Sei sulla pagina 1di 26

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

Chem Soc Rev Dynamic Article Links

Cite this: Chem. Soc. Rev., 2011, 40, 2541–2566

www.rsc.org/csr CRITICAL REVIEW


Organometallic acetylides of PtII, AuI and HgII as new generation
optical power limiting materials
Gui-Jiang Zhou*a and Wai-Yeung Wong*b
Received 31st August 2010
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

DOI: 10.1039/c0cs00094a

Within the scope of nonlinear optics, optical power limiting (OPL) materials are commonly
regarded as an important class of compounds which can protect the delicate optical sensors or
human eyes from sudden exposure to damaging intense laser beams. Recent efforts have been
devoted to developing organometallic acetylide complexes, dendrimers and polymers as high
Downloaded by UNSW Library on 17 March 2013

performance OPL materials of the next generation which can favorably optimize the optical
limiting/transparency trade-off issue. These metallated materials offer a new avenue towards a
new family of highly transparent homo- and heterometallic optical limiters with good solution
processability which outperform those of current state-of-the-art visible-light-absorbing
competitors such as fullerenes, metalloporphyrins and metallophthalocyanines. This critical review
aims to provide a detailed account on the recent advances of these novel OPL chromophores.
Their OPL activity was shown to depend strongly on the electronic characters of the
aryleneethynylene ligand and transition metal moieties as well as the conjugation chain length of
the compounds. Strategies including copolymerization with other transition metals, change of
structural geometry, use of a dendritic platform and variation of the type and content of
transition metal ions would strongly govern their photophysical behavior and improve the
resulting OPL responses. Special emphasis is placed on the structure–OPL response relationships
of these organometallic acetylide materials. The research endeavors for realizing practical OPL
devices based on these materials have also been presented. This article concludes with perspectives
on the current status of the field, as well as opportunities that lie just beyond its frontier
(106 references).

1. Introduction caused by exposure to intense accidental or hostile laser pulses.


Eventually, this has stimulated many researchers to search for
With their successful research development, various laser effective optical power limiters exhibiting fast response speed,
sources have brought about revolutionary changes and substantial good linear transparency and high linear transmittance.1–3
improvement to not only the scientific community and the Optical limiters are devices that can protect sensitive sensors
industrial world but also to our social life. Today, lasers are from in-band intrusive, possibly damaging and frequency-agile,
used extensively in many photonic, medical and military pulsed laser radiations.3,4 Thus, these devices should modulate
applications but there is an urgent need for protection against the transmitted intensity of an intense laser beam to some
lasers. Apart from the positive effects that the laser technology specified maximum which is safe to the optical sensors to be
has made to the society in the current photonic era, the side protected, but exhibit high transmittance for low-intensity
effects have also appeared because of their threats against ambient light levels. Obviously, optical limiters based on linear
human eyes, optical sensors, and sensitive optical components optical mechanisms, such as absorption, reflection and diffraction,
cannot meet the application specifications well since they
a
MOE Key Laboratory for Nonequilibrium Synthesis and Modulation would consume the required energy of the laser beam even
of Condensed Matter and Department of Chemistry, Faculty of with safe power levels.3 Encouragingly, the discovery of long
Science, Xi’an Jiaotong University, Xi’an 710049, P.R. China. transient effects in solutions of some aromatic organic molecules
E-mail: zhougj@mail.xjtu.edu.cn
b
Institute of Molecular Functional Materialsw and Department of and the reverse saturable absorption (RSA) effect of some
Chemistry and Centre for Advanced Luminescence Materials, Hong organic dyes under strong laser radiation provides us an ideal
Kong Baptist University, Waterloo Road, Kowloon Tong, Hong outlet for the design of useful optical limiters.5,6 As a result, it
Kong, P.R. China. E-mail: rwywong@hkbu.edu.hk; is desirable to develop active materials for optical power limiting
Fax: +852 3411-7348; Tel: +852 3411-7074
w Areas of Excellence Scheme, University Grants Committee (OPL) based on nonlinear optics (NLO).7–9 Currently, the
(Hong Kong). materials employed for OPL can show a variety of chemical

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2541
View Article Online

structures, such as fullerenes (e.g. C60),10–15 phthalocyanines irradiance, corresponding to a linear optical effect (i.e. obeying
(Pc),16–21 porphyrins,22–25 diacetylenes,26 nanotubes,27–29 and Beer’s law). Here, transmittance of the sample remains almost
metal–organic/cluster compounds,30–35 by virtue of their fast constant. However, when the incident irradiance increases to
response speed (in the picosecond (ps) range) and large optical reach a certain threshold, the output fluence will stay in a
nonlinearities.3 Among them, C60, metallophthalocyanines particular level which is unaffected by further increasing the
and metalloporphyrins are amongst the most effective OPL input fluence, and the NLO effect has come into play to reduce
materials, in which the heavy-metal effect increases the the fluence of the transmitted light to a safe level. Several
spin–orbit coupling (SOC) and significantly magnifies the photophysical processes such as RSA, two-photon absorption
OPL response by enhancing the intersystem crossing (ISC) (TPA), nonlinear refraction, nonlinear scattering and free
from S1 to T1 states, thereby giving greater access to the carrier absorption are possible for the OPL mechanism
highly-absorbing triplet-state manifold.36–41 However, these (Fig. 1c).1,3
benchmark OPL materials, such as fullerenes, Pc, porphyrins One of the most promising material approaches is based on
and their derivatives, are all deeply colored solids and show reverse saturable absorbers which derive their optical limiting
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

strong absorption bands in the visible-light region,4 indicating capability on the photogeneration of highly absorbing charge
their poor optical transparency. Additional efforts to over- states. For the typical five-level model in RSA, the molecules
come such deficiency would spur new developments in this of the OPL materials in the ground state (S0) initially absorb
research area. Recently, some purely organic polyynes42–44 optical energy weakly and are excited to the first singlet
and poly(1-alkyne)s45–47 were also selected for OPL testing state (S1). After the molecules in the S1 state gain a certain
and some of them were shown to be good OPL materials with population, they can absorb optical energy strongly and hop
Downloaded by UNSW Library on 17 March 2013

performance close to that of C60, but their lmax values still go to the higher singlet state (Sn) to induce the OPL effect or
beyond 400 nm, leading to inferior transparency window. undergo ISC to the first triplet state (T1) when the width of a
Thus, the molecular design of new generation OPL materials, laser pulse is long enough. Based on the assumptions that
with superior optical transparency/OPL trade-off optimization, there is no saturation, diffusion or recombination during the
remains a great challenge for practical OPL applications. laser pulse in the RSA procedures, the system can be reduced
to a three-level one. Thus, we can neglect all the excitations
higher than S1 provided that the ISC rate is fast compared
2. How does optical limiting work?
with the laser pulse duration and all of the excitations are
The typical OPL effect can be explicated by Fig. 1a and b. populated in the T1 state. When the molecules in the T1 state
When the incident irradiance upon a sample is weak, the gain a certain population, they can absorb optical energy to
relationship between input fluence and output fluence (fluence reach the higher triplet state (Tn). Thus, the strong energy
= optical energy density) is linear with increasing incident absorption accompanied with the T1 - Tn transition should

Guijiang Zhou received his Wai-Yeung Wong graduated


PhD degree from the Institute from the University of Hong
of Chemistry, Chinese Academy Kong, with both BSc (Hons)
of Sciences (CAS) in 2003. (1992) and PhD (1995) degrees
After a year as postdoctoral (with Prof. Wing-Tak Wong).
research fellow in Korea, he After his postdoctoral work at
held a postdoctoral position in Texas A&M University in
Hong Kong Baptist University 1996 with Prof. F. Albert
with Prof. Wai-Yeung Wong. Cotton, he worked for Profs
From April 2007 to September The Lord Lewis (FRS) and
2008, he was a postdoctoral Paul R. Raithby at the
fellow supported by the Ministry University of Cambridge in
of Science and Education of 1997 as a Croucher Research
Gui-Jiang Zhou Spain in the University of Wai-Yeung Wong Fellow. He joined the Hong
Murcia. In November 2008, Kong Baptist University as an
he joined the Department of Chemistry, Xi’an Jiaotong Assistant Professor in 1998 and is currently a Chair Professor in
University, where he is currently a Professor. Current research Chemistry there. His research mainly focuses on synthetic
interests include functionalized phosphorescent organometallic inorganic/organometallic chemistry and structural chemistry,
materials for optical power limiting and electroluminescence. with special emphasis on developing metallopolymers and
metallophosphors with energy functions and photofunctional
properties. He has a distinguished publication record of over
300 papers and is listed among the world’s top 1% in the ISI list
of most cited chemists. He is the recipient of the RSC Chemistry
of the Transition Metals Award in 2010, and has won the
Croucher Senior Research Fellowship and two Asian Core
Program Lectureship Awards in 2009.

2542 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 1 (a) The typical OPL effect. (b) Illustration of the concept of OPL effect. (c) Schematic diagrams for some representative OPL mechanisms.

be responsible for the observed OPL action. It can be shown direction, which then induces the OPL effect as well. This
that the OPL effect for lasers with a long pulse duration (e.g. mechanism for OPL is called nonlinear scattering.
for nanosecond (ns) lasers) is induced mainly by T1 - Tn Photogenerated free carriers can be formed in the semi-
absorption, while that for lasers with a shorter pulse duration conductors by laser irradiation. These carriers can absorb
(in the ps and femtosecond (fs) ranges) is due to the S1 - Sn extra photons to be excited to the conduction band from the
absorption (Fig. 1c(i)). valence band (i.e. free carrier absorption), which has been
Different from the RSA event, TPA is a fast nonlinear shown to induce the OPL effect. Obviously, the amount of
optical process which involves two fast absorption processes. absorbed energy by the free carriers depends on their number,
The molecules in the S0 state firstly absorb one photon to be so free carrier absorption is an accumulating nonlinear optical
excited to a virtual state, which then can absorb another process.
photon to leap to the final excitation state S1. The intense Among all the mechanisms involved in OPL, the most
energy absorption associated with the S0 - S1 transition, prevalent paradigms are RSA and TPA. To characterize
which is accomplished by absorbing two photons, will induce the RSA and TPA behavior of OPL materials, the most
the OPL effect (Fig. 1c(ii)). Usually, the S1 - Sn or T1 - Tn commonly employed technology is the Z-scan method. The
transitions can still contribute to the OPL effect under some Z-scan method was introduced in 1989 by Sheik-Bahae et al.
circumstances besides that from TPA. Actually, absorption to determine the magnitude of the excited-state absorption
from the S1 - Sn or T1 - Tn transitions does not belong to cross-section (sex), the nonlinear absorption coefficient (b)
part of the TPA process. and the sign and magnitude of the nonlinear refractive index
The OPL effect achieved by the nonlinear refraction process (n2).48,49 The typical set-up for Z-scan measurement is
is not simply due to absorption, but is caused by reducing the shown in Fig. 2. This technique involves focusing a laser beam
laser radiation received by the optical sensors through on a thin sample and detecting the transmitted light as
the optical refraction effect (including self-defocusing and the sample is scanned through the Z-direction to alter the
self-focusing) induced by the generated carriers in the materials incident fluence on it. At the same time, the energy for
under a strong optical incident fluence (Fig. 1c(iii)). the reference beam is recorded by another detector. The
When the laser radiation is strong enough, it can induce characteristic intensity variation with the incident light
light scattering centers in the medium. The scattering centers intensity is fitted to a theoretical equation to extract the
would lower the transmittance of the medium in a certain desired parameters.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2543
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

Fig. 2 The typical double-beam experimental set-up for the Z-scan measurement.

To date, achieving ultimate eye protection is still a big photonics and optical sensing for their intriguing luminescence
challenge in the OPL field. Different from optical sensor properties, electrical conductivities, NLO properties, liquid
protection, eye protection devices require the OPL materials crystallinity and photovoltaic behavior.62,63 The detailed
Downloaded by UNSW Library on 17 March 2013

to modulate the intense laser radiation of a certain wavelength photophysical investigations concerning this group of materials,
to a safe energy level for human eyes, but at the same time pass especially for some PtII acetylide compounds, indicate that
any light in the visible region (400–700 nm) that the wearer they can be efficient triplet emitters even at room temperature
may wish to see. This implies that the ideal optical limiting (rt) and their main absorption bands can be moved down
materials, especially for preventing severe eye damage, should below 400 nm with only an extremely weak absorption band
possess good optical transparency in the entire visible wavelength (absorption coefficient e o 2 dm3 mol–1 cm–1) in the visible
region. Unfortunately, the state-of-the-art OPL materials region.64–68 The efficient triplet emission at rt together with
developed, including C60, phthalocyanines, porphyrins and good transparency features would render them promising
their derivatives, exhibit unwanted visible-light absorption OPL materials according to the RSA mechanism and some
bands, which would hamper their use within eye protection encouraging results have been obtained in the past few years.
application because the absorption will cause vision loss to By a careful selection of the arylacetylide ligands, or by using
some extent. While it is known that a bathochromic shift of the different transition metal centers and tuning the molecular
ground-state absorption maximum (lmax) generally leads to an geometry, etc., organometallic acetylides, especially the PtII
increase in the third-order optical nonlinearity,50–53 the acetylide compounds, become very effective materials in terms
previously reported optical limiters have not yet achieved the of their superior OPL/transparency trade-off optimization,
nonlinearity/transparency trade-off optimization. So, we are in even better than the RSA dyes such as C60, metalloporphyrins
need of designing next generation OPL materials to address and metallophthalocyanines. So, compounds based on transition
this problematic issue. As far as the most common OPL metal acetylide chromophores are promising candidates for
mechanisms (i.e. RSA and TPA) are concerned, compounds developing OPL-active materials for eye protection and other
showing RSA behavior probably show the greatest promises optical protection devices. Herein, a survey of various alkynyl
in giving highly transparent OPL materials, since the S0 - S1 complexes of late transition metals is given and we will discuss
transition in the RSA mechanism is a weak absorption below the recent progress of this new class of OPL materials,
process, and so materials of this kind can display good optical which form an important subset of organometallic complexes
transparency at the laser wavelength used (Fig. 1c(i)). for NLO.
Therefore, if novel materials devoid of strong visible optical
absorptions can be synthesized, OPL materials fulfilling the 3. Classifications of organometallic acetylides as
bottleneck requirement of eye protection devices can be optical power limiters
obtained. However, OPL materials based on the TPA process
are less competitive in this regard, because the S0 - S1 3.1 Small molecules
transition involved in TPA usually consists of a relatively
To date, there are different kinds of organometallic acetylides
strong absorption which would endow poor transparency to
available for OPL purpose. In this section, we divide our
the materials. Furthermore, the TPA effect is typically induced
discussion according to the type of metal center, viz. PtII,
by the sub-ns laser pulse.
AuI and HgII as well as their mixed-metal combinations.
The Hagihara’s group in Japan is one of the founding
fathers of oligo- and polymetallaynes, who first reported the 3.1.1 Group 10 PtII acetylides
synthesis of oligomeric and polymeric PdII and PtII acetylide 3.1.1.1 Effect of spacer ligands. The pioneering investigations
compounds stabilized by tertiary phosphine ligands in the on the photophysical properties of a series of phenylethynyl-type
1970s.54–57 Since then, the chemistry of group 10 metal-containing PtII complexes with different conjugation lengths, 1–3, reveal
metallaynes has gained much research momentum,58–61 owing their potential in OPL research (Fig. 3).69–72 Among them, 2 has
to their anticipated applications in molecular electronics, shown its broadband OPL ability. More importantly, the

2544 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

Fig. 3 Chemical structures of phenylethynyl-type PtII acetylides in


the early stage of development. Adapted from ref. 69–72.

photophysical data correlated the effects of the ligand structure


Fig. 5 Z-scan curves for 4–9 as compared to C60 at the same linear
on the character of excited states,67,73 and the OPL behavior
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

transmittance To = 82%. Adapted from ref. 74.


accordingly. This provides vital information on how the structure
of the organic spacer ligand can be selected to give better OPL
materials of this kind. (>400 nm) due to the ligand-to-metal charge transfer (LMCT)
As aforementioned, the key ligand structure of the PtII or metal-to-ligand charge transfer (MLCT) transitions
acetylides has profound effects on their photophysical traits (Fig. 6), which indicates that the concerned ligands cannot
which have a close tie with the OPL performance. Various PtII confer good transparency to the complexes. Fortunately,
Downloaded by UNSW Library on 17 March 2013

acetylides, 4–9, with a diverse range of organic aromatic complex 5 even shows a much better OPL response than that
groups have been synthesized to characterize their OPL of C60 at To = 82%. At the same time, its absorption onset
properties (Fig. 4).74 According to the Z-scan results for these below 400 nm with no apparent absorption band within
complexes shown in Fig. 5 at 532 nm for a 10 ns laser pulse, 400–700 nm indicates its excellent optical transparency and 5
they give us a clear picture for the effect of the electronic represents a promising candidate for eye-protection-type
structure of the spacer ligand on the OPL response. PtII applications. Complex 4 also shows good transparency relative
arylacetylides with electron-donating or electron-withdrawing to 5, despite its moderate OPL performance. Due to the much
spacers 6, 7 and 8 display good OPL performance which is stronger triplet emission induced by the SOC effect of the
not far from that of the C60 solution at the same linear heavy metal ion in 5 compared with that of 4 at rt, 5 shows a
transmittance To = 82%. Upon increasing the conjugation much better OPL performance. Low temperature photo-
of the ligand and introducing an electron-donating group such luminescence (PL) spectra at 77 K also reveal that introduction
as OMe, complex 9 can show similar OPL effect to that of C60 of PtII ions into the molecules results in enhanced triplet
(Fig. 5). Despite their attractive OPL performance, complexes emission quantum yield (FT). So, their good OPL performance
6–9 all show clear absorption bands in the visible-light region is mainly attributed to the triplet RSA process for a ns laser

Fig. 4 PtII acetylides with organic ligands showing distinct electronic characters. Adapted from ref. 74.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2545
View Article Online

a clue that selection of organic ligands without electron-


donating or electron-withdrawing characters and with stronger
heavy-atom-induced SOC effect can offer a good avenue
towards designing new generation OPL materials based on
the RSA mechanism with a good compromise between OPL
response and optical transparency.
A series of symmetric PtII acetylides 10–13 with the metal
ion located at the center of the molecules for OPL character-
ization at 532 nm was developed using click chemistry
(Fig. 7).75 It was shown that the OPL response increased with
conjugation length of the organic ligand from 10 to 12. The
Fig. 6 UV-vis spectra for 4–9 in CH2Cl2 at rt. Adapted from ref. 74. output energy clamping levels for 11 and 12 (5.5 mJ and 2.5 mJ
at ca. 200 mJ, respectively) were even better than that of zinc-
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

pulse. Different from the ligands for 6–9, the organic chromo- based porphyrins, indicating their good OPL performance.75
phores for 4 and 5 do not exhibit strong electron-donating or Furthermore, complex 12 even shows better OPL response
electron-withdrawing features. The OPL results for 4–9 give us than that of 13. The same research group also ascribed the
Downloaded by UNSW Library on 17 March 2013

Fig. 7 Chemical structures of PtII acetylides containing triazole and acetonide-protected 2,2-bis(methylol)propionic acid (bis-MPA) groups.
Adapted from ref. 75.

2546 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

improved OPL performance of 12 to its longer conjugated


organic ligand coupled with the triazole end-group function-
alities, which can afford strong excited-state absorption as well
as efficient charge transfer. This kind of highly transparent PtII
acetylides can also achieve optimized optical limiting/
transparency trade-off. However, it should be noted that their
Fig. 9 Chemical structures of PtII complexes anchored with the
absorption maxima also show a bathochromic effect with 4 0 -arylterpyridyl ligand. Adapted from ref. 80.
increasing conjugation length of their chromophoric ligands.

3.1.1.2 Effect of auxiliary ligands. Besides phosphine


ligands, heterocyclic compounds containing N atoms (especially
the pyridine-type derivatives such as bipyridyl and terpyridyl
frameworks) can also serve as the auxiliary ligands for PtII
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

ions in organometallic acetylides.76–78 While phosphines


typically show strong electron-donating character, the pyridyl
derivatives exhibit electron-withdrawing features. It can be Fig. 10 Chemical structures for 4 0 -(5 0 0 0 -R-pyrimidyl)-2,2 0 :6 0 ,200 -
expected that PtII acetylides bearing these two kinds terpyridyl PtII phenylacetylide complexes. Adapted from ref. 81.
of auxiliary ligands show markedly different photophysical
properties and OPL behavior.
Downloaded by UNSW Library on 17 March 2013

Sun et al. designed and synthesized several series of PtII


alkynyls containing terpyridyl and phenyl–bipyridyl auxiliary
ligands 14–16 for OPL investigation. Fig. 8 shows a series of
PtII acetylides with terpyridyl derivatives as the auxiliary
ligands.79 Their high FT and broad transient absorption of
3
MLCT states demonstrate their great potential as broadband
OPL materials. The OPL results for a 4.1 ns laser pulse at Fig. 11 Chemical structures of 2,4-di-2 0 -pyridyl-6-p-tolyl-1,3,5-
532 nm show that they can act as effective OPL materials and triazine PtII acetylide complexes. Adapted from ref. 82
their OPL response follows the trend 14 > 15 > 16.
Complexes 14 and 15 can even give comparable figure of merit of 26 > 27 > 28) (Fig. 11) also show unsatisfactory OPL
sex/so to that of C60 at To = 70%, where sex and so refer action for a ns laser pulse at 532 nm, which are even weaker
to the effective excited-state and ground-state absorption than those of 21–25.82 This can be rationalized from the
cross-sections, respectively. UV-vis spectra for 26–28 which exhibit large absorption at
Another series of PtII acetylides with terpyridyl auxiliary 532 nm, and hence their large so.
ligands 17–20 developed by Sun’s group (Fig. 9) shows slightly Apart from the terpyridine derivatives, 4,6-diphenyl-2,2 0 -
better OPL performance than those of 14–16, on the basis of bipyridine moieties can also serve as the auxiliary ligands for
the reported sex/so values at the same To.80 Using a ns laser at PtII acetylides. Some complexes of this type (29 and 30) have
532 nm, the OPL performance followed the order 20 > 17 > been employed for OPL study (Fig. 12).83 Complex 29 shows
18 > 19. The better OPL behavior for 20 is attributed to its not only better OPL performance than 30, but also notable
longer triplet lifetime and larger sex. With smaller sex OPL response to the ps laser pulse. For a 4 ns laser pulse at
compared with that of other complexes, 19 shows the weakest 532 nm, the transmittance can be reduced to 44% for 29 and
OPL response. 51% for 30 at To = 90%, which is much better than that of a
However, similar complexes 21–25 shown in Fig. 10 produce similarly-structured PtII acetylide with a terpyridine auxiliary
weaker OPL effect relative to those listed in Fig. 8.81 Apparently, ligand.
23 and 25 give weaker OPL responses than those of 21, 22 and Recently, another series of PtII acetylides with the alkoxy-
24, which are ascribed mainly to the substantially large so. By substituted 4,6-diphenyl-2,2 0 -bipyridine ligand, 31–34, has
a similar reasoning, 26–28 (with OPL performance in the order also been studied (Fig. 13) utilizing a 4 ns laser pulse at 532
nm.84 The OPL performance order for these complexes is 31
> 32 E 33 > 34. The noticeably weaker OPL effect for 34 was

Fig. 8 Chemical structures of charged 4 0 -tolylterpyridyl arylacetylide Fig. 12 Chemical structures of PtII acetylide complexes with the
complexes of PtII. Adapted from ref. 79. 4,6-diphenyl-2,2 0 -bipyridyl ligand. Adapted from ref. 83.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2547
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

Fig. 13 Chemical structures of alkoxy-substituted 4,6-diphenyl-2,2 0 -bipyridyl PtII acetylide complexes. Adapted from ref. 84.

attributed to its large so value induced by its extended


conjugation length. Besides the factor due to FT, the so values
also give a reasonable explanation to the OPL order for 31–34.
Downloaded by UNSW Library on 17 March 2013

It was also shown that the alkoxy substituent can exert a


positive effect on the OPL performance for these reported
complexes.
In short, PtII acetylides with pyridine derivatives can be
effective OPL materials by carefully controlling the molecular
structures. Both of their photophysical and OPL investigations
for the ns laser pulse show that the 3MLCT/triplet ligand-
to-ligand charge transfer (3LLCT) excited states should be
responsible for their RSA-based OPL action. However, this
would result in strong MLCT/LLCT bands (from ca. 400 nm Fig. 15 Input–output fluence curves of OPL measurement for 35 and
to 600 nm, depending on the nature of the individual complex) 36 at 532 nm. Adapted from ref. 85.
in the visible region in their absorption spectra, together with
poor optical transparency window for these complexes.
nearly half of the fluorescence quantum yield (FF) compared
Compared with the PtII acetylides with phosphine ligands
with that of 35 (FF E ca. 0.5%), which suggests that the
mentioned in the previous section, this type of PtII acetylides
ligand chelating pattern in 36 would favor the SOC effect to
would probably not be favorable for achieving excellent
give rise to lower FF and higher FT. The stronger SOC effect
optical limiting/transparency trade-off optimization.
has also been confirmed theoretically by the much higher
contribution of the dp orbital of PtII to the HOMO in 36.
3.1.1.3 Effect of ligand chelating patterns. The metal center According to the triplet RSA mechanism, 36 would produce
for PtII acetylide complexes with phosphine-type auxiliary stronger OPL response.
ligands can be positioned either at the terminal ends (see
Fig. 4) or in a middle position (see Fig. 7). With the aim of 3.1.1.4 Effect of molecular symmetry. PtII acetylides with
evaluating the effect of ligand chelating patterns on PtII ions, phosphine auxiliary ligands usually adopt symmetric structures,
the OPL performances for 35 and 36 (Fig. 14) are compared.85 bearing two organic acetylides with identical structures
For a 10 ns laser pulse at 532 nm, 36 shows stronger arranged in a trans disposition. The asymmetric PtII acetylides
OPL response than that of 35 (Fig. 15). Their different are relatively rare. Recently, we have developed a novel series
OPL behavior can be explained by both photophysical and of PtII acetylides by tailoring the electronic properties of the
computational studies. Complex 36 (FF E ca. 0.3%) shows organic ligands and adopting symmetric/asymmetric ligand

Fig. 14 Chemical structures of fluorene-based PtII acetylides with different ligand chelating patterns. Adapted from ref. 85.

2548 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 16 Chemical structures of PtII acetylides with symmetric/asymmetric ligand configuration. Adapted from ref. 86.

configuration, such as D–Pt–D, D–Pt–A and A–Pt–A


(D = donor, A = acceptor) (37–45 in Fig. 16).86 Photo-
physical characterization of these molecules shows an
absorption maximum within the near-UV region (Fig. 17)
and emission through both fluorescence and phosphorescence
(Fig. 18). Both the photophysical and OPL behavior are
significantly affected by their molecular symmetry and the
electronic structures of the arylacetylides. For the complexes
with symmetric configuration, including 37–41, their OPL
investigations performed by Z-scan technology for a 10 ns
laser pulse at 532 nm show that complexes 37 and 38 with the
D–Pt–D structure possess much better OPL ability than that
of A–Pt–A type complexes 40 and 41 (Fig. 19). Compared with Fig. 17 UV–vis spectra of the symmetric/asymmetric PtII acetylides
the OPL results for the reference complex 39, the admirable 37–45 in CH2Cl2 at 293 K. Adapted from ref. 86.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2549
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 18 PL spectra of 37–45 in CH2Cl2 at 293 K and 77 K with the excitation wavelength at 334 or 365 nm, depending on the nature of PtII
acetylides. Adapted from ref. 86.

OPL performance of 37 and 38 gives a very clear picture for character whereas A–Pt–A structured complexes (40 and 41)
the role the electron-donating ligand and its conjugation mainly exhibit MLCT feature. The triplet states for the
length had played to enhance the OPL effect. With reference complexes with asymmetric D–Pt–A configuration show a
to 41, the much stronger OPL response for 40 can be safely ligand (D)-to-ligand (A) charge transfer (LLCT) trait, and
ascribed to its longer conjugated ligands which facilitate the accordingly their OPL behavior associated with the absorption
nonlinear process as in the case of 38. of triplet excitons would be different. It was found that the
Apart from the symmetric PtII acetylides stated above, their ligand variation to the OPL strength of these PtII compounds
asymmetric counterparts with D–Pt–A configuration were follows the order: D–Pt–D > D–Pt–A > A–Pt–A. Except for
also developed and studied for their OPL behavior at 41 and 44, all the complexes show better OPL performance
532 nm. The input–output fluence curves indicate that the than that of the benchmark C60 even at To = 92%. The sex/so
complexes adopting the D–Pt–A asymmetric structure (42–45) values fall in the range of 9–17 for 37–45 at 532 nm, which
generally show weaker OPL responses compared with their exceed that of C60 by a factor of about 3. Most of these figures
mother complex with D–Pt–D configuration (37 and 38) are also comparable to those of typical metal phthalocyanines.
(Fig. 19b and c). Owing to their higher FT caused by The optical limiting threshold (Fth) for certain complexes such
the lesser conjugation, the asymmetric complexes with the as 38 (ca. 0.05 J/cm2 at To = 92%) is lower than that of
p-accepting organoborane group give stronger OPL responses the best InPc and PbPc OPL materials (ca. 0.07 J/cm2 at
with respect to those of the analogous complexes bearing an To = 84%) currently in use even at higher To. Apart from
oxadiazole unit. their outstanding OPL performance, the absorption maxima
The photophysical studies on 37–45 indicate that most of for all of them are located below 400 nm (Fig. 17), indicating
them exhibit triplet emissions even at rt with long PL lifetimes. their attractive transparency feature in the visible-light region.
So, it is the triplet absorption that induces their OPL for the ns Although there would be a negative effect to the transparency
laser pulse. However, the drastic difference in their OPL by extending the conjugation length and employing the
behavior should originate from the different triplet state electron-accepting organic ligands, the OPL performance can
character induced by the molecular symmetry of the metal be easily tuned to approach the requirement expected for
group as well as the electronic influence of the ligand type. The practical applications. Hence, the excellent optical limiting/
results of the time-dependent density functional theory transparency trade-off optimization can also be accomplished
(TDDFT) calculations show that the complexes with the by carefully controlling the molecular symmetry of PtII
D–Pt–D structure (37 and 38) give triplet states with LMCT acetylides as well as the electronic structures of the organic

2550 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

Fig. 20 Chemical structure of the fluorene-based AuI acetylide 46.


Adapted from ref. 85.
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 21 Comparison of the OPL properties of 46 with its PtII


counterpart 35 and C60 at To = 92%. Adapted from ref. 85.

3.1.3 Group 12 HgII acetylides. The d10 AuI and HgII


centers are isolobal and isoelectronic to each other.87 So,
group 12 HgII acetylides exhibit rich photophysics and were
also developed for studying the OPL phenomenon. By
introducing HgII ions into the organic backbone, triplet
emissions were also detected from HgII acetylides.87a–c
According to the RSA mechanism, HgII acetylides can display
respectable OPL behavior, which is confirmed by the recently
reported fluorene-based HgII acetylides (Fig. 22).85 The OPL
investigation for the 10 ns laser pulse at 532 nm revealed good
responses from 47 and 48. By extending the conjugation
Fig. 19 (a) Normalized transmittance versus incident fluence length, 48 can even show better OPL performance than that
of the symmetric PtII acetylides in CH2Cl2 at To = 92%. (b) and of C60 at To = 92%, signaling the merit of these types of
(c) Comparison of the OPL performances of the asymmetric molecules as effective OPL materials (Fig. 23). Together with
complexes with their mother complexes at To = 92%. Adapted from the nice OPL behavior, their outstanding transparency
ref. 86.
(lmax = 346 nm and lcut-off = 355 nm for 47, lmax = 347 nm
ligands, which would provide another avenue to give rise to and lcut-off = 359 nm for 48) would render them very
high performance OPL materials. competitive candidates for highly transparent OPL materials.

3.1.2 Group 11 AuI acetylides. Owing to their interesting 3.1.4 Mixed-metal acetylides. While most of the reported
optoelectronic properties, group 11 AuI acetylides represent transition metal acetylides for OPL application are mono-
another research point in transition metal alkynyl chemistry.62g–l metallic in nature, a series of novel rigid-rod metal acetylides
Some fluorene-based AuI acetylide complexes were also adopted with heterometallic centers, 49–52, have been designed and
for OPL investigations.85 Encouragingly, AuI acetylide 46 synthesized (Fig. 24).85 Derived from their mother complexes
(Fig. 20) shows comparable OPL performance to that of C60 36 and 53, two groups of AuI acetylide complexes with either a
(To = 84%) at even higher To = 92% for a 10 ns laser pulse at PtII or HgII center have been developed. Their Z-scan curves
532 nm (Fig. 21). Although 46 shows a weaker OPL response as show that the heteronuclear complexes show similar OPL
compared to its PtI1 counterpart 35 due to the weaker SOC effect behavior to their parent congeners (Fig. 25) and so their
induced by the AuI ion, its excellent transparency (lmax = 361 nm OPL performance depends mainly on the identity of the
and lcut-off = 372 nm) in the visible regime would make AuI central metal ion. As mentioned in Section 3.1.1.3, the
acetylides another class of high performance OPL materials. chelating environment for the PtII ion at the center of two
For its good transparency window, the sex/so value for 46 can organic ligands will favor the SOC effect as compared to the
reach ca. 14.52, about four times larger than that of C60 under end-capped configuration at the termini. Accordingly,
the same experimental conditions. complex 49 shows a much higher FT than that of 51, which

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2551
View Article Online

Fig. 22 Chemical structures of fluorene-based HgII acetylides 47 and 48. Adapted from ref. 85.

can achieve very good optical limiting/transparency trade-off


optimization. The poorer optical transparency for 56–59
induced by the electron-donating and electron-withdrawing
groups in the organic motifs clearly shows that the selection of
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

a suitable aromatic spacer is critical in designing promising


materials and the classical way of enhancing D–A interaction
in the molecular backbone is not sufficient to obtain the
optimized optical limiting/optical transparency trade-off.
Compared with their small-molecular counterparts described
in Section 3.1.1.1, the polymetallaynes generally show stronger
Downloaded by UNSW Library on 17 March 2013

OPL responses, thanks to the conjugation effect across the


backbone of the PtII polyynes induced by interaction of their
Fig. 23 Comparison of the OPL properties of 47 and 48 with C60 at d-orbitals with the p-orbitals of the conjugated ligands
To = 92%. Adapted from ref. 85. (Fig. 28).74 Together with their inherent good film-forming
nature for device fabrication, the polymeric metallaynes
is also reflected by the almost ten times higher FF for the latter clearly indicate their advantages in developing new generation
complex (ca. 3.0% for 51) than that for the former one OPL materials.
(ca. 0.4% for 49). Based on the RSA mechanism for triplet To examine the effect of metallation on the OPL behavior,
absorption, complex 49 shows a stronger OPL response the performance of some purely organic polyynes 60–62 has
(Fig. 25). Because of the similar reason, complexes 50 and also been characterized (Fig. 29) and compared with that of
52 with terminal AuI ions produce similar OPL responses to their corresponding PtII polyynes (54–56).74 The input–output
their mother complexes 36 and 53, respectively. The photo- fluence curves show that 54–56 exhibit noticeably stronger
physical studies also indicate that introduction of terminal OPL responses than their organic counterparts 60–62
transition metal ions would not enhance the SOC effect (Fig. 30), despite the fact that the organic polyynes possess
significantly. As a result, they all show nearly identical OPL longer conjugation length and red-shifted absorption maximum.
results to their parent mononuclear metal complexes. Different from the platinated analogues, the organic polymers
60–62 give no triplet emission even at 77 K. So, it can be
3.2 Macromolecular structures concluded that addition of PtII ions to increase FT by the SOC
Apart from the small-molecular compounds, their macro- effect would represent a feasible way to enhance the OPL
molecular counterparts with good film-forming properties response of conjugated polymers, while the strategy by solely
have also been constructed for OPL investigations. The increasing the conjugation length cannot help in solving
outstanding performance of the polymeric structures observed the optical limiting/transparency trade-off optimization to
would definitely render them attractive OPL candidates. The satisfaction.
recent progress along this direction will be highlighted in this Inspired by the admirable optical limiting/transparency
section. trade-off optimization achieved in 55, a series of fluorene-
based transition metal polyynes with single or dual metal ions
3.2.1 Linear rigid-rod polymetallaynes of PtII, PdII and 63–68 (Fig. 31) were developed so that the effect of inherent
HgII. Recently, we have studied the OPL characterization of electronic character and different combinations of metal ions
a series of polymetallaynes of PtII bearing different aromatic on their OPL behavior can be identified.85 Fig. 32 depicts their
diacetylides 54–59 in the polyyne backbone under the 10 ns OPL performance for a 10 ns laser pulse at 532 nm. For the
laser pulse at 532 nm (Fig. 26).74 Their photophysical data polyynes with only one type of metal ion in the backbone
reveal that the OPL effects should be induced via the RSA (63, 64 and 65), their OPL characteristics are substantially
mechanism of either ligand-centered (54 and 55) or intra- affected by the nature of the metal ion. Polyynes 63 and 64
molecular charge transfer (ICT) triplet absorption (56–59). outperform C60 in OPL response, while 65 shows a much
From the Z-scan data at To = 82% (Fig. 27), all these weaker OPL effect (i.e. 63 > 64 > C60 >> 65). Owing to the
polymetallaynes show very decent OPL effects which are higher SOC effect of PtII ions as indicated by both its low FF
comparable to (for 54, 56 and 58) or even better than (for (or high FT) and the TDDFT data, polyyne 63 shows
55, 57 and 59) that of a C60 solution. Different from 57 and 59, obviously stronger OPL effect induced by the triplet RSA
the excellent OPL performance plus the impressive trans- process when compared with that of 64 and 65. The detectable
parency window (with lmax o 400 nm) of 55 confirm that it triplet emission of 63 at rt also provides a good support to its

2552 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 24 Chemical structures of the heterometallic acetylides 49–52 and the mother complex 53. Adapted from ref. 85.

outperformance. The ICT character of triplet states associated


with 64 is responsible for its decent OPL performance which is
much better than that of 65. Given the same organic linker, the
general OPL performance order of metal effect is PtII > HgII
c PdII. The results give us precious information for the choice
of metal ions to construct high performance OPL materials.
Regarding the transparency issue, the fluorene-based metal
polyynes of PtII and HgII (lmax = 399 nm, lcut-off = 417 nm
for 63 and lmax = 355 nm, lcut-off = 370 nm for 64) are also
preferred over their PdII congener 65 (lmax = 412 nm and
lcut-off = 442 nm).
By virtue of the fact that a fluorene-based HgII polyyne
Fig. 25 Z-scan curves for the heterometallic acetylides and their exhibits the best transparency whereas a PtII congener shows
mother complexes at To = 92%. Adapted from ref. 85. the strongest OPL response, a good balance of these two

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2553
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 26 Chemical structures of some PtII polymetallaynes bearing different aromatic diacetylide ligands. Adapted from ref. 74.

lowering FF and increasing FT as facilitated by the strong


SOC ability of PtII ions. The ability of HgII ions for improving
the optical transparency is illustrated clearly by the hypso-
chromic effect in lmax and lcut-off wavelengths for 66
(lmax = 386 nm and lcut-off = 411 nm) and 68 (lmax = 411 nm
and lcut-off = 435 nm) compared with those of 63 and 65,
respectively. Furthermore, introduction of HgII ions into the
backbone of a polyyne to form a heterometallic polymer chain
would not influence the SOC effect, which has been supported
by both photophysical and computational studies. As a result,
heterometallic polyynes containing HgII ions (66 and 68) can
show comparable or even better OPL effect with respect to
that of the non-mercury congeners (63 and 65) based on the
Fig. 27 Z-scan curves for the PtII polymetallaynes bearing different triplet RSA mechanism for a ns laser pulse. Taking the optical
aromatic diacetylide ligands as compared with that of a C60 solution at limiting/transparency trade-off optimization into consideration,
To = 82%. Adapted from ref. 74. the use of PdII ions was not encouraged to prepare hetero-
metallic polyynes.
conflicting issues is important. Hence, the preparation of From the results aforementioned, the transition metal ions
heterometallic polyynes with two types of metal centers be- can play a very critical role in affecting the OPL behavior of
came another hot focus in this research.85 For the novel the organometallic polyyne. Accordingly, PtII and HgII
mixed-metal polyynes 66–68, it was found that copolymeriza- polyynes with different metal ion contents were also prepared
tion of metal polyynes with PdII ions (67 and 68) tends to (Fig. 33).85 Their Z-scan results listed in Fig. 34 indicate that
lower the OPL activity with respect to those of their mother the OPL effect is enhanced in the presence of metal ions on
metal polyynes (63 and 64) for a 10 ns laser pulse at 532 nm, going from the purely organic polyyne 71 to the metal
while combination with PtII ions (66 FF = 0.5% and 67 polyynes 63 and 64. With reference to 63 and 64, the metal
FF = 0.6%) would generally strengthen the OPL response polyynes with lower metal ion content (69 and 70) possess
relative to those of the corresponding polyynes with homo- organic spacers with longer conjugation length. It is well
metallic centers (64 FF = 2.5% and 65 FF = 9.9%) by known that the extent of penetration of the electrons of the

2554 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 28 The comparison of optical limiting properties between the PtII-based polymetallaynes with different aromatic diacetylide ligands and their
corresponding small-molecular counterparts at To = 82%. Adapted from ref. 74.

Fig. 29 Chemical structures of some organic polyynes. Adapted from ref. 74.

Fig. 30 Comparison of the OPL performance between PtII polyynes and their purely organic counterparts. Adapted from ref. 74.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2555
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 31 Chemical structures for the fluorene-based homo- and heterometallic polyynes. Adapted from ref. 85.

organic molecular skeleton into the field gradient of the a ns laser pulse, the lower FT basically impairs the contribution
metal ion substantially affects SOC interaction. As the ligand of nonlinear triplet absorption to the OPL behavior of 69 and
conjugation length increases, the S0 - S1 transition is more 70, resulting in their inferior OPL response (Fig. 34). Also, the
localized on the conjugated organic ligand, taking on more p organic polyyne 71 shows comparable OPL performance to
- p* character, and therefore spatially away from the metal that of 69 and 70, which mainly originates from the
center, thus reducing SOC and the FT.65 This has been conjugation effect. In terms of both the transparency and
confirmed by the much higher FF values for 69 and 70 OPL issues, 71 with lmax = 419 nm is inferior to 63 and 64
(1.6% and 59.6%) compared with that for 63 and 64 (0.6% (lmax = 399 and 355 nm, respectively).
and 2.5%). The weakened SOC effect and reduced FT were Experimental data suggest that the improvement of trans-
also verified by the observation that the triplet emission peak parency window by HgII ions is caused by their conjugation-
for both 69 and 70 is absent at rt but that even disappears at breaking effect. This then results in a higher FT that is in
77 K for 70. In accordance with the triplet RSA mechanism for agreement with the energy gap law for the triplet states in

2556 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

Fig. 35 Chemical structure of 72. Adapted from ref. 85.

metal polyynes.88 This is also revealed by the lower FF for 66


than that of its parent PtII polyyne 63. Pertaining to the triplet
Fig. 32 Normalized transmittance of homo- and heterometallic RSA mechanism, it was found that the positive contribution
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

polymetallaynes and C60 versus input fluence at the same linear from the increased FT of 66 to the OPL process can nearly
transmittance (To = 92%). Adapted from ref. 85 compensate for the loss in OPL due to interruption of the
conjugation by HgII ions, and hence both 63 and 66 can give
similar OPL responses (Fig. 32).85 Yet, it is important that 66
has an improved transparency window for optimizing the
properties of OPL materials of this kind (lmax = 386 nm for
Downloaded by UNSW Library on 17 March 2013

66 versus 399 nm for 63). Apparently, such a conjugation-


breaking strategy would exert a negative side effect to weaken
the OPL response if the lost contribution from conjugation
effect cannot be compensated. So, it is worthwhile to study the
effect of conjugation length on OPL. It was shown that the
conjugation factor will become less important as the chain
length is increased up to the trimer for the HgII, PtII and their
mixed HgII/PtII systems (Fig. 35 and 36).85 The trimers 48 and
72 have shown almost identical OPL responses to their
corresponding polymeric counterparts 64 and 63, respectively.
Trimetallic complex 49 can also show nearly the same OPL
ability as that of 66. In other words, there is a certain chain
length limit for the conjugation effect to contribute to the OPL
performance which, on the other hand, might be compensated
by other effects induced by the conjugation-breaking factor.
So, a cis-configured PtII polyyne 73 (Fig. 37) was designed to
optimize further the optical limiting/transparency trade-off.89
Gratifyingly, the white-colored polymer 73 can give outstanding
Fig. 33 Chemical structures of the fluorene-based metal polyynes OPL activity analogous to that of 63 (Fig. 38a), and at the
with different metal ion contents in the backbone. Adapted from same time exhibit very attractive transparency (lmax = 364 nm
ref. 85. and lcut-off = 406 nm). The marked hypsochromic effect in its

Fig. 34 The effect of metal ion content on the OPL performance of the fluorene-based metal polyynes. Adapted from ref. 85.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2557
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

Fig. 36 The effect of conjugation length on the OPL performance of fluorene-based organometallic acetylide compounds. Adapted from ref. 85.
Downloaded by UNSW Library on 17 March 2013

Fig. 37 Chemical structure of the cis-configured polymetallayne 73.


Adapted from ref. 89.

lmax indicates that the cis-configuration does break the


conjugation path in 73. However, this conjugation-breaking
segment dramatically increases the FT, as inferred by its
much stronger triplet emission (Fig. 38b) as well as the
much lower FF (0.1%) with respect to that of 63 (0.6%).
Based on the triplet RSA mechanism for a ns laser pulse,
the loss of contribution from conjugation effect to the
OPL has been successfully compensated by the increased
FT in 73. Therefore, 73 provides a good example of obtaining
excellent optical limiting/transparency trade-off optimization
via tuning of the molecular configuration around the metal
center. Fig. 38 (a) Comparison of the OPL performance of some fluorene-
Not only metal polyynes 63 and 73, but also 64 and 66 can based metal polyynes with the benchmark dyes. (b) PL spectra for 63
reveal stronger OPL responses than those of some best ever and 73 in CH2Cl2 at rt. Adapted from ref. 89.
OPL materials, such as C60, zinc(II)tetraphenylporphyrin
(Zn–TPP) and CuPc–tBu4, even at higher To. Apart from 3.2.2 Dendrimers with PtII acetylide building blocks. Owing
their impressive OPL performance, the good optical trans- to their well-defined, unique macromolecular structure com-
parency is apparent from their light colors (Fig. 39) and pared with the linear analogues, dendrimers are attractive
UV-vis spectra (Fig. 40). The polymers 63, 64, 66 and 73 are scaffolds for a variety of high-end applications due to their
all off-white solids and can get rid of the dragged absorption highly branched and globular architecture.90 For example,
tail behind 400 nm. This is in complete contrast with C60, dendrimers can demonstrate significantly increased solubility
Zn–TPP and CuPc–tBu4 showing strong visible absorption and they also exhibit very low intrinsic viscosities.91 Remarkably,
bands, and therefore their poor optical transparencies would they can encapsulate functional molecules into their spherical
make them less competitive with the above four metallo- architecture to lead to encapsulation effect,92 which enables a
polyynes as optical limiters (Fig. 39),89 especially for the number of different applications, such as shape selective
protection of human eyes. Considering both the outstanding catalysis and anti-quenching. Furthermore, the large number
transparency and OPL activity of these metallated systems, we of capping groups would give rise to a high flexibility in
can easily devise very effective strategies to realize excellent tailoring the properties of the dendrimers,90 such as hydro-
optical limiting/transparency trade-off optimization for next phobicity and cross-linkability, etc. So, dendrimer research
generation OPL materials. has flourished rapidly in recent years.90

2558 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

Fig. 39 A photograph showing the light colors of polymetallaynes with some benchmark deeply colored OPL materials. Adapted from ref. 85
and 89.

4. Research endeavor for practical devices


Until now, most of the OPL materials were studied in the
Downloaded by UNSW Library on 17 March 2013

solution state for the evaluation of their nonlinear absorption


of light. However, materials in solutions are not desirable for
real applications. In order to develop practical protective
devices conveniently, it is preferable that the OPL materials
can be directly applied in the solid state, which is easy to
process and handle, so that the device fabrication process can
be simplified.3,94 For the time being, there are two common
strategies to prepare solid-state OPL materials either through
physically dispersing (class I) or chemically bonding (class II)
Fig. 40 Comparison of the optical transparency of some fluorene-
the OPL chromophores with a suitable matrix.95,96 Special
based metal polyynes with three traditional OPL materials
care should be taken to avoid the aggregation of OPL chromo-
(C60, Zn–TPP and CuPc–tBu4). Adapted from ref. 85 and 89.
phores (especially for class I) during the fabrication process for
preparing both classes of materials, since the aggregation
Malmström et al. had successfully developed a series of would impair the transparency as well as OPL performance
novel dendrimer-encapsulated PtII acetylides up to the fourth of the solid materials. Recently, research endeavors for practical
generation (Fig. 41) including 13, 74, 75 and 76.93 The devices also involve the use of organometallic acetylides.
OPL investigations for a ns laser pulse together with their With the aim to provide high performance solid OPL
photophysical studies have shown that these dendritic PtII materials, Malmström et al. had prepared a series of PtII
acetylides are effective broadband OPL materials exhibiting acetylides end-capped with either bis-MPA (10, 12 and 13)
the RSA mechanism. Interestingly, the OPL response is or methacrylate (77) groups (Fig. 42).97 The bis-MPA moieties
generally enhanced with increasing the size of the dendritic afford both good solubility and high compatibility with a
substituent from the first (G1) to the fourth generation (G4) poly(methyl methacrylate) (PMMA) matrix to give high
and they all show better OPL performance than that of their optical transparency. The methacrylate groups of 77 provide
mother complex 2 without the end group. The clamping levels polymerizable sites with methyl methacrylate (MMA) to bond
under the same input energy for a 532 nm laser pulse are 9.0, the OPL chromophore 77 into the PMMA matrix. All of the
7.6, 7.1, 6.3 and 6.6 mJ for 2, 13 (G1), 74 (G2), 75 (G3) and 76 OPL chromophores 10, 12 and 13 have been adopted to
(G4), respectively (Fig. 41). The values for 580 nm laser prepare class I materials by mixing the corresponding
are 9.5, 7.5, 7.2, 6.3 and 6.2 mJ for 2, 13, 74, 75 and 76, chromophore with MMA and then the polymerization process
respectively, whereas clamping levels of 18.4, 12.7, 11.9, was initiated by azobis(isobutyronitrile) (AIBN) and allowed
10.2 and 9.9 mJ, respectively, were observed for 630 nm to proceed at 50 1C for three days. Class II OPL materials were
laser. The advantage of using dendrimers as OPL materials prepared in a similar way to that of class I system except that
is related to their bulky dendritic globular architecture that 77 acted as a guest to provide cross-linkable sites with MMA
could block the triplet–triplet annihilation via site-isolation to covalently bond itself into the PMMA matrix. All the doped
effect as well as the quenching of triplet states from oxygen. solid OPL materials at a doping concentration of 50 mM show
Furthermore, all of these dendronized PtII acetylides exhibit very good transparency (To close to 90%) at wavelengths
very attractive transparency with lcut-off o 400 nm. These beyond 550 nm. If PMMA was taken as the background
results give us a hint that dendron-decorated PtII acetylides reference, To can be in the range from 98% to 99%. The class
represent a good alternative to be developed as high I solids doped with 12 and 13 together with a 77-doped class II
performance OPL materials to meet the necessary practical system were employed to evaluate their merit in OPL for a ns
requirements. laser pulse at 532 nm. Generally, the OPL response of the solid

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2559
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 41 Chemical structures of some dendronized PtII acetylides. Adapted from ref. 93.

Fig. 42 Chemical structure of 77. Adapted from ref. 97.

samples was strengthened with an increase of the PtII-acetylide chromophores in the class II solid, thus restricting structural
doping level within the solids. Regarding their OPL performance, rearrangement and relaxation. The cross-linked class II solid
the class I solids doped with 12 and 13 are almost the same also shows a lower laser damaging threshold. Hence, class I
by showing very similar clamping levels (3.9 and 3.5 mJ, solids are generally more preferred as the practical fabrication
respectively) at an incident energy level of 115 mJ. On the method to design solid OPL materials.
contrary, the 77-doped class II solid gives noticeably weaker By introducing two acrylate groups as the chemical handles
OPL response with a clamping level of 8.5 mJ at an incident to the PtII acetylide moiety, OPL chromophore 78 can be
energy of 90 mJ. By comparing with the result from the readily copolymerized with styrene through such chemical
solid doped with 13, class I materials are superior to their handles to prepare a class II OPL glass using organic
corresponding class II congeners, which has been ascribed polystyrene as the matrix (Fig. 43).98 In the same report, the
to the more pronounced rigidity experienced by the OPL authors have discussed the use of PtII acetylides containing

2560 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

Parola et al. have prepared a series of urethane–propyl-


triethoxysilane-functionalized PtII acetylides 82 and 83
(Fig. 45) and successfully incorporated them into a silica-based
system by the sol–gel process to provide some solid hybrid
OPL materials.99 This sol–gel process involves steps like acidic
hydrolysis and polycondensation of methyltriethoxysilane
(MTEOS), and the functionalized PtII acetylides then dissolve
in THF, followed by gel aging and drying to form a so-called
xerogel (i.e. a silica-based hybrid OPL glass). The procedures
used for the hydrolysis of the OPL chromophore (i.e. in situ or
ex situ hydrolysis) and the pH conditions are very critical
factors to affect the final properties of the solid hybrid
materials. The most effective way to prepare the hybrid
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

OPL glass is as follows: the starting precursor MTEOS was


hydrolyzed in an acidic medium (pH = 2.5) in the presence of
ethanol. The mixture was stirred overnight and evaporated to
the initial volume of the precursor. 3-Aminopropyltriethoxy-
silane (APTES) solution in ethanol was then added to neu-
tralize the mixture. Then, the functionalized PtII acetylides
Downloaded by UNSW Library on 17 March 2013

were added to the mixture as a solution in THF at the desired


concentration. Addition of dilute acidic water (pH = 2.6) and
stirring of the mixture for 7 days then initialized the hydrolysis.
The condensation of the PtII acetylide chromophore with the
silica-based matrix would prevent precipitation of the OPL
chromophore during the drying step. After aging, the gels were
dried between 70 and 110 1C in PTFE moulds to afford the
desired inorganic/organic hybrid glass for OPL investigations.
Fig. 43 A proposed method to prepare class II OPL glass with The OPL results for the hybrid glass with 82 at 120 mM
polystyrene as the matrix. Adapted from ref. 98. concentration show that the solid materials can serve as
effective broadband optical limiters with low clamping levels
of 0.2, 3.0, 4.5 and 7.0 mJ at 480, 532, 580 and 630 nm,
triazole (complex 12) and bis-MPA units (complex 13) in OPL respectively. The hybrid glass of 82 even shows better OPL
study. The triazole groups give a positive contribution to the performance than that of the 30 mM THF solutions of
limiting abilities of the molecules due to their low clamping acetonide dendrimers 13, 74, 75 and 76 for laser pulses at
level and high transmission in the visible wavelength region 480, 532, 580 and 630 nm. Experimental data on 82 as solid
whereas the bis-MPA units can be advantageously used to also reveal its stronger OPL response than that of the 50 mM
prepare dendritic substituents offering site isolation to the THF solution of 10 and 11 or comparable OPL performance
chromophore leading to improved clamping. The bis-MPA to that of 50 mM THF solution of 12. More importantly, it
functionalization also improves the solubility of the PtII can outperform the class II solid doped with 13 for a
acetylides in organic solvents and caters for the increase of 532 nm-laser pulse and also possess a higher damage threshold.
the number of accessible end-groups to which the attachment So, the silica-based inorganic/organic hybrid glass doped with
of matrix-compatible species takes place. the PtII acetylide OPL chromophore could provide another
Through controlled radical polymerization techniques valuable avenue towards fabrication of simple solid-state
such as atom transfer radical polymerization (ATRP), the optical limiters. Further research is, however, necessary to
matrix polymer chain can be grafted onto the surface of the elucidate the effect of glass structure on the OPL properties.
dendronized OPL PtII acetylide by the ATRP-initiator (as a Besides the linear analogue, the dendronized PtII acetylide with
multifunctionalized initiator) functionalized on the terminal of proper siloxane groups on the surface has also been made so that
the dendrimer branches to produce another kind of class II the polar siloxane periphery would afford better compatibility to
organic glass, which forms a new platform for developing the non-polar OPL chromophore with the inorganic silicon
practical OPL devices (Fig. 44).98 alkoxide matrix. In this way, high quality class I material as well
Given the results aforementioned with the target to bring as class II glass (by cross-linking the dendronized PtII acetylide
about practical OPL devices, the nature of the matrix materials with the silicon alkoxide matrix) can be obtained (Fig. 46).98
was found to be very important. For an ideal matrix material, While it is conceived that PtII–acetylide-doped organic/
it should possess the features such as high optical quality inorganic glasses are the key to fabricating solid-state OPL
and high damage threshold. Obviously, an inorganic silica- devices, the preliminary results gathered so far definitely give
based system prepared by the well-established sol–gel us valuable insights to the development of practical optical
method at low temperature is promising since it combines limiting devices derived from high performance PtII acetylide
high optical quality with good thermal stability and OPL materials with potentially optimized optical limiting/
mechanical properties. Triggered by this idea, Eliasson and optical transparency trade-off features.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2561
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A
Downloaded by UNSW Library on 17 March 2013

Fig. 44 Preparation of class II OPL glass through ATRP with functionalized dendritic PtII acetylide as the OPL active chromophore. Adapted
from ref. 98.

5. Concluding remarks and future perspectives for practical applications in the up-and-coming future. While
the need for laser safety in the laboratory is still increasing, the
The use of organometallic acetylides as new generation OPL many examples covered in this contribution highlight the
materials has been studied in detail in the past decade and richness and diversity in such a research scenario. What should
relevant recent works by a couple of research groups allow a be emphasized here is that the introduction of heavy-metal
thorough understanding of the structure–activity–function ions such as PtII and HgII into the organic polyyne backbone
relationships of this kind of OPL chromophores, which would would be a good approach for contriving novel optical limiters
give us more inspirations to develop more advanced materials with improved optical nonlinearity/transparency trade-offs.

2562 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

Fig. 45 Chemical structures of some functionalized OPL PtII acetylides adopted for preparing inorganic silica-based class II OPL glass. Adapted
from ref. 99.
Downloaded by UNSW Library on 17 March 2013

Fig. 46 The dendron-decorated PtII acetylide as OPL active chromophore for making inorganic silica-based class II OPL glass. Adapted from
ref. 98.

Some of the acetylide complexes of PtII and HgII are amongst use. Furthermore, there is a broad scope for the design and
the most efficient metal-containing OPL molecules thus far, synthesis of organometallic materials to meet the grand challenges
which even exceed in performance than the best organic in OPL research and a number of workable strategies have
performers. been devised, for instance, by tuning the electronic structure of
It is well known that the absorption of visible light by any the key organic chromophore, by changing the molecular
OPL materials will disfavor the use of these compounds as geometry and symmetry of the metal complex, by using
optical limiters for the protection of human eyes or alike, different transition metal ions (in different combinations and
because this would cause some vision loss by blocking light of compositions), as well as by adopting the polymeric and
a certain wavelength the wearer still intends to see. Comparing dendronized chemical structures, etc. Reported results certainly
the OPL performance data of the metallaynes and common indicate that these solution-processable and highly transparent
OPL materials currently in use (see Table 1), their merits as di, oligo- and polymetallaynes with the optimum combination
RSA optical limiters for protecting eyes and optical sensors of metal and ligands are attractive OPL materials that would
become apparent. Most of them are essentially off-white have enormous potential to excel in the advancement of
powders with very good optical transparencies in the visible practical appliances in the near future.
region. Given their excellent transparency in the visible regime While it was shown that extending the conjugation beyond
and their unique photophysical behavior, these organometallic two repeat units may not have significant positive impact on
acetylides have already opened up a new door to the next- the OPL of small-molecular metallaynes, higher-molecular-
generation OPL materials. They can also favorably optimize weight polymeric analogues with improved solution processability
the OPL/transparency trade-off, which has bothered the are generally desirable. The preparation of solid-state optical
researchers for a long time in the field and has been taken as power limiters, where the NLO chromophore is inserted in
an unachievable goal for the benchmark materials currently in an optically transparent matrix, has also been shown to be

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2563
View Article Online

Table 1 Comparison of the OPL performance of various organome- Table 1 (continued )


tallic PtII acetylides with some traditional state-of-the-art materials
Ftha/
Ftha/ Compound J/cm2 Tob/% Lc/mm sex/sod Ref.
Compound J/cm2 Tob/% Lc/mm sex/sod Ref.
63 0.07 92 1 19.07 85
[Tri-(n-hexyl)siloxy]InPc 0.07 84 10 16 18 64 0.11 92 1 20.81 85
[Tri-(n-hexyl)siloxy]AlPc 0.26 84 10 10 18 65 0.81 92 1 3.20 85, 87
[Tri-(n-hexyl)siloxy]GaPc 0.12 84 10 14 18 66 0.08 92 1 18.32 85, 87
Bis[tri-(n-hexyl)siloxy]SnPc 0.08 84 10 18 18 67 0.35 92 1 3.90 85, 87
PbPc(b-CP)4 0.07 62 — — 100 68 0.75 92 1 3.40 85, 87
CuPcR8e 0.30 68 10 4.18 101 69 0.14 92 1 9.76 85, 87
(R)-TMBO–CuPcf 0.3 76 1 5 102 70 0.19 92 1 13.72 85, 87
CuPc–tBu4 0.10 — — — 103 72 0.07 92 1 17.91 85, 87
CuPc–tBu4 0.13 86 1 6.20 85 73 0.08 92 1 18.62 85, 87
Br8[OSi(Hex)3]2SiNPc >1.0 80 2 2.68 104 a
Optical-limiting threshold is defined as the input light fluence at
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

[a-(2-yloxybenzoic acid)]4ZnPc 0.6 90 2 12.79 105


C60 0.18 55 2 — 13 which the output light fluence is 50%. b Linear transmittance To.
c
C60 0.19 84 1 3.89 85 Sample thickness. d sex/so = lnTsat/lnTo, where sex is the effective
2 0.9 — 2 — 70 excited-state absorption cross-section and so is the ground-state
4 — 82 1 7.84 74 absorption cross-section, Tsat is the transmittance at the saturation
5 — 82 1 11.56 74 fluence. e R = pentyloxy. f (R)-Tetrakis(2-methoxy-1,10 -binaphthalen-
6 — 82 1 10.81 74 20 -oxy)copper phthalocyanine. g The values refer to the input light
7 — 82 1 11.75 74 fluence at which the output light fluence is 90%.
Downloaded by UNSW Library on 17 March 2013

8 — 82 1 6.11 74
9 — 82 1 7.73 74
10 — 91 2 34.13 75
11 — 69 2 9.68 75 successful recently and is worthy of future examination for
12 — 80 2 19.20 75 realizing devices in practical terms. High performance organo-
14 0.048g 70 2 3.57 79 metallic optical power limiters with different chemical structures
15 0.14g 70 2 3.02 79 can be incorporated into both organic and inorganic matrices
16 1.09g 70 2 2.30 79
17 0.25g 70 2 3.70 80 by either physical mixing (e.g. by embedding) or chemical
18 0.37g 70 2 3.70 80 bonding (e.g. by covalent bonding) so that solid glasses can be
19 0.49g 70 2 3.20 80 prepared for real applications. Furthermore, it would be
20 0.052g 70 2 4.80 80
21 >1.0 75 2 o2.77 81
of much academic and industrial interest if the synthetic
22 >1.0 75 2 2.77 81 approach can be further extended to the preparation of solid
23 >1.0 75 2 2.77 81 functional materials for OPL applications in the range of
24 >1.0 75 2 2.77 81 telecommunication wavelengths (i.e. infrared region).106
25 >1.0 75 2 2.08 81
26 >2.0 70 2 1.58 82
27 >2.0 70 2 1.29 82
28 >2.0 70 2 1.17 82 Acknowledgements
29 0.1 90 2 7.79 83
30 >2.0 90 2 6.39 83 W.-Y.W. sincerely thanks all postgraduate students, post-
31 0.8 80 2 5.55 84 doctoral associates and collaborators whose names appear in
32 0.9 80 2 5.11 84
33 1.0 80 2 5.11 84 the references. This work was supported by a grant from the
34 >1.0 80 2 2.68 84 University Grants Committee of HKSAR, China (Project No.
35 0.10 92 1 14.52 85 [AoE/P-03/08]), Hong Kong Baptist University (FRG2/09-10/91)
36 0.12 92 1 11.85 85 and the Croucher Senior Research Fellowship from the
37 0.10 92 1 13.0 86
38 0.05 92 1 14.0 86 Croucher Foundation. G.-J. Zhou acknowledges the support
39 0.15 92 1 12.0 86 from Xi’an Jiaotong University (No. 08140004), the Program
40 0.19 92 1 17.0 86 for New Century Excellent Talents in University from the
41 0.68 92 1 9.0 86
42 0.12 92 1 12.0 86
Ministry of Education of China, a Research Grant from
43 0.09 92 1 11.0 86 Shaanxi Province (No. 2009JQ2008), and the National
44 0.33 92 1 7.0 86 Natural Science Foundation of China (No. 20902072).
45 0.12 92 1 11.0 86
46 0.20 92 1 9.76 85
47 0.31 92 1 22.48 85
48 0.13 92 1 17.27 85 References
49 0.13 92 1 10.68 85
1 L. W. Tutt and T. F. Boggess, Prog. Quantum Electron., 1993, 17,
50 0.11 92 1 8.26 85
299.
51 0.28 92 1 9.15 85 2 D. G. McLean, R. L. Sutherland, M. C. Brant, D. M. Brandelick,
52 0.27 92 1 10.00 85 P. A. Fleitz and T. Pottenger, Opt. Lett., 1993, 18, 858.
53 0.26 92 1 14.14 85 3 C. W. Spangler, J. Mater. Chem., 1999, 9, 2013.
54 — 82 1 6.39 74
4 T. J. Mckay, J. Staromlynska, P. Wilson and J. Davy, J. Appl.
55 — 82 1 12.99 74
Phys., 1999, 85, 1337.
56 — 82 1 7.63 74
5 C. R. Giuliano and L. D. Hess, IEEE J. Quantum Electron., 1967,
57 — 82 1 9.53 74
3, 358.
58 — 82 1 5.51 74 6 M. M. Fisher, B. Veyret and K. Weiss, Chem. Phys. Lett., 1974,
59 — 82 1 10.90 74 28, 60.

2564 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011
View Article Online

7 J. W. Perry, in Nonlinear Optics of Organic Molecules and 45 S. Yin, H. Xu, X. Su, Y. Gao, Y. Song, J. W. Y. Lam, B. Z. Tang
Polymers, ed. H. S. Nalwa and S. Miyata, CRC Press, Boca and W. Shi, Polymer, 2005, 46, 10592.
Raton, 1997, pp. 813–840. 46 S. Yin, H. Xu, X. Su, L. Wu, Y. Song and B. Z. Tang, Dyes
8 E. W. Van Stryland, D. J. Hagan, T. Xia and A. A. Said, in Pigm., 2007, 75, 675.
Nonlinear Optics of Organic Molecules and Polymers, ed. 47 S. Yin, H. Xu, M. Fang, W. Shi, Y. Gao and Y. Song, Macromol.
H. S. Nalwa and S. Miyata, CRC Press, Boca Raton, 1997, Chem. Phys., 2005, 206, 1549.
pp. 841–860. 48 M. Sheik-Bahae, A. A. Said and E. W. Van Stryland, Opt. Lett.,
9 R. L. Sutherland, Handbook of Nonlinear Optics, Marcel Dekker, 1989, 14, 955.
New York, 1996. 49 M. Sheik-Bahae, A. A. Said, T. Wei, D. J. Hagan and E. W. Van
10 L. W. Tutt and A. Kost, Nature, 1992, 356, 225. Stryland, IEEE J. Quantum Electron., 1990, 26, 760.
11 F. Henari, J. Callaghan, H. Stiel, W. Blau and D. J. Cardun, 50 N. J. Long, Angew. Chem., Int. Ed. Engl., 1995, 34, 21.
Chem. Phys. Lett., 1992, 199, 144. 51 L. T. Cheng, W. Tam, S. H. Stevenson, G. R. Meredith,
12 D. J. McLean, R. L. Sutherland, M. C. Brant, D. M. Brandelik, G. Rikken and S. R. Marder, J. Phys. Chem., 1991, 95,
P. A. Fleitz and T. Pottenger, Opt. Lett., 1993, 18, 858. 10631.
13 Y.-P. Sun, J. E. Riggs and B. Liu, Chem. Mater., 1997, 9, 1268. 52 J. W. Perry, K. Mansour, S. R. Marder, C. T. Chen, P. Miles,
14 Y.-P. Sun, G. E. Lawson and J. E. Riggs, J. Phys. Chem. A, 1998, M. E. Kenney and G. Kwag, Mater. Res. Soc. Symp. Proc., 1995,
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

102, 5520. 374, 257.


15 J. E. Riggs and Y. P. Sun, J. Phys. Chem. A, 1999, 103, 485. 53 C. E. Powell and M. G. Humphrey, Coord. Chem. Rev., 2004, 248,
16 D. R. Coulter, V. M. Miskowski, J. W. Perry, T. H. Wei, 725.
E. W. Van Stryland and D. J. Hagan, Proc. SPIE-Int. Soc. Opt. 54 Y. Fujikura, K. Sonogashira and N. Hagihara, Chem. Lett., 1975,
Eng., 1989, 1105, 42. 1067.
17 T. H. Wie, D. J. Hagan, M. J. Sence, E. W. van Stryland, 55 K. Sonogashira, S. Takahashi and N. Hagihara, Macromolecules,
J. W. Perry and D. R. Coulter, Appl. Phys. B: Photophys. Laser 1977, 10, 879.
Chem., 1992, 54, 46. 56 S. Takahashi, M. Kariya, T. Yatake, K. Sonogashira and
Downloaded by UNSW Library on 17 March 2013

18 J. W. Perry, K. Mansour, S. R. Marder, K. J. Perry, D. Jr. N. Hagihara, Macromolecules, 1978, 11, 1063.
Alvarez and I. Choong, Opt. Lett., 1994, 19, 625. 57 K. Sonogashira, S. Kataoka, S. Takahashi and N. Hagihara,
19 B. D. Rihter, M. E. Kenney, W. E. Ford and M. A. J. Rogers, J. Organomet. Chem., 1978, 160, 319.
J. Am. Chem. Soc., 1993, 115, 8146. 58 S. J. Davies, B. F. G. Johnson, M. S. Khan and J. Lewis, J. Chem.
20 M. Hanack, T. Schneider, M. Barthel, J. S. Shirk, S. R. Flom and Soc., Chem. Commun., 1991, 187.
R. G. S. Pong, Coord. Chem. Rev., 2001, 219, 235. 59 Z. Atherton, C. W. Faulkner, S. L. Ingham, A. K. Kakkar,
21 Y. Li, T. M. Pritchett, J. Huang, M. Ke, P. Shao and W. Sun, M. S. Khan, J. Lewis, N. J. Long and P. R. Raithby,
J. Phys. Chem. A, 2008, 112, 7200. J. Organomet. Chem., 1993, 462, 265.
22 W. Blau, H. Byrne, W. M. Dennis and J. M. Kelly, Opt. 60 B. F. G. Johnson, A. K. Kakkar, M. S. Khan and J. Lewis,
Commun., 1985, 56, 25. J. Organomet. Chem., 1991, 409, C12.
23 W. Su and T. M. Cooper, Chem. Mater., 1998, 10, 1212. 61 S. J. Davies, B. F. G. Johnson, J. Lewis and P. R. Raithby,
24 K. McEwan, K. Lewis, G.-Y. Yang, L.-L. Chng, Y.-W. Lee, J. Organomet. Chem., 1991, 414, C51.
W.-P. Lau and K.-S. Lai, Adv. Funct. Mater., 2003, 13, 863. 62 (a) N. J. Long and C. K. Williams, Angew. Chem., Int. Ed., 2003,
25 S. Guha, K. Kang, P. Porter, J. F. Roach, D. E. Remy, 42, 2586; (b) J. Z. Liu, J. W. Y. Lam and B. Z. Tang, Chem. Rev.,
F. J. Aranda and D. Rao, Opt. Lett., 1992, 17, 264. 2009, 109, 5799; (c) U. H. F. Bunz, Chem. Rev., 2000, 100, 1605;
26 P. Zhu, C. Yu, J. Liu, Y. Song and C. Li, Proc. SPIE–Int. Soc. E. E. Silverman, T. Cardolaccia, X. Zhao, K. Y. Kim,
Opt. Eng., 1996, 2879, 289. K. Haskins-Glusac and K. S. Schanze, Coord. Chem. Rev.,
27 B. Z. Tang and H. Y. Xu, Macromolecules, 1999, 32, 2569. 2005, 249, 1491; (d) A. S. Abd-El-Aziz, P. O. Shipman,
28 L. Vivien, E. Anglaret, D. Riehl, F. Bacou, C. Journet, C. Goze, B. N. Boden and W. S. McNeil, Prog. Polym. Sci., 2010, 35,
M. Andrieux, M. Brunet, F. Lafonta, P. Bernier and F. Hache, 714; (e) R. P. Kingsborough and T. M. Swager, Prog. Inorg.
Chem. Phys. Lett., 1999, 307, 317. Chem., 1999, 48, 123; (f) C.-H. Tao and V. W.-W. Yam,
29 X. Sun, P. Chen, X. B. Wu, J. Y. Lin, W. Ji, R. Q. Xu and J. Photochem. Photobiol., C: Photochem. Rev., 2009, 10, 130;
S. A. Hor, Appl. Phys. Lett., 1998, 73, 3632. (g) R. J. Puddephatt, Chem. Commun., 1998, 1055;
30 M. Pittman, P. Plaza, M. M. Martin and Y. H. Meyer, Opt. (h) R. J. Puddephatt, Coord. Chem. Rev., 2001, 216–217, 313;
Commun., 1998, 158, 201. (i) V. W. W. Yam, K. K. W. Lo and K. M. C. Wong,
31 G. R. Allan, D. R. Labergerie, S. J. Rychnovsky, T. F. Boggess, J. Organomet. Chem., 1999, 578, 3; (j) H. Y. Chao, W. Lu,
A. L. Smirl and L. Tutt, J. Phys. Chem., 1992, 96, 6313. Y. Li, M. C. W. Chan, C. M. Che and K. K. Cheung, J. Am.
32 S. Shi, W. Ji, J. P. Lang and X. Q. Xin, J. Phys. Chem., 1994, 98, Chem. Soc., 2002, 124, 14696; (k) W.-Y. Wong, K.-H. Choi,
3570. G.-L. Lu, J.-X. Shi, P.-Y. Lai, S.-M. Chan and Z. Lin, Organo-
33 S. Shi, W. Ji and X. Q. Xin, J. Phys. Chem., 1995, 99, 894. metallics, 2001, 20, 5446; (l) P. Li, B. Ahrens, K.-H. Choi,
34 L. Tutt and S. W. McCahon, Opt. Lett., 1990, 15, 700. M. S. Khan, P. R. Raithby, P. J. Wilson and W.-Y. Wong,
35 S. Shi, W. Ji, S. H. Tang, J. P. Lang and X. Q. Lin, J. Am. Chem. CrystEngComm, 2002, 4, 405; (m) S. Szafert and J. A. Gladysz,
Soc., 1994, 116, 3615. Chem. Rev., 2003, 103, 4175; (n) L. Sudha, M. K. Al-Suti,
36 J. W. Perry, K. Mansour, I.-Y. S. Lee, X.-L. Wu, P. V. Bedworth, N. Zhang, S. J. Teat, L. Male, H. A. Sparkes, P. R. Raithby,
C.-T. Chen, D. Ng, S. R. Marder, P. Miles, T. Wada, M. Tian M. S. Khan and A. Köhler, Macromolecules, 2009, 24, 1131;
and H. Sasabe, Science, 1996, 273, 1533. (o) T. Ren, Organometallics, 2005, 24, 4854.
37 T. M. Cooper, D. G. McLean and J. E. Rogers, Chem. Phys. 63 (a) W.-Y. Wong and C.-L. Lam, Coord. Chem. Rev., 2006, 250,
Lett., 2001, 349, 31. 2627; (b) W.-Y. Wong, J. Inorg. Organomet. Polym. Mater., 2005,
38 G. A. Kumar, J. Nonlinear Opt. Phys. Mater., 2003, 12, 367. 15, 197; (c) W.-Y. Wong, Comments Inorg. Chem., 2005, 26, 39;
39 A. Krivokapic, H. L. Anderson, G. Bourhill, R. Ives, S. Clark and (d) W.-Y. Wong, Dalton Trans., 2007, 4495; (e) W.-Y. Wong,
K. J. McEwan, Adv. Mater., 2001, 13, 652. Macromol. Chem. Phys., 2008, 209, 14; (f) W.-Y. Wong and
40 K. J. McEwan, G. Bourhill, J. M. Robertson and H. L. Anderson, P. D. Harvey, Macromol. Rapid Commun., 2010, 31, 671;
J. Nonlinear Opt. Phys. Mater., 2000, 9, 451. (g) W.-Y. Wong and C.-L. Ho, Acc. Chem. Res., 2010, 43, 1246.
41 J. S. Shirk, R. G. S. Pong, S. R. Flom, H. Heckmann and 64 T. M. Cooper, D. M. Krein, A. R. Burke, D. G. McLean,
M. Hanack, J. Phys. Chem. A, 2000, 104, 1438. J. E. Rogers, J. E. Slagle and P. A. Fleitz, J. Phys. Chem. A,
42 G. J. Zhou, S. Zhang, P. J. Wu and C. Ye, Chem. Phys. Lett., 2006, 110, 4369.
2002, 363, 610. 65 J. E. Rogers, T. M. Cooper, P. A. Fleitz, D. J. Glass and
43 G. J. Zhou, S. Zhang and C. Ye, J. Phys. Chem. B, 2004, 108, D. G. McLean, J. Phys. Chem. A, 2002, 106, 10108.
3985. 66 T. M. Cooper, B. C. Hall, D. G. McLean, J. E. Rogers,
44 G. J. Zhou, Y. Q. Liu and C. Ye, J. Appl. Polym. Sci., 2004, 93, A. R. Burke, K. Turnbull, A. Weisner, A. Fratini, Y. Liu and
131. K. S. Schanze, J. Phys. Chem. A, 2005, 109, 999.

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 2541–2566 2565
View Article Online

67 E. E. Silverman, T. Cardolaccia, X. Zhao, K.-Y. Kim, Chem., 2005, 690, 5036; (d) W.-Y. Wong, K.-H. Choi, G.-L.
K. Haskins-Glusac and K. S. Schanze, Coord. Chem. Rev., Lu and Z. Lin, Organometallics, 2002, 21, 4475.
2005, 249, 1491. 88 J. S. Wilson, N. Chawdhury, M. R. A. Al-Mandhary,
68 H. F. Wittmann, R. H. Friend, M. S. Khan and J. Lewis, M. Younus, M. S. Khan, P. R. Raithby, A. Köhler and
J. Chem. Phys., 1994, 101, 2693. R. H. Friend, J. Am. Chem. Soc., 2001, 123, 9412.
69 T. J. McKay, J. A. Bolger, J. Staromlynska and J. R. Davy, 89 G.-J. Zhou, W.-Y. Wong, Z. Lin and C. Ye, Angew. Chem., Int.
J. Chem. Phys., 1998, 108, 5537. Ed., 2006, 45, 6189.
70 J. Staromlynska, T. J. McKay, J. A. Bolger and J. R. Davy, 90 S. M. Grayson and J. M. J. Fréchet, Chem. Rev., 2001, 101, 3819.
J. Opt. Soc. Am. B, 1998, 15, 1731. 91 D. A. Tomalia, A. M. Naylor and W. A. Goddard III, Angew.
71 J. Staromlynska, T. J. McKay and P. Wilson, J. Appl. Phys., Chem., Int. Ed. Engl., 1990, 29, 138.
2000, 88, 1726. 92 S. Hecht and J. M. J. Fréchet, Angew. Chem., Int. Ed., 2001, 40,
72 J. E. Rogers, B. C. Hall, D. C. Hufnagle, J. E. Slagle, A. P. Ault, 74.
D. G. McLean, P. A. Fleitz and T. M. Cooper, J. Chem. Phys., 93 R. Vestberg, R. Westlund, A. Eriksson, C. Lopes, M. Carlsson,
2005, 122, 214708. B. Eliasson, E. Glimsdal, M. Lindgren and E. Malmström,
73 T. M. Cooper, D. M. Krein, A. R. Burke, D. G. McLean, Macromolecules, 2006, 39, 2238.
J. E. Rogers, J. E. Slagle and P. A. Fleitz, J. Phys. Chem. A, 94 R. C. Hollins, Curr. Opin. Solid State Mater. Sci., 1999, 4, 189.
Published on 24 January 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00094A

2006, 110, 4369. 95 F. Mammeri, E. Le Bourhis, L. Rozes and C. Sanchez, J. Mater.


74 G.-J. Zhou, W.-Y. Wong, D. Cui and C. Ye, Chem. Mater., 2005, Chem., 2005, 15, 3787.
17, 5209. 96 A. Priimagi, S. Cattaneo, R. H. A. Ras, S. Valkama, O. Ikkala
75 R. Westlund, E. Glimsdal, M. Lindgren, R. Vestberg, C. Hawker, and M. Kauranen, Chem. Mater., 2005, 17, 5798.
C. Lopesd and E. Malmström, J. Mater. Chem., 2008, 18, 97 R. Westlund, E. Malmström, C. Lopes, J. Öhgren, T. Rodgers,
166. Y. Saito, S. Kawata, E. Glimsdal and M. Lindgren, Adv. Funct.
76 V. W.-W. Yam, R. P.-L. Tang, K. M.-C. Wong and Mater., 2008, 18, 1939.
K. K. Cheung, Organometallics, 2001, 20, 4476. 98 R. Westlund, E. Malmström, M. Hoffmann, R. Vestberg,
Downloaded by UNSW Library on 17 March 2013

77 K. M.-C. Wong and V. W.-W. Yam, Coord. Chem. Rev., 2007, C. Hawker, E. Glimsdal, M. Lindgren, P. Norman, A. Eriksson
251, 2477. and C. Lopes, Proc. SPIE–Int. Soc. Opt. Eng., 2006, 6401,
78 W. Lu, B.-X. Mi, M. C. W. Chan, Z. Hui, C.-M. Che, N. Zhu and 64010H.
S.-T. Lee, J. Am. Chem. Soc., 2004, 126, 4958. 99 R. Zieba, C. Desroches, F. Chaput, M. Carlsson, B. Eliasson,
79 F. Guo, W. Sun, Y. Liu and K. Schanze, Inorg. Chem., 2005, 44, C. Lopes, M. Lindgren and S. Parola, Adv. Funct. Mater., 2009,
4055. 19, 235.
80 F. Guo and W. Sun, J. Phys. Chem. B, 2006, 110, 15029. 100 J. S. Shirk, R. G. S. Pong, S. R. Flom, F. J. Bartoli, M. E. Boyle
81 Z. Ji, Y. Li and W. Sun, Inorg. Chem., 2008, 47, 7599. and A. W. Snow, J. Opt. A: Pure Appl. Opt., 1996, 5, 701.
82 P. Shao, Y. Li and W. Sun, Organometallics, 2008, 27, 2743. 101 T. C. Wen and I. D. Lian, Synth. Met., 1996, 83, 111.
83 P. Shao, Y. Li and W. Sun, J. Phys. Chem. A, 2008, 112, 1172. 102 P. Wang, S. Zhang, P. J. Wu, C. Ye, H. W. Liu and F. Xi, Chem.
84 P. Shao, Y. Li, A. Azenkeng, M. R. Hoffmann and W. Sun, Inorg. Phys. Lett., 2001, 340, 261.
Chem., 2009, 48, 2407. 103 P. W. Zhu, P. Wang, W. F. Qiu, Y. Q. Liu, C. Ye, G. Y. Fang and
85 G.-J. Zhou, W.-Y. Wong, C. Ye and Z. Lin, Adv. Funct. Mater., Y. L. Song, Appl. Phys. Lett., 2001, 78, 1319.
2007, 17, 963. 104 Y. Li, D. Dini, M. J. F. Calvete, M. Hanack and W. Sun, J. Phys.
86 G.-J. Zhou, W.-Y. Wong, S.-Y. Poon, C. Ye and Z. Lin, Adv. Chem. A, 2008, 112, 472.
Funct. Mater., 2009, 19, 531. 105 Y. Li, T. M. Pritchett, J. Huang, M. Ke, P. Shao and W. Sun,
87 (a) W.-Y. Wong, L. Liu and J. X. Shi, Angew. Chem., Int. Ed., J. Phys. Chem. A, 2008, 112, 7200.
2003, 42, 4064; (b) W.-Y. Wong, Coord. Chem. Rev., 2007, 251, 106 P.-A. Bouit, R. Westlund, P. Feneyrou, O. Maury, M. Malkoch,
2400; (c) L. Liu, S.-Y. Poon and W.-Y. Wong, J. Organomet. E. Malmström and C. Andraud, New J. Chem., 2009, 33, 964.

2566 Chem. Soc. Rev., 2011, 40, 2541–2566 This journal is c The Royal Society of Chemistry 2011

Potrebbero piacerti anche