You are on page 1of 4

Journal of

Materials Chemistry A
View Article Online
COMMUNICATION View Journal | View Issue
Published on 14 January 2014. Downloaded by Memorial University of Newfoundland on 18/07/2014 09:04:05.

Efficient Co–Fe layered double hydroxide


photocatalysts for water oxidation under visible
Cite this: J. Mater. Chem. A, 2014, 2,
4136 light†
Received 28th November 2013 Sang Jun Kim,a Yeob Lee,a Dong Ki Lee,a Jung Woo Leea and Jeung Ku Kang*ab
Accepted 13th January 2014

DOI: 10.1039/c3ta14933a

www.rsc.org/MaterialsA

We report new Co–Fe LDH water oxidation photocatalysts and their and have suitable properties for photocatalyst use.8,9 For
water oxidation abilities under visible light with different ratios of example, hematite (a-Fe2O3), with a band gap of about 2 eV, has
transition metals. The layered structure containing more iron dis- been suggested as a water oxidation catalyst that is able to
played the highest efficiency, attributed to good crystallinity, harvest a wide range of the solar spectrum.10–12 However, it was
enhanced light absorbance and low charge carrier recombination of found to have a short hole diffusion length, thus resulting in the
Co–O–Fe oxo-bridges. fast recombination of holes with electrons before water oxida-
tion could occur.11–13 Cobalt oxides have also been used as co-
catalysts in water oxidation reactions because of their high
Global energy today is primarily obtained via the combustion of turnover frequencies (TOFs),14 although they were unable to
fossil fuels, which generates pollution and results in climate oxidize water by themselves.15,16 Layered double hydroxides
change.1,2 Articial photosynthesis is considered to be an (LDHs), which are composed of mixed metal oxides with the
appealing method for the regeneration of carbon fuels from chemical formula [M2+1 xM3+x(OH)2]x+[(An )x/n]x $mH2O, have
water and carbon dioxide (CO2) using solar energy.3,4 Articial also been proposed as possible photocatalysts.17–20 Structurally,
photosynthesis consists of two important steps—the rst is they contain stacked brucite-like M2+(OH)2 hydroxide layers in
water oxidation and the second is the conversion of CO2 into which some of the M2+ ions undergo isomorphous substitutions
hydrocarbon fuels (e.g., methanol) using the protons generated by M3+ ions.18,21 A feature of these structures is the linking of the
during the rst step. This mimics the natural photosynthesis metal ions to each other by oxygen atoms.22 The oxo-bridges
process in which electrons from the photosystem II oxidation (M1–O–M2) block electron–hole recombination via a metal-to-
catalyst are transferred to photosystem I for the conversion of metal charge transfer (MMCT) mechanism.23 Another advan-
CO2 into carbohydrates.5 Visible and ultraviolet light are used tage is that LDHs can be also prepared on a large scale.17,24
for articial photosynthesis. However, because the energy of LDHs based on titanium or zinc oxides have been tested for
visible light is much higher than the energy of ultraviolet light their photocatalytic abilities in water oxidation.2,17,22 Also, a Co–
at the earth's surface, the photocatalytic reaction can be per- Fe LDH was synthesized with a 2 : 1 ratio for Co2+/Fe3+ loading
formed under visible light to use solar energy at maximum precursors, but it was not very crystalline.25 Meanwhile, LDHs
efficiency. with good crystallinity can also be prepared using various
The efficiency of a photocatalyst is determined by its elec- metals and anions as well as different metal ratios. Therefore,
tronic energy band gap, electron–hole pair lifetime, and charge the exploration of photocatalytic ability as a function of the
mobility.6,7 Some transition metals are attractive as water metal ratio would be meaningful. Oxo-bridges in Fe-based LDH
oxidation catalysts because they are inexpensive and abundant, photocatalysts help to prevent the recombination of holes with
electrons, and thereby overcome the issue related to the short
a
Department of Materials Science & Engineering and NanoCentury KAIST Institute, hole diffusion length.23,26 Depending on the ratio of the two
Korea Advanced Institute of Science and Technology (KAIST), 291 Daehak-ro, metals, the number of oxo-bridges connecting different metals
Yuseong-gu, Daejeon 305-701, Republic of Korea. E-mail: jeungku@kaist.ac.kr; Fax:
can be increased, which will signicantly affect the water
+82-42-350-3310
b
Graduate School of Energy, Environment, Water, and Sustainability (EEWS), Korea
oxidation abilities of the LDHs.
Advanced Institute of Science and Technology (KAIST), 291 Daehak-ro, Yuseong-gu, Herein, we report highly efficient Co–Fe LDH water oxidation
Daejeon 305-701, Republic of Korea. E-mail: jeungku@kaist.ac.kr; Fax: +82-42-350- photocatalysts under visible light and having different water
3310 oxidation abilities with different ratios of transition metals.
† Electronic supplementary information (ESI) available. See DOI: Scheme 1 illustrates the Co–Fe LDH structure and the water
10.1039/c3ta14933a

4136 | J. Mater. Chem. A, 2014, 2, 4136–4139 This journal is © The Royal Society of Chemistry 2014
View Article Online

Communication Journal of Materials Chemistry A

2.23 (CFL2) have poor crystallinities. Fig. 1b, S3a and b and
S4a–c† show high-resolution transmission electron microscopy
(HR-TEM) images of the Co–Fe LDHs. The small round particles
in Fig. S4a† demonstrate the agglomeration of LDH particles.
The magnied image in Fig. S3b† reveals ne striped layers
indicative of a layered structure. In particular, the lattice spac-
Published on 14 January 2014. Downloaded by Memorial University of Newfoundland on 18/07/2014 09:04:05.

ings of 0.263 and 0.232 nm in Fig. 1b and S3a† correspond to


the (012) and (015) planes of Co–Fe LDH, respectively. Finally,
the energy-dispersive spectrometry (EDS) images in Fig. S3d–f
and S4d–f† conrmed that cobalt and iron were well dispersed
in the structure.
The anions located between the interlayers were identied by
Fourier transform infrared (FT-IR) spectroscopy (Fig. S5†). The
Scheme 1 Schematic illustration of a Co–Fe layered double hydroxide strong broad band at 3442 cm 1 was attributed to the stretching
structure and its catalysis of water oxidation. This reaction proceeds mode of the –OH bonds in the hydroxide layer, which are
under visible light because of its suitable bandgap.
marked as red spheres in Scheme 1, and water between the
hydroxide layers.31 Deformation of the H–O–H angle in water
oxidation reaction under visible light. Three Co–Fe LDHs were produced the characteristic small bands at 1633 and 1556
synthesized by the co-precipitation method under an inert cm 1.30,32 The peaks at 1384 and 1353 cm 1 resulted from the
atmosphere with cobalt and iron ratios of 2.23, 3.33, and 4.16 asymmetric stretching of carbonate anions in the interlayer
(Table S1†), which were denoted as CFL2, CFL3, and CFL4, region.33,34 The splitting of the CO32 peak was attributed to the
respectively. presence of both C–O and C]O bonds.22 Several metal-ion-
Fig. 1a shows the powder X-ray diffraction (XRD) patterns of related bands for M–O or M–OH stretching modes occurred at
crystals of the three Co–Fe LDHs; peaks that were characteristic low frequencies. Despite the presence of other anions such as
of an LDH structure were clearly observed.27 In the case of CFL4, NO3 in the solution, the FT-IR results suggested that the Co–Fe
the (003), (006), (012), (110), and (113) peaks (JCPDS Card no. LDHs contained CO32 in the interlayers. This is because CO32
50-0235, Fig. S1(a)†) were observed at 11.48 , 23.13 , 33.02 , has good affinity towards the hydroxide layers and water
59.05 , and 60.35 , respectively. The (003) basal plane spacing molecules in the interlayer,35 and the anion neutralizes the
was determined to be 7.7 Å (Scheme 1).27,28 The peak corre- charge in the LDH structure.
sponding to the (110) plane was very distinct, signifying a Cobalt and iron K-edge X-ray absorption near-edge structure
regular arrangement of the metal ions in the Co–Fe LDH.29,30 (XANES) spectra were obtained to determine the cobalt and iron
Meanwhile, CFL2 and CFL3 had more basal and non-basal oxidation states in the metal hydroxide layer. Fig. 1c shows the
peaks, such as (101), (009), (015), (009), (018), (0012), and (1013). cobalt K-edge XANES spectra of different cobalt oxides (CoO,
This meant that CFL2 and CFL3 had higher crystallinities than Co2O3, Co3O4) and the Co–Fe LDH. All the spectra showed a pre-
CFL4, including non-basal planes. Further experiments edge shoulder around 7710 eV, attributed to quadrupole-
(Fig. S2†) show that LDHs less than low cobalt and iron ratios of allowed electronic transitions from the 1s orbital to the 3d
orbital.36 The position of the main absorption edge around 7727
eV was similar for both Co–Fe LDH and CoO, caused by a dipole
transition from the 1s orbital to the 4p orbital. The XANES results
provide evidence that there is Co–O bonding with a Co2+ oxida-
tion state in the Co–Fe LDH. Also, the iron K-edge XANES spectra
of Fe2O3, Fe3O4, and Co–Fe LDH are plotted in Fig. 1d. In the
same manner, the pre-edge around 7113 eV and the main edge
around 7132 eV could be attributed to the 1s to 3d and the 1s to
4p transitions, respectively.37 In this graph, Fe2O3 and Co–Fe
LDH had similar main edges and shapes. This indicated that
iron was present as Fe3+ bonded with oxygen in the Co–Fe LDH.
Fig. 2a and S6† show the UV-Vis diffuse reectance spectra of
the Co–Fe LDHs, which were converted by the Kubelka–Munk
function and linear extrapolation to energy vs. F(R). All the
Co–Fe LDHs could absorb the visible light wavelengths,
especially below 550 nm. The peak below 500 nm corresponding
to the 2.25 eV band gap is due to the metal-to-metal charge
transfer (MMCT) leading to efficient water oxidation, while the
Fig. 1 (a) PXRD patterns of various Co–Fe LDHs. (b) HR-TEM image of peaks above 500 nm were attributed to the metal-to-ligand
CFL2. (c) Co K-edge and (d) Fe K-edge XANES spectra of a Co–Fe LDH charge transfer (MLCT) giving a small contribution to water
(CFL4). oxidation.17,22 The MLCT could be explained by a hexaaquo

This journal is © The Royal Society of Chemistry 2014 J. Mater. Chem. A, 2014, 2, 4136–4139 | 4137
View Article Online

Journal of Materials Chemistry A Communication


Published on 14 January 2014. Downloaded by Memorial University of Newfoundland on 18/07/2014 09:04:05.

Fig. 3 (a) Time-dependent evolved oxygen under visible light by Co–


Fe LDHs, CoO, and Fe2O3. (b) UV-Vis reflectance spectra of a Co–Fe
LDH (CFL4) suspension. The reduction of Ag+ ions into Ag nano-
particles induced a color change.

lower amounts compared to CFL2. The main reason for the


Fig. 2 (a) UV-Vis reflectance spectra of Co–Fe LDHs. (b) UPS of Co– reduced production was the number of oxo-bridges connecting
Fe LDH (CFL4). (c) Schematic model for the band structure of a Co–Fe the two different metals. In the photocatalytic reactions of
LDH having a proper band edge to produce oxygen using visible light. LDHs, the oxo-bridges play signicant roles. CFL2 had more
oxo-bridges than CFL3 or CFL4 because a greater amount of
iron atoms allowed more connections with cobalt. Furthermore,
cation, [M(H2O)6]n+, having a similar environment of transition CFL2 was more crystalline than CFL3 and CFL4, especially with
metals in the LDH.38 The cobalt hexaaquo cation ([Co(H2O)6]2+) respect to the non-basal planes which indicate the crystallinity
and the iron hexaaquo cation ([Fe(H2O)6]3+) have pink (mixed of the metal hydroxide layer. Also, CFL2 absorbed more visible
red and a little blue) and pale violet colors, respectively.39,40 light than the other Co–Fe LDHs according to the UV-Vis
Therefore, their absorption regions are green (480–550 nm), spectra. In control experiments, we found that hematite and
orange (590–630 nm), and yellow (550–590 nm). In other words, cobalt oxides produced only about 5 mmol and 0.7 mmol oxygen,
this material absorbed mainly green light from the cobalt cation respectively, under the same conditions. This indicated that the
with a small amount of yellow and orange light from cobalt and bonds formed between the Co and Fe ions in the LDH structure
iron cations. 630 nm corresponds to the absorption edge of the hindered the recombination of charge carriers, thereby
yellow color and the quantum yield of the water oxidation increasing its catalytic efficiency. The Ni–Ti and Cu–Ti LDH
reaction17,22 sharply was found to be decreased as the wave- structures reported in our previous study22 evolved about 200
length increased. Therefore, the efficient water oxidation is mmol g 1 and 125 mmol g 1 oxygen, respectively, during a 3 h
mostly attributed to the green light, while the yellow light at catalytic test. This demonstrated that the Co–Fe LDH catalysts
around 630 nm attributed to the MLCT gives a marginal for water oxidation under visible light were more than 450%
contribution to the reaction. Ultraviolet photoelectron more efficient compared to titanium-embedded structures.
spectroscopy (UPS) experiments, as seen in Fig. 2b, were used to Fig. S7† shows that the color of the Co–Fe LDH suspension
ascertain whether the Co–Fe LDHs could be used as was intensied as the irradiation time increased, which was
photocatalysts for water oxidation. The valence band maximum attributed to the reduction of Ag+ ions into Ag nanoparticles.29
(VBM) was calculated from the double tangent of the kinetic From the UV-Vis reectance spectra of a Co–Fe LDH (CFL4)
energy of the electrons released from the valence band of the suspension shown in Fig. 3b, it was apparent that most of the
Co–Fe LDH. The calculated VBM level of the Co–Fe LDH was 0.9 V light was reected at the start of the photocatalytic reaction.
vs. the normal hydrogen electrode (NHE) at pH 7.0, while the However, the reectance decreased gradually as the reaction
conduction band minimum (CBM) was 1.35 V vs. NHE at pH proceeded, implying that the Co–Fe LDH oxidized water.
7.0. The band gap energy was determined from the UV-Vis spec-
trum, and the electronic band structure of Fig. 2c suggested that Conclusions
the Co–Fe LDH would be suitable for water oxidation as its VBM
level was lower than that for water oxidation (0.818 V, H2O–O2). In summary, we prepared Co–Fe LDH photocatalysts with
The photocatalytic abilities of the Co–Fe LDHs were also various Co–Fe ratios which exhibited highly efficient water
tested under visible light (400 nm < l < 700 nm) using a 300 W oxidation under visible light. Among the three Co–Fe LDHs, the
xenon lamp tted with UV and IR lters. The total amount of sample containing more iron displayed the highest efficiency.
evolved oxygen is shown in Fig. 3a. CFL2 produced about 45 This enhanced efficiency was attributed to the Co–O–Fe bridges
mmol oxygen from water over 3 h in the presence of light and in the LDH structure, which suppressed the fast recombination
AgNO3 as an electron scavenger. Oxygen was generated at a of holes with electrons. The nature of this bonding was eluci-
uniform rate, indicating that the Co–Fe LDH generated charges dated by various experimental techniques, and the suitability of
at a steady rate upon the absorption of visible light. CFL3 and the Co–Fe LDHs as water oxidation photocatalysts was
CFL4 also produced oxygen under the same conditions, albeit in demonstrated by optical measurements. We expect that this

4138 | J. Mater. Chem. A, 2014, 2, 4136–4139 This journal is © The Royal Society of Chemistry 2014
View Article Online

Communication Journal of Materials Chemistry A

work will lay a foundation for studies that will use transition 16 N. H. Chou, P. N. Ross, A. T. Bell and T. D. Tilley,
metals to increase the photocatalytic efficiency of LDHs for ChemSusChem, 2011, 4, 1566–1569.
water oxidation. 17 C. G. Silva, Y. Bouizi, V. Fornes and H. Garcia, J. Am. Chem.
Soc., 2009, 131, 13833–13839.
Acknowledgements 18 V. Rives and S. Kannan, J. Mater. Chem., 2000, 10, 489–495.
19 R. Ma, Z. Liu, K. Takada, N. Iyi, Y. Bando and T. Sasaki, J. Am.
Chem. Soc., 2007, 129, 5257–5263.
Published on 14 January 2014. Downloaded by Memorial University of Newfoundland on 18/07/2014 09:04:05.

This research was supported by the Korea Center for Articial


Photosynthesis (KCAP) located in Sogang University (2009- 20 Z. Liu, R. Ma, M. Osada, N. Iyi, Y. Ebina, K. Takada and
0093881) and the Global Frontier R&D Program (2013-073298) T. Sasaki, J. Am. Chem. Soc., 2006, 128, 4872–4880.
on Center for Hybrid Interface Materials (HIM) funded by the 21 T. Hibino and M. Kobayashi, J. Mater. Chem., 2005, 15, 653–
Ministry of Science, ICT & Future Planning. In addition, J. K. 656.
Kang was supported by the National Research Foundation of 22 Y. Lee, J. H. Choi, H. J. Jeon, K. M. Choi, J. W. Lee and
Korea (2010-0029042, 2011-0028737, 2011-0031407, 2012- J. K. Kang, Energy Environ. Sci., 2011, 4, 914–920.
0001174). The XANES analysis was supported by PAL through 23 R. Nakamura, A. Okamoto, H. Osawa, H. Irie and
the abroad beamtime program of the Synchrotron Radiation K. Hashimoto, J. Am. Chem. Soc., 2007, 129, 9596–9597.
Facility Project under MEST and has been performed under the 24 K. Sinko, E. Manek, A. Meiszterics, K. Havancsak, U. Vainio
approval of the NSRRC (BL01C1). and H. Peterlik, J. Nanopart. Res., 2012, 14, 894.
25 H. Lin, Y. Zhang, G. Wang and J.-B. Li, Front. Mater. Sci.,
Notes and references 2012, 6, 142–148.
26 R. Nakamura and H. Frei, J. Am. Chem. Soc., 2006, 128,
1 N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 10668–10669.
2007, 104, 20142. 27 V. Rives, Mater. Chem. Phys., 2002, 75, 19–25.
2 J. L. Gunjakar, T. W. Kim, H. N. Kim, I. Y. Kim and 28 C. Jaubertie, M. J. Holgado, M. S. San Roman and V. Rives,
S. J. Hwang, J. Am. Chem. Soc., 2011, 133, 14998–15007. Chem. Mater., 2006, 18, 3114–3121.
3 M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072– 29 M. Vucelic, W. Jones and G. D. Moggridge, Clays Clay Miner.,
1075. 1997, 45, 803–813.
4 N. S. Lewis, Science, 2007, 315, 798–801. 30 J. B. Yu, Z. Jiang, L. Zhu, Z. P. Hao and Z. P. Xu, J. Phys. Chem.
5 G. W. Crabtree and N. S. Lewis, Phys. Today, 2007, 60, 37–42. B, 2006, 110, 4291–4300.
6 A. Kudo, Catal. Surv. Asia, 2003, 7, 31–38. 31 Y. Zhang, L. Q. Wang, L. J. Zou and D. F. Xue, J. Cryst. Growth,
7 A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253–278. 2010, 312, 3367–3372.
8 P. W. Du and R. Eisenberg, Energy Environ. Sci., 2012, 5, 32 K. Parida, M. Satpathy and L. Mohapatra, J. Mater. Chem.,
6012–6021. 2012, 22, 7350–7357.
9 M. C. Toroker, D. K. Kanan, N. Alidoust, L. Y. Isseroff, 33 S. Miyata, Clays Clay Miner., 1983, 31, 305–311.
P. L. Liao and E. A. Carter, Phys. Chem. Chem. Phys., 2011, 34 S. J. Palmer, R. L. Frost and T. Nguyen, Coord. Chem. Rev.,
13, 16644–16654. 2009, 253, 250–267.
10 M. J. Katz, S. C. Riha, N. C. Jeong, A. B. F. Martinson, 35 M. Z. Hussein, A. M. Jaafar, A. H. Yahaya and Z. Zainal,
O. K. Farha and J. T. Hupp, Coord. Chem. Rev., 2012, 256, Nanoscale Res. Lett., 2009, 4, 1351–1357.
2521–2529. 36 F. de Groot, G. Vanko and P. Glatzel, J. Phys.: Condens.
11 A. Duret and M. Gratzel, J. Phys. Chem. B, 2005, 109, 17184– Matter, 2009, 21, 104207.
17191. 37 P. Sipos, D. Zeller, E. Kuzmann, A. Vertes, Z. Homonnay,
12 K. Sivula, F. Le Formal and M. Gratzel, ChemSusChem, 2011, M. Walczak and S. E. Canton, Dalton Trans., 2008, 5603–
4, 432–449. 5611.
13 L. F. Xi, S. Y. Chiam, W. F. Mak, P. D. Tran, J. Barber, 38 M. delArco, V. Rives, R. Trujillano and P. Malet, J. Mater.
S. C. J. Loo and L. H. Wong, Chem. Sci., 2013, 4, 164–169. Chem., 1996, 6, 1419–1428.
14 F. Jiao and H. Frei, Chem. Commun., 2010, 46, 2920–2922. 39 B. Hathaway and C. Lewis, J. Chem. Soc. A, 1969, 1183–
15 A. J. Esswein, M. J. McMurdo, P. N. Ross, A. T. Bell and 1188.
T. D. Tilley, J. Phys. Chem. C, 2009, 113, 15068–15072. 40 I. Persson, Pure Appl. Chem., 2010, 82, 1901–1917.

This journal is © The Royal Society of Chemistry 2014 J. Mater. Chem. A, 2014, 2, 4136–4139 | 4139

You might also like