You are on page 1of 6

www.advmat.

de
www.MaterialsViews.com

COMMUNICATION
Optically Tunable Amino-Functionalized Graphene
Quantum Dots
Hiroyuki Tetsuka,* Ryoji Asahi, Akihiro Nagoya, Kazuo Okamoto, Ichiro Tajima,
Riichiro Ohta, and Atsuto Okamoto

Chemically-derived graphene quantum dots (GQDs) are quantum-confinement and edge effect of quantum-sized
emerging as optical materials for prospective applications in graphene, chemical doping is an effective way to tailor their
bio-imaging, light-emitting, and photovoltaic.[1–7] One of the electronic characteristics. For instance, Li et al. demonstrated
most notable issues is their potential for replacing commonly that direct substitution with nitrogen (N) in the GQDs lattice
used semiconductor nanocrystals including toxic/expensive induces the modulation of the chemical and electronic charac-
heavy-metals, in which the versatile tunability of optical prop- teristics of the GQDs, because of high electronic affinity of the
erties, together with solution processability, has led to various N atom.[12] Although the N-doped GQDs exhibited a blue shift
applications. However, because of the non-stoichiometric in the PL emission from their N-free counterparts, such incor-
nature of GQDs, tailor-made control of their optical proper- poration of heteroatoms in the graphene lattice would disrupt
ties is extremely challenging and is one of the key technologies sp2 hybridization of carbon (C) atoms, and tailor-made control
for putting these materials into practical use. Here, we report of optical properties is still lacking. Chemical modifications
optically tunable amino-functionalized graphene quantum dots using molecules with strong electron-donating or -accepting
(af-GQDs) with discrete molecular weights and specific edges, ability, e.g., primary amines, could also drastically impact on
extracted through the ammonia-mediated bond-scission reac- electronic characteristics of graphene,[13,14] thereby GQDs with
tion from oxidized graphene sheet. The af-GQDs exhibit clear tunable properties must be realized through the selective and
multiple colors under a single-wavelength UV excitation, and quantitative functionalization of primary amines.
their luminance quantum yields are over 40%. The optical prop- To tailor the optical properties, we have designed a novel
erties are precisely controlled only by the quantitative function- graphene nanostructure that is edge-terminated by a primary
alization of af-GQDs, supported by MALDI-TOF mass analysis amine, allowing the electronic structure to be modified system-
and density functional theory calculations. atically through the effective orbital resonance of amine moi-
The primary approach to fabricate soluble fluorescent GQDs eties with graphene core. Figure 1a presents our strategy for
was realized by either cutting graphene sheet through a con- the production of af-GQDs that starts from oxidized graphene
trolled oxidation or reduction process,[2,4,6,7] or constructing a sheets (OGSs) (see Synthetic Methods in Supporting Infor-
small precursor aromatic molecule to larger one followed by mation for experimental details; Supporting Figure S1–S6).
oxidative exfoliation.[8] The presence of finite-sized molecular We focused on a unique heterogeneous chemical struc-
sp2 domains and/or CO-related quasi-molecular fluorophores, ture of graphene oxide in which aromatic sp2 domains of
induced by adsorption of oxygen functional groups, could con- 2–3 nm, which are surrounded by linearly aligned epoxy and/or
fine π electrons and consequently give fluorescence to be dic- hydroxyl-bonded sp3 C atoms, are isolated within a continuous
tated by the nature of sp2 domains and functional groups.[9–11] sp3 C⫺O matrix. The graphitic sp2 domains in the pristine
However, despite their different oxidation states and scales, OGSs we prepared were calculated to have a mean size of
these GQDs exhibit similar blue fluorescence whose maximum ∼2.7 nm according to the empirical Tuinstra–Koenig relation-
is centred at 420–450 nm, which is mainly ascribed to the ship,[15] which is based on the Raman D/G peak-area ratio (Sup-
limited tunability of size and/or shape of sp2 domains and porting Figure S2). A well-defined GQDs must be produced by
adsorption of oxygen functional groups, arising from process extracting sp2 domains from oxidized graphene sheet via a bond
that focused only on insertion or removal of oxygen functional scission of surrounding oxygen groups. To extract sp2 domains
groups. So far, the ideal tunable PL from GQDs has not yet without deconstructing their graphitic structure, we developed
been successful, and it should be possible by enhancing the a soft self-limiting bond-scission process. In fact, OGSs were
homogeneity of local electronic modifications of graphene subjected to mild amino-hydrothermal treatment at 70–150 °C
through the precise control of sp2 domain size, edge structure, using ammonia solution, followed by thermal annealing at
and surface functionality of GQDs. In the view of remarkable 100 °C. In the treatment of OGSs, ammonia reacts with epoxy
groups to form a primary amine and alcohols by nucleophilic
substitution (Supporting Figure S3a),[16,17] and thus it enables
Dr. H. Tetsuka, Dr. R. Asahi, A. Nagoya, K. Okamoto, self-limited extraction of the sp2 domains by ring-opening of
I. Tajima, Dr. R. Ohta, A. Okamoto the epoxide and the simultaneous direct bonding of a primary
Toyota Central R&D Laboratories Inc.,Nagakute amine with a graphene edge. Consequently, size-controlled
Aichi 480-1192, Japan
GQDs that are edge-terminated with a primary amine are pro-
E-mail: h-tetsuka@mosk.tytlabs.co.jp
duced, as confirmed by the structural analyses (Figure 1b, 1c,
DOI: 10.1002/adma.201201930 Figure 2 and Supporting Figure S4, S5).

Adv. Mater. 2012, 24, 5333–5338 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 5333
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

Figure 1. a) Schematic illustration of the preparative strategy for af-GQDs. b) C1s and N1s X-ray photoelectron spectra for af-GQDs prepared at
90 °C (green line) and OGS (gray line). The pronounced decrease in the C 1s signal at 286.4 eV, which is assigned to epoxide groups, is a manifesta-
tion of the loss of epoxides through nucleophilic reaction with ammonia. This interpretation is supported by the appearance of the N 1s signal at
399.7 eV, which is assigned to primary amines (⫺NH2) bonded to the graphene. c) Transmission infrared-absorbance spectra of af-GQDs prepared
at 90 °C (green line) and OGS (gray line). New peaks at 1243, 1617, and 3300–3600 cm−1 appeared after the amino-hydrothermal treatment. These
peaks were assigned, respectively, to C⫺N in-plane, N⫺H out-of-plane, and N⫺H in-plane stretching of the amine groups. Additionally, characteristic
amide–carbonyl (⫺NH⫺CO⫺) stretching vibration was observed at 1650 cm−1, which implies the formation of amide groups through interactions
with the carboxylic groups as Lewis acids.

Transmission electron microscopy (TEM) images veri- functionalization, and hence the amino-groups might be pre-
fied the production of uniform af-GQDs with a diameter of ponderantly located in the edges of graphene layers.
∼2.5 nm (Figure 2a, 2c and Figure S5). No noticeable change The extracted af-GQDs exhibit efficient photoluminescence,
was observed in their size and shapes for the af-GQDs prepared and their fluorescence can be tuned from the violet to the yellow
at 70–120 °C, while high temperature amino-hydrothermal region thorough the functionalization of the primary amine.
treatment (>150 °C) induced the fragmentation into smaller The degree of amine functionalization, which controls the photo-
sizes (Figure S5). The corresponding atomic Force Microscope luminescence color, can be controlled simply by changing the
(AFM) image (Figure 2b and 2d) revealed that most of af-GQDs initial concentration of ammonia and the temperature of the
have a thickness of ∼1.13 nm, which corresponds to single- amino-hydrothermal treatment (Supporting Figure S4). It was
layer of functionalized graphene quantum dots.[2,6,12] Raman found that low temperature synthesis is needed for successful
spectrum also provided convincing evidence for the graphene amine functionalization. Higher temperature, i.e. >120 °C,
fragment structure of the af-GQDs. The peak at ∼1600 cm−1 (G) induces dissociation of primary amine after nucleophilic reac-
corresponds to the E2g mode of graphite and is related to the tions with ammonia, resulting lower C/N ratio. The thermal
vibration of the sp2-bonded C atoms in a two-dimensional hex- stability of amine-functional group on af-GQDs was evalu-
agonal lattice, while the peak at ∼1370 cm−1 (D) exhibits disorder ated using thermogravimetric analysis (Supporting Figure S6).
due to scattering at the edges (Figure 2e). The relative inten- Figures 3a and 3b show the emission images and fluorescence
sity of D/G band is ∼0.6, similar to that of electrochemically- spectra from af-GQDs. Aqueous suspensions of af-GQDs exhibit
derived graphene quantum dots[2,12] and few-layer graphene bright colorful luminescence under irradiation from a UV lamp
nanoribbons,[18,19] which indicates the high quality of af-GQDs. with a wavelength of 365 nm (upper images of Figure 3a); their
Figure 2f gives X-ray diffraction (XRD) patterns of the af-GQDs corresponding emission peaks vary with the quantity of amine
prepared at 90 °C and OGSs dried powders. A weak broad (002) functionalization in the range of 420–535 nm and retain a sharp
peak centred at 2θ = ∼22.7° and a disappearance of peak at full-width at half-maximum (FWHM) as narrow as ∼65–80 nm.
2θ = ∼10.6° in the spectrum for af-GQDs indicate the disor- A clear red shift of absorption band with the quantity of amine
dered stacking structures of graphene layers and the absence of functionalization was also observed in the UV-visible absorp-
oxygen functional groups in between interlayers, respectively. tion spectra of af-GQDs. For GQDs, the optical properties are
The XRD and AFM results suggest the formation of exfoli- known to be dictated by a combination of a couple of factors,
ated monolayer graphene layers.[20] Moreover, as confirmed by i.e., size, shape, and surface functionality of GQDs. Because the
X-ray photoelectron spectroscopy (XPS) and Fourier transform af-GQDs prepared at 70–120 °C have almost same shape and
infrared spectroscopy (FT-IR) analyses, the produced af-GQDs size, the tunable photoluminescence should be originated only
contain large amount of amino-groups (Figure 1b and 1c). from the difference in the quantity of functionalized amine-
Therefore, it is supposed that there is no significant basal plane groups, at least in the blue to yellow region. The emission

5334 wileyonlinelibrary.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2012, 24, 5333–5338
www.advmat.de
www.MaterialsViews.com

COMMUNICATION
passivation with polyethylene glycol sig-
nificantly enhanced the photoluminescence
QY to ∼46%, by the stabilization effect of
excitons in the GQDs.[22] Our af-GQDs also
exhibit excellent compatibility with various
organic polymers, which enables their incor-
poration into solid matrices without further
chemical modifications. The bottom images
in Figure 3a show flexible polydimethylsi-
loxane (PDMS) elastomer with a luminescent
piece of af-GQDs@PDMS hybrid patterned
in the logo of our institute. The af-GQDs
retain their fluorescent properties upon inte-
gration into polymer matrices without any
noticeable change in the shape or FWHM of
their emission spectrum peaks (Supporting
Figure S7), suggesting that no significant
surface deterioration or aggregation of the
af-GQDs occurred.
Mass spectrometry using the matrix-
assisted laser desorption/ionization reflec-
tron time-of-flight technique (MALDI-
TOF mass) revealed their narrow size
distribution and unique edge structures. The
af-GQDs prepared at 70–120 °C display similar
masses, consistent with the TEM observation
showing a rather uniform size of af-GQDs
to be ∼2.5 nm. To our surprise, regular pat-
terns with the repeating unit of m/z 74
were observed between m/z 1200 and 2500
in the mass spectrum of af-GQDs prepared
at 120 °C (Figure 4a). Amino-hydrothermal
treatment at temperatures greater than
150 °C induces breakage into smaller frag-
ments with no regular mass pattern (Sup-
porting Figure S8). To our knowledge, such
a regular pattern has never been previously
reported in a mass spectrum for graphene or
graphene oxide. Starting OGSs, purified via
Figure 2. a) TEM image of the af-GQDs prepared at 90 °C. b) AFM image of the af-GQDs same purification process without applying
prepared at 90 °C on mica substrate. The inset shows the height profile along the line. amino-hydrothermal treatment, showed a
c) The size distributions of af-GQDs. The average size is ∼2.5 nm. d) The height distributions broad mass spectrum with no regularity
of af-GQDs. The average height is ∼1.13 nm. e) Raman spectrum of the af-GQDs prepared (Figure S9) because of ‘non-selective’ frag-
at 90 °C. f) XRD patterns of the af-GQDs prepared at 90 °C and OGSs dried powders. mentation. This fact confirms the selective
fragmentation of af-GQDs with specific edges
wavelength increased with the contents of amine-groups. For and size. We attribute this repeating unit of m/z 74 to the reflec-
af-GQDs prepared at 150 °C, on the other hand, size effect may tion of their edge structure, which were verified by optimised
also make an important contribution to the observed violet geometric calculations based on molecular mechanics methods
emission, because the optical band gap should be increased using augmented MM3 parameters (details in Supporting
with decreasing sp2 domain size due to the quantum confine- Information). The mass of 74 is expected to be a six-membered
ment effects.[9,21] The photoluminescence quantum yield (QY) C ring bonded to an armchair edge of an af-GQD (schematically
calibrated against quinine sulphate as a standard sample (54% shown in the inset of Figure 4a): 74 Δm = –[loss of two-coordinate
in 0.5 M H2SO4) was found to be ∼29–19%, which decreased six-membered C ring bonded to an armchair edge (Mw = 76)] +
with quantity of amine functionalization. This is much higher [termination of two hydrogen atoms (Mw = 2)]. Here, we con-
than that of previously reported GQDs (typically <10%). The sider two types of bonds in the armchair-edged GQDs: one
high PL QYs may be attributed to the chemical nature of af- directed nearly along the axis, CCa, and another directed along
GQDs. The af-GQDs contain less carboxylic and epoxide groups the circumference, CCc (Supporting Figure S10). The optimised
acting as the non-radiative electron-hole recombination centers, geometric calculations predict that the GQDs structure without
which leads to the efficient emission. Furthermore, surface outermost CCc bonds are more stable than that without the CCa

Adv. Mater. 2012, 24, 5333–5338 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 5335
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

The different luminescence colors


should originate from a change in the
electronic structure of the af-GQDs. It has
been reported that N doping in pyridinic
sites induces blue shift in photolumines-
cence maximum because of high elec-
tronic affinity of the nitrogen atom.[12]
This behavior is inconsistent with our
tendency. To elucidate this point, we per-
formed ab initio calculations using the
plane-wave projector augmented wave
(PAW) method (details in Supporting
Information).[23,24] Based on the MALDI-
TOF mass, IR, and XPS structural anal-
yses, we modelled a graphene structure
Figure 3. a) Emission images of af-GQDs dispersed in water (upper) and af-GQD@polymer with two-coordinate six-membered C rings
hybrids (bottom) under irradiation from a 365 nm UV lamp. The af-GQDs in the upper images connected to armchair edges, as illus-
were prepared under the following conditions: 150 °C, 120 °C, 70 °C, 90 °C, and 90 °C in ×2
trated in Figure 4c. This figure also shows
conc. ammonia solution, respectively from the left. The bottom images show flexible poly-
dimethylsiloxane (PDMS) elastomer with a luminescent piece of af-GQDs@PDMS hybrid pat- the isosurfaces of the highest occupied
terned in the logo of our institute. b) Photoluminescence and selected UV/VIS absorption molecular orbital (HOMO) and the lowest
spectra of af-GQDs. The photoluminescence spectra are those of the af-GQDs shown in the unoccupied molecular orbital (LUMO)
upper image of Figure 1a. In the absorption spectra, solid and broken lines denote the spectra (details in Supporting Figure S13). The
of af-GQDs prepared at 120 °C and 90 °C, respectively. influence on the optical properties of af-
GQDs of the adsorption of primary amine
bonds, which results in one- or two-coordinate six-membered groups on their edges was investigated. A primary amine
C rings bonded to an inner armchair edge. The structure with at the edge significantly alters the whole electronic struc-
two-coordinate C ring at armchair edge is thermodynamically ture of an af-GQD; one of the degenerate HOMO orbitals
less stable than those with a one-coordinate ring, with a differ- in the hydrogen-terminated graphene molecule is lifted to a
ence in energy of 0.48 eV. The loss of two-coordinate C ring is higher energy in the af-GQD as a consequence of a strong
vital to cause the difference of m/z 74 in the mass pattern orbital interaction with the ⫺NH 2 amino group. This
(Supporting Figure S11). interaction results in a narrowing of the optical band gap
Intriguingly, the mass pattern from af-GQDs prepared at (Supporting Figure S14). The optical band gap decreases
90 °C shifted to a higher mass by precisely m/z 29 and retained with the quantity of primary amine. The amount of the
the regularity of the repeating m/z 74 unit, as shown in Figure 4b. bandgap narrowing is rather independent of the size of
The difference of m/z 29 is attributable to the attachment the molecules, confirmed by different size of models, C6H6
of additional functional groups at the two-coordinate C-ring to C36H 18 (Supporting Table S1). Looking into other func-
edge sites. FT-IR analysis showed a distinct increase in the tional groups such as ⫺CH3 and ⫺NO, we observed that
intensity of the N-H stretching signal and an appearance of the former affects neither the HOMO nor the LUMO. We
the C⫺H (⫺CH3) stretching signal from af-GQDs prepared also observed that ⫺NO induces states in the bandgap that
at 90 °C. XPS measurements also revealed that the C to N are characterized by localization in the NO group and by
molar ratio decreased to 8.3:1 for the sample prepared at 90 °C no sign of strong interaction with the original graphene
from 26:1 for the sample prepared at 120 °C. Consequently, orbitals (Supporting Figure S15). The resonance feature
the additional functional groups are assumed to at least have between the delocalized π orbital and the molecular orbital
amino group, i.e., ⫺NH2/⫺CH3 pair. The ⫺NH2/⫺CH3 pair to in the ⫺NH 2 amino group is thus an origin of both the
be attached into the two-coordinate C-ring sites, as estimated edge-treatment-dominated optical tunability and the high
from consideration of the mass difference: the attachment of quantum efficiency that is observed experimentally. In
⫺NH2/⫺CH3 pair at the two-coordinate C ring induces the addition, such a resonance orbital can lead to the process
molecular weight by 29 mass unit, while the bonding into of intersystem crossing emission as a result of enhanced
basal plane of af-GQDs makes 31 mass difference. Importantly, spin–orbital coupling. In fact, we observed extraordinary
the corresponding emission color of the af-GQDs switched long delayed luminescence of sub-second order from
to green from blue after this functionalization (shown in af-GQDs (details in Supplementary Figure S15).
the inset of Figure 4b). This result implies that the lumines- In summary, we have demonstrated novel amino-functional-
cence in af-GQDs is greatly affected by the edge structure ized GQDs with high efficiency and wide tunability of narrow
because no difference in size or shape was observed. This photoluminescence lines under a single-wavelength UV excita-
supposition is supported by the fact that blue and green tion, whose energy are controlled by their edge-termination
af-GQDs can be distinguished only using chromatographic structure with primary amine. Their superior optical properties
technique, which separates af-GQDs into different polarities should enable the use of af-GQDs in numerous applications,
depending on their degree of amino-functionalization at the including multicolor light-emitting devices, biological applica-
edge (Supporting Figure S12). tions, and photovoltaics.

5336 wileyonlinelibrary.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2012, 24, 5333–5338
www.advmat.de
www.MaterialsViews.com

COMMUNICATION
Figure 4. a) MALDI-TOF mass spectrum of af-GQDs in aqueous solution prepared at 120 °C. Note that peaks are detected as Na ion adducts ([M +
Na]+), the Na ion of which might have originated from residue in the reagents used for graphite oxidation to prepare the graphene oxide. Inset, sche-
matic representation of the edge structure of the af-GQDs. b) Comparison of MALDI-TOF mass spectra for af-GQDs prepared at 120 °C (bottom)
and 90 °C (upper). The inset shows the possible edge structures and their photoluminescence spectra excited at 350 nm. c) Schematic illustrations of
structures used for theoretical calculations with different functional groups. The isosurface represents the highest occupied molecular orbital (HOMO)
and the lowest unoccupied molecular orbital (LUMO). I: hydrogen-terminated edge structure, II: ⫺NH2/⫺CH3 pair bonded to edge. The unit of energy
is electron volt (eV).

Supporting Information [1] K. P. Loh, Q. Bao, G. Eda, M. Chhowalla, Nat. Chem. 2010, 2, 1015.
Supporting Information is available from the Wiley Online Library or [2] Y. Li, Y. Hu, Y. Zhao, G. Shi, L. Deng, Y. Hou, L. Qu, Adv. Mater.
from the author. 2011, 23, 776.
[3] V. Gupta, N. Chaudhary, R. Srivastava, G. D. Sharma, R. Bhardwaj,
S. Chand, J. Am. Chem. Soc. 2011, 133, 9960.
[4] S. Zhu, J. Zhang, C. Qiao, S. Tang, Y. Tang, Y. Li, W. Yuan, B. Li,
Acknowledgements L. Tian, F. Liu, R. Hu, H. Gao, H. Wei, H. Zhang, H. Sun, B. Yang,
The authors thank Y. Kaneko, H. Takeuchi, O. Watanabe, A. Usuki, and Chem. Commun. 2011, 47, 6858.
S. Inagaki for fruitful discussions, Y. Kato for Raman analysis, [5] X. Sun, Z Liu, K. Welsher, J. T. Robinson, A. Goodwin, S. Zaric,
N. Takahashi for XPS analysis, N. Suzuki for TEM observations, and H. Dai, Nano Res. 2008, 1, 203.
M. Horii for technical assistance. [6] D. Pan, J. Zhang, Z. Li, M. Wu, Adv. Mater. 2010, 22, 734.
[7] J. Shen, Y. Zhu, C. Chen, X. Yang, C. Li, Chem. Commun. 2011, 47,
Received: May 14, 2012 2580.
Revised: June 11, 2012 [8] R. Liu, D. Wu, X. Feng, K. Mullen, J. Am. Chem. Soc. 2011, 133,
Published online: July 25, 2012 15221.

Adv. Mater. 2012, 24, 5333–5338 © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 5337
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

[9] G. Eda, Y.-Y. Lin, C. Mattevi, H. Yamaguchi, H.-A. Chen, [17] L. Lai, L. Chen, D. Zhan, L. Sun, J. Liu, S. H. Lim, C. K. Poh, Z. Shen,
I.-S. Chen, C.-W. Chen, M. Chhowalla, Adv. Mater. 2010, 22, J. Lin, Carbon 2011, 49, 3250.
505. [18] L. Y. Jiao, X. R. Wang, G. Diankov, H. L. Wang, H. J. Dai, Nat.
[10] T. Gokus, R. R. Nair, A. Bonetti, M. Bohmler, A. Lombardo, Nanotechnol. 2010, 5, 321.
K. S. Novoselov, A. K. Geim, A. C. Ferrari, A. Hartschuh, ACS Nano [19] L. Y. Jiao, L. Zhang, X. R. Wang, G. Siankov, H. J. Dai, Nature 2009,
2009, 3, 3963. 458, 877.
[11] C. Galande, A. D. Mohite, A. V. Naumov, W. Gao, L. Ci, A. Ajayan, [20] S. Mikhailov, Physics and Applications of Graphenes–Experiments,
H. Gao, A. Srivastava, R. B. Weisman, P. M. Ajayan, Sci. Rep. 2011, InTech. Croatia (ISBN978-953-307-217-3) 2011, chap. 5.
1, 85. [21] J. Peng, W. Gao, B. K. Gupta, Z. Liu, R. R. Aburto, L. Ge,
[12] Y. Li, Y. Zhao, H. Cheng, Y. Hu, G. Shi, L. Dai, L. Qu, J. Am. Chem. L. Song, L. B. Alemany, X. Zhan, G. Gao, S. A. Vithayathil,
Soc. 2012, 134, 15. B. A. Kaipparettu, A. A. Marti, T. Hayashi, J. J. Zhu, P. M. Ajayan, Nano
[13] X. Dong, D. Fu, W. Fang, Y. Shi, P. Chen, L. J. Li, Small. 2009, 5, Lett. 2012, 12, 844.
1422. [22] Y. P. Sun, B. Zhou, Y. Lin, W. Wang, K. A. S. Fernando, P. Pathak,
[14] B. Das, R. Voggu, C. Sekhar, C. N. R. Rao, Chem. Commun. 2008, M. J. Meziani, B. A. Harruff, X. Wang, H. Wang, P. G. Luo, H. Yang,
5155. M. E. Kose, B. Chen, L. M. Veca, S.-Y. Xie, J. Am. Chem. Soc. 2006,
[15] F. Tuinstra, J. L. Koenig, J. Chem. Phys. 1970, 53, 1126. 128, 7756.
[16] M. Seredych, T. J. Bandoz, J. Phys. Chem. C 2007, 111, [23] G. Kresse, J. Furthmuller, Phys. Rev. B 1996, 54, 11169.
15596. [24] G. Kresse, J. Furthmuller, Comput. Mater. Sci. 1996, 6, 15.

5338 wileyonlinelibrary.com © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2012, 24, 5333–5338

You might also like