You are on page 1of 11

www.afm-journal.

de
www.MaterialsViews.com
FULL PAPER

Direct View on Nanoionic Proton Mobility


Wing K. Chan, Lucas A. Haverkate, Wouter J. H. Borghols, Marnix Wagemaker,
Stephen J. Picken, Ernst R. H. van Eck, Arno P. M. Kentgens, Mark R. Johnson,
Gordon J. Kearley, and Fokko M. Mulder*

the need for environmental protection


The field of nanoionics is of great importance for the development of superior multiply the need for energy-storage and
materials for devices that rely on the transport of charged ions, like fuel cells, conversion devices in which nanoionic
batteries, and sensors. Often nanostructuring leads to enhanced ionic mobili- materials can play an important role. In a
sense nanoionics is a complementary field
ties due to the induced space-charge effects. Here these large space-charge to nanoelectronics which has become an
effects occurring in composites of the proton-donating solid acid CsHSO4 indispensable technology for our society,
and the proton-accepting TiO2 or SiO2 are studied. CsHSO4 is chosen for this with major applications in computers,
study because it can operate effectively as a fuel-cell electrolyte at elevated (future) solar cells and medical diag-
temperature while its low-temperature conductivity is increased upon nano- nosis.[5] Nanoionics is concerned with
the study and application of properties,
structuring. The composites have a negative enthalpy of formation for defects
phenomena, and mechanisms related to
involving the transfer of protons from the acid to the acceptor. Very high ion transport and storage in (solid-state)
defect densities of up to 10% of the available sites are observed by neutron nanoscale systems. Within this field are
diffraction. The effect on the mobility of the protons is observed directly using systems based on solids with relatively
quasielastic neutron scattering and nuclear magnetic resonance spectros- low ionic conductivity, where the imple-
copy. Surprisingly large fractions of up to 25% of the hydrogen ions show mentation of interfaces induce significant
concentrations of charged defects, which
orders-of-magnitude enhanced mobility in the nanostructured composites of in turn can have significant effects on the
TiO2 or SiO2, both in crystalline CsHSO4 and an amorphous fraction. ionic conductivity.[6,7] Methods to induce
such nanoionic phenomena include the
blending of ionic conductors with nano-
1. Introduction
particles[8,9]
and the deposition of thin layers of different ionic
For the transformation of chemical to electrical energy in fuel conductors.[10]
cells, the chemical storage of electrical energy in batteries, In order to make feasible a sustainable hydrogen economy,
or the transformation of chemical to electrical information the chain that includes hydrogen production, storage, and back-
(chemical sensors), nanoionics plays an increasingly important conversion using fuel cells[11,12] requires significant advances on
role.[1–4] The prospects for future alternative energy sources and all fronts. Certain types of nanoionic materials have the potential

Prof. F. M. Mulder Prof. S. J. Picken


Fundamental Aspects of Materials and Energy Nano structured Materials
Department of Radiation Department of Chemical Engineering
Radionuclides, and Reactors Faculty of Applied Sciences
Faculty of Applied Sciences Delft University of Technology
Delft University of Technology Julianalaan 136, 2628 BL Delft, The Netherlands
Mekelweg 15, 2629JB Delft, The Netherlands Dr. E. R. H. van Eck, Prof. A. P. M. Kentgens
E-mail: F.M.Mulder@tudelft.nl Department of Physical Chemistry – Solid State NMR
W. K. Chan, L. A. Haverkate, Dr. M. Wagemaker IMM, Radboud University Nijmegen
Fundamental Aspects of Materials and Energy Toernooiveld 1, 6525 ED Nijmegen, The Netherlands
Department of Radiation Prof. M. R. Johnson
Radionuclides, and Reactors Institut Laue-Langevin (ILL)
Faculty of Applied Sciences BP 156, 38042 Grenoble, Cedex 9, France
Delft University of Technology Prof. G. J. Kearley
Mekelweg 15, 2629JB Delft, The Netherlands Bragg Institute, Building 87
Dr. W. J. H. Borghols Australian Nuclear Science and Technology Organisation
Institut für Festkörperforschung PMB Menai, NSW2234, Australia
Forschungszentrum Jülich GmbH
Jülich Centre for Neutron Science at FRM II, 85747 Garching, Germany

DOI: 10.1002/adfm.201001933

1364 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374
www.afm-journal.de
www.MaterialsViews.com

and their mobility is low.[21,22] In ionic materials the mobility of

FULL PAPER
to serve as ‘intermediate-temperature’ fuel-cell electrolyte
membranes. Most commercially available fuel cells operate ions strongly depends on the number of available vacant sites
using a polymer electrolyte (as in proton-exchange-membrane to which an ion can move. It is important in this respect that in
fuel cells or PEMFCs), and operate well below 100 °C to main- the high mobility phase I many vacant sites are present while
tain electrolyte humidity. Fuel cells that operate at interme- in the low mobility phase II all sites are occupied and mobility
diate temperatures (100–300 °C) are rare. However, reasons is only possible via exchange of ions.
to operate at these temperatures are manifold. Waste energy In contrast to bulk material, where electroneutrality must
(heat) coming from resistive losses is generated at an inter- be obeyed, at interfaces between ionic conductors a narrow
mediate temperature that makes it usable. Integrated systems charged zone, a so-called space-charge layer, is tolerable and
using this higher quality heat can increase overall efficiency of even thermodynamically favorable.[23] Mobile charge carriers
a fuel-cell system from roughly 50% to close to 80%. Further- of the conducting phase can physically relocate themselves
more, increased catalytic activity with less (expensive) catalysts at the interface or in the alien phase (leaving vacancies in the
is possible (Pt catalyst are used), because at these tempera- donating phase), which creates oppositely charged space-charge
tures less catalyst poisoning by CO occurs, which reduces the layers in the conducting matrix and nanoparticles. The devia-
requirements for the purity of the fuel and catalyst. Inherent to tion from charge-neutrality creates an electrical field that limits
intermediate-temperature fuel cells is a water-free conduction the magnitude and spatial extent of the carrier redistribution.
mechanism, which removes the necessity for auxiliary systems In the nanometer range, interfaces are so closely spaced that
to keep the system cool or moisturized. Due to operation at the space-charge layers, which usually extend on the nano-
higher temperature, the relatively large cooling systems that are meter scale (several times the Debye length)[24] partially overlap.
normally necessary to keep the operation temperature below The impact of the interfaces becomes increasingly significant
100 °C need not be installed; cooling with ambient-temperature until it dominates the whole material’s properties such as the
coolants is easier when operating at higher temperatures. On local ionic densities and conductivities.[10] In these materials
the other hand, when comparing operation at intermediate the overall impact of the interfacial regions increases with
temperatures with high-temperature solid-oxide fuel cells, the decreasing nanoscale dimension. This increase leads to true
materials demands are much reduced because the intermediate size effects from the nanomorphology (in contrast to trivial size
temperatures are relatively low, and startup times are reduced. effects) as an ion has a shorter diffusion path between phases
High protonic conduction at intermediate temperatures is when the dimensions are decreased.
found in solid acids.[11,13] Typical for these materials is a super- Although the space-charge effect is a widely accepted phe-
protonic phase transition after which the material shows an nomenon, the effect on the ionic densities in nanostructured
orders-of-magnitude jump in protonic conductivity.[14,15] Above materials has, to our knowledge, never been determined directly
this phase transition, these materials have been shown to be by experiments sensitive to these densities. This lack holds for
applicable as fuel-cell electrolytes.[16–18] Below this phase tran- the chosen solid acid or any other proton-conducting material.
sition nanostructuring is applied to improve the ionic conduc- Similarly, the effects of nanostructuring solid acids have been
tivity, which results in conductivity values approaching those investigated previously through macroscopic measurements,
of the superprotonic phases.[8] These materials are solid acidic but the microscopic, atomic-scale defect densities occurring or
materials that, in spite of their acidic nature, require no special the effect on the proton dynamics due to nanostructuring have
heat- or acid-resistant construction materials. Solid acids gener- not been probed directly on a microscopic level.
ally can be represented with the chemical formula: MaHb(XO4)c, Here we perform a study with direct microscopic probes of
where M is a mono- or divalent cation (e.g,. Rb, Cs), XO4 is a the hydrogen atoms/protons on the effect of nanostructuring
tetrahedral oxy-anion (e.g., SO4, SeO4, PO4), and a, b, and c are on dynamics, ionic densities, and other structural factors,
integers. which complements the macroscopic measurements performed
One particular solid acid is cesium hydrogen sulfate. in other studies. Quasielastic neutron scattering (QENS) and
Depending on temperature, CsHSO4 exhibits three crystalline nuclear magnetic resonance (NMR) spectroscopy were per-
phases, usually referred to as phases III, II, and I. Phases III and formed on the CsHSO4 solid acid and several of its nanocom-
II are low-proton-conducting phases. Bulk crystalline CsHSO4 posite samples. Both techniques directly observe the protons
phase II is a metastable phase at room temperature, which and thus provide information about the dynamics as well as
gradually reverts to Phase III. At 414K CsHSO4 goes through structure. To investigate the proton densities, neutron diffrac-
its superprotonic phase transition (from phase II to phase I),[19] tion on deuterated samples was performed. The solid acid was
which yields a higher crystal symmetry and increased lattice chosen for this study because it has considerable promise as a
dimensions. This transition allows for quasifree rotation of the fuel-cell electrolyte above its superprotonic phase transition.
SO4 tetrahedra between crystallographically identical positions, QENS and NMR measurements were performed on bulk
creating six times as many possible proton sites as there are pro- crystalline CsHSO4 and nanocomposites thereof with SiO2
tons available. Proton diffusion occurs through combined rota- and TiO2. Different proton fractions with varying mobilities
tion of the SO4 group and jumps to the next SO4 group;[20] it is a were observed between the bulk crystalline and nanocompos-
fast process that gives the material its high proton conductivity. ites. NMR on deuterated samples of the crystalline solid acid
For temperatures below 414K, CsHSO4 has a monoclinic sym- and nanocomposites with TiO2 was also performed. In the
metry in which the number of protons is equal to the number diffraction study, the bulk crystalline solid acid and the nano-
of available proton sites (phase III and II). The hydrogen atoms composite thereof with TiO2 nanoparticles in different sizes
are localized with rigid hydrogen bonds between SO4 tetrahedra and molar ratios were investigated. TiO2 was chosen as the

Adv. Funct. Mater. 2011, 21, 1364–1374 © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1365
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

bulk solid acid bulk solid acid be that the nanostructured solid acid is struc-
nano TiO2 15nm nano TiO2 15nm turally similar to the crystalline bulk sample
nano TiO2 24nm nano TiO2 24nm
with a coherently continuing crystal lattice
nano TiO2 40nm nano TiO2 40nm
around the nanofiller particles.
Rietveld analysis of the TiO2 nanocom-
intensity

intensity
posites showed a discrepancy in most of the
samples with respect to the ratio of solid acid
to nano-TiO2 filler phases, as there appeared
to be much more crystalline TiO2 present
in the composite than was added during
synthesis. Flame atomic-emission spectros-
2.7 2.8 2.9 3.0 2.75 2.80 2.85 2.90 copy (FAES) measurements combined with
d-space d-space inductively coupled plasma optical emission
spectroscopy (ICP-OES) measurements,
Figure 1. Left: Section of XRD spectra for bulk crystalline- and nanostructured solid-acid com- however, showed the theoretical ratio of
posites showing CsHSO4 peaks. Unless otherwise stated, the nanostructured composites con-
CsHSO4 to TiO2. This result is evidence that
sist of a molar ratio solid-acid versus nanoparticles of 1:2. Right: Peaks intensity normalized,
showing only limited broadening. during preparation of crystalline CsHSO4 an
X-ray amorphous phase of the solid acid is
also formed. This formation depends on the
ion-conducting nanocrystalline filler because it is crystalline presence of the nanofillers and possibly on factors like drying
and therefore can provide structural information through use rate or temperature treatments. During the investigations we
of neutron and X-ray diffraction. In a previous study, nanoscale also observed that, for the smallest nanofiller particles, the
TiO2 was shown to be prone to hydrogen uptake in pure D2SO4 amorphization progresses with time. Figure 2 shows three
acid.[25] Nanoparticulate TiO2 of a few different sizes, immersed diffraction patterns taken in chronological order of a nano-
in D2SO4 showed surprisingly large ionic densities at a specific composite solid-acid sample with 7-nm TiO2 particles. The
position in TiO2, namely [0.00, 0.75, 0.43], which corresponds upper diffraction pattern was taken directly after synthesis, the
with split positions of the deuterons inside the oxygen octahe- middle one was taken several weeks afterwards, and the lower
dron. Rietveld refinement showed occupancies at this position spectrum several months thereafter. It can be seen clearly that
of 0.045 and 0.085 for TiO2 particle sizes of 24 nm and 7 nm, the diffraction pattern in the middle has lost most of the reflec-
respectively, which implies deuteron intercalation values of 9% tions of the solid acid and the lower diffraction pattern shows
and 17%. Although CsHSO4 is less acidic than sulfuric acid we no reflections belonging to the solid-acid phase. Recrystalliza-
will see that hydrogen insertion in TiO2 still takes place and tion of the latter sample results in a diffraction pattern similar
therefore vacancies are created in the solid acid. to the upper one. This phenomenon was especially profound
Finally the experimental findings are supported by initial in the nanocomposites with the smallest nanoparticles and
simulation results using ab initio methods (Vienna ab initio leads to the conclusion that nanostructuring of the solid acid
simulation package; VASP). destabilizes the material so that it gradually becomes amor-
phous, and X-ray invisible. The first question is why the mate-
rial becomes partly amorphous in the first place. Contributing
factors are the abundant interfaces and the stresses exerted by
2. Results and Discussion them on the lattice of CsHSO4. Below we will demonstrate that
another important factor in these materials is the alteration of
2.1. X-ray Characterization of Samples and Presence
of X-ray Amorphous Phase
freshly prepared
after 2 weeks
X-ray diffraction (XRD) was performed to determine the struc- after 2 months *
ture and amount of the crystalline phases formed. Figure 1 com-
*
pares the diffraction of a part of the spectra for a bulk crystalline * * *
*
sample with several spectra of nanocomposite samples. The
a.u.

peaks for the CsHSO4 phase II are still easily observable but it
is apparent from the patterns that the peaks become less pro-
nounced (Figure 1, left) when there is less CsHSO4 in the com-
posites. However, peak-broadening due to decreased grain sizes
in the solid-acid phase did not occur either; normalizing the
intensity of certain peaks (Figure 1, right) showed no significant
broadening of the peaks of the nanocomposite sample com- 1.5 2.0 2.5 3.0
pared to the peaks of the crystalline bulk samples. The absence d-space
of could be a result of complete phase separation between solid Figure 2. Time-dependent amorphization of solid acid in nano-CsHSO4
acid and nanofiller, which seems unlikely in view of the large and 7-nm TiO2 composite. Several significant peaks belonging to the
changes in proton mobilities. A more likely explanation could solid acid phase are marked by asterisks.

1366 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
Resolution Analogous to the work of Belushkin
bulk 390K et al.[20] quantitative analysis of the spectra
raw data
bulk 463K resolution
nano 360K
uses a convolution of two Lorentzian contribu-
fit curve
nano 450K tions and a delta function with the resolution
broad peak
function (Figure 3, right). The width of the
narrow peak
background
narrowest Lorentzian shows Q2 dependence

intensity
down to small Q values (typically <0.2Å−1),
intensity

which indicates translational diffusion of the


protons (where Q is the scattering vector, a
basic quantity used in reciprocal space). The
widths plotted in Figure 4 (left) indicate that
the addition of TiO2 nanoparticles induces
a motion on a time-scale comparable to the
-0.08 -0.04 0.00 0.04 0.08 0.0 0.2 0.4
Energy transfer /meV superprotonic diffusive translational motion
Energy transfer /meV
and which is surprisingly constant over the
Figure 3. Left: QENS spectra of bulk CsHSO4 and SiO2 (7 nm) nanocomposite samples with whole temperature range measured. The SiO2
molar ratio 1:2 at Q = 0.61Å−1. The bulk sample at 390K does not show broadening relative nanoparticles result in higher mobility than
to the resolution function. Nanocomposite at 360K shows a signal comparable to that of the the TiO (green and red curves, respectively)
2
bulk at 463K. The nanocomposite at 450K shows an even broader quasielastic signal. Right:
and smaller particles have more effect on line-
Example of a raw QENS data-fit using a convolution of two Lorentzian functions with the
resolution function. broadening than bigger ones.
Up to 25% of the protons in the low-
temperature composite phases show trans-
the crystalline matrix due to the removal of protons under the lational diffusion (Figure 4, right), whereas there is no sign
influence of space-charge effects. Furthermore the progressing of mobile protons in the signal of the bulk phase at low tem-
destabilization of the solid acid resulting in the ‘growth’ of the perature (in agreement with previous reports).[20] This mobile
amorphous phase is aided by the significant mobility of the fraction will contribute significantly to the ionic conduction
material at room temperature. in the composite phase, and it constitutes a large fraction
Another effect of the nanostructuring as seen by X-ray dif- of mobile protons because it is of the same order of mag-
fraction analysis is that, when formed, phase II of CsHSO4 does nitude as the fraction of mobile protons above the super-
not relax back to phase III. The aim of this research is to study protonic transition. While impedance measurements show
the mechanism that causes enhanced proton mobility in these an enormous increase in conductivity of the nanocomposite
samples. However, sufficient care was taken to minimize the samples,[26] they cannot provide insight into the microscopic
amount of amorphous phase, to make comparisons with bulk dynamics and mechanisms of conduction. With QENS we
more straightforward. directly observed that the protons not only are much more
mobile in the nanocomposite at low temperatures, but addi-
tionally a large fraction of protons remains mobile when the
2.2. Proton Dynamics in CsHSO4 temperature is lowered.

QENS is a direct probe of the time- and


length-scale of the mobility of protons in 1.0
the material via, respectively, the measured
0.03
energy- and momentum-transfer spectra. Phase ΙΙ PhaseΙ 0.8 Phase ΙΙ Phase Ι
fraction mobile protons

Figure 3 (left) shows QENS spectra of bulk


crystalline CsHSO4 and a nanocomposite 0.6
sample of CsHSO4 mixed with SiO2 (7 nm) 0.02
Γ /meV

at several temperatures. The broadening at


0.4
the foot of the peak is a direct indication of
the mobility time-scale of the protons in the 0.01
material. As expected,[20] the spectrum of 0.2
the bulk material shows that at 390K its pro-
tons are fixed at their lattice positions, while 0.00 0.0
350 375 400 425 450 475 350 375 400 425 450 475
at 463K in the superprotonic phase I they T /K T /K
are highly mobile on the picosecond time-
scale probed. In contrast, the spectrum of Figure 4. Left: Proton mobility at Q = 1 Å−1 for four different samples from QENS measure-
the nanocomposite at 360K shows a proton ments. Right: Mobile fraction from QENS measurements. Black squares: bulk CsHSO4, red
triangles: nanocomposite CsHSO4 with 24-nm TiO2 particles, purple circles: nanocomposite
mobility almost as high as the bulk material CsHSO4 with 40-nm SiO2 particles, green hexagons: nanocomposite CsHSO4 with 7-nm SiO2
in the superprotonic phase at 463K. At 450K particles. The linewidth (Γ) is a measure used to indicate proton mobility. For jump diffusion at
2
the protons have an even higher mobility small Q values Γ is related to the diffusion constant as follows:  (Q ) = h̄ l6J Q2 = ¯h D Q 2 ,
than in the superprotonic bulk phase. where l is the jump length and τ is the residence time.

Adv. Funct. Mater. 2011, 21, 1364–1374 © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1367
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

NMR proton spin-lattice relaxation-time bulk nano TiO2 24nm


(T1) measurements at 400 MHz probe mobil- nano SiO2 7nm nano TiO2 7nm
ities to order-of-magnitude similar time- nano SiO2 40nm nano TiO2 40nm
scales to the QENS measurements and are T=20oC
therefore complementary. The 1H T1 relaxa-

intensity
tion time in bulk CsHSO4 shows a dramatic

a.u.
drop going through the superprotonic phase
bulk
transition from phase II to phase I. Using a TiO2 24nm phase II
model calculation, it was shown that in phase TiO2 24nm "amorph"
SiO2 40nm phase II
I of bulk CsHSO4T1 is directly related to the SiO2 40nm "amorph"
translational motion of hydrogen.[15,22] The
protons undergo a two-step diffusion process
that involves both a change in orientation 12 10 8 6 0 20 40 60 80 100
chem. shift /ppm t /sec
of SO4 groups as well as a transfer between mix

the SO4 groups. The low values of the T1 1


Figure 6. Left: H MAS NMR spectra of bulk and nanocomposite samples with SiO2 or TiO2
times are caused by the rapid fluctuations nanoparticles showing peaks in two different chemical shift regions. Right: T saturation meas-
1
of the nuclear dipolar interactions between urements of bulk and nanocomposite solid acid with 40-nm SiO2 and nanocomposite solid
hydrogen atoms and between hydrogen and acid with 24-nm TiO2.
Cs, which result from the three-dimensional
diffusive motion. In the bulk phase II the dif-
fusion is slowed by orders of magnitude, and the motions are the bulk crystalline phase. This result shows the inhomogeneity
characterized by intra-hydrogen-bond hopping of protons as of the mobility in the sample.
well as reorientation of the O–H…O hydrogen and chemical Further analysis of the 1H NMR spectra showed two peaks
bonds. [ 21, 27] (Figure 6 left); one at 11.0 ppm, which is expected for the solid
Because the hydrogen-bonded network in phase II
is represented by zig-zag infinite chains, these motions can be acid in phase II[29] and a second peak in a lower ppm region.
characterized as one dimensional. However, with increasing The second peak is a result of the nanostructuring, since it is
temperature the reorientations cause transfer between different absent in the bulk crystalline sample. The position and width
chains, which leads to more two-dimensional motions even of this peak is dependent on the type of nanoparticle used. SiO2
below the superionic phase transition.[27] nanoparticles result in a position around 6.0 ppm and TiO2 par-
In our experiments the measured bulk crystalline proton T1 ticles give rise to a peak around 6.6–7.4 ppm. In general the
times are in good agreement with literature (differences can be large chemical shift at 11 ppm indicates strong deshielding
explained by differences in the applied field strength).[21,22,28] in the solid acid which is caused by the electronegativity of
In agreement with the rapid diffusive signal observed in the the SO42− group. The lower chemical-shift peak observed in
neutron data, the relaxation rates of the nanocrystalline mate- the composites indicates a higher shielding, i.e., less electron
rial with 24-nm TiO2 particles (Figure 5) are, for a large frac- density withdrawn from the protons in this environment. In
tion of the protons, orders of magnitude smaller than the general this would also mean that these protons are in a less
corresponding bulk values over the whole temperature range acidic environment. In view of the presence of the amorphous
accessed. This result again shows that in the nanostructured phase, the peaks at lower chemical shift are initially attributed
composite high proton mobility is induced. However, there is to protons in this amorphous phase surrounding the nanopar-
also a fraction of the protons that show long T1 relaxation times, ticles. The shift to lower acidity could be explained by the fact
similar but still roughly twice as short as that of the protons in that fewer protons are present in the amorphous phase due to
exchange with protons at the surface or in
1.0 the nanoparticles (present as result of space-
ΙΙ Ι charge effects, see below). The lower chem-
100 ΙΙ Ι ical-shift value of the amorphous phase in the
0.8
SiO2 composite then suggests that SiO2 is a
better proton acceptor (less acidic) than TiO2
mobile fraction

0.6 (more acidic). Below it is shown that indeed


10
H T1 /sec

2
H can be observed inside the TiO2 phase, as
0.4 was previously demonstrated for TiO2 nano-
1 particles immersed in liquid D2SO4.[25]
1

0.2 Shown in the right plot of Figure 6 are


results of T1 saturation measurements of
0.1 0.0 nanocomposite solid acid with 40-nm SiO2,
250 300 350 400 450 250 300 350 400 450 24-nm TiO2, and the bulk samples showing
T /K T /K the intensities of the spectra at certain mixing
Figure 5. Left: T1 relaxation times of protons in bulk crystalline CsHSO4 and nanocomposite times. The nanocomposite samples each have
CsHSO4 with 24-nm TiO2. Right: Fraction of mobile protons. Black squares: bulk crystalline two lines representing the phase II (open
sample. Red triangles: nanocomposite sample. symbols) and the amorphous/exchanged

1368 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
Table 1. NMR proton T1 relaxation times.

Sample T1–1 Peak 1 [sec] T1–2 Peak 1 [sec] T1–1 Peak 2 [sec] T1–2 Peak 2 [sec]
Bulk Crystallin CsHSO4 42 – – –
Composite 7-nm TiO2 15 0.12 1.0 3.4 × 10−2
Composite 24-nm TiO2 10 0.62 1.0 0.20
Composite 40-nm TiO2 10 1.0 1.5 0.15
Composite 7-nm SiO2 0.50 6.0 × 10−2 4.8 × 10−2 3.1 × 10−3
Composite 40-nm SiO2 4.5 0.26 1.3 6.2 × 10−3

(closed symbols) peaks. Clearly the phase II peaks of the nano- amorphous. In this 40-nm composite SiO2 sample, the high
sized samples possess a short component and a longer one. proton mobility observed with QENS and T1 measurements
Table 1 shows the proton T1 times of the different samples. therefore can be totally attributed to the protons in the crystal-
The peak assigned to amorphous/exchanging CsHSO4 in the line phase. In conclusion, therefore, the presence of the amor-
composites (around 6–7.4 ppm) at room temperature was gen- phous phase alone does not explain the altered proton mobility.
erally well represented by two short T1 contributions; a longer Observation of the upper solid spectrum in Figure 6 (left;
one around 1 second and a much shorter one, which was 7-nm SiO2 nanocomposite) reveals a second signal in the amor-
dependent on the type of particle used. One might therefore phous phase region around 7 ppm. The origin of this signal
relate the shorter component to the protons in or at the nano- is not known precisely, but it could come from protons in the
particles, as they are observed in neutron diffraction, and the amorphous phase, while the second and larger signal is a result
longer T1 component to the protons in the amorphous CsHSO4. of protons at the surface or inside the nanoparticles.
In any case the strongly reduced T1 clearly shows the enhanced
proton mobility in the composites due to nanostructuring. In
view of this mobility one should also expect exchange of pro- 2.3. Deuterated Solid Acid
tons between different regions in the sample that can bring T1
values together and can influence spectral shapes, depending For investigation of proton densities in the solid acid using
on the exchange rates. The phase II peak (11 ppm) also shows diffraction techniques, deuterium is much more suitable than
two T1 values: a long component in the same order as the hydrogen. To investigate the motions of hydrogen in the com-
bulk phase II T1 time (roughly around 10 seconds), but, most posites in more detail the deuterium isotope is also used, as
remarkable, also a component with values on the order of the T1 this nucleus has a sizeable quadrupole moment which interacts
time found in the amorphous signal (less than 1 second). Note with the local electronic distributions.
that this short T1 time in the crystalline phase II peak is even Figure 7 (left) shows the T1 times of deuterated solid acid
slightly shorter than the long T1 time in the amorphous phase samples. Our bulk sample shows similar T1 values to those
and much shorter compared to the bulk T1 times of around 40 published in much detail.[21] In the composites there is a prom-
seconds. This result indicates that also the crystalline phase II inent fraction of the signal that has a short T1, just as for the
is influenced by the nanostructuring or rapidly exchanging pro- protonated samples.
tons with the shortest T1 protons in the sample.
The 7-nm SiO2 composite sample differs
from the other samples as it does not possess 100 1.0
a long T1 component. This difference shows ΙΙ Ι
that the nanostructuring is able to influence 0.8 ΙΙ Ι
the whole composite sample, increasing the 10
mobile fraction

mobility.
H T1 /sec

0.6
In the upper two spectra in Figure 6 (left), 1
two extreme cases of nanostructuring are
depicted. The dashed spectrum represents 0.4
2

the nanostructured solid-acid composite with 0.1


40-nm SiO2 and the solid line represents the 0.2
composite with 7-nm SiO2 nanoparticles. The 0.01
composite with 40-nm particles shows almost 0.0
no signal around 6 ppm, while the 7-nm com- 250 300 350 400 450 250 300 350 400 450
posite shows a relatively small signal in the T /K T /K
11 ppm region. This apparent randomness Figure 7. Left: Shortest observable proton-relaxation times (T ) of deuterium ions in bulk
1
in intensities of the signals is attributed to (black diamonds: taken from reference [21], green crosses: this work) and nanocomposite of
the preparation procedure. With proper care, CsDSO4 with 24-nm TiO2 nanoparticles (blue triangles). Right: Fraction of deuterium ions with
the samples can be made fully crystalline or a short T1 time, i.e., showing high mobility, in bulk and in nanocomposite CsDSO4.

Adv. Funct. Mater. 2011, 21, 1364–1374 © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1369
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

500K solid acid. From the settings of the experi-


ment it is clear that we measure both
deuterium in the amorphous/exchanged
3D phase as well as deuterium in the crys-
500K talline phase II or I (depending on the
3D
2D 400K temperature); the fraction of deuterium
400K in the crystalline phase II that has a long
T1 is suppressed because of the rapid scan
2D rate, but the fraction with a short T is not
1
300K
1D suppressed. However, discrimination in
300K the spectra between crystalline and amor-
200K phous phases is not possible on the basis
1D
of the different chemical shifts, because
of the broad lines that are present. In
-150 -100 -50 0 50 100 150 any case the temperature range between
frequency /kHz (roughly) 273K and 400K shows essen-
tially a single broad peak, as is observed
Figure 8. Left: 2H MAS NMR spectra of bulk crystalline CsDSO4 at different temperatures. Right:
Nanocrystalline CsDSO4 with 24-nm TiO2 particles. The spectra display spinning sideband patterns in the bulk sample around 385K. It is nat-
reflecting the size and symmetry of the 2H quadrupolar interaction. ural to assume similar mechanisms for
motion in the crystalline but nanostruc-
tured, as well as in the less-ordered X-ray
By comparing Figures 5 and 7 it can be seen that the T1 times amorphous part of the material, but then the transition of 1D to
of the nanocomposites are an order of magnitude shorter than 2D motions has to take place at much lower temperature than
the T1 values from the proton measurements. Three factors are before. For this reason the temperature ranges in which the
responsible for this result. First, the T1 relaxation mechanism deuterium motions in the crystalline fraction of the material
is now due to the strong fluctuating quadrupolar interactions, are mainly 1, 2, or 3D are indicated in Figure 8. In the 1D (low
instead of the mainly dipolar proton interactions. Second, the temperature) region of the nanocomposite material, the meas-
lower deuterium resonance frequency in the same applied field ured quadrupolar broadened MAS spectra may have a smaller
means that slower time-scales are probed. Third, the isotope overall width, which leads to a total spectrum with more inten-
effect causes motions to occur on longer time-scales. However, sity in the center. The lower width, i.e., lower quadrupolar
from these differences it is clear the same behavior still exists; interaction, could indicate that less-strong hydrogen bonds are
short T1 above the superprotonic transition in bulk samples, present, which lead to less asymmetry in the electron cloud sur-
and a significant fraction of the deuterium with a short T1 at all rounding the deuterium nuclei. This suggestion is consistent
temperatures measured, i.e., also below the transition tempera- with the observation of lower proton chemical shifts measured
ture. The mobile fractions found are quite similar to those from in the protonated samples. A more detailed analysis lies outside
the proton results. the scope of this paper.
Analysis of the 2H NMR spectra leads to additional insight
into the solid acid because the spectra are strongly influenced
by quadrupolar interactions that probe the surrounding of the 2.4 Deuterium Ion Intercalation in TiO2
deuterium nuclei. Figure 8 shows the deuterium MAS NMR
spectra performed on bulk and on a 24-nm TiO2-CsDSO4 com- Analogous to the model study with liquid D2SO4 and TiO2, in
posite. The temperature dependence of the bulk solid acid is which deuterium ion densities were observed in the TiO2 nano-
in complete agreement with literature.[21] Already below the particles, neutron diffraction measurements were undertaken
superprotonic transition there are specific motions on the on the nanocomposite solid-acid samples with TiO2 nanopar-
time-scale probed by the spin precession. From the analysis ticles. Space charge in the composites is modeled as interca-
of single-crystal deuterium NMR[21] it was found that around lated deuterium nuclei in the TiO2 lattice[25] and vacancies in
385K the bulk NMR spectrum alters because the deuterium the CsDSO4 and fitted as such in the neutron-diffraction data
dynamics change from motion along infinite 1D zigzag chains using GSAS. The diffraction intensities give direct informa-
to motions in a progressively more 2D network. This change tion on the position and quantity of the deuterium in TiO2
leads to narrowing of the apparent quadrupolar spectrum. and vacancies in the crystalline fraction of CsDSO4. The high
After the superprotonic phase-transition temperature the H/D quality data (Figure 9) make it possible to observe small effects
movement takes place in a truly 3D network. The experiments on the diffraction intensities caused by nanostructuring. Fitting
thus show spectra evolving from broad multi-sideband spectra the CsDSO4 phase was complicated by the induced effects of
(referred to as 1D region) to a single, broadened peak (2D), to nanostructuring, for example, vacancies in the lattice leading to
very sharp peaks with sidebands in phase I (3D). Compared to local, structural distortions, in addition to the presence of TiO2
the bulk crystalline phase, the nanocomposite solid acid shows diffraction peaks. Therefore the errors in the results extracted
a strongly modified temperature dependence. Already at 273K from these fits are relatively large for CsDSO4. As the TiO2 lat-
the sidebands disappear and a single broad peak remains. This tice hardly changes upon deuterium insertion, error margins
change occurs at a much lower temperature than in the bulk here are much smaller.

1370 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374
www.afm-journal.de
www.MaterialsViews.com

FULL PAPER
This behavior is intuitively clear since more TiO2 means that
nano more deuterium ions can be extracted from the CsDSO4, how-
TiO2 15 nm
ever, progressive extraction of deuterium from CsDSO4 will be
less energetically favorable so the average concentration of deu-
nano terium ions in TiO2 should become lower.
TiO2 24nm

2.5. Apparent Morphology


bulk
CsDSO4
Surprisingly, the addition of small nanoparticles does not seem
to impose equally small apparent grain sizes for the solid-acid
1.0 1.5 2.0 2.5 3.0 3.5 4 .0 phase, as observed from the diffraction line-widths. How-
d-spacing /Å ever, smaller sizes are noticeable upon addition of the smallest
Figure 9. Neutron-diffraction spectra of bulk and two nanostructured nanoparticles below 24 nm, as shown in Table 2. This observa-
samples (TiO2 15- and 24-nm, CsDSO4:TiO2 = 1:2), with fitted curve. tion means that the solid-acid structure continues coherently
for larger characteristic distances, in spite of the presence of
70 mol% nanoparticles dispersed through the composite. During
A fresh sample of the solid acid at room temperature (i.e., no preparation all materials were continuously mixed and the nano-
relaxation towards phase III yet) shows the monoclinic crystal- particles alone have sufficient volume to fill the volume taken up
line phase II and could be fitted well, with total wRp values of by the sample in a hard-sphere packing scheme. From our com-
around 0.02 (Table 2). As in the model study, deuterium ions bined knowledge of the presence of the amorphous phase and
were observed in the TiO2 lattice. These deuterium ions could the high intrinsic mobility throughout the materials, we postu-
only be a result of deuterium-ion migration from the solid- late a process for the formation of the nanostructured samples.
acid phase, with TiO2 acting as a hydrogen-ion acceptor. From After dissolving the solid acid, the nanoparticles are dispersed
the fitted deuterium concentrations a correlation between the in the liquid. The solvent water is slowly removed by evaporation.
number of intercalated deuterium ions and the sizes of the At a certain point the solid acid starts to solidify or crystallize on
TiO2 nanoparticles is observed and plotted in Figure 10. With nucleation sites on the nanoparticles. These solid phases may
decreasing particle size, very high deuterium-ion densities of be amorphous at the start because of perturbing interactions
up to 10% intercalation in TiO2 were found. This value is still from the TiO2/SiO2 surfaces, either because of lattice strains or
almost a factor of two lower than that of the liquid acid D2SO4 the exchange of protons between the phases. When the layer of
deuterium-ion intercalation, which is explained by the lower solid acid becomes thicker, the influence of the nearby nanopar-
acidity of the solid CsDSO4 acid than that of the liquid D2SO4 ticle surface reduces and the normal crystalline solid acid can
acid. start growing. Because this growth starts from the many dif-
The position of the intercalated deuterium ions in the TiO2 ferent nucleation centers present in the solid acid, phases only
(Table 2) was found to be around [0.00, 0.75, 0.43] with a tem- keep growing until they confront other solid material (nanofiller
perature factor around 0.04 Å2, in agreement with the previous or solid acid). The growth of the coherent lattice is facilitated by
model study and also with ab initio calculations. No constraints the high mobility in the material and the annealing that results
were needed to fit this position in the different samples, except even at room temperature, giving the relatively large coherent
in the case of the 40-nm particles. grain sizes as observed by neutron diffraction.
The effect of the deuterium ion intercalation is also Throughout this coherent phase there will be pockets in
dependent on the molar ratio between the solid acid and the which the amorphous phase has grown around the nanoparticle
nanoparticles (see Table 2). With an increasing relative molar in a core-shell fashion. The whole picture combined suggests
amount of TiO2, the number of inserted deuterium ions in that the morphology of the nanocomposites is such that the
TiO2 was found to decrease for identically sized nanoparticles, large grains of the solid acid are perforated like a Swiss cheese
while the vacancy concentration in CsDSO4 seems to increase. by small nanoparticles with a shell of amorphous solid acid

Table 2. Overview Parameters Rietveld Refinement. (SA = solid acid).

Sample wRp Fraction D In TiO2 [%] Fraction D Out SA [%] Position In TiO2 Grain Size SA [nm]
Bulk 0.0176 – – – –
Nano7 0.0165 10.6 ± 0.5 22 ± 2 0.43 ± 0.01 –
Nano15 0.0144 5.0 ±0.7 18 ± 2 0.43 ± 0.02 70 ± 6
Nano24 0.0218 3.2 ± 0.6 12 ± 2 0.43 ± 0.01 150 ± 6
Nano241 0.0190 4.3 ± 0.6 5.6 ± 2 0.43 ± 0.01 160 ± 6
Nano40 0.0202 0.36 ± 0.7 1.0 ± 2 0.43 110 ± 7
Nano401 0.0195 1.8 ± 0.7 0.9 ± 2 0.43 280 ± 7

Adv. Funct. Mater. 2011, 21, 1364–1374 © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1371
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

2.7. Ab initio Calculations


0.10
Results from first-principles DFT calculations (for further
Fraction D in TiO2

details see previous reports[30]) agree well with the determined


values for ion insertion in TiO2, approaching values to within
roughly a factor of two as illustrated in Figure 10 by the red
0.05 line. It is important to state here that, also based on theoretical
considerations, such large proton insertion densities as found
here experimentally can be justified. The essential factor that
determines the observed high proton-insertion densities is the
0.00 negative enthalpy of formation for creating defects that involve
the transfer of protons from CsHSO4 to TiO2 or SiO2.
0 10 20 30 40
Particle Size /nm
Figure 10. Fraction of deuterium ions intercalated in the TiO2 lattice as 3. Conclusions
a function of the nanoparticle size. Red line illustrates the results of the
ab initio calculations. The effects of nanostructuring the solid acid CsHSO4 with
proton-accepting TiO2 and SiO2 nanoparticles on the mobility
and mobile fraction of the charge carriers have been observed
around them. For the very small nanoparticles the X-ray data using direct microscopic probes. Large fractions of protons
above indicate progressive amorphization with time. Sufficient are mobile at all temperatures measured, and the time-scales
protons are removed from the solid acid to make it X-ray-amor- of motion are much shorter than in the bulk material. This
phous throughout. The annealing process seems to be contra- high microscopic mobility underpins the orders-of-magnitude
dictory to the amorphization described above. A possible expla- increase in conductivity below the superprotonic transition to
nation lies in the difference in time-scales of these processes, phase I.
with the amorphization occurring at a much longer timescale. Direct experimental proof of the occurrence of space-charge
Attempts to observe the morphology using (cryo)TEM were effects in the nanocomposite proton conductor CsHSO4 and
ineffective, because the samples were not stable under the elec- nanoparticulate TiO2 have been shown in the form of deute-
tron beam, showing rapidly moving objects floating around, rium-ion intercalation in TiO2 nanoparticles and deuterium
which is related to the high intrinsic mobility in these samples. depletion in CsDSO4. The scale of intercalation is quite sub-
stantial at up to 10% in 7-nm TiO2 particles. The intercalation
is size-dependent, decreasing with increasing particle size due
2.6. Induced HSO4 Movement to decrease of surface/volume ratios. Though the measure-
ments were done on a solid-acid composite, the occurrence
Besides the direct intercalation of deuterium ions in the TiO2 of space charges in other systems of proton conductors and
lattice, other effects resulting from the nanostructure in the nanoparticles is expected to be more general.
conducting solid-acid phase were studied.
By creating Fourier density maps from the
diffraction-peak intensities, selected nuclear
densities could be observed. The nuclear
densities around the nanocomposite oxygen
atoms in the SO4 tetrahedra had a much more
ellipsoidal shape than the bulk oxygen densi-
ties (Figure 11), which indicates increased
librational motions or static displacements of
the oxygen atoms in the tetrahedra. Although
we assume that there is an amorphous
layer in between the crystalline solid-acid
phase and the nanoparticles, it is clear
that due to the nanostructuring, the crys-
talline part of the solid acid is influenced
by the space-charge effects, that is, pro-
tons from the stiff hydrogen-bond network
of phase II are released conferring orien-
tational freedom on the SO4 groups. The
extraction of protons from the solid-acid
Figure 11. Left: Fourier maps of nuclear density of bulk CsHSO4. Right: Fourier maps of the
phase may also destabilize the matrix, nanocomposite with 24 nm TiO . Left and right the levels of the oxygen and Cs isosurfaces are
2
favoring amorphization as observed with the same. In the nanocomposite the O nuclear densities have more elongated shapes due to
X-ray characterization. positional and dynamic disorder.

1372 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374
www.afm-journal.de
www.MaterialsViews.com

with 2H NMR again the (deuterated) nanocomposite solid acid with 24-nm

FULL PAPER
Combination of the neutron and NMR results yields the fol-
lowing overall picture of the ionic mobility induced by nanos- TiO2 was selected for the temperature scan. The deuterated samples were
tructuring: Due to the space-charge effect, vacancies are cre- synthesized according to literature in a dry nitrogen environment.
Neutron-diffraction data on the CsDSO4–TiO2 nanocomposite samples
ated in the solid acid at temperatures below the superprotonic were collected at room temperature on the high flux diffractometer
phase-transition temperature, which allow a large fraction of the Polaris at the pulsed-spallation neutron source ISIS, Rutherford Appleton
hydrogen ions to become as mobile as they are in the superpro- Laboratories, UK. Deuterated samples were used to suppress the
tonic phase (which has abundant intrinsic vacancies) because incoherent background caused by protons. The nanocomposite samples
there are empty sites available to which ions can move. The consisted of the solid acid with the following particles: 7-, 15-, 24-,
vacancies introduce the high mobility at temperatures approx- and 40-nm TiO2. Additionally the 24-nm TiO2 composite had an extra
sample with a molar ratio of 1:1 acid:filler (referred to as nano241) and
imately 140 °C lower than observed in bulk. This behavior
the 40-nm TiO2 sample had an additional sample with a molar ratio of
also occurs in the amorphous phase, as observed using NMR 1:1/2 (nano401). Analysis was performed using the Rietveld refinement
spectroscopy. program General Structure and Analysis System (GSAS).
The crystalline part of the solid acid is influenced by the The QENS experiments were performed on fully protonated samples
space charges, as is observed from enhanced vibrational on IN5 at the high-flux reactor of the Institute Max von Laue-Paul
motions in the SO4 groups, and from the short 1H NMR T1 Langevin (ILL), with an incident wavelength of 11.5 Å, a resolution
function with a full width at half maximum of ∼6 μeV, and covering a Q
relaxation times of the bulk crystalline peak (at 11.0 ppm). A
range of 0.2 – 0.9 Å−1. Air-tight, flat, Al sample containers were used.
fully crystalline nanocomposite sample was shown to exhibit NMR measurements were performed at the solid-state NMR facility
a similar degree of mobility as a mostly amorphous nanocom- for advanced materials science, Radboud University in Nijmegen, The
posite sample, as a large fraction of the phase II peak in the Netherlands. Using a 400 MHz spectrometer with a static magnetic field
NMR spectrum showed a short T1 time. strength of 9.4 T, the 1H and 2H Larmor frequencies were 399.95 and
The space-charge effect in the crystalline bulk phase desta- 61.39 MHz, respectively. The measurements were performed over a wide
bilizes the phase as evidenced by the SO4 reorientations. This range of temperatures from 173K to 433K. A 3.2 mm air-tight zirconia
rotor was used to measure the magic-angle-spinning (MAS) 1H or 2H 1D
rotational degree of freedom may also contribute to the gradual spectra and spin-lattice relaxation times. The spinning frequencies were
time-dependent transformation of the crystalline phase to the between 14 and 17 kHz.
X-ray amorphous phase with decreasing particle sizes. Ionic conductivities were measured on thin pressed pallets of
Further analysis of the data showed the grain sizes of the dry powder. A sputtered gold coating was used for optimal contact.
solid acids to be a few times larger than the added nanoparti- Impedance spectroscopy was performed between RT and 160 °C and
cles. The nanoparticles however, are present in a large space- conductivities extracted. The results (not shown here) are fully in line
with what others have reported.[18,33–35] The superprotonic transition
filling quantity. Morphologically this leads to the conclusion
in a pure bulk CsHSO4 sample shows the three orders of magnitude
that the nanocomposites consist of relatively large solid-acid change in conductivity reported elsewhere,[26] while in the composites
grains, perforated by nanoparticles and with an amorphous the conductivity is much increased at low temperatures, which leads to
shell surrounding them. Spontaneous annealing, aided by the the strong reduction of the transition effects on the conductivity.
high intrinsic mobility, causes larger coherent domains of the
solid-acid structure.
Acknowledgements
Financial support for ISIS beam time was obtained from the Netherlands
Experimental Section Organization for Scientific Research (NWO). NWO is furthermore
thanked for financial support of the solid-state NMR facility for advanced
The solid acid cesium hydrogen sulfate was synthesized by dissolving materials science at the Radboud University in Nijmegen. This article is
cesium carbonate (Cs2CO3) in demineralized water and slowly the result of joint research in the Delft Research Centre for Sustainable
adding sulfuric acid (diluted) in the correct stoichiometry.[31] After Energy and the 3TU. Centre for Sustainable Energy Technologies. J.
recrystallization from a methanol/water mixture, the solid acid was van Os, G. Janssen, and H. Janssen are acknowledged for technical
filtered and dried in a vacuum oven. After careful drying, a part of the support with the NMR measurements. The Institute Laue-Langevin is
CsHSO4 was redissolved in demineralized water and TiO2 or SiO2 acknowledged for QENS measurement time on IN5.
nanoparticles were added in known quantities. The solvent water was
slowly evaporated under stirring. The nanocomposite samples created Received: September 15, 2010
were then carefully dried in the vacuum oven and subsequently crushed
Published online: March 24, 2011
in a mortar, resulting in a powder.
The chemicals cesium carbonate (reagentplus, 99%) and the
sulfuric acid (ACS reagent, 98%) were bought from Sigma-Aldrich. The
[1] N. Ohta, K. Takada, L. Q. Zhang, R. Z. Ma, M. Osada, T. Sasaki, Adv.
nanoparticles used were: 7-nm (Nanoamor), 15-nm (Nanoamor), 24-nm
Mater. 2006, 18, 2226.
(Evonik), and 40-nm (Evonik) TiO2, 7-nm (Aerosil300, Evonik) and
[2] R. Waser, M. Aono, Nat. Mater. 2007, 6, 833.
40-nm (AerosilOX50, Evonik) SiO2. The TiO2 nanoparticles were carefully
checked for size, polydispersity, and purity with TEM and XRD.[32] [3] V. V. Zhirnov, R. K. Cavin, Nat. Nanotechnol. 2008, 3, 377.
The QENS measurements were performed on bulk crystalline CsHSO4 [4] J. Maier, Nat. Mater. 2005, 4, 805.
and nanocomposite CsHSO4 with the following nanoparticles: 24-nm [5] High Level Group, Vision 2020 – Nanoelectronics at the Center
TiO2, 7- and 40-nm SiO2. The molar ratio of solid acid to nanoparticles of Change, vol. 6, Office for Official Publications of the European
was 1:2. Communities, Luxembourg, 2004.
The nanocomposite with 24-nm TiO2 was selected for a temperature- [6] J. Jamnik, J. Maier, Phys. Chem. Chem. Phys. 2003, 5, 5215.
dependent 1H NMR scan up to 433K and compared with the bulk [7] P. Knauth, Solid State Ionics 2006, 177, 2495.
crystalline sample. Other nanocomposite samples, (7- and 40-nm TiO2, [8] J. Otomo, H. Shigeoka, H. Nagamoto, H. Takahashi, J. Phys. Chem.
7- and 40-nm SiO2) were measured at room temperature. For comparison Solids 2005, 66, 21.

Adv. Funct. Mater. 2011, 21, 1364–1374 © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com 1373
www.afm-journal.de
www.MaterialsViews.com
FULL PAPER

[9] V. G. Ponomareva, G. V. Lavrova, L. G. Simonova, Solid State Ionics [23] J. Maier, Solid State Ionics 2002, 154, 291.
1999, 118, 317. [24] J. Maier, Prog. Solid State Chem. 1995, 23, 171.
[10] N. Sata, K. Eberman, K. Eberl, J. Maier, Nature 2000, 408, 946. [25] W. K. Chan, W. J. H. Borghols, F. M. Mulder, Chem. Commun 2008,
[11] A. I. Baranov, Crystallogr. Rep. 2003, 48, 1012. 47, 6342.
[12] H. J. Neef, Energy 2009, 34, 327. [26] S. Q. Wang, J. Otomo, M. Ogura, C. Wen, H. Nagamoto,
[13] A. V. Belushkin, M. A. Adams, S. Hull, A. I. Kolesnikov, H. Takahashi, Solid State Ionics 2005, 176, 755.
L. A. Shuvalov, Physica B 1995, 213, 1034. [27] S. Hayashi, M. Mizuno, Solid State Commun. 2004, 132,
[14] C.R. I. Chisholm, R. B. Merle, D. A. Boysen, S. M. Haile, Chem. 443.
Mater. 2002, 14, 3889. [28] Y. Daiko, S. Hayashi, A. Matsuda, Chem. Mater. 2010, 22,
[15] S. Hayashi, M. Mizuno, Solid State Ionics 2004, 171, 3418.
289. [29] M. N. Kislitsyn, Materials Chemistry of Superprotonic Solid Acids,
[16] S. M. Haile, D. A. Boysen, C.R. I. Chisholm, R. B. Merle, Nature California Institute of Technology, USA, 2009.
2001, 410, 910. [30] L. A. Haverkate, W. K. Chan, F. M. Mulder, Adv. Funct. Mater. 2010,
[17] S. M. Haile, C.R. I. Chisholm, K. Sasaki, D. A. Boysen, T. Uda, 20, 4018.
Faraday Discuss. 2007, 134, 17. [31] A. V. Belushkin, I. Natkaniec, N. M. Pakida, L. A. Shuvalov,
[18] G. V. Lavrova, M. V. Russkikh, V. G. Ponomareva, N. F. Uvarov, Russ. J. Wasicki, J. Phys. C 1987, 20, 671.
J. Electrochem. 2005, 41, 485. [32] M. Wagemaker, W.J. H. Borghols, F. M. Mulder, J. Am. Chem. Soc.
[19] A. I. Baranov, L. A. Shuvalov, N. M. Shchagina, JETP Letters 1982, 2007, 129, 4323.
36, 459. [33] J. Otomo, N. Minagawa, C. J. Wen, K. Eguchi, H. Takahashi, Solid
[20] A. V. Belushkin, C. J. Carlile, L. A. Shuvalov, J. Phys. Cond. Mat. 1992, State Ionics 2003, 156, 357.
4, 389. [34] V. G. Ponomareva, G. V. Lavrova, Solid State Ionics 1998, 106,
[21] D. Arcon, R. Blinc, J. Dolinsek, L. A. Shuvalov, Phys. Rev. B 1997, 55, 8961. 137.
[22] R. Blinc, J. Dolinsek, G. Lahajnar, I. Zupancic, L. A. Shuvalov, [35] V. G. Ponomareva, G. V. Lavrova, Solid State Ionics 2001, 145,
A. I. Baranov, Physica Status Solidi B 1984, 123, K83-K87. 197.

1374 wileyonlinelibrary.com © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Adv. Funct. Mater. 2011, 21, 1364–1374

You might also like