Вы находитесь на странице: 1из 5

By Deepak Shah, Pralay Maiti, Erica Gunn,

Daniel F. Schmidt, David D. Jiang, Carl A. Batt,


and Emmanuel P. Giannelis*
Addition of nanoclays or other nanoparticles into various
polymers to produce nanocomposites has been extensively
utilized in an attempt to enhance the mechanical, physical,
and thermal properties of polymers. While some interesting
properties have been demonstrated, the resulting nanocomposites have yet to realize their full potential. Nanoparticles
in general, and nanoclays in particular, with their nanometer
size, high surface area, and the associated predominance of interfaces in the nanocomposites, can function as structure and
morphology directors, for example stabilizing a metastable or
conventionally inaccessible polymer phase, or introduce new
energy dissipation mechanisms. Thus, what distinguishes
nanoparticles from conventional micrometer-size rigid reinforcements is that their role might not be limited to only adding stiffness to the polymer, but also to directing morphology,
as well as introducing new energy-dissipation mechanisms
leading to enhanced toughness in the nanocomposites. Herein
we demonstrate this potential by reporting a remarkable (order of magnitude) increase in toughness with a concurrent increase in stiffness in a poly(vinylidene fluoride) (PVDF)
nanocomposite.
The kinetics of crystallite growth and the details of crystallite morphology of semicrystalline polymers can be affected
by the presence of layered silicates.[1,2] Although some

[*] Prof. E. P. Giannelis, Dr. P. Maiti, Dr. E. Gunn,[+]


Dr. D. F. Schmidt,[++] Dr. D. D. Jiang
Department of Material Science and Engineering, Cornell University
Ithaca, NY 14853 (USA)
E-mail: epg2@cornell.edu
Dr. D. Shah
Department of Chemistry and Chemical Biology, Cornell University
Ithaca, NY 14853 (USA)
Prof. C. A. Batt
Department of Food Science, Cornell University
Ithaca, NY 14853 (USA)
[+] Summer research student: Department of Chemistry, Simmons
College, 300 The Fenway, Boston, MA 02115-5898, USA.
[++] Present address: Groupe BASF, Institut de Science et d'Ingnierie
Supramolculaires (ISIS)/Universit Louis Pasteur (ULP), 8, alle
Gaspard Monge, F-67083 Strasbourg, France.
[**] We thank Cornell Center for Materials Research (CCMR), AFOSR,
NASA and ONR, for financial support provided for this research
and gratefully acknowledge the use of CCMR-sponsored XRD, and
SEM/TEM facilities, as well as the time and assistance of M. Weathers (XRD) and J. Hunt (TEM/SEM). We thank R. Dieckmann for the
use of his ATR-FTIR facilities, and H. Liu for her time and assistance
in this regard. We also thank E. Reynaud, R. Krishnamoorti,
R. A. Vaia, and A. J. Lovinger for their comments and suggestions.

Adv. Mater. 2004, 16, No. 14, July 19

changes in morphology have been described in polymer/nanoparticle hybrids,[37] near-total stabilization and control of a
crystalline phase, coupled with dramatic enhancements in
materials properties, has not yet been reported.
PVDF is an important engineering plastic. It is used extensively in the pulp and paper industry due to its resistance to
halogens and acids, in nuclear-waste processing for radiation-,
and hot-acid applications, and in the chemical processing industry for chemical and high-temperature applications. It is
also used in various device applications, due to its unique
piezoelectric[810] and pyroelectric[11] properties. There are five
known crystalline forms or polymorphs of PVDF: a, b, c, d,
and e.[12] The a phase is the most common in melt crystallization, and remains the dominant crystalline form versus the b,
and c phases. The c phase does not form except at high
temperatures and pressures. Earlier reports have shown that
the a phase (chain conformationtrans-gauche trans-gauche,
tg+tg) is inactive with respect to piezo- and pyroelectric properties, while the b form (all trans) exhibits the most activity,
and is thus the focus for electromechanical and electroacoustic transducer applications. Thus, the b form has great technological utility and there have been numerous attempts to stabilize this phase. For example, the b form of the PVDF has
been obtained by careful crystallization from solution,[13] by
melt crystallization at high pressure, by application of a strong
electric field,[14] by molecular epitaxy,[15] and by preparing a
carbon-coated, highly oriented ultrathin film.[16] Earlier reports have indicated the possibility of stabilizing a phase of a
crystalline polymer in the presence of well-dispersed layered
silicates in the polymer matrix. Specifically, increases in the
amount of the c phase in Nylon 6[57] and the b phase in
PVDF[3,4] nanocomposites have previously been observed. In
this study we report the first instance of a remarkable order of
magnitude enhancement in the toughness of a nanocomposite
compared to that of the pristine polymer. The enhanced
toughness is attributed to the structural and morphological
changes induced by the nanoparticles as well as their increased mobility compared to micrometer size fillers, which
contributes to a new energy dissipation mechanism during deformation.
The nanoclays used in this study are sodium montmorillonite (NCMU) and bis(hydroxyethyl)methyltallowammonium
ion-exchanged montmorillonite (NCM). These nanoclays
have a 2-dimensional platelet like geometry and are composed of 1 nm thick layers. Polymer chains can be intercalated
between the layers from solution or a melt when favorable interactions between the polymer chains and the surface of the
inorganic host are present.[17,18] X-ray diffraction (XRD) analysis shows the formation of such an intercalated PVDF nanocomposite, referred to hereafter as PVDFNCM. A comparison of the diffraction patterns of the layered host before and
after intercalation shows an increase in the gallery spacing of
the layered silicates from 1.8 to 2.9 nm in PVDFNCM
(Fig. 1a), while it remains unchanged for the hybrid with the
unmodified layered silicate (PVDFNCU). The transmission
electron microscopy (TEM) image of the nanocomposite

DOI: 10.1002/adma.200306355

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

COMMUNICATIONS

Dramatic Enhancements in Toughness


of Polyvinylidene Fluoride
Nanocomposites via Nanoclay-Directed
Crystal Structure and Morphology**

1173

NCM
PVDFNCM

Intensity / a.u.

COMMUNICATIONS

10

2 / deg

Figure 1. a) XRD of organomodified silicate (NCM) and the corresponding nanocomposite (PVDFNCM). b) Bright-field TEM image of the PVDF
nanocomposite (PVDFNCM). The scale bar is 500 nm.

(Fig. 1b) shows good, homogeneous dispersion of the layered


silicates in the PVDF matrix. Stacks of alternating organic/inorganic layers or tactoids with an average thickness of 100 nm
can be seen.
Clear indications of a change in the crystal structure of the
polymer phase are observed in the XRD traces (Fig. 2a). In
the nanocomposites, the (020), (110), and (111) peaks corresponding to the a phase in pure PVDF disappear and are replaced by the (200/110) and (001) peaks corresponding to the
b phase.[16] This change can be observed by the addition of as
little as 2 wt.-% of the surface-modified nanoclay. The peaks
characteristic of the b phase appear with increasing intensity
as the nanoclay type is changed from pristine to surface modified. Materials based on the pristine layered silicate do not experience significant changes in crystal structure, and are comprised mostly of the a phase, with a weak b peak at 2h = 31.
Changes in the crystal structure are corroborated by Fouriertransform infrared (FTIR) spectroscopy measurements. The
a-PVDF peaks[19] at 763 and 796 cm1 can be quite clearly observed for both pure PVDF and for the PVDFNCU. A complete disappearance of these peaks is seen for PVDFNCM. Simultaneously, strong b-PVDF peaks at 840 cm1 and 500 cm1
are observed in the nanohybrids, indicating a change in the favored crystalline polymorph (Fig. 2b). Note that the pristine
layered silicate leads to an immiscible hybrid with PVDF due
to a lack of any favorable interaction between the polymer
and the silicate.

/1
10

b
PVDFNCM

00
1

20
0

PVDFNCM

PVDFNCU

PVDFNCU

1000

0
11 20
0

PVDF
11
1

Intensity / a.u.

a
2000

The mechanism by which the change from the a to b phase


occurs is believed to involve a segmental flip-flop motion,
inversion mode of chain segments, or more probably, a
combination of the two.[20] This change appears to be favored
in the presence of properly surface-modified layered silicates.
A probable mechanism for the stabilization of the b phase
could be due to the matching of the crystal lattice of the nanoclay with that of the PVDF b phase. From the XRD analysis it
is clear that the levels of both polymer and nanoparticle dispersion and interactions dictate the structural conversion of
the a phase to the b phase of PVDF. This idea is further reinforced by the fact that hybrids made using spherical amorphous silica nanoparticles, both pristine and surface treated,
did not show any changes in the structure and morphology of
the PVDF.
Polarized optical microscopy (POM) and scanning electron
microscopy (SEM) studies have provided further evidence of
the extent and selectivity of the phase transformation induced
in the PVDF matrix. For these studies, thin films (~ 30 lm) of
the pristine polymer and the nanocomposite were crystallized
on a hot stage. Figure 3 shows the SEM and POM (inset) images of the pure PVDF and the corresponding nanocomposites. The presence of large, isotropic spherulites is evident in
the POM inset (Fig. 3a), whereas the main SEM image shows
the features of a single a-spherulite (average diameter
~ 90 lm) with crystallites of lateral dimensions of the order of
2.5 lm in the pure PVDF. As reported previously, the crystallites show a significant orientation perpendicular to the chain
axis (c-axis),[21] except towards the spherulite centers. It can
also be seen that spherulites are impinging upon each other
after full solidification (inset, Fig. 3b). As mentioned previously, unmodified layered silicates are immiscible with
PVDF. Accordingly, only a partial disruption of the spherulitic structure is observed in this system, as most of the nanoparticle surface is inaccessible to the polymer matrix due to
nanoparticle aggregation. Nevertheless, the nanoparticle aggregates act as nucleating agents, and as a result the average
spherulite diameter decreases to ~ 3 lm, but the presence of
spherulites is still evident in the POM (Fig. 3b inset). SEM
images (Fig. 3b) indicate that the crystallites are arrested in
the early stages of coalescence, but that the crystalline morphology of the matrix is otherwise similar to that of the pure
PVDF. It is, however, possible to locate a few isolated b crys-

PVDF
0
10

20

30

40

2 / deg

1174

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1000

800

600
-1

Wavenumber / cm

http://www.advmat.de

Figure 2. a) XRD patterns showing development of the b phase in PVDF nanocomposites. b) FTIR curves showing
appearance of the b peak (*) in PVDF
nanocomposites.

Adv. Mater. 2004, 16, No. 14, July 19

(b)

(c)

ordered a crystallites in pristine PVDF to disordered, fiberlike b crystallites in the presence of surface-modified, layered
silicates. A direct correspondence can be drawn between the
size of the a spherulites and the extent of b phase nucleation,
with the level of nanoscale dispersion and the degree of polymer/nanoclay interactions in the nanocomposite systems. It is
also striking to note that the presence of surface modification
can mean the difference between partial and exclusive b nucleation in the PVDF matrix.
The degree of crystallinity was determined via differential
scanning calorimetry (DSC) at a scan rate of 10 C min1. The
data obtained showed no changes in either the Tm (the melting point of the crystalline regions of the semi-crystalline
polymer) or the overall degree of crystallinity for the
PVDFNCM systems, as compared with that of the pure
PVDF. Based on the DSC results, the degree of crystallinity
for the PVDF and PVDFNCM samples was calculated to be
48 % and 49 %, respectively. The melting of the a and b
phases cannot be differentiated from DSC, as their melting
points are known to be within a few degrees of each other.[22]
The thermomechanical properties were studied using dynamic
mechanical analysis (DMA). At room temperature we observe an increase of 35 % in the storage modulus for the nanocomposite system, as compared to the pure PVDF. In addition
the glass transition at 40 C remains unchanged for the neat
polymer and the nanocomposite.
Most rigid fillers produce increases in the modulus of the
composite versus that of the unfilled polymer. However, this
is generally associated with a significant decline in the elongation at break.[23] We show that the changes in polymer nanostructure and morphology have a profound impact on the
mechanical response of the PVDF nanocomposites. The properties of these systems are, to the best of our knowledge,
unique, as these materials show simultaneous increases in
both stiffness and toughness (Fig. 4), leading to a situation

6x10

4x10

2x10

COMMUNICATIONS

(a)

tallites in the SEM (Fig. 3b), which give rise to the weak b
peak observed in XRD (Fig. 2).
In contrast, the PVDFNCM nanocomposite exhibits dramatic differences. It can be observed that the sample lacks birefringence (inset, Fig. 3c). This is due to the random orientation of the smaller b crystallites, versus the larger, highly
ordered spherulitic structures previously observed. In this
case, no trace of the a spherulites is observed, either in the
POM analysis or in the SEM image. It can also be seen that
the fiber-like structures corresponding to the b crystallites
(average size ~ 0.6 lm) are quite well-defined and span the
entire observable area, further emphasizing the pronounced
transition from the a to the b phase in the presence of the appropriate surface-modified layered silicates. Thus, a distinct
transition is observed, from large spherulites made of highly
Adv. Mater. 2004, 16, No. 14, July 19

Stress / Pa

Figure 3. SEM images with POM insets showing crystal morphology of


a) PVDF, b) PVDFNCU, c) PVDFNCM.

PVDF
PVDFNCU
PVDFNCM
0
0

50

100

150

Strain / %
Figure 4. Stressstrain curves for pure PVDF, PVDFNCU, and PVDFNCM
showing the dramatic increase in elongation at break for the nanocomposite.

http://www.advmat.de

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1175

COMMUNICATIONS
1176

where the overall properties are enhanced without any corresponding trade offs (toughness here is defined as the area under the stressstrain curve). The Young's modulus increases
from 1.3 to 1.8 GPa, while the elongation at break increases
from 20 % to 140 %, giving the hybrid a toughness ~ 700 %
higher than that of the neat polymer. We have recently found
that further optimization of the processing conditions leads to
an elongation for the nanocomposites of the order of 200
250 %. With respect to macroscopic behavior during these
tensile tests, neck formation and plastic flow are evident in
the nanocomposites, whereas rapid strain hardening and brittle fracture are observed in the pure polymer. It should be
noted that in the case of pristine and surface-modified silica
as well as the pristine layered silicate, no change in toughness
was observed. This is strongly indicative of the role of dispersion and interactions between the matrix and layered nanoparticles in the resulting behavior.
While we cannot test b-PVDF in bulk form due to the difficulties in stabilizing the phase discussed earlier, we have
tested b-PVDF thin film samples prepared by axial stretching.
The films are relatively ductile in the transverse stretching direction but show much lower elongation in the longitudinal
direction, in good agreement with previous studies of b-PVDF
thin films.[24] However, the toughness of these films is about
half of that seen in the nanocomposites. When one considers
that addition of reinforcing elements typically leads to decreases in elongation, an additional decrease rather than an
increase is expected for the nanocomposites. Therefore, we
conclude that nanostucturing of the polymer induced by the
nanoparticles as well as the presence of the nanoparticles
themselves and not just the b phase formation are responsible
for the enhanced toughness. Although there is no doubt that
the transformation of the phase/morphology plays a significant part in the toughening of the nanocomposite, this is the
first reported instance of such a dramatic improvement in the
toughness of the polymer by dispersing a nanoparticle in the
matrix.
DSC of the samples following mechanical testing shows
slightly increased values of overall crystallinity (PVDF 52 %,
PVDFNCM 53 %), the increase being on the order of ~ 5 %
for both the PVDF and the nanocomposites. Although further
studies are underway to elucidate the exact mechanism responsible for this behavior, from the data obtained so far it is
quite clear that the dramatic increase in the toughness of
these nanocomposites can only be accounted for by changes
in crystal structure, morphology, and the presence of properly
surface modified layered silicates.
In the case of the PVDF nanocomposites, we postulate that,
in contrast to the neat PVDF, nucleation of the fiber-like b
phase on the faces of individual silicate layers leads to a structure much more conducive to plastic flow under applied stress.
This might give rise to a more efficient energy-dissipation
mechanism in the nanocomposites, thereby delaying crack formation. In addition, recent molecular-dynamics studies have
suggested that nanocomposites might possess increased
toughness due to the ability of nanoparticles to dissipate ener-

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

gy[25] because of their mobility under applied stress. Nanoparticles and polymer chains have comparable time scales for
motion because of their size similarity. Due to their mobility,
the nanoparticles can act as temporary crosslinks between
polymer chains, providing localized regions of enhanced
strength, which in turn can retard the growth of cracks or cavities. Similarly, energy dissipation might be further enhanced
due to the presence of smaller and hence more mobile b crystallites. Our results show that good nanoscale dispersion,
coupled with favorable polymersilicate interactions, is critical for improved toughness. Since the toughness enhancement
is observed in intercalated nanocomposites, nanoclay exfoliation in the polymer matrix might not necessarily be required
to achieve the desired property enhancements.
In conclusion, we report the first ever instance of an orderof-magnitude increase in the toughness of a polymer, with a
concurrent increase in stiffness, through nanostructural control. Addition of appropriately surface-modified nanoparticles
into a polymer can control the crystalline phase and morphology of the polymer matrix. The enhanced toughness is attributed to the structural and morphological changes induced by
the nanoparticles, as well as their increased mobility as compared to micrometer-size fillers, which contributes to a new
energy-dissipation mechanism during deformation. This study
provides a new approach for toughening of polymers and in
doing so, points the way towards a novel approach for the design of new nanocomposite materials.

Experimental
Synthesis of Nanocomposites: PVDF (Kynar 721) was provided by
the Atofina Chemical Co. as a fine powder and used as received. Unmodified sodium montmorillonite (NCMU) and bis(hydroxyethyl)methyltallowammonium ion-exchanged montmorillonite (NCM) were
obtained from Southern Clay Products. A polymer/nanoclay premix
was prepared by combining the 5 wt.-% of nanoclay and PVDF powders in a speed-mixer. This premixed powder was then melt extruded
using a DSM twin screw microcompounder at a temperature of 200 C
under flowing nitrogen. For tensile testing, standardized dog-bone
specimens were prepared via microinjection of the extruded materials
using a Microinjector at a 240 C barrel temperature and a 95 C mold
temperature, with an injection pressure of 80 psi. In order to ensure a
proper comparison, pure PVDF samples were extruded and microinjected in an identical fashion.
Characterization: X-ray diffraction spectra were collected on a
Scintag Inc. h-h diffractometer equipped with an intrinsic germanium
detector system instrument, using a Cu Ka source with a wavelength
of k = 1.54 and a scan rate of 2 min1. Infrared spectra were recorded in the reflectance mode between 400 to 1000 cm1 using a Bruker instruments. Dispersability of the nanoparticles in the matrix was
evaluated using a transmission electron microscope operated at an accelerating voltage of 100 kV. Morphological features of thin films
(30 lm) were also studied using a polarizing optical microscope. The
polarized optical microscope was fitted with a color sensitive plate
that allowed for capture of images that were then developed and analyzed. Degree of crystallinity and melting temperature were determined via differential scanning calorimetry, over a temperature range
of 50 to 200 C at a scan rate of 10 C min1. Tensile tests were performed on the microinjected dog-bone shaped samples at room tem-

http://www.advmat.de

Adv. Mater. 2004, 16, No. 14, July 19

Received: October 24, 2003


Final version: April 8, 2004

[1] P. Maiti, P. H. Nam, M. Okamoto, N. Hasegawa, A. Usuki, Macromolecules 2002, 35, 2042.
[2] P. Maiti, P., M. Okamoto, Macromol. Mater. Eng. 2003, 288, 440.
[3] L. Priya, J. P. Jog, J. Polym Sci., Part B: Polym. Phys. 2002, 40, 1682.
[4] L. Priya, J. P. Jog, J. Polym Sci., Part B: Polym. Phys. 2003, 41, 31.
[5] Q. Wu, X. Liu, L. A. Berglund, Macromol. Rapid Commun. 2001,
22, 1438.
[6] L. J. Mathias, R. D. Davis, W. L. Jarrett, Macromolecules 1999, 32,
7958.
[7] D. L. Van der Hart, A. Asano, J. W. Gilman, Chem. Mater. 2001, 13,
3781.
[8] Q. M. Zhang, H. Li, M. Poh, F. Xia, Z. Y. Cheng, H. Xu, C. Huang,
Nature 2002, 419, 284.
[9] Q. M. Zhang, V. Bharti, X. Zhao, Science 1998, 280, 2101.
[10] H. Xu, Z. Y. Cheng, D. Olson, T. Mai, Q. M. Zhang, Appl. Phys.
Lett. 2001, 78, 2360.
[11] P. C. A. Hammes, P. P. L. Regtien, Sens. Actuators, A 1992, 32, 396.
[12] G. Guerra, F. E. Karasz, W. J. MacKnight, Macromolecules 1986, 19,
1935.
[13] R. Miller, J. Polym. Sci., Polym. Chem. Ed. 1976, 14, 2325.
[14] J. Scheinbeim, C. Nakafuku, B. A. Newman, K. D. Pae, J. Appl.
Phys. 1979, 50, 4399.
[15] A. J. Lovinger, Polymer 1981, 22, 412.
[16] J. Wang, H. Li, J. Liu, Y. Duan, S. Jiang, S. Yan, J. Am. Chem. Soc.
2003, 125, 1496.
[17] R. A. Vaia, H. Ishii, E. P. Giannelis, Chem. Mater. 1993, 5, 1694.
[18] B. K. G. Theng, Formation and Properties of Clay-Polymer Complexes, Vol. 9, Elsevier, Amsterdam 1979.
[19] M. Kobayashi, K. Tashiro, H. Tadokoro, Macromolecules 1975, 8,
158.
[20] Y. Takahashi, Y. Matsubara, H. Tadokoro, Macromolecules 1982, 15,
334.
[21] F. L. Binsbergen, B. G. M. de Lange, Polymer 1968, 9, 23.
[22] W. M. Prest, Jr., D. J. Luca, J. Appl. Phys. 1978, 49, 5042.
[23] L. E. Nielson, J. Appl. Polym. Sci. 1966, 10, 97.
[24] S. Lanceros-Mendez, J. F. Mano, A. M. Costa, V. H. Schmidt,
J. Macromol. Sci. Phys. 2001, B40, 517.
[25] D. Gersappe, Phys. Rev. Lett. 2002, 89, 058 301.

______________________

Adv. Mater. 2004, 16, No. 14, July 19

Molecular-Level Insulation: An
Approach to Controlling Interfacial
Charge Transfer**
By Saif A. Haque*, Jong S. Park, Mohan Srinivasarao,
and James R. Durrant

COMMUNICATIONS

perature, using a tensile tester and a strain rate of 5 mm min1. Five


samples were tested for each specimen. The results obtained from
these experiments were within 15 % assuring good statistics.

The immobilization of redox-active or photo-active molecules on semiconductor surfaces is fundamentally important


for the development of molecular electronic devices, including
solar cells, biological and chemical sensors, and heterosupramolecular systems.[15] One such area of research attracting
particular interest at present is the functionalization of nanocrystalline metal oxide films by the adsorption of molecular
dyes. Such functionalized metal oxide films are currently under investigation for device applications ranging from dyesensitized solar cells to electrochromic displays. The attachment of molecular dyes to the metal oxide nanoparticles is
typically achieved via acidic groups, such as carboxylates or
sulfonates, covalently attached to the sensitizer dye. Such acid
groups have been shown to form ester-type linkages to metal
oxide electrodes.[6] This attachment strategy is however limited to the use of dyes with suitable acidic groups, precluding
the use of most commercially available molecular dyes. Moreover such direct covalent attachment can often result in unnecessarily strong electronic interactions between the molecular
dye and the metal oxide electrode. This has been particularly
observed for small organic dyes, where direct covalent attachment to the electrode surface as been shown to result in undesirably fast interfacial recombination of photogenerated
charge-separated species,[7,8] limiting practical device applications.
In this paper we present an alternative strategy to the immobilization of molecular dyes on nanocrystalline TiO2 electrodes. Our strategy is based upon insulated sensitizer dyes, in
which the dye molecule is encapsulated at the molecular level
by a cyclodextrin macrocycle, limiting the interaction between
the TiO2 semiconductor surface and dye molecule and there-

[*] Dr. S. A. Haque, Dr. J. R. Durrant


Department of Chemistry, Centre for Electronic Materials and Devices
Imperial College of Science Technology and Medicine
London SW7 2AZ (UK)
E-mail: s.a.haque@imperial.ac.uk
J. S. Park, Dr. M. Srinivasarao
School of Textile and Fiber Engineering and
School of Chemistry and Biochemistry, Georgia Institute of
Technology
Atlanta, GA 30332 (USA)
[**] This work was supported by the Engineering and Physical Sciences
Research Council (EPSRC). M. S. would like to thank a grant
through the National Textile Center grant (Department of
Commerce) and the National Science Foundation (DMR-096240,
CAREER) for financial support. We would also like to thank
Emilio Palomares for helpful discussions and Alex Green for synthesis of the TiO2 nanoparticles. Supporting Information is available online from Wiley Interscience or from the author.

DOI: 10.1002/adma.200400327

 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1177

Вам также может понравиться