ArticlePDF Available

Synthetically Tailored Excited States: Phosphorescent, Cyclometalated IridiumACHTUNGTRENNUNG(III) Complexes and Their Applications

Authors:

Abstract and Figures

Phosphorescent iridiumACHTUNGTRENNUNG(III) complexes are being widely explored for their utility in diverse photo- physical applications. The performance of these materi- als in such roles depends heavily on their excited-state properties, which can be tuned through ligand and sub- stituent effects. This concept article focuses on methods for synthetically tailoring the properties of bis-cyclome- talated iridiumACHTUNGTRENNUNG(III) materials, and explores the factors governing the nature of their lowest excited state.
Content may be subject to copyright.
! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2006, 12, 7970 7977
7970
CONCEPTS
DOI: 10.1002/chem.200600618
Synthetically Tailored Excited States: Phosphorescent, Cyclometalated
IridiumACHTUNGTRENNUNG(III) Complexes and Their Applications
Michael S. Lowry and Stefan Bernhard*
[a]
Introduction
Luminescent transition-metal complexes are appealing for
their utility in diverse applications.
[1–5]
The effectiveness of a
complex in a specific role is determined by its excited-state
properties and can be manipulated through synthetic modifi-
cations. Early work in this field focused predominantly on
tris-diimine ruthenium(II) complexes (e.g., [RuACHTUNGTRENNUNG(bpy)
3
]
2 +
),
but these materials offered limited color tuning capability
due to thermal population of a nonemissive metal-centered
(
3
MC) state.
[6]
IridiumACHTUNGTRENNUNG(III) complexes, on the other hand,
enable broader tuning possibilities due to their increased
ligand-field stabilization energy (LFSE) and, as a conse-
quence, less thermally accessible
3
MC states. Splitting of the
d orbitals can be further enhanced with strong-field ligands
and, as a result, bis-cyclometalated iridiumACHTUNGTRENNUNG(III) complexes
have displaced ruthenium(II) compounds at the forefront of
many photochemical and photophysical investigations.
[7–12]
In particular, the reversible electrochemistry, synthetic
versatility, and robust nature of iridiumACHTUNGTRENNUNG(III) complexes
render them appealing materials for a multitude of applica-
tions. Bis-cyclometalated iridiumACHTUNGTRENNUNG(III) complexes can be pre-
pared according to the two-step synthesis presented in
Scheme 1, in which the net charge of the mononuclear com-
plex can be determined by the nature of the ancillary ligand
(Figure 1).
[11–14]
By independently modifying the cyclometa-
lating and ancillary ligands, it is possible to endow a com-
plex with specific photophysic al and electrochemical traits
and to tune its ability to perform in areas such as organic
light-emitting diodes (OLEDs),
[13–15]
luminescence-based
sensors,
[3,16,17]
and photocatalysis.
[18]
Nature of the Excited State
The two principle transitions that are observed in the long-
lived excited state of iridiumACHTUNGTRENNUNG(III) complexes are: 1) metal-
to-ligand charge transfer (MLCT) in which an electron is
promoted from a metal d orbital to a vacant p* orbital on
one of the ligands, and 2) ligand-cen tered (LC) transitions
in which an electron is promoted between p orbitals on one
of the coordinated ligands.
[6,7]
Strong spin-orbit coupling
from the iridiumACHTUNGTRENNUNG(III) center facilitates inters ystem crossing
to energetically similar triplet states and enables the forma-
tion of an emissive, mixed (triplet) excited state [T
1
; Eq (1)]
Abstract: Phosphorescent iridiumACHTUNGTRENNUNG(III) complexes are
being widely explored for their utility in diverse photo-
physical applications. The performance of these materi-
als in such roles depends heavily on their excited-state
properties, which can be tuned through ligand and sub-
stituent effects. This concept article focuses on methods
for synthetically tailoring the properties of bis-cyclome-
talated iridiumACHTUNGTRENNUNG(III) materials, and explores the factors
governing the nature of their lowest excited state.
Keywords: iridium · luminescence · OLEDs · photo-
chemistry · structure–property relationships
[a] M. S. Lowry, Prof. S. Bernhard
Department of Chemistry, Princeton University
Princeton, New Jersey 08544 (USA)
Fax: (+ 1) 609-258-7657
E-mail: bern@princeton.edu
Scheme 1. Synthesis of a heteroleptic, bis-cyclometalated iridiumACHTUNGTRENNUNG(III)
complex. A dichloro-bridged diiridium dimer involving a cyclometalating
ligand (e.g., 2-phenylpyridine, ppy, blue) is isolated and subsequently
cleaved using an ancillary ligand (e.g., 2,2-bipyridine, bpy, red) to yield
mononuclear iridiumACHTUNGTRENNUNG(III) complexes (e.g., [IrACHTUNGTRENNUNG(ppy)
2
ACHTUNGTRENNUNG(bpy)]
+
).
Chem. Eur. J. 2006, 12, 7970 7977 ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org
7971
CONCEPTS
in which the coefficients a and b refer to the contributions
of the
3
MLCT and
3
LC states, respectively.
[6,7,19]
An excited-
state manifold (a ! b) is presented in Figure 2.
Y
T
1
¼ aY
MLCT
þ bY
LC
ð1Þ
Tremendous efforts have been made to understand the
nature of the excited state as well as the structural factors
by which it is controlled. A combination of excited-state
lifetime measurements,
[9]
time-resolved luminescence spec-
troscopy,
[8,20,21]
and computational investigations
[22,23]
have
provided significant insight into the nature of the excited
state in bis-cyclometalated iridiumACHTUNGTRENNUNG(III) complexes. For ex-
ample, Hay accurately described the HOMO and
HOMO&1 as metal and cyclometalating ligand (p) orbitals
and the LUMO and LUMO+1 as vacant (p*) ligand orbi-
tals in [IrACHTUNGTRENNUNG(ppy)
2
ACHTUNGTRENNUNG(bza)] (in which ppy= 2-phenylpyridine,
bza = 1-phenyl-1,3-butanedione) by using time-dependent
density-functional theory (TD-DFT).
[23]
Similar results have
been attained with other neutral and ionic iridiumACHTUNGTRENNUNG(III) com-
plexes.
[11,14,22,23]
Thus, it appears feasible for the excited state
of bis-cyclometalated iridiumACHTUNGTRENNUNG(III) complexes to contain
mixed MLCT–LC character (Figure 3). It is worth noting
that these calculations correspond to the absolute energy of
the associated orbitals (i.e.,
singlet transition energies for
which the
1
LC state is signifi-
cantly larger than the
1
MLCT
state) and that singlet–triplet
stabilization energies must be
considered before further pre-
dictions can be made about the
energy of the emissive
state.
[7,11]
Tuning Strategies
Excited-state mixing occurs
when sufficient overlap is pres-
ent between the
3
LC and
3
MLCT states.
[24]
Thus, it is possible to control the energy of
the lowest excited state by deliberately adjusting the energy
of metal and ligand orbitals, which can be achieved through
substituent effects
[25,26]
or by changing the ligand parent
structure entirely (e.g., ppy vs. 1-phenylpyrazole, ppz).
[14, 27]
Structural control is facilitated by the multistep synthesis
Figure 1. Bis-cyclometalated (shown in blue) iridiumACHTUNGTRENNUNG(III) luminophores. The net charge of the complex is de-
termined by the ancillary ligand(s) (shown in red): a) cationic [IrACHTUNGTRENNUNG(dFCF
3
ppy)
2
ACHTUNGTRENNUNG(dtbbpy)]
+
PF
6
&
(dFCF
3
ppy = 5-
trifluoromethyl-2-(2,4-difluorophenyl)-pyridine; dtbbpy = 4,4-di-tert-butyl-2,2-bipyridine), b) neutral [Ir-
ACHTUNGTRENNUNG(dFpz)
2
ACHTUNGTRENNUNG(CF
3
ppz)] (dFpz = 1-(2,4-difluorophenyl)-pyrazole; CF
3
ppz = 2-(5-trifluoromethylpyrazol-3-yl)-pyri-
dine), and c) anionic [TBA]
+
[IrACHTUNGTRENNUNG(ppy)
2
(CN)
2
]
&
(TBA
+
= tetrabutylammonium, CN
&
= cyanide).
Figure 2. The energetic closeness and degree of overlap between
3
MLCT
and
3
LC states results in the formation of a mixed lowest excited state
(T
1
). The excited molecule relaxes to the ground state through radiative
(k
r
) and nonradiative (k
nr
) pathways.
Figure 3. Molecular orbital diagram for [IrACHTUNGTRENNUNG(dFCF
3
ppy)
2
ACHTUNGTRENNUNG(dtbbpy)]
+
(Fig-
ure 1a) obtained from DFT calculations. The HOMO is composed pri-
marily of metal d orbitals with some contributions from ligand (ppy-
based, phenyl) orbitals, while the LUMO and LUMO+1 are localized on
separate ligands (bpy- and ppy-based, respectively). Ligand-field splitting
is eviden t between the HOMO (“t
2g
”) and LUMO+12 (“e
g
*”). The cal-
culated orbital energies are corroborated by experimental evidence
(cyclic voltammetry).
www.chemeurj.org ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2006, 12, 7970 7977
7972
S. Bernhard and M. S. Lowry
shown in Scheme 1 in which different ligands are independ-
ently added to the coordination sphere.
When both classes of ligands provide orbitals that partici-
pate in the excited-state transitions, the cyclometalating
ligand tends to be associated with the
3
LC transition and the
ancillary ligand with the
3
MLCT transition.
[23,24]
In such
cases, the excited state can be tuned directly through ligand
modifications, because each ligand is linked to a different
transition. By monitoring the effect of various ligand permu-
tations within the coordination sphere, it is possible to gain
insight into the factors that govern the photophysical and
electrochemical behavior of heteroleptic iridiumACHTUNGTRENNUNG(III) com-
plexes, and to tailor materials with specific excited-state
properties.
[11, 28]
In some excited complexes, interligand electron transfer
(ILET) from the cyclometalating ligand to the ancillary
ligand is possible. The net properties of these materials can
be controlled almost exclusively by the ancillary ligand. For
instance, Park"s group observed emission across the visible
spectrum by changing the structure of the ancillary ligand in
a series of complexes that underwent ILET prior to emis-
sion.
[29]
Campagna"s group further demonstrated that if the
structure of the ancillary ligand is modified such that some
of its p density overlaps the plane of chelation for the cyclo-
metalating ligand, another charge transfer transition (sigma-
bond-to-ligand charge transfer, SBLCT
[30]
) is feasible.
[31]
As
a result, electron density can be transferred from a carbon–
metal s bond to the ancillary ligand, and site-specific tuning
can be achieved.
When only the cyclometalating ligand contributes p* orbi-
tals at an appropriate energy for the excited-state transi-
tions, both transitions tend to be associated with the cyclo-
metalating ligand. Meanwhile the ancillary ligand (e.g., 2,4-
pentanedione, acac) plays a more passive role in determin-
ing the nature of the excited state; influencing the energy of
the metal orbitals by inductive communication through s
bonds.
[32]
Thus, even a “nonparticipating” ancillary ligand
can be used to adjust the energy of the MLCT state as well
as the degree of mixing between the states.
One of the most effective methods for tuning the energy
of the lowest excited state involves changing the degree of
conjugation in the structure of the participating (cyclometa-
lating and/or ancillary) ligands.
[11,27,28]
As the coordination
sphere becomes more diffuse, the corresponding orbitals are
stabilized. It follows, then, that the excited-state transitions
can be tuned by altering the size of the ligands and also by
localizing electron density in discrete regions of the mole-
cule or by partia lly destroying ligand aromaticity. For exam-
ple, bulky pendant groups can be used to distort a ligand
from planarity, which will destabilize its p orbitals, and, as a
result, increase the size of the associated
3
MLCT or
3
LC
transition. A critical drawback of this approach is that en-
hanced internal strain may promote nonradiative decay,
whereby higher excited-state energy is attained at the ex-
pense of other important properties (e.g., luminous intensity
and excited-state lifetime).
[33–35]
Alternatively, recent investi-
gations of iridiumACHTUNGTRENNUNG(III) complexes involving terdentate cyclo-
metalating ligands (e.g., 2,6-diphenyl-pyridine
[36,37]
and 1,3-
bis-(1-methyl-benzimidazol-2-yl)benzene
[38]
) have shown
that low-energy emission can be readily achieved in com-
plexes containing ligands with extended p systems.
[36–38]
Another promising method for tuning (and fine-tuning)
the excited-state properties of iridiumACHTUNGTRENNUNG(III) complexes in-
volves deliberate functionalization of the ligands through
the use of substituent groups. B y modifying the symmetry
and inductive influence of a ligand with different substitu-
ents, it is possible to control metal–ligand bonding as well as
ligand orbital energies and, thus, to control the nature of the
lowest excited state. Tremendous color versatility has been
achieved with iridiumACHTUNGTRENNUNG(III) luminophores in this manner
(Figure 4), and a broad range of excited-state lifetimes
(from nanoseconds to several microseconds) as well as phos-
phorescent yields (approaching 100 %) have been report-
ed.
[11, 14]
In particular, electronic effects have been considered due
to their profound influence on orbital energies as well as the
relative ease with which electron-withdrawing (e.g., -F,
-CF
3
) and electron-donating (e.g., -CACHTUNGTRENNUNG(CH
3
)
3
, -OCH
3
) groups
can be incorporated into the ligand structure. Electron-with-
drawing substituents tend to stabilize the HOMO by remov-
ing electron density from the metal, whereas donating
groups have an inverse effect.
[11,14,39,40]
This relationship is
convoluted by the fact that withdrawing groups may also
lower the energy of the LUMO (i.e., increasing the electron
affinity of the parent ligand).
[14]
Fortunately, the cyclometa-
lating and ancillary ligands can be separately substituted
with electron-withdrawing and ACHTUNGTRENNUNG-donating groups in hetero-
leptic complexes, which enables deliberate control over the
excited state.
The position of these substituents with respect to the co-
ordinating carbon of a cyclometalating ligand will strongly
influence the LFSE of the resulting complex.
[41]
For exam-
ple, electron-withdrawing groups meta to the site of coordi-
nation (and donating groups ortho or para to this position)
increase the field strength of the ligand and concomitantly
enhance d-orbital splitting. Interestingly, opposing spe ctro-
scopic trends have been observed for fluoro (-F) and tri-
fluoromethyl (-CF
3
) substituents at the same position of a
cyclometalating ring despite their similar electronic ef-
fects.
[40]
This difference has been attributed to the mesomer-
ic and inductive ability of the fluorine atom as opposed to
the purely inductive ability of the trifluoromethyl group.
Thus, it is important to consider the total impact of a sub-
stituent group when designing ligand systems for excited-
state tuning.
Due to the high number of ligand and substituent varia-
bles that influence the nature of the excited state in hetero-
leptic iridiumACHTUNGTRENNUNG(III) complexes, their electrochemical and pho-
tophysical properties are not always easy to predict a priori.
To this end, our group has developed a set of combinatorial
procedures—involving parallel synthesis and high-through-
put screening—to examine the effect of multiple structural
(and site-dependent) variables in tandem.
[11]
As our under-
standing of structure–property relationships improves, we
Chem. Eur. J. 2006, 12, 7970 7977 ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org
7973
CONCEPTS
Luminescence
have found that DFT calculations provide an excellent
method for modeling orbital configurations and extrapolat-
ing the energy of the lowest excited state.
[11,14]
These tech-
ACHTUNGTRENNUNGniques in conjunction with the other strategies presented
herein have facilitated and expedited the design of useful
materials for various optoelectronic,
[4,39]
analytical,
[3,16]
and
photocatalytic applications.
[18]
Four areas in which the highly
tunable nature of iridiumACHTUNGTRENNUNG(III) complexes have improved the
performance of inorganic materials are discussed below.
OLED applications
Transition-metal-based, organic light-emitting diodes
(OLEDs) are considered viable candidates for flat-panel dis-
plays due to their color versatility, low operation voltages,
and phosphorescent efficiency.
[42,43]
The first complex-based
OLED was assembled by Tang and VanSlyke in 1987 con-
taining an Alq
3
chromophore (q = 8-hydroxyquinoline) that
emits green light (550 nm) from a singlet excited state.
[44]
Considerable attention has since been placed on altering
device characteristics through strategic modifications to the
chromophore. In particular, complexes containing a heavy
transition metal center, such as iridiumACHTUNGTRENNUNG(III),
[45]
rutheni-
ACHTUNGTRENNUNGum(II),
[46]
or osmium(II)
[47]
are targeted for electrolumines-
cence studies because of the increased theoretical limit for
emission efficiency from triplet emitting complexes (nearly
100 %) over singlet emitters (% 25 %).
[48]
OLEDs that are constructed with neutral complexes typi-
cally consist of the luminescent chromophore embedded in
an organic matrix (e.g., 4,4-N,N-dicarbazolylbiphenyl,
CBP), sandwiched between multiple layers of charge trans-
port materials (e.g., 2,9-dimethyl-4,7-diphenyl-1,10-phenan-
throline, BCP and 4,4-bis-[N-(naphthyl-N-phenylamino)bi-
phenyl, a-NPD, respectively), and capped with a low work-
function cathode (e.g., LiF/Al) and a transparent anode
(e.g., indium–tin oxide, ITO; Figure 5a).
[49]
By systematically
adjusting the nature of the excited state in the chromophore,
Thompson and Forrest have observed emission across the
visible spectrum.
[15,28]
Similarly, Holmes and Friend fine-
tuned electroluminescence in the green-to-blue regime
through synthetic modifications to the chromophore and,
more importantly, improved the operational lifetime of the
devices through the use of more stable (i.e., less labile) an-
cillary ligands.
[50]
Neutral iridiumACHTUNGTRENNUNG(III) complexes have spur-
red tremendous interest (beyond the scope of this manu-
Figure 4. Color versatility expressed by a series of six cationic iridiumACHTUNGTRENNUNG(III) luminophores. Their structures are listed in order of increasing emission wave-
length (blue to red). The evolution of
3
LC (vibrationally structured, high-energy bands) and
3
MLCT character (structureless, low-energy bands) in the lu-
minescence spectra are indicative of a mixed excited state.
Figure 5. a) A multilayer device ensemble involving an organic material
(e.g., CBP) doped with a neutral or anionic iridiumACHTUNGTRENNUNG(III) chromophore.
This matrix is sandwiched between hole- (e.g., a-NPD) and electron-in-
jecting (e.g., Alq
3
, q = 8-hydroxyquinoline) layers as well as a hole block-
ing layer (e.g., BCP) and capped with a low work-function cathode (e.g.,
LiF/Al) and a transparent anode (e.g., ITO). b) Single-layer device ge-
ometry in which a cationic iridiumACHTUNGTRENNUNG(III) chromophore is spin-coated on an
ITO substrate and capped with a gold electrode. c) Electroluminescence
from [IrACHTUNGTRENNUNG(ppy)
2
ACHTUNGTRENNUNG(dtbbpy)]ACHTUNGTRENNUNG(PF
6
) in a single-layer, air-stable device (560 nm).
www.chemeurj.org ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2006, 12, 7970 7977
7974
S. Bernhard and M. S. Lowry
script) over the course past ten years and a concise review
of the advances in this field is overdue.
Recently, cationic iridiumACHTUNGTRENNUNG(III) complexes have emerged in
OLED applications due to their ability to electrolumines-
cence from a single-layer of neat complex sandwiched di-
rectly between two air-stable electrodes, such as ITO and
gold (Figure 5b, c).
[45]
This simple device geometry is made
feasible by the presence of mobile counterions that redis-
tribute in the presence of an applied potential and facilitate
charge injection at the electrodes. Additionally, charge-
transport occurs through a hopping mechanism, which ena-
bles low turn-on voltages (< 3 V) and device operation from
an alternating current (AC) power source.
[4]
The discovery
of novel cationic luminophores has been catalyzed by the
development of an exhaustive library of prospective materi-
als using the combinatorial procedures that were developed
in our group.
[11]
Recently, we reported yellow (560 nm),
[45]
green (531 nm)
[51,52]
and blue-green (497 nm)
[14]
electrolumi-
nescence from complexes containing a ppy–bpy backbone.
By adjusting the size (and, thus, energy) of the aromatic li-
gands, Thompson"s group achieved red electroluminescence
(635 nm) and also observed a slight hypsochromic shift
(492 nm) over these pioneering complexes.
[27]
Nevertheless,
efforts to optimize cationic systems in OLED applications
and to achieve pure blue electroluminescence (450 nm) are
ongoing.
OLED fabrication has also been successful with anionic
complexes of the form [IrACHTUNGTRENNUNG(ppy)
2
X
2
]
&1
(X = CN
&
, NCO
&
), in
which device assemblies resemble the multilayer ensembles
utilized for neutral chromophores.
[13]
In such cases, the
pseudo-halogen (X) influences the energy of the metal d or-
bitals by adjusting the LFSE and, thus, provides a viable
platform for color tuning. As such, yellow (556 nm; X=
CN
&
) and blue-green (500 nm; X= NCO
&
) emission have
been observed.
[13]
Notably, device stability and efficiency
also improved in systems with large LFSE (in which the
3
MC state is no longer accessible), but pursuit of pure blue
electroluminescence persists.
Oxygen Sensor Applications
The long-lived triplet excited state of luminescent iridium-
ACHTUNGTRENNUNG(III) complexes enables efficient energy transfer with the
triplet ground state of molecular oxygen, resulting in lumi-
nescence quenching and the formation of singlet oxygen.
[53]
Consequently, iridiumACHTUNGTRENNUNG(III) luminophores can be utilized as
oxygen probes in various medicinal, chemical, and environ-
mental sensors.
[16,54]
Oxygen concentration is quantified ac-
cording to sudden changes in the luminescent properties of
the complex and, thus, materials with high quantum yield
and long excited-state lifetime (i.e., several microseconds)
are highly desirable for facile, sensitive detection. Mixed-
ligand iridiumACHTUNGTRENNUNG(III) complexes are particularly intriguing due
to their broadly tunable excited-state properties, durability,
and high chem ical stability in the presence of singlet
oxygen.
[16]
A significant setback to the implementation of
solid-state iridiumACHTUNGTRENNUNG(III)-based sensors has been the occur-
rence of self-quenching between neighboring molecules.
One potential solutio n involves separating luminescent dyes
by embedding them in oxygen permeable polymer matrices,
but the sensitivity of such systems is limited by low loading
concentrations in order to prevent dye aggregation.
[16]
Ef-
forts to improve the loading concentration have included co-
valently binding the luminophores to the polymeric host
[55]
as well as adjusting the charge and size of the dye mole-
cules,
[56]
but investigations in this field are ongoing.
Bioanalytical Applications
The rich electrochemical and photophysical characteristics
of iridiumACHTUNGTRENNUNG(III) complexes also render them strong candi-
dates for bioanalytical applications. The intense emission
and long excited-state lifetimes of the iridiumACHTUNGTRENNUNG(III) com-
plexes enable sensitive, time-resolved detection and their
large Stokes shift helps minimize self-quenching between
dispersed materials.
[57]
More importantly, luminescent
iridiumACHTUNGTRENNUNG(III) complexes can be used to label biomaterials
due their ability to either covalently (e.g., aldehyde–amine
cross-linking)
[58]
or noncovalently (e.g., DNA intercala-
tion)
[59]
bind biological substrates. Bound complexes typical-
ly express different luminescent properties than their free
analogs due to changes in the rigidity and hydrophobicity of
the surrounding environment.
[59]
Thus, luminescent labels
provide a facile method for monitoring bioconjugation reac-
tions and quantifying the binding affinity between different
substrates.
[3,60]
A comprehensive review of the role of
iridiumACHTUNGTRENNUNG(III) and other transition-metal complexes in bioana-
lytical applications as well as methods for improving their
luminescent properties was reported by Lo and co-workers
in 2005.
[3]
Photocatalytic Water Splitting
The challenge of converting solar radiation to conveniently
usable forms of energy has been co nsidered by many re-
searchers. One promising solution involves harnessing solar
energy through a photocatalytic cycle to split water into hy-
drogen and oxygen.
[5,61,62]
This process can be conceptual-
ized as two half-reactions in which water is oxidized to mo-
lecular oxygen and reduced to molecular hydrogen. In par-
ticular, the high gravimetric energy density and clean com-
bustion products of hydrogen render it a highly viable
source of fuel and, as such, significant attention has been
placed on hydrogen production. Typical photoreduction
schemes employ a transition-metal photosensitizer in con-
junction with an electron relay to collect and store radiant
energy and convert protons into molecular hydrogen.
[5,61]
Heteroleptic iridiumACHTUNGTRENNUNG(III) complexes are appealing photo-
sensitizers due to their highly tunable properties, and, in
fact, we recently reported an iridiumACHTUNGTRENNUNG(III) photosensitizer
that exhibits tremendous improvements in hydrogen produc-
Chem. Eur. J. 2006, 12, 7970 7977 ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org
7975
CONCEPTS
Luminescence
tion over the previous state-of-the-art literature standard,
[RuACHTUNGTRENNUNG(dmphen)
3
]
2+
(dmphen = 4,7-dimethyl-1,10-phenanthro-
line).
[14]
This increase in hydrogen production has been
traced to the improved reducing strength (i.e., ability to
pass an electron to the electron relay) of iridiumACHTUNGTRENNUNG(III) com-
plexes over ruthenium(II) complexes. An overview of the
ground and excited-state redox properties of these materials
were reported in a recent publication.
[14]
Other highly promising iridiumACHTUNGTRENNUNG(III) complexes have also
been targeted by using a high-throughput technology that
was developed in our laboratory, whereby multiple photo-
sensitizers were simultaneously screened under variable re-
action conditions. A custom-built parallel photoreactor—
consisting of a series of ultra-bright light emitting diodes
(500 mW ' 10% at 465 nm) connected to a multiwell
sample holder mounted on an orbital shaker—was em-
ployed along with a Ni/Pd thin film commercial hydrogen
sensor (Figure 6).
[18]
This methodology has helped us pin-
point structural and photophysical traits that are conducive
to hydrogen productio n and also provided insight into the
interplay between the photosensitizer and electron relay in
the catalytic loop, yet considerable strides still need to be
made before a sustainable “hydrogen economy” can be de-
veloped.
Conclusions and Prospects
The nature and energy of the excited state in mixed-ligand
iridiumACHTUNGTRENNUNG(III) complexes can be manipulated by deliberate
chemical synthesis. Synthetic control is facilitated by the de-
pendence of the excited state on the orbital configurations
within a molecule as well as the relative ease by which dif-
ferent parent ligand structures can be coordinated to the
metal center. Consequently, heteroleptic iridiumACHTUNGTRENNUNG(III) com-
plexes can be tailored to express specific luminescent prop-
erties or to serve a specific functional role (e.g., cross-link-
ing) and iridiumACHTUNGTRENNUNG(III)-based materials are being explored for
a plethora of applications. The improved performance of
iridiumACHTUNGTRENNUNG(III) complexes over alternate transition metal cen-
ters as well as the straightforward manner in which their
properties can be tuned has positioned heteroleptic iridium-
ACHTUNGTRENNUNG(III) complexes at the forefront of modern photochemistry
and we can expect that they will remain there throughout
the foreseeable future.
Acknowledgements
This work was supported by the National Science Foundation (Career
Award No. CHE-0449755), the Princeton Center for Complex Materials,
which is a Materials Research Science and Engineering Center of the Na-
tional Science Foundation (DMR-9632275), and a Camille and Henry
Dreyfus Foundation New Faculty Award. We would also like to thank
George G. Malliaras and Andr eas Schrag for helpful conversations and
the contr ibution of photographs.
[1] M. J. Cook, A. P. Lewis, G. S. G. McAuliffe, V. Skarda, A. J. Thom-
son, J. L. Glasper, D. J. Robbins, J. Chem. Soc. Perkin Trans. 2 1984,
1293 1301.
[2] B. Higgins, B. A. DeGraff, J. N. Demas, Inorg. Chem. 2005, 44,
6662 6669.
[3] K. K.-W. Lo, W.-K. Hui, C.-K. Chung, K. H.-K. Tsang, D. C.-M. Ng,
N. Zhu, K.-K. Cheung, Coord. Chem. Rev. 2005, 249, 1434 1450.
[4] J. Slinker, D. Bernards, P. L. Houston, H. D. AbruÇa, S. Bernhard,
G. G. Malliaras, Chem. Commun. 2003, 2392 2399.
[5] M. Kirch, J. M. Lehn, J. P. Sauvage, Helv. Chim. Acta 1979, 62,
1345 1384.
[6] K. Kalyanasundaram, Photochemistry of Polypyridine and Porphyrin
Complexes, Academic Press, New York, NY, 1992.
[7] N. J. Turro, Modern Molecular Photochemistry, Benjamin/Cummings,
Menlo Park, CA, 1978.
[8] M. G. Colombo, A. Hauser, H. U. G#de l, Inorg. Chem. 1993, 32,
3088 3092.
[9] K. A. King, R. J. Watts, J. Am. Chem. Soc. 1987, 109, 1589 1590.
[10] D. M. Roundhill, Photochemistry and Photophysics of Metal Com-
plexes, Plenum Press, New York, NY, 1994.
[11] M. S. Lowry, W. R. Hudson, R. A. Pascal, Jr., S. Bernhard, J. Am.
Chem. Soc. 2004, 126, 14 129 14 135.
[12] C.-H. Yang, S.-W. Li, Y. Chi, Y.-M. Cheng, Y.-S. Yeh, P.-T. Chou, G.-
H. Lee, C.-H. Wang, C.-F. Shu, Inorg. Chem. 2005, 44, 7770 7780.
[13] Md. K. Nazeeruddin, R. Humphry-Baker, D. Berner, S. Rivier, L.
Zuppiroli, M. Graetzel, J. Am. Chem. Soc. 2003, 125, 8790 8797.
[14] M. S. Lowry, J. I. Goldsmith, J. D. Slinker, R. Rohl, R. A. Pascal, Jr.,
G. G. Malliaras, S. Bernhard, Chem. Mater. 2005, 17, 5712 5719.
[15] C. Adachi, R. C. Kwong, P. Djurovich, V. Adamovich, M. A. Baldo,
M. E. Thompson, S. R. Forrest, Appl. Phys. Lett. 2001, 79, 2082
2084.
Figure 6. A water-cooled parallel photoreactor with ultrabright LEDs for
studying photocatalytic reactions. IridiumACHTUNGTRENNUNG(III) complexes (green) that ex-
hibit a tremendous increas e in hydrogen production over ruthenium(II)
standards (orange) have been identif ied using this high-throughput tech-
nique (Fmppy = 5-methyl-2-(4-fluorophenyl)-pyridine; phen = 1,10-phe-
nanthroline; dphphen = 4,7-diphenyl-1,10-phenanthroline).
www.chemeurj.org ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2006, 12, 7970 7977
7976
S. Bernhard and M. S. Lowry
[16] M. C. DeRosa, D. J. Hodgson, G. D. Enright, B. Dawson, C. E. B.
Evans, R. J. Crutchley, J. Am. Chem. Soc. 2004, 126, 7619 7626.
[17] M.-L. Ho, F.-M. Hwang, P.-N. Chen, Y.-H. Hu, Y.-M. Cheng, K.-S.
Chen, G.-H. Lee, Y. Chi, P.-T. Chou, Org. Biomol. Chem. 2006, 4,
98 103.
[18] J. I. Goldsmith, W. R. Hudson, M. S. Lowry, T. H. Anderson, S.
Bernhard, J. Am. Chem. Soc. 2005, 127, 7502 7510.
[19] A. Tsuboyama, H. Iwawaki, M. Furugori, T. Mukaide, J. Kamatani,
S. Igawa, T. Moriyama, S. Miura, T. Takiguchi, S. Okada, M. Hoshi-
no, K. Ueno, J. Am. Chem. Soc. 2003, 125, 12 971 12 979.
[20] A. P. Wilde, K. A. King, R. J. Watts, J. Phys. Chem. 1991, 95, 629
634.
[21] R. J. Watts, J. Am. Chem. Soc. 1974 , 96, 6186 6187.
[22] T. H. Kwon, H. S. Cho, M. K. Kim, J. W. Kim, J. J. Kim, K. H. Lee,
S. J. Park, I. S. Shin, H. Kim, D. M. Shin, Y. K. Chung, J. I. Hong,
Organometallics 2005, 24 , 1578 1585.
[23] P. J. Hay, J. Phys. Chem. A 2002, 106, 1634 1641.
[24] M. G. Colombo, A. Hauser, H. U. G#del, Top. Curr. Chem. 1994,
171, 143 171.
[25] R. J. Watts, J. Van Houten, J. Am. Chem. Soc. 1974, 96, 4334 4335.
[26] K. Dedeian, P. I. Djurovich, F. O. Garces, G. Carlson, R. J. Watts,
Inorg. Chem. 1991, 30, 1685 1687.
[27] A. B. Tamayo, S. Garon, T. Sajoto, P. I. Djurovich, I. M. Tsyba, R.
Bau, M. E. Thompson, Inorg. Chem. 2005, 44, 8723 8732.
[28] S. Lamansky, P. Djurovich, D. Murphy, F. Abdel-Razzaq, H.-E. Lee,
C. Adachi, P. E. Burrows, S. R. Forrest, M. E. Thompson, J. Am.
Chem. Soc. 2001, 123, 4304 4312.
[29] Y. You, S. Y. Park, J. Am. Chem. Soc. 2005 , 127, 12 438 12 439.
[30] P. Didier, I. Ortmans, A. Kirsch -De Mesmaeker, R. J. Watts, Inorg.
Chem. 1993, 32, 5239 5245.
[31] F. Neve, M. La Deda, A. Crispini, A. Bellus ci, F. Puntoriero, S.
Campagna, Organometallics 2004, 23, 5856 5863.
[32] J. Li, P. I. Djurovich, B. D. Alleyne, M. Yousufuddin, N. N. Ho, J. C.
Thomas, J. C. Peters, R. Bau, M. E. Thompson, Inorg. Chem. 2005,
44, 1713 1727.
[33] K. Nakamaru, Bull. Chem. Soc. Jpn. 1982, 55, 2697 2705.
[34] A. Juris, V. Balzani, P. Belser, A. von Zelewsky, Helv. Chim. Acta
1981, 64, 2175 2182.
[35] M. S. Lowry, S. Bernhard, Princeton University, Princeto n, NJ, un-
published results, 2005.
[36] A. J. Wilkinson, A. E. Goeta, C. E. Foster, J. A. G. Williams, Inorg.
Chem. 2004, 43, 6513 6515.
[37] M. Polson, S. Fracasso, V. Bertolasi, M. Ravaglia, F. Scandola, Inorg.
Chem. 2004, 43, 1950 1956.
[38] T. Yutaka, S. Obara, S. Ogawa, K. Nozaki, N. Ikeda, T. Ohno, Y.
Ishii, K. Sakai, M.-a. Haga, Inorg. Chem. 2005, 44, 4737 4746.
[39] I. R. Laskar, T.-M. Chen, Chem. Mater. 2004, 16, 111 117.
[40] P. Coppo, E. A. Plummer, L. De Cola, Chem. Commun. 2004, 1774
1775.
[41] T. Tsuzuki, N. Shirasawa, T. Suzuki, S. Tokito, Adv. Mater. 2003, 15,
1455 1458.
[42] E. Holder, B. M. W. Langeveld, U. S. Schubert, Adv. Mater. 2005, 17,
1109 1121.
[43] S. Forrest, P. Burrows, M. Thompson, IEEE Spectrum 2000, 37, 29
34.
[44] C. W. Tang, S. A. VanSlyke, Appl. Phys. Lett. 1987, 51, 913 915.
[45] J. D. Slinker, A. A. Gorodetsky, M. S. Lowry, J. Wang, S. Parker, R.
Rohl, S. Bernhard, G. G. Malliaras, J. Am. Chem. Soc. 2004, 126,
2763 2767.
[46] H. Rudmann, S. Shimada, M. F. Rubner, J. Am. Chem. Soc. 2002,
124, 4918 4921.
[47] S. Bernhard, X. Gao, G. G. Malliaras, H. D. AbruÇa, Adv. Mater.
2002, 14, 433 436.
[48] C. Adachi, M. A. Baldo, M. E. Thompson, S. R. Forrest, J. Appl.
Phys. 2001, 90, 5048 5051.
[49] I. R. Laskar, S.-F. Hsu, T.-M. Chen, Polyhedron 2005, 24, 189 200.
[50] C. S. K. Mak, A. Hayer, S. I. Pascu, S. E. Watkins, A. B. Holmes, A.
Kçhler, R. H. Friend, Chem. Commun. 2005, 4708 4710.
[51] J. D. Slinker, C. Y. Koh, G. G. Malliaras, M. S. Lowry, S. Bernhard,
Appl. Phys. Lett. 2005, 86, 173 506 173 508.
[52] S. T. Parker, J. D. Slinker, M. S. Lowry, M. P. Cox, S. Bernhard, G. G.
Malliaras, Chem. Mater. 2005, 17, 3187 3190.
[53] M. C. DeRosa, P. J. Mosher, C. E. B. Evans, R. J. Crutchley, Macro-
mol. Symp. 2003, 196, 235 248.
[54] Y. Amao, Y. Ishikawa, I. Okura, Anal. Chim. Acta 2001, 445, 177
182.
[55] M. C. DeRosa, P. J. Mosher, G. P. A. Yap, K.-S. Focsaneanu, R. J.
Crutchley, C. E. B. Evans, Inorg. Chem. 2003, 42, 4864 4872.
[56] G. Di Marco, M. Lanza, A. Mamo, I. Stefio, C. Di Pietro, G. Romeo,
S. Campagna, Anal. Chem. 1998, 70, 5019 5023.
[57] K. K.-W. Lo, D. C.-M. Ng, C.-K. Chung, Organometallics 2001, 20,
4999 5001.
[58] K. K.-W. Lo, C.-K. Chung, N. Zhu, Chem. Eur. J. 2003, 9, 475 483.
[59] K. K.-W. Lo, C.-K. Chung, N. Zhu, Chem. Eur. J. 2006, 12, 1500
1512.
[60] K. K.-W. Lo, C.-K. Chung, T. K.-M. Lee, L.-H. Lui, K. H.-K. Tsang,
N. Zhu, Inorg. Chem. 2003, 42, 6886 6897.
[61] K. Kalyanasundaram, J. Kiwi, M. Gr$tzel, Helv. Chim. Acta 1978,
61, 2720 2730.
[62] C. V. Krishn an, B. S. Brunschwig, C. Creutz, N. Sutin, J. Am. Chem.
Soc. 1985, 107, 2005 2015.
Published online: August 25, 2006
Chem. Eur. J. 2006, 12, 7970 7977 ! 2006 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org
7977
CONCEPTS
Luminescence
... Heteroleptic compounds incorporating (C^N) and (N^N) ligands, as in [Ir(ppy) 2 (bpy)] + complexes, benefit from independent tunability of HOMO and LUMO levels via ligand architecture. [9][10][11][12][13][14][15][16][17][18][19][20] Yet, the relationship between structural/environmental perturbations and changes in the intricate excited-state manifold of [Ir(ppy) 2 (bpy)] + complexes are hard to capture with existing spectroscopies, making it commensurately difficult to understand all the factors that modulate their photoredox activity. Tuning of the redox potentials, absorption energy, excited-state lifetime, and quantum yield, are well known criteria, but other factors can play an important role. ...
Preprint
Full-text available
We report evidence of excited-state ion pair reorganisation in a cationic iridium (III) photoredox catalyst in 1,4-dioxane. Microwave-frequency dielectric-loss measurements allow us to assign both ground and excited-state molecular dipoles and excited-state polarizability volumes. These measurements show significant changes in ground-state dipole moment between [Ir[dF(CF_{3})ppy]_{2}(dtbpy)]PF_{6} (10.74 Debye) and [Ir[dF(CF_{3})ppy]_{2}(dtbpy)]BAr^{F}_{4} (4.86 Debye). Photoexcitation of each complex results in population of highly mixed ligand centered and metal-to-ligand charge transfer states. Relaxation to the lowest lying excited-state leads to a negative change in the dipole for [Ir[dF(CF_{3})ppy]_{2}(dtbpy)]PF_{6}, and a positive change in dipole for [Ir[dF(CF_{3})ppy]_{2}(dtbpy)]BAr^{F}_{4}. These observations are consistent with a sub-nanosecond reorganization with the PF_{6}^{-} counter-ion. Taken together, these observations suggest contact-ion pair formation in [Ir[dF(CF_{3})ppy]_{2}(dtbpy)]PF_{6}. The ion pair reorganisation we observe with the PF_{6}^{-} may modify both the thermodynamic potential available for electron transfer and inhibit oxidative catalysis, providing a possible mechanism for recently observed trends in similar complexes.
... In recent years, cyclometalated iridium(III) complexes, such as Ir(ppy) 3 (ppy = 2-phenylpyridine) and [Ir(ppy) 2 (N ∧ N)] + , have been used as PSs in various photocatalytic reactions, such as H 2 evolution ( Goldsmith et al., 2005;Lowry and Bernhard, 2006), CO 2 reduction ( Thoi et al., 2013;Bonin et al., 2014;Chen et al., 2015;Rao et al., 2017Rao et al., , 2018, and organic synthesis (Figure 2; Prier et al., 2013;Schultz and Yoon, 2014;Shaw et al., 2016), even though the absorption by Ir(ppy) 3 and [Ir(ppy) 2 (N ∧ N)] + is relatively weak in the visible region. We already reported that [Ir(piq) 2 (dmb)] + (piq = 1-phenylisoquinoline, dmb = 4,4 ′ -dimethyl-2,2 ′ -bipyridine), which has a stronger absorption in the visible region (ε 444nm = 7,800 M −1 cm −1 ) than Ir(ppy) 3 , acts as a PS for CO 2 reduction without forming decomposed species that catalyze CO 2 reduction (Figure 2; Kuramochi and Ishitani, 2016). ...
Article
Full-text available
A cyclometalated iridium(III) complex having 2-(pyren-1-yl)-4-methylquinoline ligands [Ir(pyr)] has a strong absorption band in the visible region (ε444nm = 67,000 M⁻¹ cm⁻¹) but does not act as a photosensitizer for photochemical reduction reactions in the presence of triethylamine as an electron donor. Here, 1,3-dimethyl-2-(o-hydroxyphenyl)-2,3-dihydro-1H-benzo[d]imidazole (BI(OH)H) was used instead of the amine, demonstrating that BI(OH)H efficiently quenched the excited state of Ir(pyr) and can undergo the photochemical carbon dioxide (CO2) reduction catalyzed by trans(Cl)-Ru(dmb)(CO)2Cl2 (dmb = 4,4′-dimethyl-2,2′-bipyridine, Ru) to produce formate as the main product. We also synthesized a binuclear complex combining Ir(pyr) and Ru via an ethylene bridge and investigated its photochemical CO2 reduction activity in the presence of BI(OH)H.
Preprint
Full-text available
We report evidence of excited-state ion pair reorganisation in a cationic iridium (III) photoredox catalyst in 1,4-dioxane. Microwave-frequency dielectric-loss measurements combined with accurate calculations of dipolar relaxation time allow us to assign both ground and excited-state molecular dipole moments in solution and determine the polarizability volume in the excited-state. These measurements show significant changes in ground-state dipole moment between [Ir[dF(CF 3 )ppy] 2 (dtbpy)]PF 6 (10.74 Debye) and [Ir[dF(CF 3 )ppy] 2 (dtbpy)]BAr F 4 (4.86 Debye). Photoexcitation of each complex results in population of highly mixed ligand centered and metal-to-ligand charge transfer states with enormous polarizability. Relaxation to the lowest lying excited-state leads to a negative change in the dipole moment for [Ir[dF(CF 3 )ppy] 2 (dtbpy)]PF 6 , and a positive change in dipole moment for [Ir[dF(CF 3 )ppy] 2 (dtbpy)]BAr F 4 . These observations are consistent with a sub-nanosecond reorganization with the PF 6 ⁻ counter-ion, which cancels the dipole moment of the lowest lying excited-state, a process which is absent for the BAr F- 4 counter-ion. Taken together, these observations suggest contact-ion pair formation between the cationic metal complex and the PF 6 ⁻ anion and, at most, solvent-separated pairing with BAr F- 4 . The dynamic ion pair reorganisation we observe with the PF 6 ⁻ counter-ion may substantially modify both the thermodynamic potential available for electron transfer and kinetically inhibit oxidative catalysis, as the anion moves to cover the positively charged end of the molecule, providing a possible mechanistic explanation for recently observed trends in the catalytic activity of these complexes as a function of anion identity in low-polarity solvents. These tunable ion-pair dynamics could prove to be a valuable tool for tailoring the reactivity of both new and extant photocatalysts.
Article
Full-text available
Intermolecular energy transfer processes typically involve an exothermic transfer of energy from a donor site to a molecule with a substantially lower-energy excited state (trap). Here, we demonstrate that an endothermic energy transfer from a molecular organic host (donor) to an organometallic phosphor (trap) can lead to highly efficient blue electroluminescence. This demonstration of endothermic transfer employs iridium(III)bis(4,6-di-fluorophenyl)-pyridinato-N,C2′)picolinate as the phosphor. Due to the comparable energy of the phosphor triplet state relative to that of the 4,4′-N,N′-dicarbazole-biphenyl conductive host molecule into which it is doped, the rapid exothermic transfer of energy from phosphor to host, and subsequent slow endothermic transfer from host back to phosphor, is clearly observed. Using this unique triplet energy transfer process, we force emission from the higher-energy, blue triplet state of the phosphor (peak wavelength of 470 nm), obtaining a very high maximum organic light-emitting device external quantum efficiency of (5.7±0.3)% and a luminous power efficiency of (6.3±0.3)lm/W. © 2001 American Institute of Physics.
Article
New five complexes of the type of [RuL3−x(dmby)x]X2(x=1,2,3, L=2,2′-bipyridyl or 1,10-phenanthroline, dmby=3,3′-dimethyl-2,2′-bipyridyl, X=halide ion) have been synthesized in order to investigate the effects of two methyl groups of dmby on the absorption and emission spectra, luminescence quantum yields, and lifetimes. Values of the radiative and nonradiative rate constants have been calculated from these data at 77 K. Although the absorption and emission maxima and the lifetimes are not much affected by the dmby ligand substitution, the molar extinction coefficients and emission quantum yields are decreased compared with trischelated complexes of the parent bipyridyl or phenanthroline ligands. At 25 °C the emission yields of the complexes containing dmby decrease by 3–4 orders of magnitude than at 77 K. Possible causes of the decrease in the quantum yields are discussed.
Article
Several complexes of Ir(III) containing both the bidentate N-coordinating ligand 2,2'-bipyridine (bpy) and the N,C-orthometalating ligand 2-phenylpyridine (ppy) have recently been prepared; these include the two species Ir(ppy)â(bpy)/sup +/ (A) and Ir(ppy)(bpy)â/sup 2 +/ (B). The former was prepared from the dichloro-bridged dimer, (Ir(ppy)âCl)â, by modification of the procedure of Nonoyama while the latter was obtained by reaction of cis-(Ir(bpy)â(OSOâCFâ)â) (CFâSOâ) with ppy in refluxing 2-ethoxyethanol. The purity of the complexes was monitored with thin-layer chromatography using silica gel plates and 1:1:1 acetone/methanol/water mixtures for elution. Samples of the complexes used in these studies showed only one component in thin-layer chromatography. While only one isomer of B is possible, there are three possible isomers of A. Data from ¹H and ¹³C NMR experiments indicate that A has Câ symmetry. The NMR spectrum indicates, as does thin-layer chromatography, that only a single isomer of A is present with no detectable impurities due to a mixture of isomers. While X-ray structural data for A are lacking, structural data for related complexes suggest that A is the isomer with cisoid metal-carbon bonds and bpy metal-nitrogen bonds transoid to the metal-carbon bonds and bpy metal-nitrogen bonds transoid to the metal-carbon bonds. These species were prepared in order to probe further the effects of metal-carbon bonding on energy-transfer processes and electron-transfer reactions of metal complexes. Emission spectroscopic studies reported here reveal unusual and distinct intramolecular energy-transfer behavior in these complexes. Whereas dual emission from the former is observed in glasses at 77 K, a single emission is observed in the latter.
Article
The electrochemical and UV-visible absorption and emission properties of new heteroleptic ortho-metalated complexes of Rh(III) and Ir(III) are reported and compared to those of similar complexes studied previously in the literature. These novel compounds contain two ortho-metalating ligands (ortho-C-deprotonated forms of 2-phenylpyridine, PPY) and a polyazaaromatic ligand such as 1,4,5,8-tetraazaphenanthrene (TAP) or 1,4,5,8,9,12-hexaazatriphenylene (HAT). The electrochemical behavior, studied by cyclic voltammetry, shows the presence of two reversible reductions, both on a TAP or HAT ligand, and of an irreversible oxidation. Interestingly, contrary to the Rh complexes studied previously in the literature, the TAP- and HAT-Rh compounds present emissions clearly characteristic of distorted CT states to the TAP or HAT ligand, as shown by the unstructured emission spectra at room temperature and 77 K, by their hypsochromic shifts from fluid solutions to rigid glasses at 77 K, and by the luminescence lifetimes. On the basis of these data and due to the irreversible behavior of the oxidation wave, these CT states for Rh complexes are described in terms of sigma-bond-to-ligand (Rh-C sigma bond) charge-transfer (SBLCT) transitions. The iridium HAT complex, like the other similar Ir complexes examined in the literature, shows a dual emission at 77 K in a frozen glass, indicated by two bands in the luminescence spectrum with two distinct lifetimes. They are discussed as originating from two excited states: one metal-to-ligand charge-transfer state (3MLCT), [Ir(IV) HAT-]*, and one SBLCT state, both with a charge transfer to the accepting HAT pi* orbital.
Article
We demonstrate very high efficiency electrophosphorescence in organic light-emitting devices employing a phosphorescent molecule doped into a wide energy gap host. Using bis(2-phenylpyridine)iridium(III) acetylacetonate [(ppy)2Ir(acac)] doped into 3-phenyl-4(1′-naphthyl)-5-phenyl-1,2,4-triazole, a maximum external quantum efficiency of (19.0±1.0)% and luminous power efficiency of (60±5) lm/W are achieved. The calculated internal quantum efficiency of (87±7)% is supported by the observed absence of thermally activated nonradiative loss in the photoluminescent efficiency of (ppy)2Ir(acac). Thus, very high external quantum efficiencies are due to the nearly 100% internal phosphorescence efficiency of (ppy)2Ir(acac) coupled with balanced hole and electron injection, and triplet exciton confinement within the light-emitting layer. © 2001 American Institute of Physics.
Article
An optical oxygen sensor based on the luminescence intensity changes of tris(2-phenylpyridine anion) iridium(III) complex ([Ir(ppy) 3 ]) immobilized in fluoropolymer, poly(styrene-co-2,2,2-trifluoroethyl methacrylate) (poly(styrene-co-TFEM)) film was developed. The luminescence intensity of [Ir(ppy) 3 ] in poly(styrene-co-TFEM) film decreased with increase of oxygen concentration. The ratio I 0 /I 100 (a sensitivity of the film, where I 0 and I 100 represent the detected luminescence intensities from a film exposed to 100% argon and 100% oxygen, respectively) value of [Ir(ppy) 3 ] in poly(styrene-co-TFEM) film is estimated to be 15.3. A Stern–Volmer plot of the [Ir(ppy) 3 ] film exhibited considerable linearity (r 2 = 0.998) and the Stern–Volmer constant, K SV is estimated to be 0.02 Torr −1 , indicating that a highly sensitive oxygen sensor is developed using [Ir(ppy) 3 ] immobilized in poly(styrene-co-TFEM) film. © 2001 Elsevier Science B.V. All rights reserved.
Article
A novel electroluminescent device is constructed using organic materials as the emitting elements. The diode has a double‐layer structure of organic thin films, prepared by vapor deposition. Efficient injection of holes and electrons is provided from an indium‐tin‐oxide anode and an alloyed Mg:Ag cathode. Electron‐hole recombination and green electroluminescent emission are confined near the organic interface region. High external quantum efficiency (1% photon/electron), luminous efficiency (1.5 lm/W), and brightness (≫1000 cd/m<sup>2</sup>) are achievable at a driving voltage below 10 V.
Article
Oxygen quenching of the luminescence of mononuclear and dinuclear Ir(III) cyclometalated complexes immobilized in the pPEGMA matrixes has been studied. Linear Stern-Volmer plots, even when experiments at different emission wavelengths have been performed, were evidenced. Despite the different luminescence lifetimes of the chromophores in the absence of quencher, similar Stern-Volmer slopes have been calculated. This behavior was tentatively interpreted by taking into account the size and charge of the chromophores. Increased sizes and lower charges seem to enhance the sensitivity of the systems. Such findings could be of interest for the design of new solid-state luminescent oxygen sensors with improved performance.