e for long-life rechargeablesodium-ion batteries: effect of oxygen vacancy inMoO3�x

Yifei Li,†a Dandan Wang,†a Qinyou An,a Bush Ren,a Yaoguang Ronga and Yan Yao*ab

Layered molybdenum trioxide (MoO3) has rarely been studied as an

electrode material for sodium ion batteries due to its low electronic

conductivity and irreversible phase transitions. Here we demonstrate

that MoO3�x, with a well-controlled oxygen vacancy, exhibits near

theoretical capacity, excellent rate capability, and 2000 stable cycles

with minimal capacity loss. The oxygen vacancy in MoO3�x is

responsible for the two-plateau voltage profile, in contrast to the

sloping feature observed in a-MoO3. This work highlights the impor-

tance of oxygen vacancies in enabling long-life rechargeable sodium-

ion batteries.

The cost of lithium is expected to soar as it powers the nextgeneration of electric vehicles due to limited resources,rendering the development of low-cost alternatives to lithiumion batteries of great importance.1 The earthly abundance andlow cost of raw materials has become of unparalleled superi-ority.2 In this respect, ubiquitous distribution of sodium in theearth's crust has stimulated extensive research interest towardssodium-ion batteries.3–6

Molybdenum oxides have enabled a wide range of applica-tions in gas sensing, catalysts, electronic devices and lubricantsdue to their low cost, resource abundance and nontoxicity. Inparticular, a layered molybdenum trioxide (a-MoO3) has beenstudied as an attractive electrode in rechargeable lithium-ionbatteries due to its high theoretical specic capacity of 372mA hg�1, corresponding to two electrons reaction per formula.7–12

However, the poor cycling stability so far has limited its appli-cation as a battery electrode for two reasons. One is the lowelectronic and ionic conductivity, and the other is the irrevers-ible phase transition related to intralayer Li insertion.7 Strate-gies were developed in the past to overcome such challenges: (i)

Engineering and Materials Science and

n, Houston, Texas, 77204, USA. E-mail:

), University of Houston, Houston, Texas,

is work.

2–5405

nanosizing particle size to reduce the Li ion diffusion distance,(ii) carbon coating to increase the electronic conductivity,9–13

and (iii) pre-intercalating Li, Na or K ions in layered MoO3 tosuppress intralayer Li insertion.14–17

Oxygen decient MoO3�x was reported to have fast interca-lation and excellent electrochemical performance in lithium-ion batteries.18 According to Law et al.,19 oxygen vacancies inMoO3�x led to increased electronic conductivity. Goodenoughet al. showed that reduced MoO2 had metallic conductivity asthe Mo 4d orbitals split into t2g and eg orbitals and one extraelectron per metal cation in MoO2 partially populated thehigher t bands.20 Therefore, partially reduced MoO3�x is bene-cial for electrochemical applications.21–24 To the best of ourknowledge, MoO3�x so far has not been investigated as anelectrode material in sodium-ion batteries.

In this work, we demonstrate partially reducedMoO3�x grownon exible carbon cloth and investigate its electrochemicalperformance as a sodium-ion electrode material (Fig. 1),exploiting the reducing power of carbon cloth when heated withMoO3. The electrochemical characterizations reveal improvedelectrochemical performance with a reversible capacity close tothe theoretical specic capacity of MoO3 (156.3 mA h g�1) as wellas a signicantly improved cycling stability (92% capacityretention aer 2000 cycles at 1000 mA g�1).

Fig. 1 Schematic diagram of partially-reduced MoO3�x deposited onflexible carbon cloth to enable facile ion diffusion and electronconduction.

This journal is © The Royal Society of Chemistry 2016

Table 1 Deconvolution of the XPS spectra of Samples 1–3

Sample 1 Sample 2 Sample 3

Mo6+ 3d3/2 260 8792 56633d5/2 4700 14 241 8812

Mo5+ 3d3/2 1700 440 03d5/2 6600 440 0

Mo4+ 3d3/2 1300 0 03d5/2 3705 0 0

Valence state 5.00 5.94 6.00x in MoO3�x 0.50 0.03 0

Communication Journal of Materials Chemistry A

Publ

ishe

d on

11

Mar

ch 2

016.

Dow

nloa

ded

by U

nive

rsity

of

Hou

ston

on

05/0

6/20

16 1

7:28

:27.

View Article Online

To study the effect of the oxygen vacancy inMoO3�x, Samples1–3 were prepared as following. Carbon cloth (Fuel Cell EarthCCP40) was cleaned and punched to make discs of 1 cm indiameter. Sample 1 was prepared by annealing carbon cloth and0.2 g MoO3 powder in a corundum crucible at 690 �C for 30 min.Molybdenum oxide was deposited on the carbon cloth andbecame partially reduced (�5 mg). Sample 2 was prepared byannealing Sample 1 at 350 �C for 2 h in air. Sample 3 wasprepared by annealing Sample 1 at the higher temperature of450 �C for 2 h in air. All samples were characterized using X-raydiffraction (XRD, Rigaku MiniFlex 600) with Cu Ka radiation(l ¼ 1.5406 A), X-ray photoelectron spectroscopy (XPS, VGMultiLab 2000), and scanning electron microscopy (SEM;Gemini LEO 1525).

Fig. 2a shows the XRD patterns of Samples 1–3 as well as theMoO2 and MoO3 standards. Monoclinic MoO2 (JCPDS no. 01-086-0135, a ¼ 5.6096 A, b ¼ 4.8570 A, c ¼ 5.6259 A and b ¼120.9120�) has a distorted rutile structure that has two O siteslabeled as O1 and O2 in Fig. 2c. In comparison, orthorhombica-MoO3 is a stable bilayered structure (JCPDS no. 00-001-0706,a ¼ 3.9540 A, b ¼ 13.8250 A, c ¼ 3.6940 A and a ¼ b ¼ g ¼ 90�)with the space group Pnma.25 Each bilayer consists of two sub-layers of distorted octahedral MoO6 (Fig. 2c). There are threeoxygen sites O1, O2 and O3 in the MoO3 structure.19 O1 oxygenis bonded to oneMo atom, O2 is asymmetrically placed between2 Mo centers with a two-fold symmetry, while O3 is locatedbetween two Mo centers in one sublayer and another Mo centerin the neighboring sublayer. Comparing with the XRD stan-dards, Sample 1 is mainly the monoclinic MoO2 phase, Sample3 is mainly the orthorhombic a-MoO3 phase, while Sample 2 isa mixture of the MoO3 and MoO2 phases.

In order to quantify the concentration (x) of the oxygenvacancy in the MoO3�x samples, we conducted XPS measure-ments and the results are shown in Fig. 2b and summarized inTable 1. Deconvolution of the Mo 3d spectra in Sample 1 revealssimultaneous existence of peaks from Mo4+, Mo5+ and Mo6+.

Fig. 2 (a) XRD patterns of Samples 1–3 and standard PDFs formonoclinic MoO2 and orthorhombic MoO3. (b) XPS of Samples 1–3. (c)Crystal structures of MoO2 and a-MoO3 with different oxygen atomshighlighted.

This journal is © The Royal Society of Chemistry 2016

Mo4+ consists of two well-dened peaks located at 229.4 eV forMo4+ 3d5/2 and 232.6 eV for Mo4+ 3d3/2.26,27 Peaks at 235.1 and231.6 eV for Mo5+ and 236.0 and 232.9 eV for Mo6+ exist withsmaller percentages. Through tting the peak area ratio in theXPS spectra, we nd that the average valence state of Mo inSample 1 is 5.0 and that x is 0.5 in MoO3�x. In Sample 2, thepeaks corresponding to Mo6+, located at 236.0 eV and 232.9 eV,become dominant, while the Mo5+ peaks are suppressed andthe Mo4+ peaks disappear. The average valence state of Mo inSample 2 is 5.94 and thus x is 0.03 in MoO3�x. In Sample 3, theXPS spectrum shows a sharp Mo 3d5/2 peak at 232.9 eV and aMo3d3/2 peak at 236 eV, suggesting x is 0 in Sample 3.28,29 Therefore,it is clear that carbon cloth reduces MoO3 to MoO2.5 duringsynthesis, while thermal annealing in air partially oxidizes Moback and results in a mixture of Mo5+ and Mo6+.

Fig. 3a–c show the SEM images of Samples 1–3, in whichcarbon bers are clearly visible. At intermediate magnication(Fig. 3d–f), MoO3�x nanoplates uniformly cover the surface ofthe carbon bers. They are several micrometers in length anda few hundred nanometers in thickness (Fig. 3g–i). The nano-sized MoO3�x plates would allow fast kinetics for Na-ion diffu-sion due to the short diffusion distance. In addition, theconductive carbon ber network is retained very well during theannealing process. All three MoO3�x samples show similarnanoscale morphology, therefore allowing us to compare elec-trochemical performance solely based on the differences inoxygen vacancies in the structure.

Fig. 3 (a, d and g) SEM images for Sample 1. (b, e and h) SEM images forSample 2. (c, f and i) SEM images for Sample 3. Insets show the opticalimages of Samples 1–3.

J. Mater. Chem. A, 2016, 4, 5402–5405 | 5403

Journal of Materials Chemistry A Communication

Publ

ishe

d on

11

Mar

ch 2

016.

Dow

nloa

ded

by U

nive

rsity

of

Hou

ston

on

05/0

6/20

16 1

7:28

:27.

View Article Online

To shed light on the effect of oxygen vacancy concentrationon the electrochemical performance, the three samples weremeasured using CR2032 coin cells with sodium metal as thecounter electrode and 1MNaClO4 in propylene carbonate as theelectrolyte. No additional conductive agent or binder was used.Fig. 4a–c display the discharge–charge voltage proles of thethree samples at the 1st, 50th, and 100th cycles at a low rate(current density of 100 mA g�1). Sample 1 shows negligiblecapacity at the rst cycle, and slowly increases to �60 mA h g�1

aer 100 cycles. Interestingly, the voltage prole develops a two-plateau characteristic, which is quite different from the slopingprole typically observed in orthorhombic a-MoO3 aer the rstcycle.30 The sloping prole suggests the amorphous structure ofa-MoO3 aer the rst cycle, while the plateaus observed inSample 1 indicate the ability of the oxygen vacancies to stabilizethe crystalline structure of MoO3�x during cycling. Sample2 also shows a two-voltage plateau but with a higher capacity of156 mA h g�1, close to the theoretical specic capacity ofMoO3�x. For Sample 3, the capacity shows a capacity as high as156 mA h g�1 in the rst cycle. However only 40% capacityretention is observed aer 50 cycles, which is comparable to theperformance of orthorhombic a-MoO3.10 These results indicatethat a 3% oxygen vacancy (Sample 2) is optimal for both highcapacity and stable cycling. Too many (Sample 1) or too few(Sample 3) oxygen vacancies result in a reduced electrochemicalperformance.

The rate capabilities of Samples 1–3 are further tested atvarious current densities (50 to 1000 mA g�1) within the voltagewindow of 1.0–4.0 V vs. Na/Na+ (Fig. 4d). Sample 2 delivers the

Fig. 4 (a–c) Galvanostatic charge–discharge profiles of Samples 1–3at a current of 100 mA g�1. (d) The rate performance of Samples 1–3.(e) Cycling performance of Samples 1–3 at a current of 100 mA g�1,with discharge and charge capacities. (f) Long-term cycling perfor-mance of Samples 1–3 at a current of 1000 mA g�1.

5404 | J. Mater. Chem. A, 2016, 4, 5402–5405

highest capacity among the three samples with its dischargecapacity of 176.6, 122.9, 98.4 and 79.2 mA h g�1 at currentdensities of 50, 200, 500 and 1000 mA g�1, respectively. Fig. 4eand f show the cycling performance of the three samples at 100and 1000 mA g�1, respectively. With an initial capacity drop at100 mA g�1, the capacity of Sample 2 increases over cycling andeventually reaches 156.3 mA h g�1 aer 200 cycles, correspond-ing to 0.84 sodium ions per MoO3�x. The excellent stability inSample 2 arises from a balancedmixture of the conductiveMoO2

phase and a Na-intercalating MoO3 phase. A uctuation in thespecic capacity during cycling is observed and is possibly due tothe uneven electrolyte wetting of the phases of MoO2 and MoO3.This phenomenon has been previously observed in amorphousMnOx–carbon nanocomposites,31 and in a-Fe2O3/b-MnO2 nano-rods32 during lithium-ion storage. Sample 1 approachesa specic capacity of 64mA h g�1 aer 80 cycles. Sample 3 has aninitial specic capacity of 157 mA h g�1, but decays very fast withonly 63.4 mA h g�1 aer 100 cycles. A similar decay is well knownfor MoO3 in lithium-ion batteries due to the irreversible phasetransition and intralayer trapping of Li ions.8,33–37 At a currentdensity of 1000 mA g�1 (Fig. 4f), Sample 2 still shows the bestperformance. The discharge capacity of Sample 2 increases to92.6 mA h g�1 aer 500 cycles and then shows minimal capacityloss over 2000 cycles, with 92% capacity retention compared tothe capacity at the 500th cycle.

Conclusions

In summary, partially reducedMoO3�x grown on exible carboncloth has been synthesized via a facile thermal reductionmethod. The obtained MoO3�x with well-controlled oxygenvacancies exhibits near theoretical capacity, good rate capabilityand much enhanced cycling stability. Moreover, it presents verysmall capacity loss over 2000 cycles at a high rate (1000 mA g�1).Further investigations of XRD and XPS reveal that the bestperforming electrode features a 3% oxygen vacancy in thestructure of MoO3�x. This study demonstrates the importanceof controlling the oxygen vacancy in optimizing materials forlong-life rechargeable sodium-ion batteries.

Acknowledgements

We acknowledge the support from the National Science Foun-dation (CMMI-1400261).

Notes and references

1 S.-W. Kim, D.-H. Seo, X. Ma, G. Ceder and K. Kang, Adv.Energy Mater., 2012, 2, 710–721.

2 B. Dunn, H. Kamath and J. M. Tarascon, Science, 2011, 334,928–935.

3 K. B. Hueso, M. Armand and T. Rojo, Energy Environ. Sci.,2013, 6, 734–749.

4 D. Kundu, E. Talaie, V. Duffort and L. F. Nazar, Angew. Chem.,Int. Ed., 2015, 54, 3431–3448.

5 S. Li, Y. Dong, L. Xu, X. Xu, L. He and L. Mai, Adv. Mater.,2014, 26, 3545–3553.

This journal is © The Royal Society of Chemistry 2016

Communication Journal of Materials Chemistry A

Publ

ishe

d on

11

Mar

ch 2

016.

Dow

nloa

ded

by U

nive

rsity

of

Hou

ston

on

05/0

6/20

16 1

7:28

:27.

View Article Online

6 Y. Li, Y. Liang, F. C. Robles Hernandez, H. Deog Yoo, Q. Anand Y. Yao, Nano Energy, 2015, 15, 453–461.

7 X. J. Wang, R. Nesper, C. Villevieille and P. Novak, Adv. EnergyMater., 2013, 3, 606–614.

8 T. Brezesinski, J. Wang, S. H. Tolbert and B. Dunn, Nat.Mater., 2010, 9, 146–151.

9 W. Tang, L. Liu, Y. Zhu, H. Sun, Y. Wu and K. Zhu, EnergyEnviron. Sci., 2012, 5, 6909–6913.

10 M. Baldoni, L. Craco, G. Seifert and S. Leoni, J. Mater. Chem.A, 2013, 1, 1778–1784.

11 L. Noerochim, J. Z. Wang, D. Wexler, Z. Chao and H. K. Liu,J. Power Sources, 2013, 228, 198–205.

12 W. Tang, Y. Zhu, Y. Hou, L. Liu, Y. Wu, K. P. Loh, H. Zhangand K. Zhu, Energy Environ. Sci., 2013, 6, 2093–2104.

13 Y. Liu, B. H. Zhang, S. Y. Xiao, L. L. Liu, Z. B. Wen andY. P. Wu, Electrochim. Acta, 2014, 116, 512–517.

14 J. G. Simmons, G. S. Nadkarni and M. C. Lancaster, J. Appl.Phys., 1970, 41, 538–544.

15 L. Q. Mai, B. Hu, W. Chen, Y. Y. Qi, C. S. Lao, R. S. Yang,Y. Dai and Z. L. Wang, Adv. Mater., 2007, 19, 3712–3716.

16 Y. Dong, X. Xu, S. Li, C. Han, K. Zhao, L. Zhang, C. Niu,Z. Huang and L. Mai, Nano Energy, 2015, 15, 145–152.

17 Z. Hu, C. Zhou, M. Zheng, J. Lu, B. Varghese, H. Cheng andC.-H. Sow, J. Phys. Chem. C, 2012, 116, 3962–3967.

18 P. Meduri, E. Clark, J. H. Kim, E. Dayalan,G. U. Sumanasekera and M. K. Sunkara, Nano Lett., 2012,12, 1784–1788.

19 D. O. Scanlon, G. W. Watson, D. J. Payne, G. R. Atkinson,R. G. Egdell and D. S. L. Law, J. Phys. Chem. C, 2010, 114,4636–4645.

20 J. B. Goodenough, Prog. Solid State Chem., 1971, 5, 145–399.21 B. Hu, L. Mai, W. Chen and F. Yang, ACS Nano, 2009, 3, 478–

482.22 Y. Xu, R. Yi, B. Yuan, X. Wu, M. Dunwell, Q. Lin, L. Fei,

S. Deng, P. Andersen, D. Wang and H. Luo, J. Phys. Chem.Lett., 2012, 3, 309–314.

This journal is © The Royal Society of Chemistry 2016

23 Y. Shi, B. Guo, S. A. Corr, Q. Shi, Y.-S. Hu, K. R. Heier,L. Chen, R. Seshadri and G. D. Stucky, Nano Lett., 2009, 9,4215–4220.

24 Y. Sun, X. Hu, W. Luo and Y. Huang, ACS Nano, 2011, 5,7100–7107.

25 N. A. Chernova, M. Roppolo, A. C. Dillon andM. S. Whittingham, J. Mater. Chem., 2009, 19, 2526–2552.

26 X. Xiao, Z. Peng, C. Chen, C. Zhang, M. Beidaghi, Z. Yang,N. Wu, Y. Huang, L. Miao, Y. Gogotsi and J. Zhou, NanoEnergy, 2014, 9, 355–363.

27 J. Swiatowska-Mrowiecka, S. de Diesbach, V. Maurice,S. Zanna, L. Klein, E. Briand, I. Vickridge and P. Marcus,J. Phys. Chem. C, 2008, 112, 11050–11058.

28 T. H. Fleisch and G. J. Mains, J. Chem. Phys., 1982, 76, 780–786.

29 J. G. Choi and L. T. Thompson, Appl. Surf. Sci., 1996, 93, 143–149.

30 M. E. Spahr, P. Novak, O. Haas and R. Nesper, J. PowerSources, 1995, 54, 346–351.

31 J. Guo, Q. Liu, C. Wang and M. R. Zachariah, Adv. Funct.Mater., 2012, 22, 803–811.

32 X. Gu, L. Chen, Z. Ju, H. Xu, J. Yang and Y. Qian, Adv. Funct.Mater., 2013, 23, 4049–4056.

33 H. Liu, Y. Wang, W. Yang and H. Zhou, Electrochim. Acta,2011, 56, 1392–1398.

34 F. Tanguy, J. Gaubicher and D. Guyomard, Electrochim. Acta,2010, 55, 3979–3986.

35 S. Jouanneau, A. Verbaere, S. Lascaud and D. Guyomard,Solid State Ionics, 2006, 177, 311–315.

36 X. Xu, Y.-Z. Luo, L.-Q. Mai, Y.-L. Zhao, Q.-Y. An, L. Xu, F. Hu,L. Zhang and Q.-J. Zhang, NPG Asia Mater., 2012, 4, e20.

37 Q. An, F. Lv, Q. Liu, C. Han, K. Zhao, J. Sheng, Q. Wei, M. Yanand L. Mai, Nano Lett., 2014, 14, 6250–6256.

J. Mater. Chem. A, 2016, 4, 5402–5405 | 5405


Recommended