Sei sulla pagina 1di 5

COMMUNICATIONS

(5 nm diameter) are thus obtained with a rather narrow size distribution


(around 13 %). The particles are highly stable. By deposition of a drop on a
TEM carbon grid, with a filter paper underneath on a carbon substrate, nanocrystals self-organize in a 2D hexagonal network or 3D compact structure,
depending on the nanocrystal density [18].
Received: July 31, 2000
Final version: October 4, 2000

[1] P. C. Ohara, W. M. Gelbart, Langmuir 1998, 14, 3418.


[2] V. Kurrika P. M. Shafi, I. Felner, Y. Mastai, A. Gedanken, J. Phys. Chem.
B 1999, 103, 3358.
[3] A. Oron, S. H. Davis, S. G. Bankoff, Rev. Mod. Phys. 1997, 69, 931.
[4] V. V. Yaminsky, K. Thuresson, B. W. Ninham, Langmuir 1999, 15, 3683.
[5] M. P. Pileni, Langmuir 1997, 13, 3266.
[6] C. Petit, P. Lixon, M. P. Pileni, J. Phys. Chem. 1990, 94, 1598.
[7] JEOL electron microscope (100CX).
[8] Optical micrographs were obtained with a Nachet NS-400.
[9] H. E. Huppert, Nature 1982, 300, 427.
[10] A. M. Cazabat, F. Heslot, S. M. Troian, P. Carles, Nature 1990, 346, 824.
[11] H. Bnard, Rev. Gen. Sci. Pure Appl. 1900, 11, 1261.
[12] M. Block, Nature 1956, 178, 650.
[13] J. R. A. Pearson, J. Fluid. Mech. 1958, 4, 489.
[14] E. L. Koschmieder, J. Fluid Mech. 1967, 30, 9.
[15] T. Ondaruhu, J. Millan-Rodriguez, H. L. Mancini, A. Garcimartin, C.
Perez-Garcia, Phys. Rev. E 1993, 48, 1051.
[16] S. J. VanHook, M. F. Shatz, W. D. McCornick, J. B. Swift, H. L. Swinney,
Phys. Rev. Lett. 1995, 75, 4397.
[17] Organic Solvents, Physical Properties and Methods of Purification,
Techniques of Chemistry, Vol. II (Eds: J. A. Riddick, W. B. Bunger, T. K.
Sakano), Wiley, New York 1986.
[18] A. Taleb, C. Petit, M. P. Pileni, Chem. Mater. 1997, 9, 950.

A Novel Approach to Functionalized


Nanoparticles: Self-Crosslinking of
Macromolecules in Ultradilute Solution**
By David Mecerreyes,* Victor Lee, Craig J. Hawker,
James L. Hedrick, Andreas Wursch, Willi Volksen,
Teddie Magbitang, Elbert Huang, and Robert D. Miller*
Nanomaterials have been intensely studied due to their
peculiar properties and key role in the implementation of
nanotechnology. In particular, nanoparticles provide an important brick for nanoconstruction with a broad range of medical, mechanical, and electronic applications.[1] Although particles with sizes ranging from 50 nm to several micrometers
are commercially available via micro- and mini-emulsion
polymerization techniques,[2] the preparation of smaller nanoparticles is challenging. An original approach developed by
Wooley[3] and others leads to spherical particles with sizes
ranging between 15 and 50 nm by utilizing the chemical fixation of self-assembled micellar structures based on block
copolymers. An unimolecular approach involving the intro-

duction of branching sites provides a synthetic route toward


particulate macromolecules. The synthesis of new hyperbranched polymers and dendritic macromolecules exemplifies
this approach.[4] Another route developed by Antonietti and
others involves the introduction of crosslinking sites rather
than branching points leading to the formation of p-gels or
nanosponges.[5] Considerable attention has been devoted to
these materials as models for studying networks. In this work,
we describe the preparation of unimolecular particles by
intramolecular self-crosslinking of functional polymers leading to nanoparticles with sizes ranging from 3 to 15 nm and
discuss their properties and applications.
It is well known that macromolecules with pendant acrylate
functionalities can be crosslinked in concentrated solution
using a radical initiator such as 2,2-azo-bis-isobutyronitrile
(AIBN) generating covalent links between different chains
leading ultimately to a three dimensional network (Scheme 1a).
In this paper, we have investigated cases where the radical
crosslinking reaction is performed in ultradilute solution.
Under these conditions, the covalent links are exclusively
formed between segments of the same polymer chain leading
to an unimolecular particle (Scheme 1b).
Synthetically, the preparation of these polymeric nanoparticles involves two steps. The first involves the preparation of
potentially crosslinkable macromolecules. Our studies utilize
linear copolymers containing pendant acryloyl or methacryloyl
functionalities. Scheme 2a shows two examples of the controlled synthesis of macromolecules each bearing pendant
crosslinkable functionalities. Aliphatic polyesters such as 1 can
be obtained in one step by living ring-opening copolymerization (ROP) of 4-acryloyloxy caprolactone with either e-caprolactone or L,L-lactide.[6] The polystyrene copolymers bearing
methacryloyl pendant groups 2 can be prepared by subsequent
chemical modification of poly(styrene-co-hydroxyethyl methacrylate) copolymers previously prepared by nitroxide
mediated living free radical polymerization.[7] The second strategy is particularly attractive due to the variety of copolymers
containing pendant hydroxyl groups that can be prepared.
The unimolecular particles are subsequently prepared by
self-crosslinking in dilute solution using a radical initiator

[*] Dr. D. Mecerreyes, Dr. R. D. Miller, V. Lee, Dr. C. J. Hawker,


Dr. J. L. Hedrick, Dr. A. Wursch, Dr. W. Volksen, T. Magbitang,
Dr. E. Huang
IBM Almaden Research Center
650 Harry Road, San Jose, CA 95120-6099 (USA)
E-mail: rdmiller@almaden.ibm.com

[**] DM thanks the Govierno Vasco for a fellowship. Partial funding from
NIST-ATP Cooperative Agreement 70NANB8H4013 and CPIMA is
gratefully acknowledged.

204

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2001

Scheme 1.

0935-9648/01/0302-0204 $ 17.50+.50/0

Adv. Mater. 2001, 13, No. 3, February 5

COMMUNICATIONS

Scheme 2.

such as AIBN (Scheme 2b). After reaction of 1, unimolecular


particles 3 were recovered quantitatively by precipitation in
hexanes and characterized by infrared (IR) and nuclear magnetic resonance (NMR) spectroscopy and size exclusion chromatography (SEC).
Table 1 lists the results obtained from a number of different
macromolecules bearing either acrylate or methacrylate pendant groups. The self-crosslinking reaction was confirmed by
1
H NMR spectroscopy by monitoring the disappearance of
the vinyl signals associated with the (meth)acrylate pendant
group. As an example, Figure 1 shows the 1H NMR spectra of
both the precursor polymer (entry 1a) and the isolated, crosslinked nanoparticle. For the latter, the decrease in the signal
intensities of the vinyl resonances of the acrylate is consistent
with the radical vinyl addition. The extent of the crosslinking
reaction is estimated to range from 72 to 92 % for the different precursors as shown in the conversion data of Table 1.
This level of conversion is substantial for polymer analogous
transformations.
Interestingly, SEC of the crosslinked nanoparticles shows
that the hydrodynamic volumes of the particles as measured
by the elution time and related to polystyrene standards

Fig. 1. 1H NMR spectra (CDCl3) of poly(e-CL-co-acryloyloxyCL) (entry 1a),


Table 1: a) as synthesized, b) after intramolecular crosslinking.

(Table 1), are significantly smaller than those of the starting


linear macromolecules. This result is in agreement with the
expected decrease in the size of the macromolecule due to the
formation of new internal covalent bonds. In order to provide
some insight into the self-crosslinking reaction, the internal
covalent bonds of the PMMA-based (PMMA = poly(methylmethacrylate)) unimolecular nanoparticle 5 were eliminated

Table 1. Characteristics of the precursor polymers and self-crosslinked nanoparticles.


Precursors
Entry

Polymer

1a
1b
1c
1d
1e
1f
1g
1h

Poly(e-CL-co-acryloyloxyCL)
Poly(e-CL-co-acryloyloxyCL)
Poly(MMA-co-methacroyloxyEMA)
Poly(MMA-co-methacroyloxyEMA)
Poly(styrene-co- methacroyloxyEMA)
Poly(styrene-co- methacroyloxyEMA)
Poly(styrene-co- methacroyloxyEMA)
Poly(styrene-co- methacroyloxyEMA)

Facry [a] Mn [b]


SEC
0.15
15200
0.07
9900
0.08
22000
0.23
32000
0.08
29000
0.16
32000
0.05
88000
0.10
108000

Nanoparticles
Mw/Mn Tg [C]
1.35
1.08
1.32
1.35
1.34
1.29
1.45
1.60

68
62
77
60
98
90
105
95

Conv.
[%] [c]
72
88
90
88
92
85
85
85

Mn [d]
SEC
11600
9000
14500
15600
20000
15000
57000
47000

Mw/Mn
1.35
1.10
1.21
1.25
1.34
1.20
1.50
1.50

Tg [e]
[C]

45
120

124

125
130

Rh [f]
[nm]
4.3
3.8
5.5
6.5
6.2
5.4
13.1
9.2

[a] Molar fraction of (meth)acrylic units in the copolymer. [b] Number average molecular weight measured by SEC using poly(styrene)
calibration standards. [c] Conversion of (meth)acrylic units as measured by 1H NMR. [d] Number average molecular weight measured
by SEC using poly(styrene) calibration standards. [e] Glass transition temperature as measured by DSC. [f] Hydrodynamic radius as
measured by DLS in THF.

Adv. Mater. 2001, 13, No. 3, February 5

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2001

0935-9648/01/0302-0205 $ 17.50+.50/0

205

COMMUNICATIONS

by transesterification using methanol in the presence of


Ti(OEt)4 leading to the linear PMMA analog 6 (Scheme 3).
The transformation was confirmed by 1H NMR spectroscopy.

Scheme 3.

The size exclusion chromatograms of the original precursor


4, the unimolecular particle 5 and the PMMA analog 6 are all
shown in Figure 2. As expected the hydrodynamic volume of
the crosslinked nanoparticle 5 is smaller than the original precursor 4. Furthermore, the transesterified PMMA analog 6
returns to approximately the size of the original macromolecule prior to self-crosslinking confirming the relationship
between the formation of internal covalent bonds and the
decrease in molecular volume.

reaction (between acrylic units of different chains). Since the


intra- and inter-reactions are chemically the same, the fraction
of intra- versus inter-reaction will only depend on the relative
probability of encountering another internal acrylate relative
to that associated with intermolecular reaction. This will
exclusively depend on the concentration of the polymer solution and the polymersolvent relationship.
As demonstrated in the synthesis of cyclic polymers in
ultradilute solutions, the concentration at which intra- and
intermolecular reactions are equally probable is given by the
equation: Ceq = (3/2pr2)3/2M/NA where r2 is the mean
square end-to-end distance of the chain, M is the molecular
weight of the chain and NA is Avogadro's number.[7] This indicates that unimolecular particles should be formed from the
functionalized precursors at concentrations below Ceq
whereas three dimensional networks will be produced at concentrations above Ceq. At intermediate concentrations both
inter- and intra-reactions should occur. In order to confirm
this hypothesis, copolymer 1b (Table 1) was reacted at various
concentrations by heating at 65 C in the presence of AIBN.
At low concentrations (<105 M), unimolecular nanoparticles
are obtained and the SEC trace (Fig. 3) shows a shift to smaller hydrodynamic volumes. Increasing the polymer concentration results in the appearance of a higher molecular weight
fraction at shorter elution times due to intermolecular reactions. Finally, as expected, an insoluble gel is produced from
reaction at functional groups concentrations above 5 102 M.

Fig. 2. SEC of 4 poly(methyl methacrylate-co-methacryloyloxyethyl methacrylate) (entry 1d) precursor Mn = 32 000 g/mol, Mw/Mn = 1.35; 5 poly(methyl
methacrylate) crosslinked nanoparticle Mn = 15 600 g/mol, Mw/Mn = 1.20; and 6
poly(methyl methacrylate) linear analog Mn = 31 000 g/mol, Mw/Mn = 1.35.

A further indication of the presence of the internal covalent


bonds in the nanoparticles was provided from the glass transition temperature (Tg) measured by differential scanning calorimetry (DSC). The particles with low crosslinking densities
show an increased Tg relative to the precursor polymers
reflecting decreased segmental chain mobility. Those particles
with a higher degree of cross-linking (entries 1a, 1d, 1f,
Table 1) show no measurable glass transition temperature by
DSC. For example, the Tg value of a polystyrene-based nanoparticle derived from 1e shifts from 75 C for the linear precursor to 130 C for the nanoparticle with 8 % crosslinking
while no Tg is observed for the nanoparticle derived from 1f
with 16 % crosslinking functionality. This progressive increase
and eventual disappearance of the Tg is consistent with general observations for crosslinked materials suggesting that crosslinked nanoparticles behave more like a compact particle than
a free linear macromolecule.
The formation of self-crosslinked nanoparticles is based on
a favorable competition between an intramolecular reaction
(involving acrylic units of a single chain) and intermolecular

206

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2001

Fig. 3. SEC of poly(e-caprolactone-co-4-acryloyloxycaprolactone) (entry 1b,


Table 1) after crosslinking reaction in toluene at different initial concentrations
of polymer precursor 1b. The concentrations are given as mol/L of acrylic functional groups as calculated from 1H NMR. The SEC traces correspond to
a) [Acryl] = 4 103 M, b) [Acryl] = 2 103 M, c) [Acryl] = 2 104 M, and
d) is the initial polymer 1b before reaction.

Characterization of the nanoparticles by dynamic light scattering (DLS) in tetrahydrofuran (THF) reveals that the sizes
are very small (Rh = 3.813.1 nm, Table 1). To our knowledge,
this range of sizes, which is characteristic of dendrimers and
some micelles, has never been reached before for a nanoparticle. Furthermore, the results indicate that the size of the particle increases with the molecular weight of the functionalized
precursor. This provides some external control over the nano-

0935-9648/01/0302-0206 $ 17.50+.50/0

Adv. Mater. 2001, 13, No. 3, February 5

COMMUNICATIONS

particle size, since both the molecular weight of the precursors


and the amount of pendant reactive functionality can be regulated by the use of controlled polymerization techniques.[8]
Furthermore, the physicalchemical properties of the nanoparticles themselves such as solubility, biodegradability, and
surface functionality, is defined by the choice of the original
comonomers, stoichiometries, and polymerization techniques.
While it is still early to define a broad range of applications
for nanoparticles prepared in this fashion, one of our interests
is their potential use as sacrificial pore generators for templating porosity in thin films thereby lowering the film dielectric
constant.[9] Our approach depends on the tendency of the
nanoparticles to uniformly disperse with minimal aggregation
in a low dielectric constant thermosetting matrix. After vitrification of the matrix, the nanoparticle can, in principle, be
thermally degraded leaving behind voids which ideally will
reflect the size and distribution of the original dispersed nanoparticles (Scheme 4).

Fig. 4. FESEM picture of a nanoporous poly(methylsilsesquioxane) generated


by curing a hybrid (80 %-MSSQ/20 %-PMMA-based nanoparticle entry 1d).

preparation of nanoparticles with very small sizes. Added


advantages of this method are that the size and chemical
nature of the nanoparticles are easily controlled. The utility of
such nanoparticles in the generation of nanoporous thin films
using the particles as sacrificial porogens has been demonstrated. Other technological applications for such particles
can be envisioned and are under investigation.

Experimental
Scheme 4.

Preliminary results were obtained using thin films hybrids


of methyl silsesquioxane (MSSQ) and the PMMA-based
nanoparticles prepared by spin coating techniques. During a
thermal ramp (5 C/min) to 450 C, the matrix first vitrifies
and then the sacrificial polymer is subsequently removed.
Porous MSSQ thin films are obtained in this process based on
the observed substantial decrease in refractive index relative
to the densified material. The refractive index for pure MSSQ
is 1.38 while hybrids containing either 20 wt.-% or 30 wt.-%
of nanoparticles produce films with refractive indexes of 1.31
and 1.28, respectively, after nanoparticle burn-out.
Figure 4 shows a field emission scanning microscopy (FESEM) cross section of the nanoporous MSSQ with 20 % porosity. The pores appear smaller than 10 nm and are uniformly
dispersed through the film. The average pore size calculated
from small angle X-ray scattering measurements was 7.2 nm,
which compares favorably with the size of the crosslinked
nanoparticles as measured by DLS (Rh = 6.5 nm). The measured dielectric constant (100 kHz) of a MSSQ sample with
20 % porosity is 2.1 showing a considerable reduction from
that of the dense, fully cured MSSQ film (2.8).
In conclusion, we have demonstrated that the intramolecular self-crosslinking of functionalized macromolecules under
ultradilute conditions provides a versatile approach for the

Adv. Mater. 2001, 13, No. 3, February 5

Materials: Toluene and THF were dried by refluxing over sodium and distilled under nitrogen prior to use. The e-caprolactone (eCL) was dried over
CaH2 and distilled under reduced pressure. The samples were stored under a
dry nitrogen atmosphere until use. All other chemicals were purchased from
Aldrich Chemical Co. and used without further purification.
Synthesis of Crosslinkable Macromolecules: ROP Using Al(OiPr)3: Poly(ecaprolactone-co-4-acryloyloxycaprolactone) random copolymers were prepared
as reported elsewhere [6].
Synthesis of Poly(methyl methacrylate-co-hydroxyethyl methacrylate) Copolymers by Atom Transfer Radical Polymerization (ATRP): In a typical experiment, methyl methacrylate (9 eq.), hydroxyethyl methacrylate (1 eq.), ethyl
2-bromoisobutyrate (0.03 eq.), and NiBr2(PPh)3 (0.03 eq.) were added to a
glass reactor. The flask was then closed with a three-way stopcock and a cycle
of vacuum-nitrogen purges was repeated three times to remove the oxygen.
After reaction at 85 C for 14 h, the polymer was recovered in 85 % yield by
precipitation in hexanes after dissolution in THF; SEC (Mn = 22 000 g/mol, Mw/
Mn = 1.32); 1H NMR (CDCl3) d 4.06, 3.79, 3.56, 2.260.8(m).
Synthesis of Poly(styrene-co-hydroxyethyl methacrylate): A mixture of the
hydrido alkoxyamine (2,2,5-trimethyl-3-(1-phenylethyoxy)-4-phenyl-3-azahexane) (1 eq.), styrene (90 eq.), and 2-hydroxyethyl methacrylate (10 eq.) was
degassed in three freeze/thaw cycles, sealed under argon and heated at 125 C
for 10 h. The viscous mixture was then dissolved in THF and precipitated in
cold hexanes as a white powder in 7090 % yield; SEC (Mn = 9000 g/mol, Mw/
Mn = 1.15); 1H NMR (CDCl3) d 7.25.6, 33.5, 11.8.
General Procedure for Reaction of Methacrylic Acid with Hydroxyethyl
Methacrylic Pendant Groups: A poly(methyl methacrylate0.9-co-hydroxyethyl
methacrylate0.1) (Mn = 22 000 g/mol, Mw/Mn = 1.30) (3 g, 3 mmol of hydroxyethyl units), methacrylic acid (0.534 g, 6 mmol), and triphenylphosphine (1.5 g,
6 mmol) were dissolved in 30 mL of dry THF. To this mixture (1.21 g, 6 mmol)
of diisopropyl azodicarboxylate (DIAD) were added drop-wise. After 14 h
reaction at room temperature the polymer was quantitatively recovered by precipitation in hexanes; SEC (Mn = 22 000 g/mol, Mw/Mn = 1.30); 1H NMR
(CDCl3) d 6.13, 5.61, 4.30, 4.66, 3.56, 2.260.8(m).
General Procedure for the Preparation of Self-Crosslinked Nanoparticles: In
a typical experiment, a mixture of 0.2 g of poly(methyl methacrylate0.92-comethacryoyloxyethyl methacrylate0.08) (Mn = 22 000 g/mol, Mw/Mn = 1.32)
(0.2 mmol of methacrylic groups) and AIBN (0.02 mmol) were dissolved in

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2001

0935-9648/01/0302-0207 $ 17.50+.50/0

207

COMMUNICATIONS

500 mL of toluene. The mixture was degassed by bubbling N2 for 15 min and
heated for 10 h at 65 C. After reaction, the solvent was evaporated and the
polymer was dissolved in 2 mL of THF and precipitated in an excess of hexanes
(0.18 g, 90 %); SEC (Mn = 14 500 g/mol, Mw/Mn = 1.21); 1H NMR (CDCl3)
d 4.30, 4.66, 3.56, 2.260.8(m).
Characterization: 1H NMR and 13C NMR spectra were recorded with a Bruker AM 250 (250 MHz) spectrometer. SEC was carried out on a Waters chromatograph connected to a Waters 410 differential refractometer, using polystyrene of known molecular weight as the calibration standards. Four 5 lm Waters
columns (300 7.7 mm) connected in series in order of increasing pore size
(100, 1000, 105, and 106 ) were used with THF as elution solvent. Glass transition temperatures were measured in a modulated DSC 2920 TA Instrument
using a 4 C/min heating rate. DLS measurements were performed using a 2 W
Lexel model 95 argon ion laser (k = 514.5 nm) at a power of 500 mW. The scattered light intensity was measured at an angle of 90.0. Samples of 10 mg/mL in
HPLC grade THF were prepared in test tubes which were previously cleaned in
acetone in an ultrasonic bath. The samples were transferred via a syringe filter
(Whatman Anatop 10 0.2 lm) into a second test tube. Data acquisition lasted
3 min and was stored on a Brookhaven Instruments BI-9000 correlator. The
temperature was held constant at 25 C by a Neslab circulating bath.
Received: June 19, 2000
Final version: September 25

[1] Y. Xia, B. Gates, Y. Yin, Y. Lu, Adv. Mater. 2000, 12, 693.
[2] K. Landferter, W. Bechthold, S. Forster, M. Antonietti, J. Macromol. Rapid
Commun. 1999, 20, 81.
[3] Q. Zhang, E. E. Remsem, K. L. Wooley, J. Am. Chem. Soc. 2000, published on the web. K. B. Thurmond, T. Kowalewski, K. L. Wooley, J. Am.
Chem. Soc. 1997, 119, 6656. A. Guo, G. Liu, J. Tao, Macromolecules 1996,
29, 2487.
[4] J. M. Frechet, Science 1994, 263, 1710. G. R. Newkome, Advances in Dendritic Macromolecules, JAI Press, Greenwich, CT 19941995, Vols. 12.
[5] a) M. Antonietti, Angew. Chem. Int. Ed. 1998, 27, 1743. b) V. A. Davankov,
M. M. Ilyin, M. P. Tsyurupa, G. I. Timofeeva, L. V. Dubrovina, Macromolecules 1996, 29, 8398.
[6] D. Mecerreyes, R. D. Miller, J. L. Hedrick, P. Lecomte, C. Detrembleur,
R. Jerome, J. Macromol. Rapid Commun., in press.
[7] D. Benoit, V. Chaplinski, R. Braslau, C. J. Hawker, J. Am. Chem. Soc.
1999, 121, 3904. K. Matyjaszewski, in Controlled Radical Polymerization,
ACS Symp. Ser. 1998, 685.
[8] K. Roovers, D. Toporowski, Macromolecules 1983, 16, 483.
[9] R. D. Miller, Science 1999, 286, 421. C. Nguyen, K. Carter, C. J. Hawker,
J. L. Hedrick, R. Jaffe, R. D. Miller, J. F. Remenar, H. Rhee, P. Rice,
M. Toney, M. Trollsas, D. Yoon, Chem. Mater. 1999, 11, 3080. J. L.
Hedrick, R. D. Miller, C. J. Hawker, K. Carter, W. Volksen, D. Yoon,
M. Trollsas, Adv. Mater. 1998, 10, 1049.

Solvatochromism of a Copper(II)
(Tetramethylethylenediamine)(acetylacetonate)+
Complex Encapsulated in EMT Zeolite Cages**
By Julia L. Meinershagen and Thomas Bein*
The inclusion of the copper chelate CuII(N,N,N,N-tetramethylethylenediamine)(acetylacetonate)+ in the nanopores
of hexagonal faujasite type zeolites has been studied for optical sensing applications. Optical vapor sensing of analytes has

[*] Prof. T. Bein, Dr. J. L. Meinershagen


Department of Chemistry
Ludwig Maximilians Universitt
Butenandtstr. 5-13, E, D-81377 Munich (Germany)
E-mail: tbein@cup.uni-muenchen.de

[**] The authors acknowledge Dr. D. E. De Vos for measuring the solution
phase EPR spectra and for many helpful discussions.

208

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2001

recently been the focus of several research groups;[14] while


many sensing methods are currently available, size selective
sensors are not as widespread. The aim of our work is to combine the shape selectivity of zeolites with the chemical sensitivity of solvatochromic dyes to produce highly selective optical sensors. Recent work in our lab has shown promising
results for these types of dye/zeolite ensembles.[5] We now
extend our work to include a new type of dye.
Copper chelates with mixed diamine and b-diketonate ligands have been studied extensively in the literature as solvent polarity indicators.[610] These compounds are strongly
solvatochromic due to the subtle structural changes about the
copper center. The energy of the visible dd transition band is
influenced by the structure of the coordination center, which
is extremely sensitive to solvent attack. Specifically, the complex with mixed N,N,N,N-tetramethylethylenediamine and
acetylacetonate ligands (Cu(tmen)(acac)+) has shown exceptional solvatochromic properties (Fig. 1). This phenomenon is
what has led to the compound being used as a basicity indicator with Gutmann's donor number (DN) scale, which is a sol-

Fig. 1. Structure of Cu(tmen)(acac)BPh4.

vent independent representation of the donor ability of a molecule.[7,11] The Cu(tmen)(acac)+ complex shows impressive
color changes in going from weakly donating to strongly
donating solvents. For example, a pink color is observed for a
solution of Cu(tmen)(acac)BPh4 in the non-donating solvent
dichloroethane (DN = 0), whereas in the strongly donating
solvent piperidine (DN = 52.3) the solution is green. Noting
the strong solvent effects in the solution phase, it was anticipated that by ion exchanging the chelate into zeolites, an optical sensor related to solvent basicity was feasible. The goals of
this work were to relate geometrical changes of the copper
complex to optical responses. Electron paramagnetic spectroscopy proved to be an invaluable tool for better understanding
of the axial coordination and resulting changes in the d orbital
energy levels. Finally, demonstration of optical responses and
sensing properties were shown.
The zeolite used was the large pore hexagonal faujasite type
zeolite EMT. EMT is comprised of hexagonally packed sodalite cages connected by double six rings giving rise to 6.9
7.2 windows leading into 13 14 hyper-cages,[12] with an
average crystal size of 10 lm. These small windows leading to
large cages are ideal for encapsulating a wide range of functional molecules.
Cu(tmen)(acac)ClO4 was prepared according to a published
procedure.[6] The compound was recrystallized and the perchlorate anion was exchanged for tetraphenylborate (BPh4)
in order to assure ion exchange, excluding the large BPh4
anion which can not penetrate into the zeolite pores.[8,13] The
purity was determined by infrared (IR) and chemical ioniza-

0935-9648/01/0302-0208 $ 17.50+.50/0

Adv. Mater. 2001, 13, No. 3, February 5

Potrebbero piacerti anche