Sie sind auf Seite 1von 6

www.advmat.

de
www.MaterialsViews.com

COMMUNICATION
Hydrogenated TiO2 Nanocrystals: A Novel Microwave
Absorbing Material
Ting Xia, Chi Zhang, Nathan A. Oyler, and Xiaobo Chen*

Dipole rotation and ferromagnetic resonance are well-known property changes in the hydrogenated TiO2 nanocrystals caused
mechanisms normally responsible for microwave absorption by the hydrogenation.
due to the alignment of the polar groups in the materials to Here hydrogenation promoted the growth of pure TiO2
the electromagnetic field in the microwave field[1–3] or the anatase nanocrystals and a phase transition into a mixture
strongly interacting electron spins in the ferromagnet in the of anatase and rutile phases at the temperature as low as
ferromagnetic filler particles (e.g., iron or cobalt) dispersed 450 °C (Figure 1A).[20] Normally phase transition from anatase
in a matrix.[1–3] TiO2 is not among the traditional microwave to rutile TiO2 nanocrystals requires annealing at temperatures
absorbing material (MAM) candidates due to its poor response above 600 °C.[8] The crystalline grain size was estimated with
in the microwave region. Here, we demonstrate that hydro- the Scherrer equation from the XRD pattern:[21] 19.0 nm for
genation successfully turns this inactive MAM into an exciting pure TiO2 nanocrystals, 45.0 nm for the anatase phase and
MAM with an innovative collective-movement-of-interfacial- 50.0 nm for the rutile phase in hydrogenated TiO2 nanocrys-
dipole (CMID) mechanism. This study thus opens up new con- tals. Hydrogenation also induced amorphous phases as seen
cepts for MAM innovations, expands the applications of TiO2 from the broad photoluminescence background in the treated
and other oxide materials into new areas beyond photocatal- TiO2 nanocrystals beyond the characteristic anatase Raman
ysis,[4–8] battery[5,9] and solar cells,[5,8–10] and will probably attract features at around 142.8, 201.4, 399.1, 520.0, and 642.4 cm−1
the attention of materials scientists. (Figure 1B).[7,8] The disappeared Raman features and the
Gigahertz radar MAMs are very important in many civil (e.g., apparent XRD rutile features suggested rutile-anatase interior-
communications, noise reduction, signal and data protection) exterior structures of the hydrogenated TiO2 nanocrystals, as
and military (e.g., the stealth coating in B-2 bomber) applica- Raman is more sensitive to the composition on the surface.[7]
tions.[11–18] Traditional candidates include ferroelectric ceramics The growth and surface amorphization from hydrogena-
and carbonaceous materials.[3,11–18] Ferroelectric ceramics exam- tion were obvious from the transmission electron microscopy
ples are ferrites and ferromagnetic materials,[11–15] barium and (TEM) images and the fast Fourier transforms (FFT) results:
lead zirconate titanate.[3] However, due to high reflection coef- 18 nm pure TiO2 nanocrystals (Figure 1C) with clearly-defined
ficients at the interface with air, ceramic ferroelectrics cannot lattice fringe edges (Figure 1D) and clear diffraction spots in
be used as such.[3] Carbonaceous materials include carbonyl- the FTT (Figure 1C) vs. 50 nm hydrogenated TiO2 nanocrystals
iron[17,18] and carbon nanotubes[11,16,18] with permittivity values (Figure 1E) with an inner well-defined lattice fringe wrapped
of 10–16 in the frequency range of 2–18 GHz[10] and concern of with an outer 2-nm amorphous layer (Figure 1F) and a diffrac-
flame resistance. Pure TiO2 is not among the traditional MAM tion cloud around the central spot in the FTT (Figure 1E).[26,27]
candidates, due to its poor response in the microwave gigahertz The microwave absorption properties are evaluated with
region although it has a good static dielectric constant.[1,6] the complex permittivity and permeability values: the real part
Hydrogenation has recently emerged as an innovative ε′ and μ′, related to the stored electrical and magnetic energy
approach in effectively modifying the properties of materials, within the medium, the imaginary part ε′′ and μ′′, related to the
and has shown great ability in TiO2 nanocrystals for intro- dissipation (or loss) of electrical and magnetic energy, and their
ducing dramatic structural changes[7,19–21] and large improve- ratios, the dielectric and magnetic dissipation factors tgδε =
ment in their optical,[5,7] electronic,[5,7] photocatalytic,[5,7] ε′′/ε′ and tgδμ = μ′′/μ′, which provide a measure of how much
lithium ion storage capability[22–24] and field emission.[25] As power is lost in a material versus how much is stored.[1–3,11–18]
we report here, hydrogenated TiO2 is an excellent MAM can- We dispersed the nanocrystals in melting paraffin wax, and
didate benefiting from an innovative CMID mechanism which cast the mixture into a ring for the microwave measure-
induces the collective interfacial polarization amplified micro- ments.[11] The anatase TiO2 nanocrystals/epoxy composites
wave absorption (CIPAMA) in hydrogenated TiO2 nanocrystals, displayed a relatively constant value of 6.1–6.5 for ε′ in the fre-
distinct from the traditional dipole or ferromagnetic resonance quency range of 1.0–18.0 GHz, while the hydrogenated TiO2
mechanisms. This CMID mechanism is based on the dramatic nanocrystals/epoxy composites showed much higher ε′ values
(Figure 2A). The ε′ decreased gradually from 25.7 at 1.0 GHz
to 17.4 at 14.50 GHz, then increased to 20.0 at 16.7 GHz, and
T. Xia, C. Zhang, Prof. N. A. Oyler, Prof. X. Chen decreased to at 17.7 at 20.0 GHz. As the ε′ value of the epoxy
Department of Chemistry was 2.5 at X-band frequency,[11] the ε′ value of the hydrogen-
University of Missouri – Kansas City ated TiO2 nanocrystals was thus estimated to be about 4.3 times
Kansas City, Missouri 64110, USA
E-mail: chenxiaobo@umkc.edu as that of the pure TiO2 nanocrystals in the frequency range
of 1.0–18.0 GHz. Apparently, hydrogenated TiO2 nanocrystals
DOI: 10.1002/adma.201303088 showed much higher efficiency in storing the electric energy

Adv. Mater. 2013, 25, 6905–6910 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 6905
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

o
o
A Hydrogenated TiO2
Pure TiO2
B
o

Intensity / a.u.

Intensity / a.u.
oo
o o o
oo
* oo o o
* * * * * *
Hydrogenated TiO2 nanocrystals
Pure TiO2 nanocrystals
100 200 300 400 500 600 700
o
o oo

20 30 40 50 60 70 80 500 1000 1500 2000 2500 3000


o -1
2Theta / Wavenumber / cm

Figure 1. Structural and morphologic properties. (A) XRD patterns, and (B) Raman scattering spectra of pure anatase and hydrogenated TiO2 nanocrys-
tals. (o) anatase phase, (*) rutile phase. TEM images of pure (C,D) and hydrogenated (E,F) TiO2 nanocrystals, with insets showing the FFT transforms.

inside, possibly due to they being more polarizable in the elec- As the epoxy has a ε′′ value of 0–0.02 at X-band frequency,[11]
tromagnetic field. Carbon nanotubes or Ni-coated, or Ag-filled thus hydrogenated TiO2 nanocrystals had 70.1 times higher
carbon nanotubes have shown permittivity value of 10–16 in ε′′ values as that of pure TiO2 nanocrystals in the frequency
the microwave range of 2–18 GHz.[11] Thus the high ε′ values range of 1.0–18.0 GHz. This suggests that hydrogenated TiO2
(17–25) of the hydrogenated TiO2 nanocrystals hereby dem- nanocrystals were more efficient in dissipating the electrical
onstrate their great potential as a new material for microwave energy under the electromagnetic field as well. Meanwhile,
absorption. after hydrogenation, the dielectric dissipation factor tgδε was
TiO2 nanocrystals/epoxy composites showed relatively improved over 15 times, on average, in the microwave region
constant ε′′ values of 0.03–0.3, in the frequency range of of 1.0–18.0 GHz (Figure S1A, Supporting Information),
1.0–18.0 GHz, but hydrogenated TiO2 nanocrystals/epoxy demonstrating that hydrogenated TiO2 nanocrystals were
composites had much higher ε′′ values in the same fre- more efficient in dissipating electrical energy than pure TiO2
quency range (Figure 2B). The ε′′ value first decreased gradu- nanocrystals.
ally from 5.9 at 1.0 GHz to 4.7 at 14.50 GHz, then increased Hydrogenation induced lower μ′ values (Figure 2C), or
to 10.6 at 17.50 GHz, and decreased to at 10.0 at 20.0 GHz. less susceptibility to the electromagnetic field, and higher

6906 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2013, 25, 6905–6910
www.advmat.de
www.MaterialsViews.com

COMMUNICATION
10
24 A B
8

16 6

"
'
Hydrogenated TiO2
Pure TiO2 4
8 Hydrogenated TiO2
2 Pure TiO2

0
0
4 8 12 16 4 8 12 16
Frequency / GHz Frequency / GHz
1.1 0.6

C D
1.0
0.4 Pure TiO2
Hydrogenated TiO2

"
'

0.9
0.2
Pure TiO2
Hydrogenated TiO2
0.8

0.0
4 8 12 16 4 8 12 16
Frequency / GHz Frequency / GHz

Figure 2. Comparisons of microwave absorption performance. Frequency dependence of (A) the real part (ε′) and (B) the imaginary part (ε′′) of the
complex permittivity (ε), (C) the real part (μ′) and (D) the imaginary part (μ′′) of the complex permeability (μ), of pure anatase and hydrogenated
TiO2 nanocrystals.

μ′′ values in the low and high frequency ranges (1.0–5.7 and suspected to be responsible for the improvement of hydrogen-
16.4–20.0 GHz) but lower μ′′ values in the median frequency ated TiO2 nanocrystals, as they are either the polar groups or
range (5.7–16.4 GHz) (Figure 2D). The change of the mag- they can induce polarization. The X-ray photoelectron spectro-
netic dissipation factor tgδμ followed the same trend as μ′′ scopy (XPS) survey spectra were similar for both pure and
(Figure S1B). This suggested that the hydrogenated TiO2 hydrogenated TiO2 nanocrystals (Figure 3A), the O 1s core-
nanocrystals were more efficient in dissipating magnetic energy level XPS spectra displayed the hydrogenated TiO2 nanocrystals
in the frequency regions, but less efficient in the median fre- had less OH content from the smaller OH band at 532.5 eV
quency region. (Figure 3B) besides the lattice O2− peak around 530.0 eV,[7,8]
Apparently, hydrogenation mainly induced the improvement and the Ti2p XPS showed almost identical pattern from lattice
of the response of the electrical component in the microwave Ti4+ ions with peaks centering at 458.7, 464.3, and 472.2 eV
region. To better understand this phenomenon, we conducted (Figure 3C).[7,8,28]
static-state and low-frequency electrical impedance measure- The lower OH concentration in hydrogenated TiO2 nanocrys-
ments, as the dielectric constant depends on the frequency. The tals was confirmed with Fourier transform infrared spec-
results (Figure S2) suggested that hydrogenated TiO2 nanocrys- troscopy (FTIR) results (Figure 3D) with no OH absorption
tals show higher electrical resistivity than pure TiO2 nanocrys- bands, compared to the obvious OH bands in the pure TiO2
tals at static state or lower-frequency (<700 mHz) modulated nanocrystals: a smaller and narrower band around 1640 cm−1
electric field, and lower electrical resistivity at higher-frequency from O–H bending of molecularly physisorbed water and a
(>700 mHz). Apparently hydrogenated TiO2 nanocrystals larger and broader band around 3250 cm−1 due to the stretch
responded more to the modulated electric field at higher fre- region of the surface hydroxyl groups and molecularly chem-
quencies, consistent with their higher microwave response. isorbed water.[29,30] The lack of OH in the hydrogenated TiO2
Dipole rotation is the mechanism normally responsible for nanocrystals was confirmed further with nuclear magnetic
microwave absorption due to the alignment of the polar groups resonance spectroscopy (NMR) results (Figure 3E) with much
in the materials to the electromagnetic field in the microwave smaller or barely seen peaks, compared to the obvious large
field,[1–3,11,15] besides the ferromagnetic resonance caused by and broad peak in the pure TiO2 nanocrystals. One argument
strongly interacting electron spins in the ferromagnet in the for this could be that, the detected NMR signals could have
ferromagnetic filler particles (e.g., iron or cobalt) dispersed been reduced by the microwave absorption properties of hydro-
in a matrix for microwave absorption.[1–3,12–15] Thus, hydroxyl genated TiO2, as imagined from the increased electrical con-
(OH) groups, Ti3+ ions or oxygen vacancies are the first species ductivity of the hydrogenated TiO2 nanocrystals in the lower

Adv. Mater. 2013, 25, 6905–6910 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 6907
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

2-
O
A O1s XPS B
Intensity / a.u.

Intensity / a.u.
O
OH
Ti

C from atmosphere

O 1s

Pure TiO2 nanocrystals Pure TiO2


Hydrogenated TiO2 nanocrystals Hydrogenated TiO2

1000 800 600 400 200 0 538 534 530 526

Binding Energy / eV Binding Energy / eV

4+

Ti2p XPS
Ti
C D

Reflectance / %
Intensity / a.u.

OH OH

4+
Ti
4+
Ti

Ti 2p3/2, sat Ti 2p1/2


Ti 2p3/2 Pure TiO2
Pure TiO2 Hydrogenated TiO2
Hydrogenated TiO2
476 471 466 461 456 451 4000 3000 2000 1000
-1
Binding Energy / eV Wavenumber / cm

OH
E F
Intensity / a.u.

Intensity / a.u.

Pure TiO2 nanocrystals


Pure TiO2 Hydrogenated TiO2 nanocrystals
Hydrogenated TiO2

30 20 10 0 -10 -20 1.96 2.03 2.10


Chemical Shift / ppm g factor
Figure 3. Examination of polar groups, Ti3+ ion and oxygen vacany defects. (A) Survey, (B) O 1s and (C) Ti 2p core-level XPS, (D) FTIR, (E) NMR and
(F) ESR spectra of pure anatase and hydrogenated TiO2 nanocrystals.

frequencies shown in the impedance spectra from 10 Hz to the lack of Ti3+ or oxygen vacancy in the hydrogenated TiO2
100 kHz (Figures S2C and S2D) and in the higher frequencies nanocrystals. So, OH, Ti3+ or oxygen vacancy was unlikely to
in the microwave from 1 GHz to 18 GHz (Figures 2A and 2B). be responsible for the excellent microwave absorption of the
However, this suspicion seems unlikely since the probe tuning hydrogenated TiO2 nanocrystals.
required for the two TiO2 samples was essentially identical. We propose here that the enhanced microwave absorption
As XPS is a surface-probe technique with the information of the hydrogenated TiO2 nanocrystals is due to the CMID in
depth of 1–2 nm, we further used electron spin resonance the crystalline/disordered and anatase/rutile interfaces, which
spectroscopy (ESR) to see if Ti3+ or oxygen vacancy was respon- induces the CIPAMA as illustrated in Figure 4. Interfacial
sible for the microwave absorption for the hydrogenated TiO2 polarization occurs whenever there is a buildup of a charge
nanocrystals, as the information depth of the ESR can be up to at a boundary between two regions or materials.[1–3,15] Crystal
a few centimeters. Both hydrogenated and pure TiO2 nanocrys- defects, such as dislocations or grain boundaries, electronic
tals did not show observable signals (Figure 3F): suggesting defects or phase boundaries all can contribute to the interfacial

6908 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2013, 25, 6905–6910
www.advmat.de
www.MaterialsViews.com

COMMUNICATION
rutile 8%). As the anatase/rutile ratio may affect the interfacial
polarization effect, i.e. on the anatase/rutile interfaces, we are
currently conducting comprehensive research to further reveal
the detailed relationship between the microwave absorption
performance and the anatase/rutile ratio by varying the hydro-
genation treatment time or temperature as well as the relative
importance of the anatase/rutile interface and the crystalline/
disordered interface on the microwave absorption of hydrogen-
ated TiO2 nanocrystals.
In summary, we have shown that hydrogenation of TiO2
nanocrystals can induce excellent microwave absorption perfor-
mance, due to the CMID on the nanometer scale. The CMID
and the induced CIPAMA will inspire further developments of
other exciting MAM materials. And hydrogenation expands the
applications of TiO2 nanocrystals into new areas beyond photo-
catalysis, battery, and solar cells.

Experimental Section
Figure 4. Proposed mechanism of collective interfacial polarization
Pure anatase and hydrogenated TiO2 nanocrystals were prepared
amplified microwave absorption (CIPAMA). The collective movements of
as follows. Briefly, we prepared crystalline TiO2 nanoparticles with
interfacial dipoles (CMID) at the anatase/rutile and crystalline/disordered
a precursor solution consisting of titanium butoxide, ethanol and
interfaces amplify the response to the incoming electromagnetic field and
deionized water. The solution was maintained at room temperature for
thus induce enhanced microwave absorption performance. The positions
24 hours under stirring and then dried at 120 °C overnight. The dried
of the electronic band structure of the disordered layer are assumed to lie
white powders were calcinated in air at 500 °C for 2 hours in flowed
between those of anatase and rutile phases. CBE: conduction band edge,
air to enhance their crystallinity. The pure TiO2 nanocrystals were then
VBE: valence band edge.
placed in a sealed sample chamber under vacuum for 1 hour and then
hydrogenated in a 5.0 bar H2 atmosphere at about 450 °C for 4 hours to
polarization.[1–3,15] Charge transfers and accumulation are obtain the hydrogenated TiO2 nanocrystals.
expected on the rutile/anatase and crystalline/disordered inter- The XRD was performed using a Rigaku Miniflex XRD machine with
Cu Kα as the X-ray sources (wavelength = 1.5418 Å). The Raman spectra
faces as shown Figure 4, as rutile and anatase TiO2 have dif-
were collected on an EZRaman-N benchtop Raman spectrometer with
ferent electronic band structures and dielectric constants, inter- the excitation wavelength of 785 nm. The spectrum range was from
facial band bending, and that hydrogenated TiO2 nanocrystals 100 cm−1 to 3100 cm−1. The spectrum collection time was 4 seconds
have been shown previous to possess p-type characteristics,[7] and was averaged over three measurements to improve the signal-
and they also displayed some structural characteristics between to-noise ratio. Both Raman spectra were normalized to the Raman peak
anatase and rutile.[19] The assembled movements of collective at 142.8 cm−1. The TEM study was performed on a FEI Tecnai F200 TEM.
The electron accelerating voltage was at 200 kV. A small amount of
interfacial dipoles at these interfaces amplify the response to
sample was dispersed in water, dropped onto a thin holey carbon film,
the incoming electromagnetic field and thus induce enhanced and dried overnight before TEM measurement.
microwave absorption performance. To verify this hypoth- The complex permittivity and permeability of pure and hydrogenated
esis, we measured the microwave absorption properties of TiO2 nanocrystals were measured at the frequency range of
pure rutile TiO2 nanocrystals with size around 55 nanom- 2–18 GHz using a HP8722ES network analyzer. The TiO2 nanocrystals
eters (Figure S3) and a mechanically mixed anatase/rutile were dispersed in melting paraffin wax, and the mixture was cast into
TiO2 nanocrystal mixture (anatase 92%, rutile 8%, Figure S4) a ring mold with thickness of 2.0 mm, inner diameter of 3 mm, and
outer diameter of 7 mm. The content of TiO2 nanocrystals was 60 wt%,
with similar composition and sizes to the hydrogenated TiO2 and the testing was performed at 0 °C. For comparison, the complex
nanocrystals. Pure rutile TiO2 nanocrystals showed similar permittivity and permeability of pure rutile TiO2 (from Alfa Aesar) and
microwave absorption (Figures S5-S10) as pure anatase TiO2 a mechanically mixed anatase/rutile TiO2 nanocrystals mixture with
nanocrystals. The mechanically mixed anatase/rutile mixture similar composition (anatase 92%, rutile 8%) were measured under
displayed higher ε′ values (Figure S5) than pure anatase or the same conditions. The anatase/rutile TiO2 nanocrystals mixture was
rutile TiO2 due to increased anatase/rutile interface, but much obtained by manually grinding suitable amount of rutile and anatase
TiO2 nanocrystals together.
lower values than the hydrogenated TiO2 nanocrystals, which
The DC and frequency-dependent resistivity of the pure and
apparently had many more interfacial junctions on the nano- hydrogenated TiO2 nanocrystals were measured with a Biologic
meter scale not only due to the larger anatase/rutile interface potentiostat/EIS instrument. The samples were pressed into small
but also the much larger crystalline/disorder interfacial area. pellets with diameter of 10.0 mm, and thickness of 1.0 mm, using an
The slightly increased microwave absorption of mechani- IR hand press. The pellets were sandwiched between two stainless
cally mixed anatase/rutile mixture over that of pure anatase or steel disks for the resistivity measurements. The frequency-dependent
resistivity was measured between 200 kHz and 100 mHz. The voltage
rutile also hinted that the main contribution of the microwave
modulation applied was 500 mV.
absorption of the hydrogenated TiO2 nanocrystals was from The surface chemical bonding and valence band information of the
the crystalline/disorder interface after the hydrogenation reac- samples was obtained with XPS on a PHI 5400 XPS system equipped
tion, as both samples had similar compositions (anatase 92%, with a conventional (non-monochromatic) Al anode X-ray source with

Adv. Mater. 2013, 25, 6905–6910 © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 6909
www.advmat.de
www.MaterialsViews.com
COMMUNICATION

Kα radiation. Small amounts of TiO2 nanocrystals were pressed onto [7] a) X. Chen, L. Liu, P. Y. Yu, S. S. Mao, Science 2011, 331, 746;
conductive carbon tape for XPS measurements. The binding energies b) H. Pan, Y.-W. Zhang, V. B. Shenoy, H. Gao, J. Phys. Chem. C
from the samples were calibrated, with respect to the C 1s peak, from 2011, 115, 12224; c) J. Lu, Y. Dai, H. Jin, B. Huang, Phys. Chem.
the carbon tape at 284.6 eV. The FTIR spectra were collected using a Chem. Phys. 2011, 13, 18063; d) Z. Zheng, B. Huang, J. Lu, Z. Wang,
Thermo-Nicolet iS10 FT-IR spectrometer with an attenuated total X. Qin, X. Zhang, Y. Dai, M.-H. Whangbo, Chem. Commun. 2012,
reflectance (ATR) unit. TiO2 nanocrystals were pressed onto the ZnSe 48, 5733; e) Z. Wang, C. Yang, T. Lin, H. Yin, P. Chen, D. Wan, F. Xu,
crystal of the ATR unit, and the measurement were performed in air at F. Huang, J. Lin, X. Xie, M. Jiang, Adv. Funct. Mater. 2013, 10.1002/
room temperature. adfm.201300486.
All 1H NMR experiments were performed on an 8.45 tesla Oxford [8] a) X. Chen, S. S. Mao, Chem. Rev. 2007, 107, 2891; b) X. Chen, Chin.
magnet with a three channel Tecmag Apollo NMR console, using a
J. Catal. 2009, 30, 839; c) D. A. H. Hanaor, C. C. Sorrell, J. Mater. Sci.
home-built, doubly tuned, MAS solid-state NMR probe with a 3.2 mm
2011, 46, 855.
stator assembly from Revolution NMR. The magic angle spinning
[9] C. H. Sun, X. H. Yang, J. S. Chen, Z. Li, X. W. Lou, C. Li, S. C. Smith,
rate was 8 kHz, the proton excitation pulse during the Bloch decay
experiment was ≈50 kHz, and each spectrum was acquired with G. Q. (Max) Lu, H. G. Yang, Chem. Commun. 2010, 46, 6129.
256 scans. The spectra were externally referenced to adamantane. NMR [10] M. Grätzel, Nature 2001, 414, 338.
spectra were acquired of the both samples as well as an empty rotor to [11] D.-L. Zhao, X. Li, Z.-M. Shen, Composites Sci. Technol. 2008, 68, 2902.
measure the hydrogen present. The contribution of the empty rotor was [12] A. N. Yusoff, M. H. Abdullah, J. Magnetism Magnetic Mater. 2004,
then subtracted to obtain the information from the samples which was 269, 271.
further scaled by the mass of samples to give a quantitatively accurate [13] Y.-B. Feng, T. Qiu, C.-Y. Shen, X.-Y. Li, IEEE Trans. Magnetics 2006,
spectrum. The ESR spectra were collected on a Benchtop Micro-ESRTM 42, 363.
machine, at a microwave frequency of 9.70 GHz at room temperature [14] C.-H. Peng, H.-W. Wang, S.-W. Kan, M.-Z. Shen, Y.-M. Wei,
without light irradiation. The ESR data were calibrated in relation S.-Y. Chen, J. Magnetism Magnetic Mater. 2004, 284, 113.
to g = 2.0066 of 2,2,6,6-tetramethyl-4-hydroxylpiperidine-1-oxyl. The [15] N. Bowler, IEEE Trans. Dielectr. Electr. Insul. 2006, 13, 703.
scanning time was 30 s per sweep, and the spectrum was averaged after [16] L. Deng, M. Han, Appl. Phys. Lett. 2007, 91, 023119.
30 sweeps. Quartz ESR tubes with an inner diameter of about 5.8 mm [17] Y. Qing, W. Zhou, F. Luo, D. Zhu, J. Magnetism Magnetic Mater.
were used. 2009, 321, 25.
[18] Z. Fan, G. Luo, Z. Zhang, L. Zhou, F. Wei, Mater. Sci. Engin. B 2006,
132, 85.
[19] a) X. Chen, L. Liu, Z. Liu, M. A. Marcus, W.-C. Wang, N. A. Oyler,
Supporting Information M. E. Grass, B. Mao, P.-A. Glans, P. Y. Yu, J. Guo, S. S. Mao, Sci.
Supporting Information is available from the Wiley Online Library or Rep. 2013, 3, 1510; b) L. Liu, P. P. Yu, X. Chen, S. S. Mao, D. Z. Shen,
from the author. Phys. Rev. Lett. 2013, 111, 065505.
[20] A. Naldoni, M. Allieta, S. Santangelo, M. Marelli, F. Fabbri,
S. Cappelli, C. L. Bianchi, R. Psaro, V. Dal Santo, J. Am. Chem. Soc.
2012, 134, 7600.
Acknowledgements [21] a) Z. Lu, C.-T. Yip, L. Wang, H. Huang, L. Zhou, ChemPlusChem
X.C. thanks the support from College of Arts and Sciences, University 2012, 77, 991; b) T. Xia, X. Chen, J. Mater. Chem. A 2013, 1, 2983;
of Missouri – Kansas City (UMKC), University of Missouri Research c) G. Li, Z. Zhang, H. Peng, K. Chen, RSC Adv. 2013, 10.1039/
Board and the generous gift from Dow Kokam. X. C. thanks the helpful c3ra41858h.
discussion and help on the measurement of ESR spectra from Professor [22] a) J.-Y. Shin, J. H. Joo, D. Samuelis, J. Maier, Chem. Mater. 2012, 24,
Thomas C. Sandreczki at UMKC, Department of Chemistry and 543; b) G. Wang, H. Wang, Y. Ling, Y. Tang, X. Yang, R. C. Fitzmorris,
Mr. James White from Active Spectrum, Inc. C. Wang, J. Z. Zhang, Y. Li, Nano Lett. 2011, 11, 3026.
[23] a) T. Xia, W. Zhang, W. Li, N. A. Oyler, G. Liu, X. Chen, Nano
Received: July 6, 2013 Energy 2013, 10.1016/j.nanoen.2013.02.005; b) T. Xia, W. Zhang,
Published online: September 17, 2013 J. Murowchick, G. Liu, X. Chen, Adv. Energy Mater. 2013, 10.1002/
aenm.201300294.
[24] a) L. Shen, E. Uchaker, X. Zhang, G. Cao, Adv. Mater. 2012, 24,
6502; b) Q. Zhang, E. Uchaker, S. L. Candelaria, G. Cao, Chem. Soc.
[1] a) A. R. Von Hippel, Dielectric materials and applications Artech Rev. 2013, 42, 3127.
House, Boston, 1995; b) P. Y. Yu, M. Cardona, Fundamentals of [25] W.-D. Zhu, C. W. Wang, J. B. Chen, D. S. Li, F. Zhou, H. L. Zhang,
Semiconductors: Physics and Materials Properties. Berlin: Springer, Nanotechnology 2012, 23, 455204.
2001. [26] B. Fultz, J. Howe, Transmission electron microscopy and diffractometry
[2] R. F. Soohoo, Microwave magnetics, Harper & Row Publishers, New of materials, Springer, New York, 2009.
York, 1985. [27] P. E. Champness, Electron diffraction in the transmission electron
[3] V. M. Petrov, V. V. Gagulin, Inorg. Mater. 2001, 37, 93. microscope, Taylor & Francis, 2001.
[4] a) A. Fujishima, K. Honda, Nature 1972, 238, 37; b) E. Borgarello, [28] a) X. Chen, C. Burda, J. Am. Chem. Soc. 2008, 130, 5018; b) X. Chen,
J. Kiwi, E. Pelizzetti, M. Visca, M. Grätzel, Nature 1981, 289, 158. P.-A. Glans, X. Qiu, S. Dayal, K. E. Smith, C. Burda, J. Guo, J. Elec-
[5] X. Chen, C. Li, M. Grätzel, R. Kostecki, S. S. Mao, Chem. Soc. Rev. tron Spectrosc. Relat. Phenom. 2007, 162, 75.
2012, 41, 7909. [29] J. Zou, J. Gao, F. Xie, J. Alloys Compd. 2010, 497, 420.
[6] a) U. Diebold, Surf. Sci. Rep. 2003, 48, 53; b) M. A. Henderson, Surf. [30] G. Li, L. Li, J. Boerio-Goates, B. F. Woodfield, J. Am. Chem. Soc.
Sci. Rep. 2011, 66, 185. 2005, 127, 8659.

6910 wileyonlinelibrary.com © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Mater. 2013, 25, 6905–6910

Das könnte Ihnen auch gefallen