You are on page 1of 7

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER www.rsc.org/materials | Journal of Materials Chemistry

Synthesis, characterization and optical limiting properties of a gallium


phthalocyanine dimer
Helmut Bertagnolli,*a Werner J. Blau,b Yu Chen,c Danilo Dini,d Martin P. Feth,a Sean M. O’Flaherty,b
Michael Hanack*d and Venkata Krishnana
Received 13th August 2004, Accepted 20th October 2004
First published as an Advance Article on the web 24th November 2004
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

DOI: 10.1039/b412546k

A gallium phthalocyanine dimer with a direct gallium–gallium bond, [tBu4PcGa]2?2dioxane (2),


has been synthesized and characterized by Extended X-ray Absorption Fine Structure (EXAFS)
spectroscopy, UV/Vis, IR, NMR and MALDI-MS. By EXAFS the gallium–gallium distance
in 2 was found to be 2.46 Å. Dimer 2 exhibits reverse saturable absorption with significantly
reduced saturation intensity with respect to the starting material tBu4PcGaCl (1) as determined by
means of optical limiting experiments.

Introduction [tBu4PcIn]2?2tmed by a Wurtz coupling reaction of tBu4PcInCl


in the presence of N,N,N9,N9-tetramethylethylenediamine
Optical limiting (OL) is an important application of nonlinear (tmed). The role of the neutral tmed ligand is mainly to
optics, useful for the protection of human eyes, optical promote maximum coordinative saturation at the indium
elements and optical sensors from intense laser pulses. center. By EXAFS measurements the In–In distance was
Phthalocyanines (Pcs) are materials that optically limit found to be 3.24 Å.4,5 The OL response is generally enhanced
nanosecond light pulses in a fairly wide range of the UV/Vis by dimerization.4j,k
spectrum via excited state absorption processes. The nonlinear Similarly, the synthesis and molecular structure of organo-
optical absorption mechanism of phthalocyanines in the gallium compounds with a Ga–Ga bond are also an interesting
optical region comprised between Q- and B-bands involves area of study.6 A number of novel organogallium compounds
the population of excited states which absorb more effectively with a gallium–gallium bond, for example, Ga2R2(m-O2CCH3–
than the ground state. This gives rise to the phenomenon of O,O9)2 [R 5 CH(SiMe3)2] 7, [Ga2Cl4(dioxane)2]x 8, [(Me3Si)2-
Reverse Saturable Absorption (RSA) as a consequence of C(Ph)C(Me3Si)NGaCl]2 9 and [(Me3Si)2HC]4Ga2,10 have been
multiphoton absorption.1 It has been shown that phthalocy- reported so far. Herein we report on the synthesis, structural
anine compounds exhibit RSA because of the occurrence of characterization and nonlinear optical properties (OL pro-
intersystem crossing from the lowest excited singlet state (S1) perties), of a gallium phthalocyanine dimer with a direct
to the lowest triplet state (T1) and the subsequent increase bond between central atoms,11 i.e. [tBu4PcGa]2?2dioxane (2)
in the population of the strongly absorbing T1 state with (Scheme 1). Dimer 2 is prepared in 83% yield by the reaction of
nanosecond dynamics.2–4 soluble tBu4PcGaCl (1)4a with activated magnesium in freshly
Axial substitution on phthalocyanines generally increases dried THF and in the presence of 1,4-dioxane. The resulting
their solubility and, therefore, reduces molecular aggrega- dimer exhibits an enhancement of OL response at the
tion.4c,4d In terms of OL effectiveness in the visible range, same wavelengths when compared to the parent monomer
axial substitution can improve the efficiency of excited state tBu4PcGaCl (1).
absorption through the inhibition of the decay of the excited
state formed in a nonlinear optical regime.1a,4a,c–e,k Moreover,
Results and discussion
axial substituents in Pcs influence favourably nonlinear
optical absorption for the presence of a dipole moment For most of the phthalocyanines and naphthalocyanines, it is
perpendicular to the macrocycle in the axially substituted very difficult to obtain suitable single crystals required for
phthalocyanines.4c,d,o The exploitation of the chemical reac- X-ray structural determination. In such cases, EXAFS
tivity of the bond M–Cl [M 5 Ga3+, In3+] and MLO [M 5 Ti4+, (extended X-ray absorption fine structure) spectroscopy is
etc.] bonds can allow the preparation of a series of highly considered to be a powerful technique for the determination of
soluble axially substituted and bridged Pc complexes.4f,k the local structure of a specific atom, regardless of the state of
Organometallic compounds containing metal–metal single the sample. EXAFS provides information on the coordination
bonds between the metallic elements of Group 3A (Al, Ga, In, number, the nature of the scattering atoms surrounding the
Tl) are unusual but can be of interest in the field of material absorbing (excited) atom, the interatomic distance between the
sciences.5 Recently4f,j we have prepared the first dimeric absorbing atom and the back-scattering atoms and the Debye–
indium phthalocyanine complex with an indium–indium bond Waller factor, which accounts for the disorders due to static
displacements and thermal vibrations.12 In this study, the
*h.bertagnolli@ipc.uni-stuttgart.de (Helmut Bertagnolli) formation of the gallium–gallium bond in 2 was initially
hanack@uni-tuebingen.de (Michael Hanack) confirmed by EXAFS spectroscopy. The experimentally

This journal is ß The Royal Society of Chemistry 2005 J. Mater. Chem., 2005, 15, 683–689 | 683
View Article Online

Table 1 EXAFS determined structural data of [tBu4PcGa]2?2dioxane


(2)
Na r/Åb s/Åc

Ga–NPyrrole/ODioxane 5 2.00 ¡ 0.02 0.067 ¡ 0.007


Ga–C 8 3.00 ¡ 0.03 0.071 ¡ 0.016
Ga–NAza 4 3.33 ¡ 0.04 0.081 ¡ 0.026
Ga–Ga 1 2.46 ¡ 0.03 0.095 ¡ 0.016
a
N 5 coordination number. b r 5 interatomic distance between the
absorber and the back-scatterers. c s 5 Debye–Waller factor with
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

calculated standard deviations.

Fig. 2 EXAFS structure of [tBu4PcGa]2?2dioxane (2).


Scheme 1
combined shell with a coordination number of five was
determined and fitted EXAFS functions of 2 are shown in k determined at a distance of about 2.00 Å. This includes the
space as well as by Fourier transforms in real space in Fig. 1. four back-scattering nitrogen atoms from the pyrrole and the
The structural parameters along with the estimated errors one back-scattering oxygen atom from the dioxane ligand.
arising from the refinement analysis are summarized in Table 1. Another four back-scattering nitrogen atoms from the aza
The molecular structure of 2 is given in Fig. 2. In the fitting nitrogen atoms were found at 3.33 Å distance. Further, an
procedure, the intramolecular coordination numbers were additional shell consisting of eight carbon atoms, having
fixed to the known values for the ligands around the gallium gallium–carbon distances at 3.00 Å was observed. The fit was
atom in the complex. The other parameters including significantly improved by having a gallium–gallium distance
intermolecular coordination number, interatomic distances, with a coordination number of one at a distance of 2.46 Å.
Debye–Waller factor and energy zero value were determined For comparison, the reported gallium–gallium bond
by iterations. The analysis of the data shows the contribution lengths in some organogallium compounds are 2.38 Å for
of the phthalocyanine macrocycle in the spectra. Owing to the Ga2R2(m-O2CCH3–O,O9)2,7 2.38 Å for [Ga2Cl4(dioxane)2]x,8
similar back-scattering behavior of nitrogen and oxygen, 2.45 Å for [(Me3Si)2C(Ph)C(Me3Si)NGaCl]2,9 and 2.56 Å for
occurring at nearly the same distance from the metal atom, a [(Me3Si)2HC]4Ga2.10 When the coordination number for

Fig. 1 Experimental (solid line) and calculated (dotted line) EXAFS functions (a) (k range: 2.6–15.0 Å21, fit index: 32.66, DE0: 13.83 eV) and their
Fourier transforms (b) for [tBu4PcGa]2?2dioxane measured at the Ga K-edge at 10 367 eV.

684 | J. Mater. Chem., 2005, 15, 683–689 This journal is ß The Royal Society of Chemistry 2005
View Article Online

gallium was iterated, it resulted in a value of 0.7, which was


close to unity, and therefore the coordination number was
fixed to one. This result is consistent with an arrangement of
the complex as a dimer. In addition, it should be noted that
the back-scattering behavior of Ga (Z 5 32), Cl (Z 5 17) and
I (Z 5 53) are different, thus these elements can be well
distinguished by EXAFS spectroscopy, especially when these
back-scatterers occur in a distance-range of 2–3 Å. We did not
find any evidence that this compound is tBu4PcGaI, a possible
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

by-product.
The effect of the metal ions on the phthalocyanine ring in
the metallophthalocyanines has been widely studied.13 To the
best of our knowledge, the shortest interplanar ring spacing in
gallium phthalocyanines was found to be 3.34 Å in the case of
Ga(Pc)Cl for a parallel stacking arrangement.14a When such a
short interplanar ring spacing is assumed in the present case,
the intermolecular metal–metal distance should be about
3.35 Å, but the EXAFS results indicate the gallium–gallium
distance to be 2.46 Å, which supposes that the two gallium
atoms in the molecule are projected inwards of the phthalo-
cyanine ring and that the gallium center is located 0.45 Å out
of the plane formed by the four coordinated nitrogen atoms. In
addition, when a rigid phthalocyanine molecule with an
average cavity diameter of about 3.85 Å, which is not
deformed by the metal atom is assumed, the shortest metal–
nitrogen distance occurs at 1.97 Å.15,16 But the present results Fig. 3 UV/Vis spectra (a) and photoluminescence spectra (b,
indicate a gallium–nitrogen and a gallium–oxygen distance of lex 5 355 nm) of tBu4PcGaCl (1) and [tBu4PcGa]2?2dioxane (2) in
2.00 Å, which also supposes that the gallium center is located chloroform.
0.45 Å out of the plane. This value is in good agreement with
the reported out-of-plane value of 0.439 Å in crystalline to the change of the molar extinction coefficient due to the
PcGaCl.14a Further, in a rigid phthalocyanine molecule, the different number of absorbing Pc rings per molecular unit4o in
shortest distance between the metal center and the nearest passing from the monomer to the Ga–Ga bridged dimer.
carbon occurs at 2.96 Å, while in the present case it results a Photoluminescence spectra in dilute chloroform solutions
gallium–carbon distance of 3.00 Å. From these values an out- have also been obtained under the same conditions of the
of-plane location of gallium equal to 0.54 Å was found, which spectra of Fig. 3(a) when the wavelength of excitation was
is in agreement with the value determined from the shortest 355 nm [Fig. 3(b)]. In Pcs, such a wavelength of excitation
gallium–nitrogen separation values. Since the complex was corresponds to the energy necessary for the occurrence of
found to be a dimer, it is believed that the t-butyl groups in the the transition HOMO (21)A LUMO, which originates the
periphery of the macrocycles have no steric hindrance with B-band absorption [Fig. 3(a)].17 Upon emission, both com-
each other and the two phthalocyanine macrocycles are pounds 1 and 2 display a Stokes shift of the emission peak with
arranged in a staggered conformation. In any case, the exact respect to the location of the Q-band absorption [Fig. 3(b)].
geometry of these groups cannot be determined by EXAFS Different to the absorption spectra, dimerization influences
spectroscopy since a mixture of regioisomers is present. the energy of emission more strongly, since the extent of the
Further, the sterically less crowded staggered conformation Stokes shift is larger for the monomer with respect to the
is preferred when interplanar spacing is less than 3.5 Å, dimer. Maximum emission peaks for 1 and 2 are found at 713
whereas the eclipsed conformation occurs at larger spacings.14b and 700 nm, respectively. Such a difference could be ascribed
The linear absorption spectra of tBu4PcGaCl (1) and dimer to the more rigid structure of the dimer with respect to the
2 are presented in Fig. 3(a). The similarity of the UV/vis monomer. This means that the dimer relaxes radiatively from
spectra of 1 and 2 shows that dimerization does not affect the electronic excited state p* in a higher vibrational level.18
considerably the spectral profile. In fact, the positions of As shown in Fig. 4, the main differences in the IR spectra
the maxima of optical absorption are practically unaltered between 1 and 2 are: (a) the absence of the absorption band of
(693 and 695 nm for the Q-band absorption of 1 and 2, the Ga–Cl stretching mode at 351 cm21 in the IR spectrum of
respectively). This implies that the two Pc rings in the Ga–Ga the dimer; (b) a striking increase in the relative intensity, and
bridged dimer do not modify considerably the electronic broadening, of some characteristic absorption peaks in 2 with
distribution of each other. The extent of electronic conjugation respect to tBu4PcGaCl (1); and (c) some new absorption peaks,
in both compounds 1 and 2 is practically the same since e.g. at 802 and 393 cm21 and others, are observed.
compound 2 has a gallium–gallium single bond of s-type with The molecular peak of [tBu4PcGa]2?2(1,4-dioxane) (2) was
no presence of delocalized electrons. As expected, the main found in the Matrix Assisted Laser Desorption (MALDI) mass
differences between the spectral features of 1 and 2 are related spectrum. This appeared at m/z 5 1789 (calc. m/z 5 1788.8).

This journal is ß The Royal Society of Chemistry 2005 J. Mater. Chem., 2005, 15, 683–689 | 685
View Article Online

times. For the NLO characterization of Pcs the Z-scan


technique has been used.21 Herein, the Z-scan in the open-
aperture mode21 was employed to investigate the nonlinear
optical absorptive properties of the phthalocyanine com-
pounds tBu4PcGaCl (1) and [tBu4PcGa]2?2dioxane (2). In this
technique the transmittance of a focused Gaussian laser beam
is monitored as a function of the position of the sample passing
through the focus of the collimated beam. The nonlinear
absorption coefficient (bI) was calculated as a function of
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

incident focal intensity (I0).


All open-aperture Z-scans performed in this study, from
both compounds, exhibited a reduction in the transmission
about the focus of the lens. This was typical of an induced
positive nonlinear absorption of the incident light, in this case
attributed to excited state absorption. It was noticed that the
effective nonlinear absorption coefficient bI, in all cases, was
not stationary with respect to the intensity at the focus. The
nonlinear absorption coefficient bI plotted as a function of the
Fig. 4 IR spectra of (top) tBu4PcGaCl (1) and (bottom) incident focal intensity I0 is depicted in Fig. 5 for 1 and 2,
[tBu4PcGa]2?2dioxane (2).
where each data point on the plot represents an independent
open-aperture Z-scan of the compound in question and the
There was no evidence that this peak resulted from an in-situ solid lines are sketched merely as guides to the eye. It can
by-product formed under the MALDI-MS experimental clearly be seen that bI reduces in magnitude with increasing
conditions. For the starting material tBu4PcGaCl (1), we did focal intensity I0 in the figure. Despite this, effective third
not observe the formation of tBu4PcGa–GaPctBu4 under the order nonlinear absorption coefficients bI can be estimated by
same experimental conditions. suitably interpolating the data in Fig. 5 at certain on-focus
tBu4PcGaCl (1) is a mixture of four structural isomers.4a,19 intensities. For example, bI values interpolated for an irradia-
We have been the first to separate the isomers of tetrasub- tion value of 0.5 GW cm22 (arbitrarily chosen) are quoted as a
stituted Pcs and undertake detailed and extensive NMR sample case study in Table 2. It can be seen that both
studies probing the influence of isomers on the NMR spectra compounds exhibit similar nonlinear absorption coefficients
of the Pcs.20 The NMR spectrum of an isomeric mixture is with 2 tending to exhibit a larger bI than 1 at lower I0 values.
somewhat different from the pure isomers, but it is still The normalized transmission (TNorm) against the incident
possible to draw definite conclusions about the structure. The fluence F per pulse [F 5 E/pw(z)2, where E is the incident
main reason that we took a tetrasubstituted and not an energy per pulse and w(z) is the beam waist as a function of z in
octasubstituted Pc as a mode compound is the higher solubility the experiment] was analyzed to further investigate the non-
of the tetrasubstituted over the octasubstituted Pc.19 As linear absorption. The nonlinear absorption coefficient a(F,
expected, the 1H NMR spectrum of 2 measured in CDCl3 is FSat, k),4j where a(F, FSat, k) # a0(1 + F/FSat)21(1 + kF/FSat)
somewhat more complicated than that of 1 because of the ring- was used to fit the normalized transmission as a function of
current effect caused by the two macrocycles and the possible this energy density to a superposition of all open-aperture
existence of rotational isomers. In 2, the aromatic proton
signals are broadened, and appear at d 8.8–9.3 ppm (multi-
plet), and 8.2–8.3 ppm (multiplet), respectively, and the tBu
protons at 1.9 ppm. The proton signals for free 1,4-dioxane
appear at d 3.3 ppm (s, 8H), whereas the corresponding proton
signals for the coordinated 1,4-dioxane ligands in 2 are found
at d 3.7 and 4.2 ppm, respectively, due to the asymmetry of the
1,4-dioxane coordinated with the gallium atoms. The observed
downfield shift would normally not be expected; here it may be
explained, that the coordinating molecules do not fit into the
cone of the ring-current.4f In the 13C NMR spectrum
the aromatic carbon signals of 2 are very similar to those of
the starting compound 1, and can easily be assigned to the
appropriate carbon atoms in the macrocycle. In comparison
with free 1,4-dioxane, whose 13C NMR signals appear at
66.6 ppm (for four equivalent carbon atoms), the carbon
signals of the coordinated 1,4-dioxane moved downfield Fig. 5 Plot of effective nonlinear absorption coefficient bI against the
slightly and are located at d 67.9 and 68.2 ppm, respectively. on-focus beam intensity I0 for both compounds. Each data point
In phthalocyanine materials, optical nonlinearities are of represents an independent open-aperture Z-scan and the solid lines are
interest as they can be large with short (less then 1 ps) response intended as a guide to the eye.

686 | J. Mater. Chem., 2005, 15, 683–689 This journal is ß The Royal Society of Chemistry 2005
View Article Online

Table 2 Summary of the nonlinear optical properties for tBu4PcGaCl (1) and [tBu4PcGa]2?2dioxane (2) in toluene
Compound c/g L21 a0/cm21 I0/GW cm22 bI/cm W21 FSat/J cm22 k (sex/s0)

1 0.5 1.1 0.5 (3.2 ¡ 0.6) 6 1028 27.0 ¡ 1.0 13.5 ¡ 0.4
2 0.5 1.3 0.5 (3.5 ¡ 0.7) 6 1028 8.9 ¡ 0.4 10.4 ¡ 0.3

datasets for each compound. This expression assumes an ideal significant that the reduction by a factor of 1.3 for the Ga
three level absorption system where F represents the fluence, compounds presented here. This difference, however, can be
FSat the saturation fluence and k the ratio of the excited to partially attributed to the particularly low linear absorption
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

ground state absorption cross sections sex/s0. The parameters coefficient of the tBu4PcInCl molecule, which allows it to
k (realistically sex and a0 were measured) and FSat were treated exhibit a large k coefficient compared to other Pcs.4l
as free constants in the fitting. The plots of TNorm against
energy density are depicted in Fig. 6 and the a0, k and FSat Conclusions
values for each compound are also presented in Table 2. It was
found that compound 2 exhibits a significantly lower FSat than A gallium phthalocyanine dimer 2 with a direct gallium–
that of 1, by approximately a factor of 3. The ratio of their gallium bond was prepared by the reaction of soluble
excited and ground state absorption cross sections is similar tBu4PcGaCl with activated magnesium turnings in freshly
with 1 exhibiting a k coefficient approximately 1.35 times that dried THF in the presence of 1,4-dioxane in 83% yield. From
of 2. Thus, dimerization of 1 in the manner described here to the EXAFS measurements the intermolecular gallium–gallium
yield in 2 is clearly a viable method of tuning the saturation distance was determined at 2.46 Å and the gallium metal center
energy density (FSat) of the material, reducing it by a factor of is supposed to be located 0.45 Å out-of-plane from the
3. This reduction in FSat is coupled with a slight increase of bI phthalocyanine ring. Dimer 2 exhibits a significantly reduced
at low incident intensity and with approximately equivalent saturation intensity, determined via optical limiting experi-
bIs at higher intensities. The k coefficient is reduced by ments, relative to the starting material tBu4PcGaCl (1). This
approximately 22% after dimerization of 1 to 2 but this can reduction in the saturation energy density is coupled with a
probably be mostly attributed to the increase in linear slight increase of nonlinear absorption coefficient (bI) at low
absorption coefficient by approximately 18% over the same incident intensity and with approximately equivalent bIs at
molecular modification. higher intensities. Further investigations of the nonlinear
Recently we have reported on the optical limiting of a Pc optical properties and other interesting photophysical proper-
dimer with a direct In–In bond.4j In this study we compared ties of thin films of compounds 1 and 2 are currently in
the starting tBu4PcInCl with the new [tBu4PcIn]2?2tmed progress.
compound. Similar to the results presented in this study for
the Ga–Ga Pc dimer 2 we found that dimerization of the Experimental
starting single In unit reduced FSat by a factor of approxi-
mately 2.5 (comparable to a factor of approximately 3 for the General
Ga–Ga dimer presented here). The In–In Pc dimer also The operations for synthesis prior to the termination reaction
exhibited a slightly lower FSat of (24.2 ¡ 0.8) J cm22 than that were carried out under purified dry nitrogen. Solvents were
of compound 2 presented here. The reduction of k for the In purified, dried and distilled under dry nitrogen. FT-IR: Perkin-
Pcs by a factor of about 2.2 after dimerization is far more Elmer Spectrum 1000; UV/Vis: Shimadzu UV-365; MALDI-
MS: Bruker Biflex (preparation method of sample: the samples
were dissolved in CH3CN, then mixed 1 : 1 (0.8 ml each) with a
methanol solution of a-cyano-4-hydroxycinnamic acid as
matrix); 1H-, 13C-NMR: Bruker AC 250 (1H: 250.131 MHz,
13
C: 62.902 MHz).
The EXAFS measurements of the complex were performed
at the Ga K-edge at 10 367 eV at the beamline X1.1 of the
Hamburger Synchrotron Radiation Laboratory (HASYLAB)
at DESY, Hamburg. The complex was measured with a
Si(111) double crystal monochromator at ambient conditions.
The positron energy was 4.45 GeV and the beam current was
about 120 mA. Data were collected in transmission mode with
ion chambers filled with nitrogen gas. Energy calibration was
monitored using 20 mm thick iridium metal foil having the L3-
edge at 11 215 eV. The sample was prepared as a pellet of a
mixture of the complex with polyethylene. The data were
Fig. 6 Plots of normalized transmission against incident pulse energy analyzed with a program package especially developed for
density for (a) tBu4PcGaCl (1), and (b) [tBu4PcGa]2?2dioxane (2). The the requirements of amorphous samples.22 The program
solid lines are theoretical curve fits. AUTOBK23 was used for the removal of background and

This journal is ß The Royal Society of Chemistry 2005 J. Mater. Chem., 2005, 15, 683–689 | 687
View Article Online

the program EXCURV9224 was used for the evaluation of the HASYLAB, Hamburg for the support for synchrotron
EXAFS function. Curved wave theory was used for the data radiation experiments.
analysis in k space and the resulting EXAFS function was
weighted with k3. The mean free path of the scattered electrons Helmut Bertagnolli,*a Werner J. Blau,b Yu Chen,c Danilo Dini,d
Martin P. Feth,a Sean M. O’Flaherty,b Michael Hanack*d and
was calculated from the imaginary part of the potential (VPI
Venkata Krishnana
set to 24.00), the amplitude reduction factor AFAC was fixed a
Institut für Physicalische Chemie, Universität Stuttgart,
at 0.8 and an overall energy shift DE0 was introduced to give a Pfaffenwaldring 55, 70569 Stuttgart, Germany.
best fit to the data. E-mail: h.bertagnolli@ipc.uni-stuttgart.de; Fax: +49-(0)711-6854443;
Tel: +49-(0)711-6854450
All Z-scan experiments described in this study were b
Materials Ireland Polymer Research Centre, Department of Physics,
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

performed using 6 ns pulses from a Q switched Nd:YAG Trinity College Dublin, Dublin 2, Republic of Ireland
c
laser. The beam was spatially filtered to remove the higher Department of Chemistry, East China University of Science and
order modes and tightly focused with a 9 cm focal length lens. Technology, 130 Meilong Road, Shanghai 200237, People’s Republic of
China
The laser was operated at its second harmonic, 532 nm, with a d
Institut für Organische Chemie, Universität Tuebingen, Auf der
pulse repetition rate of 10 Hz. The samples were prepared Morgenstelle 18, D-72076 Tuebingen, Germany.
at 0.5 g L21 in toluene followed by gentle agitation for ca. E-mail: hanack@uni-tuebingen.de; Fax: +49-(0)7071-295268;
Tel: +49-(0)7071-2972432
20 minutes in a low-power (60 W) sonic bath to ensure
complete and uniform dispersal. All subsequent measurements
were performed in quartz cells with a 1 mm path length. References
1 (a) J. W. Perry, K. Mansour, I. Y. S. Lee, X. L. Wu, P. V. Bedworth,
Synthesis of gallium phthalocyanine dimer 2 C. T. Chen, D. Ng, S. R. Marder, P. Miles, T. Wada, M. Tian and
H. Sasabe, Science, 1996, 273, 1533–1536; (b) L. W. Tutt and
To a suspension of magnesium turnings (0.5 g, 20.57 mmol), T. F. Boggess, Prog. Quantum Electron., 1993, 17, 299–338; (c)
previously activated by means of an ultrasonic bath under M. Calvete, G. Y. Yang and M. Hanack, Synth. Met., 2004, 141,
231–244.
reduced pressure, in anhydrous 1,4-dioxane (20 ml, Aldrich
2 J. W. Perry, K. Mansour, S. R. Marder, K. J. Perry, D. Alvarez
product) was added a solution of tBu4PcGaCl 4a (0.3 g, and I. Choong, Opt. Lett., 1994, 190, 625–627; N. J. Turro, Modern
0.36 mmol) in anhydrous 1,4-dioxane (30 ml) and a catalytic Molecular Photochemistry, University Science Books, Sausalito
amount of iodine (10 mg, 0.04 mmol) under argon. After (CA, USA), 1991.
3 H. S. Nalwa and J. S. Shirk, in Phthalocyanines: Properties
refluxing at 110 uC for 48 h, the reaction solution was and Applications, C. C. Leznoff and A. B. P. Lever, eds., VCH
transferred into a pressure equalizing fritted glass funnel Publishers, Inc., New York, 1996, vol. 4, pp. 79–181.
containing a small quantity of glass wool to hold back excess 4 (a) Y. Chen, L. R. Subramanian, M. Barthel and M. Hanack, Eur.
solid magnesium. The obtained filtrate was evaporated to J. Inorg. Chem., 2002, 1032–1034; (b) M. Hanack and
H. Heckmann, Eur. J. Inorg. Chem., 1998, 367–373; (c) J. S. Shirk,
dryness under high vacuum. To further remove any inorganic R. G. S. Pong, S. R. Flom, H. Heckmann and M. Hanack, J. Phys.
impurity, the crude product was dissolved in dry toluene, Chem. A, 2000, 104, 1438–1449; (d) M. Hanack, T. Schneider,
stirred, filtered (this procedure was repeated several times), M. Barthel, J. S. Shirk, S. R. Flom and R. G. S. Pong,
dried at 60 uC in vacuo to afford a blackish-green solid Coord. Chem. Rev., 2001, 219–221, 235–258; (e) Y. Chen,
L. R. Subramanian, M. Fujitsuka, O. Ito, S. O’Flaherty,
powder in 83% yield (0.264 g). MALDI-MS (m/z): calc. for W. J. Blau, T. Schneider, D. Dini and M. Hanack, Chem. Eur.
C104H112Ga2N16O4: 1788.8; found: 1789. UV/Vis (CHCl3): J., 2002, 8, 4248–4254; (f) Y. Chen, M. Barthel, M. Seiler,
lmax(nm) 5 693, 664(sh), 625, 356. PL (CHCl3, lexc 5 355 nm): L. R. Subramanian, H. Bertagnolli and M. Hanack, Angew. Chem.,
Int. Ed., 2002, 41, 3239–3242; (g) Y. Chen, S. M. O’Flaherty,
lmaxfluor (nm) 5 707; FTIR (KBr, cm21): 2962(ms), 2922(m), M. Fujitsuka, M. Hanack, L. R. Subramanian, O. Ito and
2857(m), 1779(w), 1723(w), 1613(m), 1506(m), 1482(m), W. J. Blau, Chem. Mater., 2002, 14, 5163–5168; (h) D. Dini,
1458(m), 1407(m), 1394(m), 1363(m), 1333(ms), 1260(vs), M. Barthel and M. Hanack, Eur. J. Org. Chem., 2001, 20,
1199(m), 1144(m), 1087(vs), 1051(s), 1022(vs), 967 (w), 3759–3769; (i) M. Hanack, D. Dini, M. Barthel and S. Vagin,
Chem. Record, 2002, 2, 129–148; (j) Y. Chen, M. Fujitsuka,
929(ms), 9167(w), 893(w), 863(m), 803(vs), 764(m), 749(s), S. M. O’Flaherty, M. Hanack, O. Ito and W. J. Blau, Adv. Mater.,
737(m), 694(m), 670(m), 601(w), 566(w), 531(m), 518 (vw), 2003, 15, 899–902; (k) G. Y. Yang, M. Hanack, Y. W. Lee,
442(w), 390(m). 1H NMR (CDCl3): d/ppm 5 1.9(m, 72H, tBu), Y. Chen, M. K. Y. Lee and D. Dini, Chem. Eur. J., 2003, 9, 2758;
(l) S. M. O’Flaherty, S. V. Hold, M. J. Cook, T. Torres, Y. Chen,
8.2–8.3(m, 8H, H-1), 8.8–9.3(m, 16H, H-2, 20), 3.7 and 4.2(m,
M. Hanack and W. J. Blau, Adv. Mater., 2002, 15, 19–32; (m) G. De
H-dioxane). 13C NMR (CDCl3): d/ppm 5 32.1 (tBuCH3), la Torre, P. Vazquez, F. Agullo-Lopez and T. Torres, J. Mater.
36.2(CMe3), 118.9–119.6(C-20), 122.6–123.0(C-2), 128.5(C-1), Chem., 1998, 8, 1671–1683; (n) G. De la Torre, P. Vazquez,
133.5–134.1(C-3), 135.8–136.6(C-30), 150.8–151.7(C-4,40), F. Agullo-Lopez and T. Torres, Chem. Rev., 2004, 104, 3723–3750;
(o) M. Barthel, D. Dini, S. Vagin and M. Hanack, Eur. J. Org.
154.2–154.5(C-10). Chem., 2002, 3756–3762.
5 W. Uhl, Coord. Chem. Rev., 1997, 163, 1–32; M. J. Taylor and
P. J. Brothers, in Chemistry of Aluminium, Gallium, Indium, and
Acknowledgements Thallium, A. J. Downs, ed., Blackie-Chapman-Hall, London, 1993,
pp. 111–247.
This research was supported by Deutsche
6 G. H. Robinson, Chem. Commun., 2000, 2175–2176.
Forschungsgemeinschaft (Ha 280/165-1) and EU project 7 W. Uhl, T. Spies and R. Koch, J. Chem. Soc., Dalton Trans., 1999,
HPRN-CT-2000-00020. Y. C. thanks the Alexander 2385–2392.
von Humboldt Foundation of Germany for a fellowship. 8 P. Wei, X. W. Li and G. H. Robinson, Chem. Commun., 1999,
1287–1288.
The Dublin group acknowledges support by the Irish
9 K. S. Klimek, C. Cui, H. W. Roesky, M. Noltemeyer and
Higher Education Authority and Enterprise Ireland, and H. G. Schmidt, Organometallics, 2000, 19, 3085–3090.
lab assistance from Ms S. V. Hold. We also thank 10 W. Uhl and T. Spies, Z. Anorg. Allg. Chem., 2000, 626, 1059–1064.

688 | J. Mater. Chem., 2005, 15, 683–689 This journal is ß The Royal Society of Chemistry 2005
View Article Online

11 J. M. Barbe and R. Guilard, in The Porphyrin Handbook, 17 D. Eastwood, L. Edwards, M. Gouterman and J. Steinfeld,
K. Kadish, K. M. Smith and R. Guilard, eds., Academic Press: J. Mol. Spectrosc., 1966, 20, 381–390; M. Goutermann, in
San Diego, 1999, vol. 3, ch. 19; C. Ercolani, J. Porphyrins The Porphyrins, R. Dolphin, ed., Academic Press, New York,
Phthalocyanines, 2000, 4, 340–343; R. Caminiti, M. Donzello, 1977, vol. 3, pp. 1–157.
C. Ercolani and C. Sadun, Inorg. Chem., 1999, 38, 3027; 18 S. Dhami, A. J. De Mello, G. Rumbles, S. M. Bishop, D. Phillips
R. Caminiti, M. Donzello, C. Ercolani and C. Sadun, Inorg. and A. Beeby, Photochem. Photobiol., 1995, 61, 341–346;
Chem., 1998, 37, 4210–4213; A. Capobianchi, A. M. Paoletti, M. L. Spaeth and W. R. Sooy, J. Chem. Phys., 1998, 48,
G. Pennesi, G. Rossi, R. Caminiti and C. Ercolani, Inorg. Chem., 2315–2323; T. H. Huang and J. H. Sharp, Chem. Phys., 1982, 65,
1994, 33, 4635–4640. 205–216.
12 H. Bertagnolli and T. S. Ertel, Angew. Chem., Int. Ed. Engl., 1994, 19 M. Hanack and M. Lang, Adv. Mater., 1994, 6, 819–833.
33, 45–66. 20 M. Sommerauer, C. Rager and M. Hanack, J. Am. Chem. Soc.,
13 D. R. Tackley, G. Dent and W. E. Smith, Phys. Chem. Chem. 1996, 118, 10 085–10 093.
Published on 24 November 2004. Downloaded by FAC DE QUIMICA on 20/03/2014 19:49:21.

Phys., 2001, 3, 1419–1426. 21 M. Sheik-Bahae, A. A. Said, T. H. Wei, D. J. Hagan and E. W. Van


14 (a) K. J. Wynne, Inorg. Chem., 1984, 23, 4658–4663; (b) Stryland, IEEE J. Quantum Electron., 1990, 26, 760–769.
K. J. Wynne, Inorg. Chem., 1985, 24, 1339–1343. 22 T. S. Ertel, H. Bertagnolli, S. Hückmann, U. Kolb and D. Peter,
15 L. H. Vogt, Jr., A. Zalkin and D. H. Templeton, Inorg. Chem., Appl. Spectrosc., 1992, 46, 690–698.
1967, 6, 1725–1730. 23 M. Newville, P. Livins, Y. Yakobi, J. J. Rehr and E. A. Stern,
16 C. Fietzek, M. Seiler, B. Görlach, P. Schütz, U. Weimar, Phys. Rev. B, 1993, 47, 14 126–14 131.
M. Hanack, C. Ziegler and H. Bertagnolli, J. Mater. Chem., 24 S. J. Gurman, N. Binsted and I. Ross, J. Phys. C, 1986, 19,
2002, 12, 2305–2311. 1845–1861.

This journal is ß The Royal Society of Chemistry 2005 J. Mater. Chem., 2005, 15, 683–689 | 689

You might also like