Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Coordination Chemistry Reviews 293–294 (2015) 19–47

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

On the design of highly luminescent lanthanide complexes


Jean-Claude G. Bünzli a,b,∗
a
École Polytechnique Fédérale de Lausanne, Institut des Sciences Chimiques et Ingénierie, BCH 1402, CH-1015 Lausanne, Switzerland
b
Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, People’s Republic of China

Contents

1. Lanthanide luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.1. Historical landmarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.2. Mechanisms of lanthanide luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.2.1. Interconfigurational 4fn ↔4fn 5d1 transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.2.2. Charge-transfer transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.2.3. Intraconfigurational f–f transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.3. Applications of lanthanide luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2. Optimization of energy transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1. Stating the problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2. Theoretical modelling for EuIII complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3. Role of triplet states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4. Role of singlet states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.5. Role of charge-transfer states and photo-induced electron transfers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5.1. Intraligand charge transfer states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5.2. Ligand-to-metal charge-transfer states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.5.3. Metal-to-ligand charge-transfer states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.5.4. Relationship with redox properties (photoinduced electron transfers). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3. Minimizing radiationless vibrational de-activation processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1. Quenching by OH vibrations and hydration number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2. Influence of weak intermolecular interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.3. Vibrational quenching of NIR-luminescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.4. Revisiting the energy gap law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4. Can the radiative lifetime be tuned? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5. Highly luminescent LnIII -containing coordination compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.1. Visible-emitting ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.2. Near-infrared emitting ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.3. Dual emissive ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

a r t i c l e i n f o a b s t r a c t

Article history: Presently, phosphors and luminescent materials for lighting, telecommunications, displays, security inks
Received 4 September 2014 and marking, as well as for probes in biosciences represent one third of the total value of the lanthanides
Received in revised form 20 October 2014 used worldwide. If optical glasses and laser materials are added, this figure is close to 40%, explaining the
Accepted 30 October 2014
large interest that the scientific community is devoting to such materials. The present review focuses on
Available online 7 November 2014
the design of highly luminescent lanthanide complexes and discusses all aspects needing optimization.
Reference is made to the mastering of the various energy migration processes in luminescence sensi-
Keywords:
tization by organic ligands, to minimizing non-radiative deactivation mechanisms, as well as to other
Lanthanide
Luminescence
parameters such as the radiative lifetime, the refractive index, and the benefit of inserting luminescent

∗ Correspondence to: École Polytechnique Fédérale de Lausanne, Institut des Sciences Chimiques et Ingénierie, BCH 1402, CH-1015 Lausanne, Switzerland.
E-mail address: jean-claude.bunzli@epfl.ch

http://dx.doi.org/10.1016/j.ccr.2014.10.013
0010-8545/© 2014 Elsevier B.V. All rights reserved.
20 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Design complexes into inorganic-hybrid structures. Comparative tables list the most luminescent complexes
Energy transfer emitting in the visible and near-infrared ranges and the best chromophores are pointed out.
Radiationless transitions © 2014 Elsevier B.V. All rights reserved.
Quantum yield
Lifetime
Radiative lifetime
Luminescent complexes
Model calculations

1. Lanthanide luminescence Welsbach demonstrated in 1891 that addition of 1% CeO2


to ThO2 leads to much brighter white light emission upon
A search for lanthanide luminescence and associated terms in heating; he therefore created the incandescent gas mantle,
any bibliographic database or web search engine will typically also known as the Auer light. In fact since visible emission
return tens of thousands of articles, pointing to the popularity is larger than expected from the black body curve in these
of the field. Indeed, applications as diverse as lighting, telecom- mantles, while infrared emission is minimized, it seems that
munications, road marking, or immunoassays rely on lanthanide thermoluminescence is operating in addition to incandes-
luminescence. In this review article, we intend to highlight salient cence, so that the resulting complex phenomenon is termed
features of the emission of light by lanthanide coordination com- candoluminescence. Modern versions of these mantles are
pounds, with special emphasis on the requirements needed to still in use today.
design highly emissive materials. It is noteworthy that detection (ii) The red phosphor. Another important discovery was made at
of lanthanide-containing single particle or molecule is at hand, the turn of the 20th century when Georges Urbain studied
using relatively simple instrumentation; in particular, it has been the luminescence of EuIII ions diluted in various matrices
demonstrated for Ln-doped oxide nanoparticles [1], upconverting [12]. This eventually led to the discovery of the extraordinary
nanocrystals [2,3], and, quite recently, for an europium chelate bright, orange-red emitting phosphor Y2 O3 :Eu (4–6 mol%)
[4]. This is useful for multicolour barcoding [5,6] and in biologi- that is close to the ideal prime-colour red-emitter. This phos-
cal sensing and imaging: for instance, a single nanoparticle can be phor, with quantum yield close to 1 [13], is a real nature’s
used to monitor the production of reactive oxygen species inside gift. It has been used in fluorescent lamps and cathode-ray
live cells [7]. tubes since the early 1960s and is still providing brilliant red
Emission of light by chemical compounds or materials stems colour for light-emitting diodes, displays of all kind, includ-
from two different mechanisms: incandescence, or black-body ing flat-panel televisions, despite the large variety of potential
emission, which does not depend on the chemical nature of the substitute materials tested so far, but with limited success.
material but only on its temperature, and luminescence or “cold (iii) Solving the puzzle of lanthanide optical spectra. In 1937, J.H. Van
light emission” [8] that involves quantified energy levels in the Vleck deciphered the long intriguing and obscure nature of
sample. Luminescence can be excited in different ways and specific the sharp optical transitions observed for most of the trivalent
terms are used accordingly; to name but a few: thermolumines- lanthanide ions [14]. He stated that these transitions are for-
cence is consecutive to heating of the sample, liberating energy bidden and postulated that the most intense ones are electric
trapped in defects; photoluminescence results when the sample dipole in nature and caused by a “distortion of the elec-
is irradiated with photons; electroluminescence is produced upon tronic motion by crystalline fields”. He also established that
excitation of the sample in an electric discharge or in an elec- this “distortion” arises only if the ligand field does not have
tric field; chemiluminescence (bioluminescence) arises from the an inversion centre, otherwise orbital mixing cannot occur.
energy released by a chemical (biological) reaction. In the follow- Probabilities for quadrupolar, magnetic dipole, and electric
ing, however, we exclusively concentrate on photoluminescence, dipole transitions were estimated. Finally, Van Vleck pointed
abbreviated luminescence, and on f–f transitions. Depending on the out that interaction between vibrations and electron motion
emission mechanism, luminescence can be divided into two pro- results in additional vibronic lines in the spectra and/or in
cesses: fluorescence is defined as a fast, spin-allowed phenomenon broadening of some lines. In a way, this contribution can be
while phosphorescence is a slow, spin-forbidden emission. Lan- compared with the order brought in the forest of elements by
thanide ions display fluorescence, phosphorescence or, often, both the Periodic Table and it paved the way for Diecke’s diagram,
types of luminescence [9]. ligand-field modelling, and Judd–Ofelt theory.
(iv) The antenna effect. Molar absorption coefficients (ε) of
1.1. Historical landmarks trivalent lanthanide ions are small, usually in the range
0.1–10 M−1 cm−1 so that even if quantum yields (Q)
As for many other elements, spectroscopy, particularly atomic reach sizeable values, the overall luminosity (L = ε × Q) of
spectroscopy but, also, luminescence of salts and oxides, has played LnIII -containing luminescent compounds remains small (typ-
a major role in the discovery and precise identification of the ically  10 M−1 cm−1 ) if samples are excited into the f–f
series of elements named lanthanoids or lanthanides (Ln, Z = 57–71) transitions. In 1942, S.I. Weissman [15] discovered that light
between 1803 (cerium) and 1907 (lutetium; promethium was arti- emission in lanthanide complexes with organic ligands can
ficially synthesized in 1947) [10]. The spectroscopic properties of also be triggered when excitation is performed into the ligand
lanthanide ions are unique due to the radial extension of the 4f electronic levels. Energy is then funnelled onto the metal ion
orbitals being smaller than the expansion of filled 5s2 and 5p6 excited states by intramolecular energy transfer. The energy
sub-shells. This feature confers Ln ions a special status in photon- transfer mechanism is highly complex and will be dealt with
ics with respect to light generation, amplification, and conversion in details in Section 2. This finding, usually referred to as the
[11]. In our opinion, the field has been shaped by eight important antenna effect or luminescence sensitization, proved to have
discoveries that are briefly described in this section. a major impact on the design of Ln-containing luminescent
compounds and materials since luminosity can now easily
(i) The Auer light. The first industrial application of lanthanides reach 104 –105 M−1 cm−1 , being essentially determined by the
has taken advantage of cerium emission when Carl Auer von molar absorption coefficients of the surrounding ligands.
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 21

(v) The Judd–Ofelt theory. In 1962, two authors, B.R. Judd [16] resulting in much improved signal-to-noise ratios, and, last
and G.S. Ofelt [17] independently submitted, within a month, but not least, little or no photobleaching. In addition, cum-
two manuscripts describing a similar model for the empiri- bersome elimination of (low activity) radioactive wastes
cal parameterization of the intensities of lanthanide optical is avoided. Detection limits in the fM range are attain-
spectra. The theory considers a free ion with a well-defined able. The technology that involves a two-step protocol, met
single ground-state electronic configuration perturbed by a immediate success [27] and was later simplified thanks to
static electric ligand-field (LF) generated by the chemical Förster resonant energy transfer allowing one-step homo-
environment of the ion. While purely magnetic dipole tran- geneous assays [28]; nowadays, immunoassays based on
sitions can be calculated in a straightforward way, electric lanthanide luminescence are ubiquitous in medical labora-
dipole intensities are fitted to a set of three “Judd–Ofelt” (JO) tories. Furthermore a logical extension is to apply similar
parameters, ˝2 , ˝4 , and ˝6 that allow comparing covalence protocols to bioimaging thanks to time-resolved lumines-
contributions to the Ln-ligand bonds. Several assumptions are cence microscopy [29,30].
assumed to derive expressions leading to the determination (viii) Erbium-doped fibre amplifier. The ErIII ion has a rich elec-
of these parameters from 4f–4f absorption spectra: (i) elec- tronic configuration with many equally separated levels,
tronic states are completely degenerate in quantum number allowing the design of multi-level lasers in addition to its
J, (ii) the energy difference between the 4fn and 4fn−1 5d1 elec- role in upconversion. In 1986, the first tuneable continuous-
tronic configurations is equal to the difference in the average wave and Q-switched erbium-doped fibre laser was described
energies of the 4f and 5d sub-shells, (iii) all LF sub-levels are [31] which, combined with the earlier work of Digonnet
equally populated, and (iv) the material is isotropic. Despite and Shaw on tuneable optical fibre couplers [32], led to
all hypotheses involved, the theory is remarkably accurate, the development of erbium-doped fibre amplifiers (EDFAs)
except for some ions, such a TbIII for instance. Once JO param- and planar waveguides together with associated optical-fibre
eters are known, transition probabilities between two states telecommunication networks [33]. In EDFAs, ErIII is excited
with J and J quantum numbers, AJ ,J , can be evaluated, as either at 980 nm, on the 4 I11/2 level, or at 1480 nm, on a
well as the radiative lifetime  rad = 1/krad (vide infra) and the metastable ligand-field sub-level of the 4 I13/2 electronic level.
branching ratio ˇJ ,J [18]. This results in a laser effect with emission in the range
(vi) Upconversion. Emission of light in compounds and materi- 1530–1550 nm depending on the fine-tuning of the materi-
als usually follows Stokes’ law which states that emitted als into which ErIII is doped. The 980-nm band has higher
photons have a smaller energy than excitation photons. In absorption cross-section, so it is generally used when low-
a way this is equivalent to say that the efficiency of the noise performances are required; the transition is narrow
luminescence process cannot be larger than one. In 1966, and therefore wavelength-stabilized laser sources are needed
F. Auzel suggested that energy transfer can occur between to pump energy into it. The 1480-nm band features a lower
neighbouring Ln ions that are both in their excited states absorption cross-section but is broader, which makes it ideal
[19]; this sequential energy transfer is more efficient than for high-power amplifiers. Many EDFAs use a combination
excited-state absorption (or 2-step absorption) and rep- of both pump wavelengths. Owing to restrictions imposed
resents therefore a major mechanism in upconversion, a by silica-based EDFAs, namely low absorption cross-section
process in which two (or more) low-energy photons are of ErIII and concentration quenching preventing the use of
combined resulting in the emission of one higher-energy metal-ion concentrations larger than 1020 ions per cm3 , alter-
photon. Initially developed for infrared quantum counters, native matrices and materials are presently investigated.
upconverting materials were soon found to have other appli- This includes co-doping YbIII for larger cross-sections and
cations. In particular, ytterbium, which has a relatively large designing polymer fibres containing antenna ligands, a dif-
absorption coefficient at 980 nm (around 10–12 M−1 cm−1 ) ficult task since C H vibrations have to be eliminated as far
is an ideal sensitizer for upconverted luminescence of ErIII as possible because they easily quench ErIII luminescence;
and HoIII (green and red) or TmIII (UV and blue) [20]. YbIII - important progress has recently been made towards this goal
and LnIII -doped materials (Ln = Ho, Er, Tm) have been there- [34]. Other extensions feature the integration of both ErIII -
fore used in the manufacture of lasers and, also, in security doped waveguide amplifier and laser on a single chip [35]
inks and markers. Present developments include upconvert- and the fabrication of microstructures with very high gain
ing nanoparticles (or nanophosphors, UCNPs) which are very [36].
useful in bioanalyses [21], bioimaging [22], drug release [23],
barcodes and counterfeiting tags [5], as well as, possibly, 1.2. Mechanisms of lanthanide luminescence
solar energy conversion [11,24]. The opposite phenomenon
to upconversion, downconversion (or quantum cutting) in The electronic configuration of trivalent lanthanide ions is
which one high-energy photon is transformed into two [Xe]4f n (n = 0–14); some divalent ions have similar f-electronic con-
lower-energy ones, was documented in 1997 [25] and has figuration [Xe]4f n + 1 (n = 3, 5, 6, 9, 12, 13; Ln = Nd, Sm, Eu, Dy, Tm,
potential applications in lighting and energy conversion Yb), while the electronic configuration of the others is [Xe]4f n 5d1
processes. (n = 0, 1, 2, 7, 8, 10, 11, 14; Ln = La, Ce, Pr, Gd, Tb, Ho, Er, Lu).
(vii) Time-resolved luminescence immunoassays. At the beginning of The 4f n configurations generate a wealth of electronic levels char-
the 1980s, an original idea came from the city where the first acterized by three quantum numbers, S, L, and J, and that are
rare-earth element, yttrium, was identified, Turku in Finland: further split by weak ligand field effects, on the order of 102 cm−1 ,
E. Soini and I. Hemmilä proposed replacing radioactive- since the 4f orbitals are shielded: for instance, EuIII possesses 3003
based immunoassays by time-resolved luminescent assays ligand-field (or Stark) sub-levels. When d-orbitals are implied,
using a lanthanide chelate conjugated to some specific anti- the ligand field generated by the surrounding ligands is much
body as bioprobe [26]. Not only are these analyses taking larger (103 –104 cm−1 ). The arduous task of finding, calculating and
advantage of intramolecular energy transfer to excite the assigning energy levels in transition metal containing compounds
metal ion but, also, they fully exploit specific properties of started to be undertaken in the 1940s, particularly by Racah, and
lanthanide luminescence: wavelength discrimination thanks enabled quantitative fits of the LF sub-levels for the entire lan-
to large ligand-induced Stokes’ shifts, time discrimination, thanide series to be performed thanks to the work of B.G. Wybourne
22 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

48 2D
λ / nm 5/2

225 250 300

44
ε / 102 M-1cm-1

E / 103 cm-1
4
2D
3/2
2 40
0
48 44 40 36 32
E / 103 cm-1 2F
2 7/2

2F
0 5/2

Fig. 1. Left: absorption spectrum of [Ce(H2 O)9 ]3+ doped into lanthanum ethylsul-
fate; redrawn from Ref. [43]. Right: corresponding energy scheme.

Fig. 2. Emission colours of EuII - and CeIII -doped thiosilicates (1 mol%).


[37]. The first synopsis of the 4fn energy levels of all trivalent lan- Reproduced with permission from Ref. [45]; © 2010 MDPI publishers.
thanides doped in various crystals and spanning the entire UV,
visible and NIR spectral ranges was given by G.H. Dieke [38]. The so that 9 DJ ↔7 F6 transitions are spin forbidden while those between
“Dieke diagram” has now been expanded to include higher-energy 7 D and 7 F are spin-allowed. Therefore if TbIII is coordinated to soft
J 6
levels [39]. In addition, theoretical work by Judd and Ofelt [40] nitrogen ligands, the latter transitions can be red-shifted as low as
allowed describing, comparing, and predicting electronic spectra of 31 000 cm−1 . For a listing of 5d1 4fn−1 ←4fn transitions of LnIII ions
lanthanide ions in isotropic crystals, glasses, and coordination com- (Ln = La Dy, Er Yb) into inorganic matrices, see refs. [46,47].
pounds. Presently, the spectroscopy of lanthanide ions with respect
to both energy levels and transition intensities is fairly well under- 1.2.2. Charge-transfer transitions
stood and will not be developed here. The reader is referred to one Ligand-to-metal charge-transfer (LMCT) transitions are essen-
of the several review articles dealing with the subject [9,18,40,41]. tial in sensitizing the luminescence of lanthanide-containing
There are three types of optical transitions in lanthanide com- inorganic phosphors but as for 5d–4f transitions, their energy is
pounds: allowed f↔d and charge-transfer transitions, as well as large and corresponding bands usually appear in the VUV and UV
forbidden f↔f transitions. ranges [46]. Easily reducible trivalent ions such as SmIII , EuIII , and
YbIII have the lowest LCMT transitions: 200–250 nm for 2p(O)→4fn
1.2.1. Interconfigurational 4fn ↔4fn 5d1 transitions charge transfer [47], but this can be modulated by the nature of
For the free ions (LnIII or LnII ), these transitions lie usually in coordinated ligands and, in the case of EuIII , transitions with ener-
the VUV or UV spectral ranges. Due to the extension of the 5d gies as low as 20–25 000 cm−1 have been reported for dinuclear
orbitals, ligand-field splitting can however push some of the Stark complexes with calix[8]arene, with molar absorption coefficients
sub-levels deep into the visible; this is particularly true for CeIII around 720–740 M−1 cm−1 . The presence of such a low-lying CT
and EuII compounds. Spectra are difficult to interpret in that a good state induces mixing with the 4f-states and results in larger oscil-
deal of their intensity rests in vibronic components and both 4f and lator strengths for the 4f–4f transitions [48]. In the case of YbIII , the
5d LF splitting have to be taken into account. As a consequence, LMCT state has energy blue-shifted by ∼5000 cm−1 compared with
f↔d transitions are broad, intense, energy-tuneable, and feature EuIII , which approximately corresponds to the difference in normal
large Stokes’ shifts while the 4fn 5d1 excited state lifetime is short, reduction potentials of these ions (∼0.7 eV, 5600 cm−1 ). The role of
typically in the 10–100 ns range. As an example, the absorption LMCT states in sensitizing or quenching LnIII luminescence will be
spectrum of [Ce(H2 O)9 ]3+ in water (idealized D3h symmetry) is discussed in Section 2.5, but we would like to mention here that a
displayed in Fig. 1 along with the corresponding energy scheme, few CeIV compounds emit light upon excitation into O(2p)→Ce(4f)
exemplifying the (J + 1/2) ligand-field splitting of the d- and f- CT transition, yielding a transient reduced CeIII species, e.g. the
orbitals. Note the scale change due to the much larger 5d-splitting Lindqvist polyoxolanthanate [Ce(W5 O18 )2 ]8− [49].
compared with 4f-splitting. At low temperature, two luminescence Metal-to-ligand charge-transfer (MLCT) states are very com-
bands are observed at 30 000 and 32 000 cm−1 for [Ce(H2 O)9 ]3+ mon in d-transition compounds but they are rarely observed
doped into lanthanum ethylsulfate [42], assigned to emission from with lanthanides because they lie at very high energy, except for
2D 2
3/2 to F7/2,5/2 . The corresponding lifetime in water is 45 ns [43]. CeIII which can easily be oxidized into Ce(IV): in tris(pyrazine-
The tuneability of d–f luminescence is demonstrated in a series 2-carboxylato)cerium(III), a weak absorption band at 388 nm
of EuII -doped inorganic matrices in which emission maxima range observed in mixed CH3 CN/DMF solvent is thought to arise from
from 330 to 520 nm, albeit thanks to some mixing with conduction- a MLCT transition from 4f1 to the ␲* orbital of the pyrazine-
band states of the host material [44]. When sulfides are used as 2-carboxylate ligand, known as being an electron acceptor;
matrices, the emission wavelength can even be extended to 650 nm conversely, the weak blue-green luminescence exhibited by this
[45]. The versatility of such materials is demonstrated in Fig. 2 in compound has been attributed to the MLCT state [50]. TbIII is also
which the coordinates of several EuII - or CeIII -doped thiosilicates a potential candidate for MLCT transitions.
are reported on the CIE (Commission Internationale de l’Éclairage)
trichromatic diagram. The TbIII ion (4f8 ) represents an interest- 1.2.3. Intraconfigurational f–f transitions
ing case. Since the number of valence electrons is larger than 7, Description of the interaction between photons and matter
both low- and high-spin 4f7 5d1 electronic configurations can be considers the former behaving as waves comprised of two perpen-
generated depending on the surrounding ligands. The high-spin dicular fields, electric and magnetic, oscillating in time. When a
configuration (9 DJ ) has lower energy than the low-spin one (7 DJ ) photon is absorbed, its energy is transferred to an electron which
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 23

Table 1
Selection rules and approximate relative intensities Irel for f–d and f–f transitions [53].

Transition Parity S |L| |J| Irel

ED, 4f–5d Opposite 0 ≤1a ≤1c 0.01–1


Forced ED, f–f Same 0 ≤6a (2,4,6 if L or L = 0) ≤6c (2,4,6 if J or J = 0) 10−4
MD, f–f Same 0 0 ≤1c 10−6
EQ Same 0 ≤2b ≤2d 10−10
ED vibr, f–f Same 0 ≤6a (2,4,6 if L or L = 0) ≤6c (2,4,6 if J or J = 0) 10−7 –10−10
a
L = 0↔L = 0 forbidden.
b
L = 0↔L = 0,1 forbidden.
c
J = 0↔J = 0 forbidden.
d
J = 0↔J = 0,1 forbidden.

is then lifted into an orbital with higher energy. The absorption is equation assumes that the absorption bands are symmetrical, i.e.
promoted by operators linked to the nature of light: the odd-parity either Gaussian or Lorentzian. If not, Eq. (3) has to be replaced with:
electric dipole (ED) operator, the even-parity magnetic dipole (MD)   
and electric quadrupole (EQ) operators. Laporte’s parity selection 1036 9n ε(˜)
D(exp) = · (2J + 1) · · d˜ (4)
rule states that levels with same parity cannot be connected by 108.9 · XA (n2 + 2)
2 ˜
electric dipole transitions; as a consequence f–f transitions are
forbidden by the ED mechanism. However, under the influence of As for absorption, emission of light through f–f transitions is
a ligand-field, non-centrosymmetric interactions result in the mix- achieved by either ED or MD mechanisms and the selection rules
ing of electronic states of opposite parity, which somewhat relaxes detailed in Table 1 apply. Important parameters characterizing the
the selection rules and the transitions become partially allowed; emission of light from an LnIII ion are the lifetime of the excited
they are called induced (or forced) electric dipole transitions. Some state  obs = 1/kobs and the quantum yield Q:
induced ED transitions are highly sensitive to minute changes in
the LnIII environment and are called hypersensitive (for a listing, number of emitted photons
Q = (5)
see [51]). number of absorbed photons
On the other hand, magnetic dipole transitions are allowed, but
their intensity is weak; in 4f–4f spectra however they frequently The quantum yield depends on the rate at which the excited
have intensity of the same order of magnitude as induced electric level is depopulated, kobs , and on the radiative rate krad :
dipole transitions. Quadrupolar transitions are also parity allowed,
but they are weak and have been rarely identified unambiguously. krad 
Ln
QLn = = obs (6)
Mathematical treatment of the parity mixing by the ligand-field kobs rad
perturbation leads to the selection rules for f–f transitions listed in
Table 1. Selection rules are derived under several hypotheses which where subscript and superscript “Ln” mean that excitation has been
are not always completely fulfilled in real compounds so that the performed directly into the 4f excited state and the quantity defined
terms “forbidden” and “allowed” transitions should be understood in Eq. (6) is called the intrinsic quantum yield (sometimes, inter-
as “low probability” and “high probability” transitions. According nal quantum efficiency). The rate constant kobs is the sum of the
to Judd–Ofelt theory the dipole strength in esu2 cm2 (=1036 debye2 ) rate constants characterizing the various radiative and radiation-
of an induced ED f–f transition between states  and   is given less deactivation processes:
by: 
i
kobs = krad + knr (7)

DED = e2 · ˝ · |
 ||U  ||  |2 (1) i

=2,4,6
Determination of QLnLn , which is difficult to measure experimen-

in which e is the electric charge of the electron, ˝ are phenomen- tally, is frequently conducted using Eq. (6), requiring evaluation of
ological Judd–Ofelt parameters, expressed in cm2 and calculated the radiative lifetime:
from the absorption spectrum ε(˜), and U are the irreducible ten- 2

sor forms of the ED operator. The bracketed expressions in Eq. (1) 1 644 ˜ 3 n(n2 + 2)
= krad = DED + n3 DMD (8)
are dimensionless doubly reduced matrix elements which are tab- rad 3h(2J + 1) 9
ulated [52] and independent of the ion environment.
Oscillator strengths for magnetic dipole transitions can be cal- Except in few cases, this calculation is not trivial and errors may
culated from the following equation where h is Planck’s constant: occur, including those pertaining to the hypotheses made within
Judd–Ofelt theory. In particular, it has been assumed that the emit-
 e·h
2 ting and receiving levels are (2J + 1)-fold degenerate or, if split by
DMD = · |
 ||L + 2S||  |2 (2) ligand-field effects, that all the sub-levels are equally populated.
4 ·  · me · c
This is obviously not true and in the case of TbIII and ErIII this may
The experimental dipole (or oscillator) strength is defined as: lead to errors up to 20–100%.
On the other hand, if the absorption spectrum corresponding
  
1036 9n to an emission spectrum is known, which may be the case when
D(exp) = · (2J + 1) · · ε(˜)d˜ (3) the luminescence transitions terminate onto the ground level, the
108.9 · ˜ mean · XA (n2 + 2)
2
radiative lifetime can be simply calculated from the following equa-
tion where NA is Avogadro’s number:
with XA being the fractional   of the initial state while
population
˜ mean is given by ˜ mean = ˜ · ε(˜)d˜/ ε(˜)d˜; (2J + 1) is the degen- 
1 8cn2 ˜ 2 (2J + 1)
eracy of the initial state and the expression involving the refractive = 2303 × ε(˜)d˜ (9)
index n is known as Lorentz’s local-field correction. The above rad NA (2J  + 1)
24 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Table 2 telecommunications and laser materials. The best known laser


Experimental intrinsic quantum yields, observed and radiative lifetimes of
is the neodymium YAG laser emitting at 1.06 ␮m, which can be
M3 [Ln(dpa)3 ] (M = Na or Cs) samples in solution (0.1 M Tris–HCl, pH 7.4) and solid
state at 295 K; 2 are given between parentheses [56]. doubled (532 nm), tripled (355 nm), or quadrupled (266 nm) and
which is seen in uses as disparate as laser pointers, military target
Ln
Sample QLn /%  obs /ms  rad /msa
designation, rangefinders, and weapons, experimental setups for
(i) (ii) (iii) controlling nuclear fusion, or medicine, in oncology, ophthalmol-
−2
Eu, 1.8–3.7 × 10 M 41(2) 1.7(0.1) 4.1(3) 3.15 4.0 ogy or thermotherapy. Another well-established use of lanthanide
Eu, solid stateb 68(4) 1.8(0.1) 2.6(2) – 2.7 luminescence is in time-resolved bioanalyses and bioimaging. In
Tb, 2.0 × 10−2 M c
1.74(1) – 1.0 – parallel, security inks and counterfeiting tags are more and more
Tb, solid stateb 72(5) 1.36(2) 1.9(1) – – relying on lanthanide luminescence, particularly since the advent
Yb, 4.04 × 10−2 M c
2.23(1)d – – 1.31(2)
of UCNPs. Potential new applications include solar energy conver-
a
(i) experimental, Eq. (6), (ii) from JO theory, Eq. (8), (iii) from Eqs. (10) for Eu and sion [24], luminescence thermometry [59], damage and pressure
(9) for Yb.
b
sensors based on mechanoluminescence [60], optical refrigeration
Refractive index = 1.517.
c
Determination not feasible. for electronic devices [61], and quantum information processing
d
In ␮s. [62].
Until the turn of the 20th century, the search for lanthanide-
containing luminescent materials has essentially proceeded
In the special case of EuIII for which the 5 D0 →7 F1 transition has through a trial-and-error strategy with only some semi-
pure magnetic origin [54], a convenient simplified equation can be quantitative concepts and rules to guide the investigators. But the
derived [55]: development of powerful methods of calculation with good param-
1
I  eter sets for the 4f orbitals, the re-visitation of some initial concepts
tot
= AMD,0 · n3 (10) underlying lanthanide spectroscopy, as well as systematic stud-
rad IMD
ies of the wealth of quantitative quantum yield and lifetime data
with AMD,0 being equal to 14.65 s−1 and (Itot /IMD ) being the ratio reported in the literature are slowly changing the attitude of the
of the total integrated emission intensity from the Eu(5 D0 ) level to scientists and engineers so that the design of highly luminescent
the 7 FJ manifold (J = 0–6), to the integrated intensity of the 5 D0 →7 F1 materials and probes is becoming more programmed. Some of the
transition. challenges and potential answers are described in the next section
Examples of radiative lifetime calculations based on the vari- but one always has to keep in mind that each application has special
ous equations described above are collected in Table 2. The limit of requirements and that each lanthanide ion is a special case!
Judd–Ofelt approach is evident for TbIII (error > 100%). On the other
hand, Eqs. (9) and (10) yield trustworthy results; in particular, the
radiative lifetime for the YbIII complex is in line with those reported 2. Optimization of energy transfer
for other coordination environments, 1.2 ms for [Yb(dtpa)]2− [56]
and 0.7–0.75 ms for complexes with benzoxazole-substituted 8- In this section, we focus exclusively on f–f luminescence and
hydroxy-quinolinates [57]. decipher the various parameters to be optimized upon the assump-
The radiative lifetime depends on the LnIII ion, the specific tion that luminescence sensitization is achieved by the chemical
excited state level giving rise to luminescence, the chemical envi- environment into which the emissive ion is embedded, inorganic
ronment of the metal ion, and the refractive index of the medium; or organic.
it can therefore vary considerably even for a given emitting level,
e.g. between <1 and 14 ms for Eu(5 D0 ).
2.1. Stating the problem
1.3. Applications of lanthanide luminescence
A highly simplified schematic diagram of energy transfer pro-
According to various reports and estimates, luminescent mate- cesses occurring in a luminescent lanthanide material is presented
rials account for about one third (∼1.5–2 billion US$ in 2012) of in Fig. 3. In a first step, the surrounding ligand(s) absorb light and
the commercial value of the rare earths extracted annually, while reach an excited state that is often a singlet excited state in com-
representing ∼7–8% of their tonnage (8–9000 metric t equivalent pounds with organic ligands. In a second, composite step, energy
rare-earth oxides in 2012). A characteristic is that these high value- is funnelled onto a donor state (for example a triplet state) with,
added materials only constitute a fraction of the weight of the ideally, a relatively long lifetime and is then transferred onto one
products they valorize. Typically, a fluorescent bulb contains less or several LnIII excited states. Finally, after suitable internal conver-
than 1 g of Ln phosphors, a LED lamp or a cell phone has less than sion to an emissive state, metal-centred luminescence is emitted.
0.1 g of rare earths, and a luminescent immunoassay frequently Consequently, a new quantum yield has to be defined, different
from QLnLn , the overall quantum yield Q L :
requires less than 10 ␮g of luminescent lanthanide. Both d–f and Ln
f–f luminescence have found applications. The former in scintilla-
tors for the detection of ionizing radiation and in white-light LEDs L ILn (E)
QLn = (11)
under the form of a yellow phosphor, Y3 Al5 O12 :CeIII (0.033 mol%). IL (A)
Long-persistence materials that feature light emission up to several
hours after excitation commonly rely on broad d–f emission from where ILn (E) represents the number of photons emitted by the
EuII [58]. They are seen in emergency and safety signals, includ- metal ion and IL (A) the number of photons absorbed by the
ing road marking, in dials and displays, and textile printing; more ligand. This quantum yield can at most be equal to the intrinsic
recently they have been introduced into alternative-current LEDs quantum yield QLn Ln described in Eq. (6). Absorption being ligand-

and they are tested for applications in bioimaging and energy stor- centred while emission is Ln-centred, there is usually a large
age. wavelength shift between absorption and emission, the so-called
Luminescence arising from f–f transitions is the basis of ligand-induced Stokes’ shift (or Richardson’s shift [53]), not to be
numerous applications [11]. Previously mentioned lamp and confused with the frequently misused Stokes’ shift (which is very
display phosphors constitute the largest category, followed by small for LnIII ions, due to the inner nature of 4f orbitals). The
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 25

(1 S) operate via dipolar-multipolar interactions. This does not mean


that singlet states do not contribute to the energy transfer process
(see Section 2.4): their importance has been recognized as soon
as 1969, particularly for TbIII sensitization when the ligand triplet
state is lower than the emissive 5 D4 level [64]. As a consequence,
the scheme of Fig. 3 is oversimplified and in reality one should take
into account more than one donor state. Other important levels are
charge-transfer states, either internal to the ligand (e.g. 1 ILCT in
push–pull ligands) or characteristic of the complex, LMCT or MLCT
states. Internal charge-transfer states may serve as energy relay in a
complex cascade of energy transfers [65]. They are often mixed with
singlet states, rendering interpretation of singlet-state excitation
delicate; the same hold for 1 LMCT states [66]. Moreover, in hete-
rometallic edifices featuring both d- and f-transition metal ions,
d levels and 1.3 MLCT states can be convenient feeding levels for
LnIII excitation [67,68]. Finally, in 4f–4f heterometallic compounds,
very efficient transfers between f-levels are operative which can be
exploited to excite a given LnIII ion; examples are TbIII -to-EuIII and
YbIII -to-LnIII (Ln = Ho, Er, Tm) transfers. One important point to real-
ize is that most discussions about energy transfer processes usually
consider energy levels of one ligand, whereas wavefunctions of the
Fig. 3. Simplified scheme depicting the antenna effect, with associated photophys- entire complex edifice should be taken into account.
ical parameters; right: rule of thumb for minimizing vibrational quenching. Key:
Radiationless de-activation processes. Since they open new routes
A = absorption, E = emission, D = donor state.
for de-exciting the metal ion, these processes can be recognized
by their influence on the excited-state lifetime, which is short-
sensitization efficiency, which characterizes the set ligand(s)/LnIII ened. Both vibrations and photo-induced electron transfers being
ion can now be defined as: temperature-dependent, dependence of  obs on temperature is also
L
QLn diagnostic of their presence. The right-hand side of Fig. 3 depicts

sens = Ln
(12) de-activation through vibrations and their overtones. The smaller
QLn
the number of phonons needed to bridge the energy gap, the most
Therefore the following experimental photophysical parame- efficient is the luminescence quenching and as a rule of thumb, and
ters are needed to fully characterize the process: QLn Ln , Q L ,  , ,
Ln obs rad ideally, the chemical environment of the emissive metal ion should
and
sens . be devoid of vibrations with energy larger than 1/6 of the energy
From a conceptual point of view, maximizing the luminos- gap. For NIR-emitting ion, this requirement extends to the second
ity of lanthanide-containing materials is fairly straightforward: (i) coordination sphere as well. To avoid quenching by charge-transfer
the metal-ion surrounding must be highly absorbing with large states, the only possibility is to modify the metal-ion coordina-
molar absorption coefficients, (ii) losses in energy transfers have tion environment in such a way that the energy of the CT state is
to be minimized and (iii) radiationless deactivation of the excited increased largely above the emissive state of the metal ion, which
metal ion should be minimized. From a practical viewpoint, how- requires chemical modification of the ligand(s).
ever, the situation is far from being simple. The ligand-to-LnIII
energy transfer process is very complicated in that it may involve 2.2. Theoretical modelling for EuIII complexes
several different mechanisms, exchange, dipolar-dipolar, dipolar-
multipolar, as well as several electronic levels, both from the Luminescent compounds containing EuIII ions are among the
ligand(s) and from the metal ion. Nevertheless, on can say that, most studied because of the intense red emission of this ion. From
overall, the sensitization process starts to be well mastered and a theoretical point of view this ion is also interesting because it
many highly luminescent complexes and materials have been has non-degenerate ground (7 F0 ) and main emissive (5 D0 ) states.
reported during the past decades. Regarding radiationless deactiva- On the other hand, its luminescence is sensitive to the influence of
tion, the experimental command on this aspect is more complicated LMCT states and it presents somewhat uncommon spectroscopic
because not only vibrations intervene but, also, back energy trans- features: (i) the first excited spin-orbit level, 7 F1 , lies only about
fer processes as well as quenching by charge-transfer states that are 300 cm−1 above the ground state, so that at room temperature its
not always easy to identify. The problem of vibrational deactivation population is around 25–30% [54]; this is important because while
is particularly acute for ions emitting in the red and near-infrared the 5 D0 ←7 F0 transition is highly forbidden, selection rules on J
since the energy difference between the emissive and receiving point to 5 D0 ←7 F1 being MD allowed; (ii) moreover the 5 D1 excited
LnIII states can easily be bridged by high-energy vibrations (O H, state is located a mere 1700 cm−1 above 5 D0 , so that it represents
N H, C H, C O, etc.) and their overtones. Finally, the influence an important receiving state in the energy transfer process, since
of the polarizability of Ln-ligand bonds and of the refractive index both 5 D1 ←7 F0 and 5 D1 ←7 F1 are MD allowed. Theoretical 3-step
on the radiative lifetime (see Eqs. (8)–(10)) is only qualitatively modelling of the intricate energy transfers in EuIII complexes with
understood [63]. organic ligands has been initially proposed by G.F. de Sá et al. [69]
Nature of the donor state(s). The overall energy migration paths and further refined by R.O. Freire et al. [70], who designed a friendly
in the complex represent a kinetic scheme and what has really to be user computer package. The model consists in three steps and deals
maximized are the rates of transfers between the initially excited essentially with EuIII complexes with organic ligands, but it will
and the donor (D) ligand state(s), as well as between D and the certainly be extended to other ions in the future since one example
LnIII levels. Therefore it is desirable that D is long-lived and conse- already exists for its application to a SmIII dinuclear helicate [71].
quently, triplet states (3 T, 3 MLCT) are often more likely to be major
actors in the process than singlet states. Triplet states usually trans- Step 1. The geometry of the complex is optimized within the
fer energy through an exchange mechanism, while singlet states frame of the AM1 semi-empirical molecular orbital model. The
26 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Fig. 4. (Left) Partial energy diagram showing electronic levels implied in energy transfer and relevant parameters. (Middle) Overlap between ligand emission and LnIII
absorption spectra. (Right) Definition of the parameters needed for the kinetic model.
Reprinted from Ref. [51].

bonding in lanthanide complexes having predominantly elec- integral JDA between the emission spectrum of the donor (ligand)
trostatic character, it is simulated within a central-potential and the absorption spectrum of the acceptor (EuIII ion) and (ii)
model, known as the Sparkle model. Suitable parameterization the distance between the donor and the acceptor, rDA , according
of Sparkle/AM1 with Gaussian functions has been achieved by to an exponentially decreasing function, e−ˇ·rDA . In the model it
reproducing known structures of several EuIII complexes [69,72], is expressed by the following equation:
and has then been rapidly extended to other LnIII ions; presently, 4
parameters are available for the entire series, from LaIII to LuIII [73]. 8 e2
4f |L
F
˛ J  ||S||˛J
ex 2
WET =
Other versions of the Sparkle model such as Sparkle/PM7 [74] or 3 (2J + 1)G · r 4
DA
Sparkle/RM1 for larger molecules, or even metal-organic frame- 2
 

works [75], have been tested satisfyingly, although Sparkle/AM1
tends to describe best the organic part of the complexes [76]. ×
ϕ| Z (k) sm (k)|ϕ (15)

Step 2 implies calculating the energy and transition moments of m k

the ligand excited states with the INDO/S-CI model implemented


in the ZINDO program. The sparkle is replaced by a +3e point
 r 3.5
0

4f |L = (16)
charge to match the LnIII charge and despite this purely ionic model rDA
excluding any covalent contribution to the ligand-metal bonding,
4
calculated and experimental electronic spectra are in satisfying where
4f |L is a distance-dependent radial overlap where , the
agreement. Other authors have relied on more sophisticated meth- radial overlap integral between the 4f orbitals and the valence
ods for calculating triplet-state energies in complexes, for instance orbitals of the ligating atoms, is about 0.05, and r0 is the short-
ab-initio complete active space self-consistent field (CASSF) or est Ln X ligating distance [79]; G is the degeneracy of the
time-dependent density functional calculations (TD-DFT) [77,78]. ligand ground state, J is the total angular momentum quan-
The final step 3 is devoted to developing a kinetic model including tum number of the LnIII ion, S is the spin operator of the LnIII
analytical expressions for the transfer rates. The rates of transfer, ion, z is the z-component of the ED operator, sm (m = 0,±1) is
WET , are expressed within the general theory proposed by Fermi a spherical component of the spin operator (the index k runs
assuming the validity of Born–Oppenheimer approximation: over the electrons of the ligand). Note that the e−ˇ·rDA distance
 is contained
dependence in
 the right-hand summation in which
2 the ϕ k z (k)sm (k) ϕ matrix elements are calculated by
WET = |
˛ J  ϕ|H| ˛Jϕ |2 · F (13) INDO/S-CI method. The rDA distance is in fact the distance from

the LnIII ion to the region of the ligand in which the donor state is
In this expression,  = h/2, with h being Planck’s constant, H is the
localized; it is evaluated with the following equation in which ci
Hamiltonian operator and other parameters are defined in Fig. 4.
is the molecular orbital coefficient of atom i contributing to the
The temperature-dependent factor F contains a summation over
ligand donor state and ri (L) is the distance from this atom to the
Franck–Condon factors, as well as the dependence of the energy
LnIII ion:
transfer on the energy gap; it is sometimes referred to as the energy 
mismatch factor. Assuming a Gaussian shape for the absorption and c2 · ri (L)
rDA = 
i i
(17)
emission bands, as well as wLn  wL , which is always true for c2
i i
f–f transitions, this factor can be expressed as:

The calculated transfer rates are in the range 107 –108 s−1 for
1 ln 2 E
2 3 T→Eu(5 D ) and ≈107 s−1 for 1 S→Eu(5 L6 ). In the initial model
1
F= exp − ln 2 (14) 4
WL  WL
4f |L in Eq. (15) was replaced by a squared distant-dependent
screening factor, (1 − 0 )2 , which led to overestimate these rates
Energy transfer mechanisms taken into account are: to 1010 –1011 s−1 for the former and ≈109 s−1 for the latter. How-
ever, the rates estimated for Dexter’s mechanism remain much
a. Exchange (Dexter) mechanism. After excitation of the ligand sin- faster than typical intra-configurational 4f–4f decay rates so that
glet state and intersystem crossing, the electron in the triplet conclusions reached in initial works are not altered [79].
state is transferred onto one of the excited LnIII states. Simul- b. Dipole–dipole (Förster) mechanism. After excitation of the ligand
taneously, an electron from the highest occupied 4f orbital is 1 S state and intersystem crossing to the donor 3 T state, the

transferred onto the ligand, filling up the gap created by the ini- excited electron returns to its initial orbital while transition
tial photoexcitation. The rate of transfer depends on the overlap dipole moments of the ligand and EuIII ion couple resulting in
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 27

excitation of the latter. The rate of energy transfer via this mecha- This mechanism is particularly effective if ˝2 is large and rDA
nism depends on (i) the spectral overlap JDA defined above, (ii) the relatively short (<5–6 Å). The other two mechanisms need large
luminescence quantum yield of the donor QD , (iii) the lifetime of ˝4 and ˝6 parameters to contribute substantially to the overall
the donor excited state, (iv) the relative orientation of the donor transfer process.
and acceptor transition dipoles , and (v) the distance between
the donor and the acceptor according to a rDA −6 dependence. The Selection rules are embedded in the reduced matrix elements
corresponding equation reads as follows: which appear in Eqs. (15), (18) and (22):

4 e2 SL (1 − 1 )2    2 |J − J  | ≤  ≤ J + J  (J = J  = 0 excluded)
˝ ˛ J  U ()  ˛J
dd (24)
WET = F (18)
 (2J + 1) G · r 6
DA  for dipole–2 pole and dipole–dipole mechanisms and
where SL is the dipole strength associated with the ␸→␸ J = 0, ±1 (J = J  = 0 excluded) (25)
transition in the ligand, (1 − 1 )2 is the squared Sternheimer
shielding factor due to the filled 5s2 5p6 sub-shells, ˝ are JO for the exchange mechanism.
intensity parameters ( = 2, 4, 6), U() is a unit tensor opera- Finally, the kinetic model expresses the variation in the popu-
tor the reduced matrix elements of which are tabulated [52], lation of the levels implied under steady-state approximation (see
and (1 − 1 )2 is a distance-dependent squared screening factor right part of Fig. 4 for definitions):
[79]. ∂
i  
Experimentally, the efficiency of energy transfer is calculated =− kij
i + kji
j = 0 (26)
∂t
from the lifetimes in the presence,  obs , and in the absence,  0 , of j=1 j=1
transfer and can be related to the D–A distance and to the critical
j=
/ i j=
/ i
Förster distance R0 (corresponding to 50% transfer):
obs 1 The system of equations is solved by Runge–Kutta method and

et = 1 − = 6
(19) the theoretically estimated quantum yield is given by:
0 1 + (rDA /R0 )
Atot ·
j
R06 = 8.78 × 10−25 (2 · QD · n−4 · JD ) [cm−6 ] (20) Q = (27)
 ·
i

F(˜) · ε(˜) · ˜ −4 d˜
JD =  (21) here Atot refers to the sum of the radiative and non-radiative rates
F(˜)d˜ (=1/ obs ) and  is the pumping rate in photons/s.
The model involves quite a few approximations and requires
n is the refractive index and the isotropic limit to  is 2/3; F(˜) is
estimating several parameters, such as non-radiative decay rates
the luminescence spectrum recorded on an energy scale.
between LnIII excited levels (on the order of 106 s−1 ) or the
c. Dipole–multipole (2 ) mechanisms. These mechanisms are less
lifetime of the donor state(s) at the temperature at which the
known, but they can nevertheless contribute greatly to the
transfer occurs. Another point is that the ZINDO procedure largely
energy transfer process [64]. Their general expression is:
overestimates the energy of the singlet state when the excitation
2 e2 SL    2 window is chosen such as to reproduce closely the triplet-state
 ˛ J  U ()  ˛J
mp
WET = F (22)
 (2J + 1)G energy. Overall, however, the model yields reasonably good
 results, often within less than ±20% of the experimental value, as
With   being defined as: shown in Table 3. In fact sets of rate constants are often optimized
2
to reproduce the experimental quantum yield [71] so that the

r  2 predictive power of the model is still somewhat limited. On the
 = ( + 1)
||C () || (1 −  )2 (23)
(+2) 2 other hand, once adequate sets of rate constants are at hand, one
(rDA )
clearly sees by which main route the energy is transferred onto the
where
r  is the radial expectation value of r for 4f electrons metal ion. In most cases, the predominant path for EuIII complexes
( = 3), C() is a Racah tensor operator,  ’s are screening factors involves the triplet state of the ligand, transfer from the singlet
for the 5s2 5p6 subshells, and  = 2, 4, 6. Therefore the distance state being much less efficient. This is illustrated in Fig. 5 for the
dependence is (1/rDA )8 for the dipole–quadrupole mechanism. dimeric complex (NBu4 [Eu(L1)])2 (Scheme 1): transfer from 1 S is

Table 3
Selected data for EuIII complexes in the solid state at room temperature: experimental triplet (E0 –0 (3 T)) or LMCT (ELMCT ) state energies, calculated and experimental quantum
L
yields (QLn , %). Relevant chemical formulae are given in Scheme 1.

Compound E0 –0 (3 T) (ELMCT )/cm−1 L


Calc. QLn /% L
Exp. QLn /% Ref.

[Eu(tta)3 (TPPO)2 ] 20 661 73 73 [85]


[Eu(tta)2 (NO3 )(TPPO)2 ] 21 645 65 68 [85]
[Eu(btfa)3 (H2 O)2 ] 21 595 19 22 [86]
[Eu(btfa)3 (phen)] 21 633 50 38 [87]
[Eu(btfa)3 (phenNO)] 19 600 56 65 [86]
[Eu(btfa)3 (bpy)] n.a. 57.6 43 [87]
[Eu(btfa)3 (4,4 -bpy)(EtOH)] 20 276 46 37.8 [88]
d-U-[Eu(btfa)3 (4,4 -bpy)]a 21 473 52.5 51.0 [88,89]
(NBu4 [Eu(L1)])2 21 320 51 54 [76]
[Eu(L2)3 (H2 O)2 ] 20 200 (24 753) 24.3 20.5 [90]
[Eu(L3)3 (H2 O)2 ](H2 O)2 20 554 (24 474) 10.5 2.3 [90]
[Eu(nta)3 (t-DAB)] 19 650 (17 800) b
1.6 × 10−2 [91,92]
[Eu(L4)3 ](ClO4 )3 c 21 300 (<25 000) 6.7 × 10−5 2 × 10−3 [93,94]
a
In a di-ureasil host.
b
No quantitative determination but the quantum yield is reported to be very small [92].
c
1 mM solution in anhydrous acetonitrile.
28 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

R1 Ph
O
CF3 Hhfa
Ph P Ph
S
O N N
Htta O
TPPO Phen
F3C

Ph Hbtfa
R1 =
Hnta N N
N N
bpy
O
Hdbm PhenNO
N N
O

O O 4,4'-bpy
HO
Me
N O Me N C C N Me
O H H
3
HO t-DAB
O H2DP3C1
O O P
2 O 2
EtO
HO2C N CO2H DPPOT
H2dpa N N
HO N
F3C CF3
O N N

HO L4
O
O
O
R1
O O N
O N
O
OH
OH N R2
O HL5x
F 3C H4L1 CF3
OH H
H
O N N O

OH HO
OH
R = H HL2 O O
R Se R R
R = Cl HL3
O NHMe NHMe
H2L6x

Scheme 1. Chemical formulae of the ligands for complexes listed in Table 3 and Figs. 6 and 7.

40 negligible; moreover, in line with the above discussion, transfer


S from 3 T occurs almost equally onto 5 D1 (∼44%) and 5 D0 (∼56%)
[76]. Involvement of the 5 D1 state in the transfer process can be
30 detected experimentally by recording the time evolution of the
Energy /103 cm-1

5 D and 5 D luminescence: if 5 D contributes to the population of


1 0 1
5 D , its intensity should decrease with time and then the rise time
20 0
104 of 5 D0 should match the decay time of 5 D1 . Such determinations
are still scarce [80–82]. Indeed, detecting weak 5 D1 luminescence
10 is not easy, especially, when the EuIII ion is coordinated to carbonyl
groups or water molecules the vibrations of which (C O stretch or
S HOH bending) fairly well match the 5 D1 –5 D0 gap [83,84]. It is to
0
Ligand All rate constants in
be stressed here that evidencing 5 D1 luminescence is not a proof
that this level is populated through ligand energy transfer since
Fig. 5. Schematic energy level diagram, energy transfer processes and transfer rate it may be simply be populated by thermal excitation from 5 D0
constants for (NBu4 [Eu(L1)])2 . given the small energy difference between these two levels. Two
Reproduced with permission from Ref. [76], © 2014 American Chemical Society. convincing examples of 5 D1 populating 5 D0 have been recently
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 29

Fig. 6. Decay of the 5 D1 luminescence and building up of the 5 D0 emission in


Na3 [Eu(DP3C1)3 ] complex in the solid state at room temperature, exc = 320 nm.
Reproduced with permission from Ref. [82], © 2013 The Royal Society of Chemistry.

published featuring [Eu(NO3 )3 (DPPOT)3 ] [81] (see Section 2.5)


and a tris(dipicolinate) complex in which a coumarin sensitizer is 0.8
coupled to the 4-position of the pyridine ring, Na3 [Eu(DP3C1)3 ] Q LEu
(see Scheme 1): in the latter, the decay of the 5 D1 luminescence
0.7
intensity perfectly matches the growing up of emission from 5 D0
as shown in Fig. 6; both decay and rise times are around 1.3 ␮s [82].
Data listed in Table 3 reveal interesting features. Firstly, they
0.6
point to the detrimental quenching of water molecules bound
0.5 5D 5D
in the inner coordination sphere. Secondly, when examining the 1 2
series of chelates with btfa− it becomes evident that the nature
of the ancillary ligand, as well as the matrix into which the 0.4
complex is embedded, is critical and can be used to tune the photo-
physical properties. Thirdly, when comparing [Eu(L2)3 (H2 O)2 ] 0.3
with [Eu(L3)3 (H2 O)2 ](H2 O)2 , for instance, one realizes that water 1.5 2.0 2.5 3.0 3.5 4.0 4.5
molecules bound in the second coordination sphere have also
E0-0(3T)-E(5D0) / 103 cm-1
a large quenching effect since the quantum yield predicted for
[Eu(L3)3 (H2 O)2 ](H2 O)2 is far larger than the experimental one. The
last four entries feature complexes in which charge-transfer states 0.4 Q LTb
have been identified; they will be discussed in Section 2.5.
Improving such model calculations may come from the consid- 0.3 5D
eration of ligand relaxation in the excited state which occurs on 4
the time scale of a vibration that is several orders of magnitude 0.2
faster than the energy transfer processes. High-level calculations
(state-specific and state-averaged CASSCF, allied with extended 0.1 b
multi configurations) taking this fact into consideration recently
succeeded in explaining why energy transfer in [Eu(tta)3 (phen)] is 0.0 a
more efficient than in [Eu(tta)3 (bpy)] [95].
19 20 21 22 23 24 25
2.3. Role of triplet states E0-0(3T) / 103 cm-1

Triplet states play a central role in energy migration between Fig. 7. (Top) Quantum yields of EuIII tetrakis(␤-diketonates) in acetonitrile versus
organic ligands and LnIII ions, and vice versa, by either participat- triplet-state energy [80]. (Middle) Quantum yields of solid state samples of
[Eu(L5x)3 ] complexes versus the 3 T–5 D0 energy difference; black squares: alkyl sub-
ing in the sensitization process, when the lowest energy triplet
stituents; red circles: aryl substituents [96–98]. (Bottom) Quantum yields of TbIII
state, with 0-phonon energy, E0–0 (3 T), is located above the emit- chelates [Tb(L6x)2 ]− in 0.1 M Tris buffer containing 0.2% DMSO versus energy of the
ting state, or by quenching the metal-centred excited state when triplet state.
it lies below it. The few systematic works published so far have Redrawn from Ref. [99].
focused on the relationship between E0–0 (3 T) and the quantum
yield of visible-emitting EuIII and TbIII chelates. Their main findings microsecond after the excitation pulse and then disappearing to
are summarized below. the benefit of 5 D0 emission, in line with the latter level being pop-
In an early, seminal, study published in 1970, S. Sato and M. ulated by 5 D1 . When the triplet state lies between 5 D1 and 5 D0 ,
Wada reported the quantum yields and lifetimes of EuIII and TbIII direct transfer to the latter occurs, but if 3 T is almost resonant with
tris and tetrakis(␤-diketonates) [80], along with time-resolved 5 D , back transfer considerably reduces the quantum yield. Finally,
0
experiments to demonstrate the mechanism of energy transfer and when E0–0 (3 T) < E(5 D0 ), no sensitization occurs. For TbIII chelates,
the role of the 5 D1 level. They showed that the quantum yield the optimum energy gap 3 T–5 D4 is about 2400 cm−1 , which is low-
reaches a maximum when E0–0 (3 T) is about 1200 cm−1 above the ered to 800 cm−1 when the temperature is decreased to 100 K.
5 D level (or 3750 cm−1 above the 5 D level, see Fig. 7, top). The Although much less detailed in terms of analysis of the energy
1 0
largest quantum yield is obtained for [Eu(tta)4 ]− : 55% in acetoni- transfer mechanisms, a study of 41 luminescent EuIII and TbIII
trile at room temperature, E0–0 (3 T) = 20 300 cm−1 . Time-resolved aminocarboxylates in water by M. Latva et al. [100] is often
spectra for chelates for which E0–0 (3 T) > E(5 D1 ) clearly show 5 D1 cited. For EuIII , the situation is complex because both 5 D1 and
emission dominating the luminescence spectrum during the first 5 D can play the role of receiving states and, moreover, data
2
30 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

display a large spread. QEu L values tend to increase when 3 T is

above 5 D1 , in line with energy transfer on the Eu(5 D1 ) level and


coherent with selection rules for Dexter mechanism. With increas-
ing E0–0 (3 T) energy, the quantum yield then slightly decreases
followed by another increase when the triplet state becomes res-
onant with the 5 D2 level; in summary, favourable situations are
when the E0–0 (3 T) − E(5 D0 ) gap is in the ranges 2500–2800 cm−1 or
4000–6000 cm−1 . For TbIII chelates, quantum yields are the largest
when the energy difference between the triplet state and the emis-
sive Tb(5 D4 ) level is around 2000 cm−1 . Authors are using these
results as phenomenological rules of thumb for tailoring the coor-
dination environment of luminescent compounds. However, one
has to realize that the quantum yields reported for EuIII chelates
are mostly smaller than 25% (with one exception at 38%), pointing
to quenching by diffusing water molecules. In addition, consider-
ing only one parameter is an oversimplification given that several
levels may be simultaneously involved in the transfer process. The
influence of charge transfer states and of potential hydration of the
complexes is also neglected in this approach.
Another investigation deals with a series of 16 solid-state
samples of 9-coordinate EuIII complexes with benzimidazole-
Scheme 2. Chemical formulae of ligands for the complexes described in Section 2.4.
substituted pyridine-2-carboxylates, [Eu(L5x)3 ] (Scheme 1) that
have their coordination sphere saturated and display quantum
yields up to 70% [97,98]. The benzimidazole moiety has been dec- the successful interpretation of energy transfer in many EuIII and
orated in R1 with alkyl or aryl groups, and in R2 with F, Cl, Br, Me, TbIII chelates with a singlet-triplet-LnIII path, as ascertained by
or OR substituents. The correlation with the triplet state energy in the wealth of publications dealing with the design of lumines-
the range 18 800–21 300 cm−1 is relatively smooth, at least for lig- cent LnIII compounds with organic ligands based on this concept.
ands bearing alkyl substituents in R1 (Fig. 7, middle) and shows that Secondly, since singlet states have very short lifetimes, their impli-
quantum yields are sizeable when the 3 T–5 D0 energy difference is cation in the transfer process is not simple to demonstrate, unless
in the range 2000–4000 cm−1 in line with the conclusions reached fast spectroscopic equipment is at hand. Finally, singlet-state trans-
for ␤-diketonates and polyaminocarboxylates. fer is often only partial [81] and its contribution to the overall
Finally, an investigation of the antenna effect for TbIII transfer difficult to decipher. Theoretical modelling (see Section
chelates has been conducted with para-substituted 2- 2.2) however shows that these states can participate in the trans-
hydroxyisophthalamide ligands H2 L6x [99]. Energies of ligand fer, especially that there are often several of them with suitable
absorption and fluorescence maxima as well as of the lowest energies for transfer on higher lanthanide excited states that are
triplet state are in linear correlation with the Hammett parame- quite numerous. There are only a few examples for which sin-
ters, increasing with the ␲-withdrawing ability. Time-dependent glet state transfer has been specifically taken into consideration
density functional theory (TD-DFT) calculations also show that the in the overall energy transfer path because usually only indirect
singlet-triplet gap decreases with increasing Hammett parameters, proof can be found. For instance, modelling of the energy transfer
favouring intersystem crossing. Data for the quantum yields are between tryptophan moieties of calcium-binding proteins and TbIII
displayed in Fig. 7 (bottom) and feature S-shape behaviour, with a ions explicitly involved the spectral overlap between tryptophan
sharp increase in QTb L values when E 3 −1 above fluorescence band and metal-ion absorption spectra [101].
0–0 ( T) is >1200 cm
the emitting 5 D4 level. Two complexes, however, are out of the Detailed and quantitative spectroscopic studies dealing with
correlation; their ligands bear NO2 (a) and Br (b) substituents. the involvement of 1 S states in the energy-transfer process by
For the former complex the reason is clear in that the TD-DFT lifetime measurements and/or triplet-triplet (3 T–3 T) transient
calculations point to the presence an ILCT state probably lying absorption spectroscopy (TAS) and related time-resolved evo-
lower than the 3 T state and quenching luminescence. As for the lution are still scarce. In one recent example, no triplet state
brominated ligand that has E0–0 (3 T) only 180 cm−1 lower than the involvement has been postulated in the sensitization of EuIII lumi-
chlorinated one (QTb L = 0.30), no definite explanation has been put nescence in Na3 [Eu(DP3C1)3 ] (Scheme 1) because time evolution
forward by the authors who suspect a heavy atom effect. of 3 T–3 T TAS for both EuIII and GdIII complexes is the same at
In conclusion for both EuIII and TbIII chelates, if the ligand triplet room temperature with a lifetime of 2.9 ± 0.5 ␮s and, moreover,
state is the main donor state, it should lie at least 1500 cm−1 above EuIII emission from 5 D1 starts a few ns after the initial excita-
the emitting level to avoid too much back transfer and the ideal tion pulse [82]. A more complex system consisting in an EuIII
energy gap lies between 2000 and 4000 cm−1 . This is however a complex with a DOTA-appended 1,8-napthalimide ligand H3 L8
mere guideline and not a rule, since the presence of charge-transfer (Scheme 2) and its GdIII counterpart, has been investigated both
states (ILCT, LMCT) can substantially change this picture. in water and in acetonitrile. Aggregation takes part so that spec-
There are no such systematic studies for the other LnIII ions. tra are concentration-dependent, which makes interpretation quite
Some of them have much more complex electronic structures so involved. Nevertheless, from TAS and lifetime data, the authors
that extrapolation of the above rules of thumb should be made with concluded that both singlet and triplet state sensitization was oper-
care. ative, the former with a much larger rate constant (∼109 s−1 ) than
the latter (∼104 s−1 ). This came to a surprise since usually 3 T trans-
fer is much faster than 1 S transfer. In this system, however, the
2.4. Role of singlet states 1 S–EuIII transfer rate occurs on a time scale similar to the intersys-

tem crossing and, moreover, the overlap between EuIII absorption


The role of singlet states in ligand-to-metal energy transfer and triplet emission is much smaller compared with singlet emis-
is largely ignored for three main reasons. The first one lies in sion [102].
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 31

26

24 singlet N
1LMCT

isc
22
E / 103 cm-1
5D
2

20
5D
1 N N
18 triplets
5D
0 N N N

N N
L7 tta EuIII L7

Fig. 8. (Left) Proposed ligand–EuIII energy migration paths through singlet (red lines) [103], triplet (yellow lines) [104], or 1 LMCT (blue lines) [66] states in [Eu(tta)3 (L7)].
(Right) Formula of the ancillary ligand L7.

That positive identification of a singlet-state-mediated trans- transfer, an explanation substantiated by temperature-dependent


fer is difficult is exemplified by the investigation of a bright measurements. Thus the 1.8 ns lifetime corresponds to the lifetime
yellow ␤-diketonate ternary complex with dipyrazolyl-triazine, of the 1 LMCT state which feeds the 5 D1 level.
[Eu(tta)3 (L7)] (see Fig. 8) in toluene [103]. The ancillary ligand pos- Other evidences for singlet-state excitation of lanthanide ions
sesses a 1 S state located at 24 600 cm−1 in the complex and which are non-observation of the triplet state (e.g. absence of ligand phos-
has 1 ILCT character. Room-temperature time-resolved lumines- phorescence for the GdIII complex) [105] or localization of the
cence spectra after excitation into this level display initial emission triplet state below the emitting state, especially for TbIII chelates,
from the ligand singlet state with lifetime in the range 1.3–1.8 ns; as exemplified for dinuclear macrocyclic Schiff base complexes
Eu(5 D1 ) emission is also seen with a rise time of 1.8 ns and a decay [106,107], a TbIII ZnII heterometallic cryptate [108], or the TbIII
time of 0.39 ␮s. Both signals then disappear to the benefit of the complex with 2-mercaptobenzothiazole [109].
5 D emission which grows with a rise time of 0.39 ␮s and a decay Sensitization of the luminescence of NIR-emitting LnIII ions
0
time of 0.48 ms. At low temperature (77 K), ligand phosphorescence is more prone to involve energy transfer from 1 S states, partic-
is detected with E0–0 = 20 800 cm−1 and lifetime of 3.9 s, while 5 D0 ularly for the ions having several excited electronic levels above
emission has a decay time of 0.65 ms. Considering these facts and 30 000 cm−1 . For instance, energy transfer in polylysine dendrimers
the large decrease in ligand emission quantum yield from 75% for bound to NdIII and ErIII ions has been assigned to the fluorescent 1 S
the free ligand to 1% in the EuIII complex, the authors concluded that state of the built-in dansyl chromophores [110]. However, the 3 T
sensitization of EuIII luminescence is essentially achieved through route is not excluded and can be operative in parallel. Important
the singlet state (Fig. 8, red arrow). This conclusion has been chal- parameters to be considered in addition to the spectral overlap are
lenged soon after and data rationalized in an alternative way as the rate of the intersystem crossing process and selection rules for
follows (yellow mechanism in Fig. 8) [104]. The triplet state of energy transfer. The behaviour of dansyl and lissamine substituents
tta (E0–0 ≈ 21 000 cm−1 ), which is known to sensitize efficiently that sensitize NdIII luminescence in a tris(carboxylate) complex
EuIII luminescence by transfer to both 5 D1 (5 D1 ←7 F0,1 ) and 5 D0 with derivatized ligands fitted with these dyes, but not ErIII and
(5 D0 ←7 F1 ) levels through the exchange mechanism (J = 0,±1) is YbIII emission, can be rationalized taking these parameters into
almost resonant with the triplet state of L7 (20 800 cm−1 ). Therefore account [111]. Complexes with derivatized boron dipyrromethene
the energy migration path 1 ILCT(L7)–3 T(L7)–3 T(tta)–Eu(5 D1 ,5 D0 ) (Bodipy) ligands, [Ln(L9)3 ] (Ln = Er, Yb), are also sensitized via
becomes feasible: the transfer from 3 T(tta) to Eu(5 D1 ,5 D0 ) is fast but Bodipy singlet state: no triplet state phosphorescence is observed,
3 T(tta) is populated by the slow decaying 3 T(L7) state and therefore even for the GdIII complexes and the lowest 1 S fluorescence
the apparent lifetime of 3 T(tta) is lengthened accordingly, which band overlaps the Er(4 F9/2 ←4 I15/2 ) absorption band while the
explains why 5 D0 luminescence still occurs 1 s after the excita- 1 S-Yb(2 F −1 [112].
5/2 ) energy gap amounts to only 5700 cm
tion pulse, indicating that the EuIII emissive level continues to be
populated in this time frame by the triplet states. The second inter- 2.5. Role of charge-transfer states and photo-induced electron
pretation points to the importance of considering the rates at which transfers
energy migrates between states and not only the energy of the
implied levels! The [Eu(tta)3 (L7)] complex was subsequently re- Several types of charge transfer states may be implied in sensi-
investigated by ultrafast spectroscopic techniques, giving attention tizing or quenching LnIII -centred luminescence: (i) ligand-centred
to the initial fast quenching of the ligand excited singlet state [66]. charge-transfer states such as ILCTs which are usually singlets, as
The study revealed that ligand fluorescence occurs in two distinct well as singlet and triplet MLCT states present in transition metal
phases with decay times 8.5 ps and 1.8 ns. A third mechanism was containing chromophores, and (ii) LMCTs and MLCTs between the
therefore invoked postulating the presence of a spectroscopically ligand(s) and the lanthanide ion.
silent singlet state with charge transfer character, 1 LMCT (blue path
on Fig. 8). The shorter decay time of the ligand fluorescence thus 2.5.1. Intraligand charge transfer states
corresponds to a 1 S(L7)→1 LMCT(L7,Eu) fast transfer with estimated In fact, ILCTs have not always been identified because they
98% efficiency. The longer lifetime is assigned to delayed fluores- sometimes are simply the lowest singlet excited state and a the-
cence due to thermally activated 1 LMCT(L7,Eu)→1 S(L7), by back oretical approach is needed to determine the composition of the
32 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Scheme 3. Ligands generating ILCT and LMCT states documented in Section 2.5.

corresponding wavefunctions in order to positively identify an ILCT to 20–22 000 cm−1 in some other complexes [48,66]. For EuIII
contribution. Optical transitions of ILCTs also often overlap with for instance, strong quenching of luminescence usually occurs if
the lowest *␲←␲ absorption band. To the best of our knowledge E(LMCT) < 22–24 000 cm−1 , whereas positive contribution to the
the role of ILCT states has been specifically pointed out for the sensitization of the LnIII ion luminescence occurs when the LMCT
first time by V.F. Zolin and collaborators who studied EuIII and state has higher energy [66]. In fact, the influence of an LMCT
TbIII pyridine-carboxylates [113]. Similar states have subsequently state on the quantum yield depends on its energy with respect
been identified in several classes of compounds, such as dansyl-N- to both ligand and metal-ion levels and on the importance of the
methylaminobenzoates [114], 2:1 complexes with calix[8]arenes crossing over between this state and the excited and ground term
[115], carbazole-functionalized ␤-diketonates [116], aromatic ␤- manifolds of the LnIII ion, i.e. in the case of EuIII , the 5 D and/or 7 F
diketonates bearing amine groups [117], EuIII chloride adducts with spectroscopic terms.
phenanthroline [118] and 2,2 -bipyridine [65], complexes with A theoretical approach has been worked out for EuIII consid-
benzoxazole-substituted 8-hydroxyquinolines [57], 4-benzyloxy ering two-electron Coulomb and exchange interactions and from
benzoates [119], substituted [2.2]paracyclophanes [120], or which selection rules may be derived, transfer rates estimated, and
tetrathiafulvalene-amido-2-pyridine-N-oxide [121] for example. quantum yield calculated [92,126]. The model taken into consider-
The role of the ILCT state varies depending on the nature of the ation is represented on the left hand side of Fig. 9. Energies of the
LnIII ion and of other bound ligands, as well as on its energy. triplet and singlet states have been set at 20 000 and 30 000 cm−1 ,
It may function as the main energy donor state, selected exam- respectively. Decay rates were chosen as follows: 108 s−1 for S1 →T
ples being [Ln(L10)3 (tpy)] (Ln = Nd, Er) [114], L11[Eu(tta)4 ] [122], intersystem crossing and for LMCT→S0 transfer; 106 s−1 for T→S0
Na[Yb(L12x)4 ] [57], [Yb(hfa)3 (L13)2 ] [121], [Eu(L14)3 (bpy) [120]], intersystem crossing, S1 →S0 internal conversion, and nonradiative
or [Er(hfa)3 (L15)] [123], see Scheme 3 for ligand structures. The decay rates in the EuIII ion, with the exception of 5 D0 →7 FJ for which
ILCT states are usually located at low energy, so that excitation of 103 s−1 was selected; the total radiative rate of 5 D0 was also set to
the LnIII ions can be achieved in the visible, a definite advantage 103 s−1 . The transfer rates involving ligand and LMCT states were
for several applications. In addition, the dipolar nature of ligands estimated according to the procedure outlined in Section 2.2 tak-
generating ILCTs makes them amenable to multiphoton excitation, ing into consideration both the J mixing between 7 F0 and 7 F2 (on
another benefit, particularly when it comes to luminescent bio- the order of 5%) and the 7 F1 population at room temperature (30%).
probes [124]. Sometimes, ILCT states simply act as relays between Three different situations have been modelled.
the singlet and the triplet states, as demonstrated for adducts of
europium chloride with phenanthroline and bipyridine [65,118], (i) In addition to ligand-to-EuIII energy transfer, another
or between singlet and LMCT states [117]. When the energy of the transfer occurs between the metal ion and the LMCT
ILCT state is too close to the emitting level or to a donor ligand state. In this scheme, luminescence is strongly quenched
level, it has a quenching effect, see for instance [Tb(L16)3 (solv)n ] when 5000 < E(LMCT) < 20 000 cm−1 , the charge transfer state
(solv = DMSO, H2 O) [119]. pumping out the population of 5 D1 and 5 D0 ; on the other hand,
if the LMCT state lies above 5 D4 (>27 600 cm−1 ) the quantum
2.5.2. Ligand-to-metal charge-transfer states yield behave as in a complex without LMCT state. It is note-
Luminescence sensitization of EuIII and, to a lesser extent worthy that in the high energy range, the best quantum yield is
SmIII , TmIII , and YbIII , by LMCT states has been investigated in obtained when the LMCT state lies between 5 D1 and 5 D2 point-
details, especially for inorganic compounds since these states lie ing to the importance of 5 D1 in the energy transfer process.
far above the emitting states and are commonly used to excite (ii) In addition to ligand-to-EuIII energy transfer, another transfer
luminescent ions in lamp and display phosphors. When it comes occurs between the LMCT state and the ligand states. Three
to complexes with organic ligands, LMCT states play equivocal values of the overlap integral between the wavefunctions of
role in that their energies can be considerably lowered by covalent the charge transfer state and of the ␲ states have been tested,
contributions to the bonding, down to 27 800 cm−1 in aromatic
ct | = 0.01, 0.05, and 0.1. The effect of the LMCT state is
carboxylate and nitrato complexes with sulfoxides [125], or even proportional to this integral. Two quenching domains are
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 33

There is, however, at least one pertinent case, the tris complexes
with 2-mercaptobenzothiazole, [Ln(L17)3 ] (Ln = Sm, Eu, Tb, Tm, see
Scheme 3) two of them only displaying luminescence, EuIII (triplet
state energy transfer) and TbIII (singlet state energy transfer) [109].
The 3 T state lies at 18 200 cm−1 and estimates of LMCT state ener-
gies for SmIII and EuIII by non-empirical density functional theory
yields 12 000 and 7000 cm−1 , respectively [127]. The SmIII (LMCT)
energy is therefore in the middle of the quenching range, explaining
the absence of luminescence for this ion while emission of EuIII is lit-
tle affected, the EuIII (LMCT) energy lying in the middle of the sharp
increase in quantum yield with increasing E(LMCT). Explanation for
the difference in luminescent properties between TbIII and TmIII is
different and pertains to a larger distortion of the triplet state with
respect to the fundamental singlet state in the TmIII complex: in
the presence of such a distortion, the zeroth vibrational level of the
excited triplet state is located below the intersection of the sin-
gle configurational curve of this state with the 4f manifold, so that
energy transfer will be operative only as long as higher vibrational
states are populated [127].
Going back to the last four entries of Table 3, the LMCT state
in [Eu(L2)3 (H2 O)2 ] and [Eu(L3)3 (H2 O)2 ](H2 O)2 (E > 24 500 cm−1 )
does not have a quenching effect: the low values of QLn L can be

explained entirely by the presence of the inner- and outer-sphere


water molecules. On the contrary, [Eu(nta)3 (t-DAB)] is indeed
non-luminescent, as predicted, because the LMCT is located at
17 800 cm−1 , just above the emissive Eu(5 D0 ) level and can eas-
ily deactivate both the 5 D0 (17 300 cm−1 ) and 5 D1 (19 000 cm−1 )
levels [91]. Only weak luminescence has been reported for
[Eu(L4)3 ](ClO4 )3 , E(LMCT) < 25 000 cm−1 . Modelling for this com-
pound reproduces the experimental situation when a LMCT state
with energy close to 1 S is introduced. It is noteworthy that when the
LMCT state is shifted to higher energy in [Eu(NO3 )3 (L4)(MeOH)], a
reasonable value for QLn L is restored [93]. The presence of CT state

at around 22 800 cm −1 has also been invoked for explaining the


weak luminescence exhibited by the dimeric complex [Eu2 (L18)6 ]
(Scheme 3) [128].

2.5.3. Metal-to-ligand charge-transfer states


For most of the LnIII ion, these states lie at very high energy,
Fig. 9. Effect of LMCT on the quantum yield of EuIII : (Top) energy diagram used in typically >50 000 cm−1 , with the exception of CeIII (see Section
the model (redrawn after Ref. [92]). (Right) quantum yield versus LMCT energy for 3 1.2) and some inorganic TbIII materials [129], so that they do not
values of the overlap integral between CT and ␲ wavefunctions
ct | = 0.01 (full
play a role in either energy transfer processes or emission spectra.
line), 0.05 (dashed line), and 0.10 (dotted line).
Nevertheless, there are several literature entries wrongly invoking
Reproduced with permission from Ref. [92] © American Institute of Physics 2005.
MLCT (and without experimental proof of their existence) in energy
transfer schemes, particularly for EuIII and TbIII complexes.
observed, which are associated with the exhaustion of the
populations of the S1 and T1 levels, respectively. Minima in
2.5.4. Relationship with redox properties (photoinduced electron
the quantum yield occur when E(LMCT) is about 1500 cm−1
transfers)
smaller than E(S1 ) or E(T1 ).
The quenching effect of LMCT states described above consid-
(iii) The third case combines the previous two ones, taking into
ers these states simply as additional electronic levels implied in
account forward and backward transfers involving (i) the
the overall energy migration process. However depending on ther-
LMCT state and the ligand 1 S and 3 T states, (ii) the LMCT state
modynamic conditions, the presence of a LMCT state may lead to
and the LnIII excited states, and (iii) the ligand 1 S and 3 T states
reduce the *LnIII centre into *LnII , resulting in the disappearance
and the LnIII excited states. The variation of the quantum
of the f–f transitions. The process is called photoinduced elec-
yield with E(LMCT) is depicted on the right part of Fig. 9.
tron transfer (PET) and occurs when the electrochemical reduction
There are two quenching domains, the first one corresponding
potential of the ligand excited state Er (L•+ /L*) is more negative than
to E(LMCT) in the range 5000–20 000 cm−1 and leading to
the reduction potential of LnIII ; it has largely been documented for
almost complete obliteration of the luminescence. Evalua-
EuIII (−0.35 V vs SCE). The driving force of the process, GPET , can
tion of the energy transfer rates shows the ligand-to-LMCT
be estimated as follows:
rates being higher by at least two orders of magnitude than •+
EuIII -to-LMCT transfer rates. The second quenching range GPET = −Er (L /L∗ ) − E(L∗ ) − Er (EuIII /EuII ) (28)
corresponds to LMCT energies becoming close to the energy
where E(L*) is the energy of the excited singlet or triplet state trans-
of the excited singlet state. Such a case has been documented
ferring the electron.
for [Eu(L4)3 ](ClO4 )3 (Scheme 1) [93].
Such a mechanism is operating in the europium complex with
The low-energy domain for which quantum yields are insen- diphenylphosphanoethane, [Eu(NO3 )3 (DPPOT)3 ] (Scheme 2) which
sitive to the LMCT state is difficult to explore experimentally. is thermodynamically stable in MeCN/CHCl3 , with log ˇ13 = 20.8
34 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Table 4
Photophysical parameters for DPPOT and [Eu(NO3 )3 (DPPOT)3 ] in CHCl3 /MeCN (20/80, v/v) at room temperature [81].

Parameter DPPOT Eu(DPPOT)3 Tb(DPPOT)3 Parameter Eu(DPPOT)3

QF /% 20 2.5 4.5 ± 0.5  (5 D1 )rise /␮s 0.5 ± 0.1


 S /nsa 0.31 0.088 n.a.  (5 D1 )decay /␮s 15 ± 2
 T /␮s 4.5 ± 0.5 0.5 ± 0.1 1.7 ± 0.2  (5 D0 )rise /␮s 10 ± 1

ISC /% n.a. 7 n.a.  (5 D0 )decay /ms 3.08 ± 0.03
a
Average lifetime (3-exponential decay).

[81]. Steady-state and time-resolved measurements in nanosec- In the first case, an electron-rich moiety functioning as electron
ond, microsecond, and millisecond time frames of both ligand donor transfers one electron into the HOMO of the antenna when
and europium luminescence yield the photophysical parameters it is linked to an analyte (Fig. 11, bottom), preventing energy trans-
reported in Table 4 and the following overall picture of the various fer on the LnIII ion. In the absence of interaction, the transfer does
energy migration paths operative in this complex is sketched on not occur and luminescence sensitization can proceed (Fig. 11, top).
Fig. 10. There are numerous examples of this case in bioanalyses; the PET
Both the fluorescence quantum yield and lifetime of the ligand switch is typically a benzene derivative which sees its properties
singlet state are considerably reduced in the complexes. In par- altered by interaction with a biochemical molecule, e.g. a protease
allel, luminescence from the Eu(5 D1 ) level is recorded with a rise such as microsomal leucine aminopeptidase. In the reported analy-
time of 0.5 ␮s exactly corresponding to the lifetime of the triplet sis, the quantum yield of the EuIII and TbIII luminescence decreases
state in the complex, confirming the transfer onto 5 D1 . Moreover, from 6.8 and 4.8%, respectively to 0.01, leading to highly specific
analysis of time-resolved fluorescence shows that energy is also and sensitive time-resolved assays.
transferred from 1 S, in the proportion of 20%, the remaining 80% Similar turn-off probes have been proposed for various analyses:
being transferred from 3 T. With time, Eu(5 D1 ) luminescence fades
of to the benefit of Eu(5 D0 ) emission which occurs with a rise time (i) Detection of N-acetyltransferase (NAT) activity in recombinant
of 10 ± 1 ␮s, in reasonable correlation with the Eu(5 D1 ) decay time enzymes or cell lysates; the PET process is suppressed by inter-
(15 ␮s), pointing to population of the emissive 5 D0 level by 5 D1 . action with NAT which catalyses N-acetylation of the aniline
Despite the long lifetime of Eu(5 D0 ), 3.08 ms, indicating a good moiety of the chromophore, suppressing the PET process. As a
protection against nonradiative deactivation in the coordination result, TbIII luminescence is switched on [132].
sphere, and the ideal location of the triplet state (21 000 cm−1 ) (ii) Numerous reports deal with LnIII complexes with terpyridine,
above the Eu(5 D1 ) state (19 000 cm−1 ), the quantum yield of the which have been rendered responsive by adding PET function-
EuIII complex is poor, 1%, reflecting the presence of a very effi- ality. These complexes are interesting because the triplet state
cient quenching process. The electrochemical potential of DPPOT of terpyridine (∼22 200 cm−1 ) is able to sensitize both EuIII
is −2 V versus SCE and application of Eq. (28) gives estimated ther- and TbIII luminescence and the tpy excited state can be eas-
modynamic driving forces of −1.52 and −0.25 eV for the singlet and ily quenched by electron-rich substituents. A few examples are
triplet excited states, respectively. Deactivation of the ligand singlet described here. For instance, a EuIII probe for the detection of
state through the PET process seems therefore thermodynamically nitric oxide in living plants and tissues is based on an o-diamine
quite feasible and is mostly responsible for the poor quantum yield switch reacting with NO to form benzotriazole in the presence
of the Eu(5 D0 ) emission. of oxygen [133]. Reactive oxygen species (ROS) such as • OH and
Photoinduced electron transfers are useful in the design of lumi- ClO− are detected with both EuIII and TbIII luminescence thanks
nescent analytical probes and sensors [130]. In particular, LnIII to a p-aminophenoxy substituent which is hydrolysed upon
luminescence can be switched off or on depending on the absence reaction with these species [134]. Another system involves
or presence of PET involving either the chromophoric ligand or the ratiometric TbIII /EuIII analysis of intracellular biothiols with a
metal ion. 2,4-dinitrobenzene-sulfonyl switch onto which the tpy moiety
transfers an electron [135]. On the other hand, the probe spe-
cific for singlet oxygen and using an anthracene chromophore
which gets oxidized upon reaction with this species does not
work thanks to a PET mechanism, but rather, the anthracene
triplet state acts a sink preventing tpy to transfer energy onto
the metal ion [136].

Fig. 11. Control of LnIII luminescence by a PET switch acting on the chromophoric
Fig. 10. Energy migration paths in [Eu(NO3 )3 (DPPOT)3 ]. antenna; the yellow circle symbolizes interaction with the analyte.
Redrawn after Ref. [81]. Redrawn after Ref. [131].
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 35

ksep F• +⋅⋅⋅Ln2+ 3. Minimizing radiationless vibrational de-activation


kdif kPET processes
F* + Ln3+ {1F*/Ln3+} {F• +/Ln2+}
k-dif ΔGPET
kbt F + Ln3+ In the 1960s it was recognized that dissolving LnIII salts in D2 O
rather than in water results in much brighter emission for EuIII
Scheme 4. Quenching mechanisms of aromatic hydrocarbons by LnIII ions (Ln = Eu,
Yb); F denotes the fluorescent hydrocarbon [137].
and TbIII but not for GdIII [139] and that the effect depends on the
energy gap between the emitting and receiving levels. Careful study
of the phenomenon in water, acetonitrile, and associated deuter-
ated counterparts led to two important conclusions: (i) quenching
[Ln(hfa)3(1An*-PB)] of the luminescence is due to high-energy vibrations (namely OH
CF3
N stretch for water) and (ii) coordinated water molecules act inde-
O
pendently [140]. Quenching is achieved through several phonons
Ln [Ln(hfa)3(An•+-PB•-)]
from different bound molecules and/or through overtones of the
O
N N implied vibration and therefore this multiphonon de-activation is
CF3
3 [Ln(hfa)3(3An*-PB)] temperature-dependent. Moreover, the radiationless deactivation
rate constant is highly sensitive to the LnIII -ligand distance. The
nature of the interaction is essentially dipolar-dipolar so that it can
[Ln(hfa)3(L19)] [Ln*(hfa)3(An-PB)]
be modelled with Förster’s mechanism [141]. In fact, minimizing
Scheme 5. (Left) Chemical formula for [Ln(hfa)3 (L19)], Ln = Nd, Gd, Er, Yb. (Right) radiationless deactivation is equivalent to maximizing the intrinsic
quantum yield QLn Ln .
proposed mechanism for sensitizing LnIII NIR luminescence; An denotes anthracene
and PB pyridylbenzimidazole units [138].

3.1. Quenching by OH vibrations and hydration number

The higher the vibrational overtone needed to bridge the energy


Regarding the second case, an example of a PET process involv- difference between the emitting level and the highest sub-level
ing LnIII ions is the quenching of aromatic hydrocarbons by EuIII and of the receiving multiplet, Eg , the less likely is the quenching
YbIII ions. Two mechanisms are involved with switching between phenomenon to occur. Relationship between the luminescence life-
them when GPET lies in the range −1.4 and −1.6 eV. When times, Eg , and the number of phonons needed to bridge the energy
GPET < −1.6 eV, the quenching effect of the LnIII ion occurs through gap is illustrated in Table 5 in which ˜ (O H) = 3600 cm−1 and ˜
long-distance electron transfer and yields a geminate radical ion (O D) = 2200 cm−1 . The resulting correlation is commonly referred
pair (Scheme 4); the effective quenching distances are 5.0 and 8.0 Å to as the “energy-gap law” and is widely used when explaining
for YbIII and EuIII , respectively. When GPET > −1.4 eV, quenching differences in quantum yields between LnIII ions in series of iso-
occurs by exciplex formation [137]. morphous compounds (see Fig. 3 for a rule of thumb).
Finally, photoinduced electron transfer may play a role in Quantitative estimates of the contribution of different groups
the sensitization of NIR LnIII luminescence as demonstrated for (O H, C H3 , C D) to the deactivation of several LnIII salts and com-
tris(␤-diketonate) adducts with the chelating ancillary ligand L19 plexes (Ln = Nd, Sm, Eu, Tb, Dy, Er, Yb) have been made by detailed
(Scheme 5). The anthracene moiety acts as a sensitizer for the analysis of luminescence decay curves [142]. Data listed in Table 5
luminescence of NdIII , ErIII , and YbIII , according to a complex mech- clearly point to the dramatic quenching effect of water molecules,
anism. Quenching of the ligand fluorescence occurs in parallel but particularly for the NIR-emitting ions with small Eg , as well as,
is also seen for complexes with non-luminescent ions, GdIII or ZnII . for the latter, the detrimental effect of C H vibrations.
This suggests that binding of a metal ion to the pyridylbenzim- The effect of O H vibrations on both quantum yield and lifetime
idazole unit strongly increases the electron attraction capability serves to the determination of the number of coordinated water
of this moiety by stabilizing its LUMO. Intraligand photoinduced molecules q, indispensable information when developing contrast
electron transfer consequently takes place from the 1 S state of the agent for magnetic resonance bioimaging. Several phenomenologi-
anthracene moiety to the coordinated diimine of the pyridylbenz- cal equations have been proposed, based on lifetime measurements
imidazole group. Rapid back electron transfer follows, generating in water and deuterated water as well as on the assumptions that (i)
the anthracene 3 T state which subsequently acts as energy donor O D oscillators contribute little to deactivation and (ii) that all the
for the emitting LnIII ions, as demonstrated by transient absorption other non-radiative deactivation paths are the same in water and in
measurements [138]. deuterated water and can henceforth be assessed by measuring the

Table 5
Quenching of LnIII luminescence by high-energy OH and CH3 vibrations: lifetimes [141] and compiled nonradiative rate constants [142] versus the energy gap. Lifetimes are
for dilute solutions of perchlorates or triflates at room temperature.

Ln Eg /cm−1 Nr of phonons Lifetime/␮s knr (OH)/ knr (CH3 )/


−1 a
OH OD H2 O D2 O s s−1 b
6
Gd( P7/2 ) 32 100 9 15 2300 n.a. n.a. n.a.
Tb(5 D4 ) 14 800 4 7 467 3800 100–200 12
Eu(5 D0 ) 12 200 3–4 5–6 108 4100 400–500 40
Yb(4 F5/2 ) 10 300 3 4.5 0.17c 3.95 3.5 × 105 3.6 × 104
Dy(4 F9/2 ) 7850 2–3 3–4 2.6 42 2 × 104 9 × 103
Sm(4 G5/2 ) 7400 2 3 2.7 60 1.8 × 104 9 × 103
Er(4 I11/2 ) 6600 2 3 n.a. 0.37 (3–5) × 108 3.3 × 106
Nd(4 F3/2 ) 5400 1–2 2–3 0.031 0.14 n.a. n.a.
a
Distance: 2.2–2.5 Å.
b
Distance: 3.6–3.9 Å.
c
Estimated from quantum yields in water and deuterated water and from  obs (D2 O).
36 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Table 6
Phenomenological parameters for estimating hydration numbers of LnIII ions from excited state lifetime measurements in H2 O and D2 O, see Eq. (29).

Ln A B (Eu, Tb: ms; Yb: ␮s)a Accuracy Conditions Ref.

1.11 0.31 ±0.1 25 solutions, q ≤ 6 aminocarboxylates [144]


Eu 1.11 0.30 + 0.44nOH + 0.99nNH + 0.075nCONH ±0.1 8 solutions, q ≤ 3 aminocarboxylates [145]
1.2 0.25 + 1.20nNH + 0.075nCONH b
Solutions 17 q ≤ 1 and 3 q ≤ 3 cyclen derivativesc [141]

4.2 0 ±0.5 26 solutions and solids q ≤ 9 [146]


Tb
5.0 0.06 b
Solutions 13 q ≤ 1, 1 q ≤ 3 cyclen derivativesc [141]

Yb 1.0 0.20 b
11 solutions 10 q ≤ 1, 1 q ≤ 3 cyclen derivativesc [141]
a
nx is the number of X H oscillators in the first coordination sphere or of amide groups.
b
Estimated to ±0.2–0.3.
c
Including edta and dtpa complexes.

lifetime in the deuterated solvent. When applying these equations, de-activate the excited levels, in particular C H vibrations and large
it is important to make sure that quenching by solvent vibrations corrections have to be implemented.
is by far the most important deactivation process in the molecule. The best way to minimize vibration-induced deactivation pro-
If other temperature-dependent phenomena (e.g. phonon-assisted cesses is to design a rigid metal-ion environment, devoid of
back transfer or charge transfer) are operating, these relationships high-energy oscillators and protecting the LnIII ion from solvent
become unreliable. This has often been observed with TbIII [143]. interaction. Such an environment contributes to reduce collision-
The general form of these relationships is: induced deactivation in solution. Further protection may be gained
by inserting the luminescent edifice into micelles, a strategy used
q = A × (kobs − B) with in bioanalyses [150].
(29)
kobs = kH2 O − kD2 O = 1/(H2 O) − 1/(D2 O)
3.2. Influence of weak intermolecular interactions
where A and B are phenomenological Ln-depending (and somewhat
ligand-depending) parameters determined using series of com- Curiously, the presence of water molecules in the inner coor-
pounds with known hydration numbers. Parameter A describes the dination sphere does not always lead to important quenching.
inner-sphere contribution to the quenching, parameter B, which Recent reports have indeed demonstrated a considerable weaken-
has the same units as k, contains both the outer-sphere con- ing of the quenching ability of O H vibrations if the coordinated
tribution of closely diffusing solvent molecules and a corrective water molecules are involved in strong intra- or inter-molecular
factor accounting for the presence of other deactivating vibra- H-bonding. This has been invoked for a series of polymeric metal-
tions, e.g. N H or C H vibrations in the vicinity of the LnIII ion. organic frameworks Ln(H2 O)8 ⊂[Ln2 (L20)4 ] (Scheme 6): two LnIII
If these X H groups have exchangeable hydrogen atoms, lifetime ions are located at the vertices of a cage formed by the four
will be lengthened upon D substitution; amide O CN H oscillators 1,2,4-triazole-bridged bis(carboxylate) ligand while [Ln(H2 O)8 ]3+
also contribute to radiationless deactivation, but to a lesser extent. is encapsulated in the resulting cavity. When the NdIII edifice is
The most reliable experimental relationships for H2 O are given in treated with a EuIII solution, the central NdIII ion is exchanged for
Table 6, along with an estimate of their accuracy and the conditions EuIII and emission from [Eu(H2 O)8 ]3+ is seen because the bound
under which they have been determined. They are phenomenolo- water molecules strongly interact with the cage (H O· · ·H and
gical relationships so that it is very important to use them with H N· · ·H), which reduces their quenching ability [151].
the same type of compounds for which they were calibrated and TbIII is less sensitive to quenching by O H oscillators than
for q values not exceeding the calibration range. The smaller q is, other LnIII ions thanks to its larger energy gap and several TbIII
the more precise the determination is. Strongly H-bonded water complexes and metal-organic frameworks featuring metal-bound
molecules may also have smaller quenching effect (see next sec- water molecules in the inner coordination sphere display intense
tion). luminescence. Some examples are given in Table 7, the most aston-
Less convincing and more empirical relationships based on ishing being compounds with ligands L21–L23 which have an
the sole measurement of H2 O have been suggested for poly- overall quantum yield >80% despite two bonded water molecules
aminocarboxylates and for NdIII , SmIII , EuIII , TbIII , and DyIII [147]. per TbIII ion. In all these compounds, bound water molecules are
Using the equations described in Table 6 with  obs (D2 O) set equal involved in a complex network of H-bonds. The more important
to the observed lifetime measured on the hydrated sample at 77 K is quenching role of water molecules for EuIII compared with TbIII
also dubious because it is not granted that all vibrational quenching is exemplified in [Tb2 (L22)(H2 O)4 ]∞ : after undergoing thermal
is switched off at this temperature. Finally, establishing q for NdIII is treatment at 120 ◦ C, the quantum yield of the TbIII metal-organic
really problematic; two equations have been described [148,149] framework increases from 89% to 96%, while similar treatment of
which have to be used with care because other vibrations also the EuIII compound leads to an improvement from 31% to 77%.

Table 7
Examples of highly luminescent TbIII coordination polymers with bound water molecules in the first coordination sphere. See Scheme 6 for chemical formulae of the ligands.

Complex E(1 S)/cm−1 E(3 T)/cm−1  obs /msa L


QLn /%a Ref.

[Tb(L21)3(H2 O)2 ]∞ n.a. n.a. n.a. 90 [154]


[Tb2 (L22)(H2 O)4 ]∞ n.a. n.a. 1.33 89 [155]
[Tb2 (L23)6 (H2 O)4 ]∞ 30 220 23 640 1.02 82 [156]
[Tb(L24)6 (H2 O)4 (DMF)]∞ 34 840 23 300 3.0 75 [157]
[Tb(L26)3 (EtOH)(H2 O)]∞ 31 150 23 150 0.92 72 [158]
{[Tb2 (L27)2 (L28)(H2 O)2 ](DMF)}∞ n.a. n.a. n.a. 64 [159]
[Tb(L29)3 (EtOH)(H2 O)]∞ 30 580 24 510 1.16 60 [160]
[Tb(L25)3 (H2 O)2 ]∞ 31 250 23 640 0.79 59 [158]
a
For solid state samples at room temperature.
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 37

CO2H HO O HO2C CO2H

HO2C N N N CO2H
N
N NH2 H2L20
N N N
N
H6L22
N
CO2H HL21

CO2H CO2H

HO2C

N
R
O
N CO2H
O HO2C
O
H
HL23 R = Et HL25
H2L24
HCMe2 HL26
CO2H
CO2H

CH3

CO2H
CO2H O O
H CO2H

H2L27 H2L28 HL29

Scheme 6. Carboxylic acid ligands.

In some instances, dehydration of a coordination polymer nearest-neighbour C H oscillators are the biggest contributors to
results in lower quantum yield, e.g. in the coordination poly- ErIII quenching (kCH = 91 ms−1 for hfa− ), in addition, C H bonds in
mer with benzene-1,4-dicarboxylate, [Tb2 (L28)3 (H2 O)4 ]∞ which the subsequent coordination spheres have also to be replaced: if a
sees its quantum yield decrease from 43% to 26% upon lifetime of 0.1 ms is targeted (a minimum value for applications in
removal of the water molecules [152]. Other documented telecommunications) any C H oscillator within a sphere of 20 Å of
examples are the 3-dimensional porous coordination polymers the emitting ion has to be removed. However, full deuteration is dif-
{[Ln2 (fumarate)2 (oxalate)(H2 O)4 ]·4H2 O}∞ (Ln = Eu, Tb) which lose ficult to achieve, so that it may not be a viable solution for materials
their luminescent properties upon dehydration [153]. to be used in telecommunications. In the case of NdIII the situation
is even worse since the energy gap is smaller compared with ErIII ,
3.3. Vibrational quenching of NIR-luminescence kCH increases to 600 ms−1 for hfa− [166]. Halogenation instead of
deuteration has been proposed to remedy the situation [167] and,
NIR-luminescence, e.g. from NdIII , ErIII , YbIII , has important indeed, fluorination or chlorination of the organic ligands yields
applications in lasers, telecommunications, security inks, and bio- materials with interesting photophysical properties [161,168–170]
sciences. Albeit, when linked to organic ligands, which in principle bringing the field of polymer waveguide amplifiers closer to practi-
should allow easier solution processing of the complexes for fab- cal applications [34]. Another strategy consists in inserting ErIII ions
ricating devices, these ions tend to have very poor quantum yields in confined inorganic cavities such as zeolites [171,172]. Indeed, the
because of vibrational quenching, mainly through C H oscilla- problem of luminescence deactivation through vibrations is so seri-
tors. As a consequence, this quenching phenomenon has received ous that perhalogenation may not be always the ultimate solution:
detailed attention, based on the assumption that it can be described other relatively energetic vibrations such as C C, C O, or even C O
by dipole–dipole (Förster) mechanism [142]. Adequate modelling and C C, also induce non-negligible radiationless deactivations.
has to take into account the distance between the emitting cen- Quantitative and systematic studies on luminescence quenching
tre and the quenching oscillator, as well as the Franck–Condon by C H vibrators have been carried out for macrocyclic complexes
overlap between the wavefunctions of the 4f excited state and with dota [141] and the cryptand (bpy·bpy·bpy) [173] (Scheme 7)
overtones of the deactivating oscillator [161]. The distances can by selective deuteration of given C H groups. In the case of NdIII ,
be obtained either from crystallographic data or from an analy- the quenching effect of C H vibrators, although important, seems
sis of the lanthanide-induced shifts and relaxation times in NMR to be less effective than in the hfa− complex mentioned above. On
experiments [162]. the other hand, the quenching rate constant of YbIII obtained for the
A first, obvious strategy is to replace C H oscillator with C D axial C H located in the ring methylene groups and lying at approx-
vibrators. Enhanced emission by deuteration of [Nd(hfa)3 (H2 O)2 ] imately the same distance from the metal ion (≈3.5–3.7 Å) is exactly
has been demonstrated in 1996 by Y. Hasegawa et al. [163,164], the same for both macrocyclic complexes, which casts confidence
but modelling of the phenomenon had to wait the work of W.P. in these determinations by two different groups on two different
Gillin and coworkers who quantified the quenching of the lumi- systems. Representative quenching rate constants are collected in
nescence in [Er(hfa)3 (H2 O)2 ] and hydrogenated and deuterated Table 8. As a follow up of this work, highly luminescent cryptates
Cs[Er(hfa)4 ] [165]. A major conclusion of this work is that if the with fully deuterated [bpy·bpy·bpy O] (30 C D bonds) have been
38 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Scheme 7. Ligands used in the study of the quenching of LnIII NIR luminescence and for upconversion.

isolated and the crystal structure of the LuIII compound determined These two examples, again, point to the importance of considering
[174]. One arm of the cryptand contains N-oxide pyridine moieties the kinetics of energy migration when designing luminescent LnIII
instead of pyridine groups. Solutions in CD3 CN display lifetimes of compounds.
3.3, 4.2 (mean), and 79 ␮s for NdIII , ErIII , and YbIII , respectively. In
deuterated methanol, the lifetime of the YbIII cryptate increases to 3.4. Revisiting the energy gap law
91 ␮s, corresponding to an estimated intrinsic quantum yield QYb Yb

in the range 6–7%. The energy gap law is a first approximation approach to the
The detrimental effect of C H vibrations is further exempli- problem of vibrational deactivation of LnIII excited states in that
fied in a series of NdIII chalcogenides containing phenyl sulfide or harmonic oscillators are taken into consideration and, moreover,
selenide and dimethyl ether, pyridine, or tetrahydrofuran solvation usually one “major” vibration is introduced into the following equa-
molecules; in this series, the number of NdIII ions increases from tion, the one with the largest energy, ωmax :
1 to 8 and 17. As a result, the number of C H oscillators per NdIII
knr = A · e−(B·Eg /ωmax ) (30)
ion, nCH/Nd , decreases from 20 to 12.8 and 10.8; in addition some
oscillators are positioned further away from the metal centre. This The parameters A and B are empirical and fitted to the char-
translates into the quantum efficiency QNd Nd increasing from 9 [175]
acteristics of the system under study. The approach is indeed
to 16 [176] and 35% [177], respectively. A further modification of oversimplified for the following reasons: (i) anharmonicity of the
the molecular design leading to the cluster (py)10 Nd7 Se21 (HgSePh) oscillator should be introduced, (ii) compounds sometimes have
Nd = 43% [178]. The latter
confirms the trend: nCH/Nd = 8.5 and QNd more than one quenching vibrational modes, and (iii) since LnIII ions
yield is still substantially lower than the one observed for Nd(4 F3/2 ) have numerous electronic levels, more careful consideration of var-
in La2 S3 glass, which is near unity at low activator concentration ious resonances between energy differences and vibrational modes
[179], but this systematic work demonstrates the necessity of min- should be made. Another approach has therefore been proposed,
imizing the number of C H oscillators in the surrounding of the the “inductive-resonant” mechanism: transfer from the excited
NIR-emitting ion if high performances are sought. LnIII level is assumed to principally occur through dipole–dipole
On the other hand, some coordination compounds with ligands interaction (Förster’s energy transfer) [183]. The nonradiative deac-
not or minimally halogenated have surprisingly good NIR-emitting tivation rate constant is then expressed as [184]:
properties, for instance [Er(L30)3 ] (Scheme 7) in which intersystem 161.9 · kr · 2
crossing in the ligand is as fast as 1011 s−1 while 3 T-to-Er(4 I11/2,13/2 ) knr = × SOI (31)
2 · n4 · NA · r 6
transfer occurs with kET = 1010 s−1 ; the isc process is unexpectedly
fast and efficient, which cannot be explained on the sole basis of with kr being the radiative rate constant,  is an orientation factor
the heavy-atom effect and which opens interesting perspectives with isotropic limit equal to 2/3, n is the refractive index, NA Avo-
[180]. As a matter of fact, thin films of ErIII 8-hydroxyquinolinate gadro’s number, and r the distance between the LnIII ion and the
[Er(Q)3 ] display NIR-to-Vis ErIII upconversion [181]. Another recent oscillator; finally SOI is a Franck–Condon spectral overlap of the
breakthrough is the demonstration of ErIII upconversion in a molec- form:

ular trinuclear bimetallic helicate with an non-deuterated organic
ligand in which a sequential energy transfer operates between the SOI = In (˜) · εvibr (˜) · ˜ −4 d˜ (32)
two long-lived excited CrIII ions and the sandwiched ErIII ion [182].
In is the normalized emission spectrum expressed in energy units
Table 8 and εvibr is the molar absorption coefficient of the implied vibra-
Selected rate constants for the quenching of NdIII , ErIII , and YbIII luminescence by tion. Several parameters featured in Eqs. (31) and (32) are very
C H oscillators. Refer to Scheme 7 for chemical formulae. difficult to either calculate or determine experimentally. Neverthe-
Complex Oscillator Distance/Å kXH /ms−1 Ref. less, C. Doffek et al. [184] have recently proposed a comprehensive
approach to this problem, using Morse functions and estimates
Cs[Nd(hfa)4 ] C H 4.7 600 [166]
of SOI based on the recording of vibrational overtone spectra.
[Nd(bpy·bpy·bpy)] CH2 (eq) 3.6 73 [173] Experimental data come from the deuterated [bpy·bpy·bpy O]
CH2 (ax) 4.4 22 [173]
cryptates the structures of which have been apprehended through
CH(ar) 5.5 3.9 [173]
lanthanide-induced shift analysis. The work is quite involved and
Cs[Er(hfa)4 ] C H 4.7 91 [165] detailed but the calculated deactivation rates kobs = 1/ obs for NdIII
C D 4.7 2.4 [165]
(3.5 × 105 s−1 ), ErIII (2.5 × 105 s−1 ), and YbIII (1.1 × 104 s−1 ) per-
[Yb(dota)]− Axial ring C H 3.7 11 [141] fectly match the experimental values. Quenching rate differences
NCH2 CO n.a. 13 [141]
between the various C (H/D) oscillators have also been deciphered.
[Yb(bpy·bpy·bpy)] CH2 (eq) 3.5 12.8 [173] In turn these differences allowed the authors to explain why if
CH2 (ax) 42 11.0 [173] PrIII ⊂[bpy·bpy·bpy] is deuterated on the pyridine rings, the faint
CH(ar) 5.4 4.3 [173]
luminescence from the 1 D2 level is not improved ( obs = 157 ns
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 39

for 2 deuterated rings, vs 150 ns without deuteration): a negative this is in line with the short radiative lifetime in ␤-diketonates (with
kpy = −45 ms−1 is calculated while benzyl oscillators behave nor- O6 N2 or O8 chemical environments) one would have expected a
mally, with kb = +788 ms−1 . The explanation lies in the fact that larger covalent contribution from the softer N-donor atoms. More-
the second harmonic of the C D vibrations (∼6900 cm−1 ) almost over a clear relationship between  rad and the nephelauxetic effect,
perfectly matches the energy gap (∼7000 cm−1 ), in contrast to the as estimated from the energy of the 5 D0 →7 F0 transition, could not
first overtone of the C H vibration (∼6200 cm−1 ) [185]. be evidenced. For YbIII complexes, a similar trend is observed with
The new approach considering not only the smallest energy  rad (N3 O6 ) = 0.83 ms and  rad (N6 O3 ) = 0.93 ms. It is evident that
gap between the emissive level and the next-lower sub-level of more experimental data and theoretical calculations are needed
the receiving manifold, but energy gaps matching overtone ener- to clarify this point.
gies is revealing more adequate, for instance for SmIII for which
4G 6 nd overtone
5/2 – F9/2 corresponds exactly to the energy of the 2
5. Highly luminescent LnIII -containing coordination
of a C H vibration and of the 3rd overtone of a C D oscillator. A
compounds
similar situation prevails for DyIII (4 F9/2 –6 F5/2 ) and the model can
of course be extended to the other emitting LnIII ions [186].
The design of highly luminescent compounds must compel
with several requirements which may be different depending on
4. Can the radiative lifetime be tuned? the emission spectral range and on the end use of the mate-
rial. Molecular designers will of course try to optimize both
sens
Rewriting equation (12) to explicit the overall quantum yield: and QLn Ln by finding the antenna most suitable for the targeted
III
Ln ion and by tailoring a rigid coordination environment, devoid
L Ln
QLn =
sens · QLn (33) of high-energy vibrations. The latter aspect is of highest impor-
tance in the case of NIR-emitting ions as described above. In
shows that both quantities on the right hand side must be opti-
addition, specific needs are required for specific applications. For
mized to get highly luminous compounds. Optimization of the
instance, luminescent materials for light-emitting diodes, solar
sensitization efficiency has been dealt with in Section 2 while
energy conversion, telecommunications, or luminescent coatings
the preceding section was concerned with maximizing the second
Ln , by minimizing vibrational de-activation. However, look- should display high thermal stability and easy processing (if pos-
term, QLn
sible, solution processing), while luminescent bioprobes should be
ing back into Eqs. (6)–(10), it becomes clear that we have another
both thermodynamically stable and kinetically inert in live environ-
handle for tuning the intrinsic quantum yield: the radiative life-
ment and, at the same time, they should be non-cytotoxic, unless
time associated with the emitting electronic level. This lifetime
destruction of cancerous cells is desired. Solubility may also be an
essentially depends on two features, the refractive index and the
important issue.
composition of the inner coordination sphere of the luminescent
In the following tables, we list selected highly luminescent
ion [63].
coordination compounds reported in the literature. Comparison is
If we focus on EuIII (5 D0 ), Eq. (10) points to an n−3 dependence,
not always easy in view of the heterogeneity of the reports and,
meaning that even small variations in the refractive index can
sometimes, the scarcity of experimental details. Moreover, deter-
result in relatively large differences in  rad , as reflected in Table 2
mination of quantum yields is a real conundrum and many artefacts
for Na3 [Eu(dpa)3 ]: the radiative lifetime increases from 2.6 ± 0.2 to
may interfere. The most reliable ways are the comparative method
4.1 ± 0.3 ms in going from the solid-state sample (n = 1.517) to the
(with, ideally, two standards) and the absolute method using an
aqueous solution (n = 1.338), in line with the estimated increase
integrating sphere, but in any case, it is difficult to reach accuracy
 rad (H2 O) =  rad (sol.) × (1.517/1.338)3 = 2.6 × 1.46 = 3.8 ± 0.3 ms
better than 10%, even if reproducibility is good. Regarding solutions,
[56]. A similar dependence has been demonstrated for Yb(2 F5/2 )
the comparative method requires optically dilute samples: A < 0.05,
[57]. Therefore doping a luminescent compound into a matrix with
or A < 0.5 if both sample and reference are excited at the same
relatively large refractive index will decrease its radiative lifetime
wavelength, otherwise a correction for inner-filter effect has to be
and provided all other radiative and nonradiative deactivation
applied; in turn dilute solutions have the disadvantage of favouring
paths remain the same, the intrinsic quantum yield will become
complex dissociation, a fact rarely taken into account, particularly
larger. A LnIII complex with organic ligands has commonly n ∼ 1.5,
when highly coordinating solvents such as DMSO or DMF are used.
so that if doped into inorganic matrices such as Y2 O3 (n = 1.9),
Stability constant determination should be carried out in parallel
YAl5 O12 (YAG, 1.8), YVO4 (2.0), or SrTiO3 (2.4) one may expect a
to photophysical investigation when dealing with solutions.
decrease in  rad of about 50, 40, 60, and 75% with concomitant
increase in QLn Ln . In addition, diluting the sample often leads to

less concentration quenching. With organic matrices, the range 5.1. Visible-emitting ions
of available refractive indices is more limited, e.g. for the popular
PMMA, n = 1.49, but the beneficial dilution effect remains, in Table 9 reports overall quantum yields for the emission of
addition to energy transfer from the matrix itself, depending on PrIII , SmIII , EuIII , TbIII (also see Table 7), and DyIII that are the
the polymer. most commonly used ions in visible-emitting probes and mate-
The influence of the chemical environment is more intricate rials. It is noteworthy that several LnIII ions emit both in the
to decipher. Conceptually, the more mixing into the 4f-orbitals, visible (up to 750 nm) and in the NIR, so that the classifica-
the less forbidden will the f–f transitions be. That is, one should tion made here is somewhat arbitrary. Several types of samples
try to implement strong and polarizable bonds. Systematic stud- are included in Table 9: solid-state samples, solutions, disper-
ies relating  rad to the coordination environment of the LnIII ion are sions in PMMA and in hybrid materials. Most of the researchers
still scarce. One example deals with 9-coordinate mononuclear and designing visible-emitting LnIII -based materials commonly turn to
dinuclear EuIII complexes with coordination environments Nx O9−x red-emitting EuIII and green-emitting TbIII complexes since the
essentially arising from benzimidazole pyridine carboxylate lig- luminescence of these ions is less affected by the presence of
ands [63]. Valence-bond analysis of the available crystal structures high-energy vibrators, and, consequently, they present the largest
points to a larger contribution from the O-donor atoms compared quantum yields.
with the N-donor atoms and indeed, the radiative lifetime varies in An amazing series of various antennae have been designed
the series N3 O6 (2.7 ms) < N4 O5 (3.3 ms) < N6 O3 (4.7 ms). Although and shown to work pretty well, including the simplest ones:
40 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Table 9
Selected overall quantum yields of visible-emitting LnIII complexes at room temperature; see Schemes 1, 6 and 8 for ligand formulae. For TbIII , see also Table 7.
L
Compound State/solvent ex /nm QLn /%a Ref.

[Pr(hfa)3 (pyz)2 ] CHCl3 340 1.3 [187]


[Pr(S2 CNEt2 )3 (phen)] CH3 CN 362 0.9(5) [188]
[Pr2 (L31)3 ] H2 O + 1% CH3 CN 361 0.22 [189]
[Pr(Tf2 N)3 ] Bmpyr Tf2 N 444 0.13(3)b [190]
[Pr(L32)3 (HNOct3 )3 ] MeOH 322 1 × 10−2 [191]
[Pr(tta)3 (H2 O)2 ] DMSO 410 6 × 10−3 [192]

[Sm2 (L42)(DME)2 ]∞ Solid 330 17 [193]


[Sm2 (L33)3 (phen)] CH3 CN 334 13 [194]
[Sm(S2 CNEt2 )3 (phen)] CH3 CN 356 11(2) [188]
Sm IL Al (8.3%) c Solid 304 10 [195]
[Sm(L32)3 ](HNOct3 )3 MeOH 322 6.6 [191]
[Sm(dbm)3 (phen)] Solid n.a. 5.7(6) [196]
{Sm2 (L37)2 (L28)(H2 O)2 }·DMF Solid 280 5.2 [159]
NBu4 [Sm(L1)] Solid 330 3.4(5) [197]

Eu IL Al (9.5%) c Solid 303 96 [195]


[Eu(tta)3 (DBSO)] Solid 370 85 [86]
[Eu(L35)3 ] Solid 254 80 [198]
[Eu(tta)3 (DBP)] 1% in PMMA 405 80 [199]
[Eu(hfa)3 (DPEPO)] PMMA (% n.a.) 340 85 [200]
[Eu(nta)3 (DMSO)2 ] Solid 400 75 [201]
[Eu(L36)3 ] Solid 330 71 [97]
[Eu(L44)3 ] d MeOH 360 55 [202]

[Tb(L37)3 ] Solid 254 100 [198]


[Tb(L38b)]− H2 O n.a. 95 [203]
[Tb2 Cl6 (␮-bpy)(py)6 ]∞ Solid 310 86 [204]
[Tb(NO3 )3 (bpm)2 ] Solid 336–338 80(1) [205]
[Tb(L39)] H2 O 350 61 [206]

Dy IL Al (9%) c Solid 298 47 [195]


[Dy(L40)(phen)(H2 O)4 ]Cl·5H2 O Solid 325 19.0 [207]
[Dy(ppi)3 (DPEPO)] Solid 275 12(2) [208]
[Dy(L43)phen(OH)]∞ Solid 365 11 [209]
[Dy(NO3 )3 (L41)] CH3 CN/CHCl3 350 7.6 [210]
[Dy(L32)3 ](HNEt3 )3 H2 O 322 7.1 [191]
K3 [Dy(NTA)2 (H2 O)] Solid 325 6.3(2) [211]
[Dy(NO3 )3 (bpm)2 ] Solid 336–338 5.1(1) [205]
a
Whenever available, experimental errors are given between parentheses.
b Pr
QPr (excitation in f–f transition).
c
Hybrid material: (LnW10 O36 )9− IL@Al2 O3 , IL = 1-methyl-3-propionyloxy imidazolium bromide; % corresponds to the mol% of Ln polyoxometalate.
d
This complex has an amazingly large luminosity: 31 350 cm−1 M−1.

in solid state QLn L as high as 100% was obtained for TbIII ben- ligand worth mentioning is 2,2 -bipyrimidine (bpm) which
zoate [198] or 80–85% for EuIII ternary ␤-diketonates [86,200]. sensitizes the luminescence of EuIII , TbIII , and DyIII to a large extent:
Apart from ␤-diketonates, known for being highly luminescent, [Ln(NO3 )3 (bpm)2 ] feature quantum yields of 70, 80, and 5.1%,
benzimidazole-substituted pyridine-2-carboxylic acids HL36 respectively [205].
(Scheme 8) can also serve as efficient sensitizers of EuIII lumines- SmIII and DyIII complexes exhibiting orange/red and yel-
cence with quantum yields up to 71% in solid state and 52% in low/green emission, respectively, are less luminescent, and the
CH2 Cl2 solution [97]. highest quantum yields only reach 17% for the dimeric SmIII com-
In water, overall quantum yields are usually much lower: plex with L42 [193] or 19% for a ternary DyIII oxydiacetate complex
for EuIII , top QLnL values lie in the range 20–30%, for exam- with phenanthroline in solid state [207]; however, an amazing 47%
ple, for complexes with dipicolinate [212] or ditopic ligands yield has been recorded for the inorganic–organic hybrid mate-
with bis(benzimidazole)pyridine core [213], or “pybox” deriva- rial Dy IL Al (9%) [195]. In aqueous solutions, quantum yields for
tives [214]. For TbIII the best values of overall quantum yield these ions are small [71,220–224] and usually reach only a few
in water are usually around 60% [206] although QLn L as high as percent. The largest values are reported for complexes with H3 L39
95% has been reported for a TbIII polyaminocarboxylate fitted (Scheme 8): 1% (SmIII ) and 3% (DyIII ) [206]. Another example of the
with a 4-carbamoyl-substituted pyridine antenna H4 L38b [203] benefit obtained by using polymers is revealed by [Dy(NO3 )3 L41],
but this performance has never been confirmed nor matched. which exhibits a quantum yield of 7.6% in CH3 CN/CHCl3 [210].
In order to increase luminescence efficiency, LnIII complexes can The design of other visible-emitting coordination compounds
be inserted into polymer matrices [158,199,215], essentially for with PrIII (450–750 nm), HoIII (640 and 975 nm) or TmIII (480 nm)
decreasing energy migration between luminescent ions (con- is even more challenging. Quantitative data for such complexes
are occasional and values of QLn L do not exceed 1% even for
centration quenching). Interesting increases in luminosity are
documented: for example, the EuIII complex with a carbazole- solid-state samples. The largest values reported to date for PrIII
substituted ␤-diketone sees its quantum yield increasing from 47% are for [Pr(hfa)3 (pyz)2 ] (1.3% in CHCl3 ) [187] and a ternary
in CH2 Cl2 solution to 84% in PMMA matrix [215]. Similarly, to over- dithiocarbamate in CH3 CN (0.9%) [188]. In aqueous solution, the
tris(␤-diketonate) [Pr2 (L31)3 ] has QLn L = 0.22%, which can be
come the problem of lower quantum yields in aqueous solutions,
LnIII complexes can be inserted into micelles or nanoparticles: QEuL further increased 2.7-fold upon addition of tri(n-octyl)phosphine
between 55% and 75% have been obtained in this way [216–219] oxide [189]. The two other ions have quantum yields lower than
(this aspect is not covered in this review). One simple chromophoric 0.1%: 1×10−2 % (HoIII ) and 6×10−2 % (TmIII ) for [Ln(L32)3 (HNOct3 )3 ]
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 41

R O OH
N HO O O
N N
HO OH
N N N
O O H2L31 O O
N N
H3L42

N N
N COOH
N N HOOC
N N H2L32 COOH
N N HOOC
N H H N
R=H H4L38a O O
CHONH2 H4L38b HO OH
O O O O
F F N O O
H2L43
F F Ar
HN NH NH
F F
O O O
H2L33
PO(OH)Me
CO2H OH OH HO Me(HO)OP N
N N
N O O O
N N
HO2C CO2H HN HN NH Ar
N
H3L34 N H3L44
H3L39
O
Ar PO(OH)Me
HOOC O COOH OMe
OH
H2L4
HL35 0 Ar = MeO

0.66x 0.34x OMe


C8H17
N O NH
HO2C N
D D D D
(CH2)5
N O
HL36 D D
CO2H N N
HO2C CO2H
N N H2L45
N
N N
H O O
Ln N-
HL37 pyz O3N NO3 F S S F
[Ln(NO3)3L41] NO3
F O O F
N F F
N N Tf2N -
N
N
N N N+
bpm
bmpyr+
N O
O
S O
DBP DBSO DME

Scheme 8. Ligands for the design of highly luminescent visible-emitting LnIII complexes.

[191], 2×10−3 % (HoIII ) and 6×10−2 % (TmIII ) for [Ln(hfa)3 (pyz)2 ] by quantum cutting (downconversion) with yields in the range
[187] or in the 10−5 % range for complexes with tropolonate 120–190% [228] so that this ion has not been included in this
[225]. Deuterating the ligand and working in deuterated solvents section.
however increases the quantum yield by several orders of mag- In summary, present status of antenna design is such that quan-
nitude: a partial quantum yield of 0.12%, with a luminosity of tum yields for visible-emitting solid-state samples in the range
30 M−1 cm−1 has been reported for the 1 G4 →3 H6 transition of 50–70% and 70–90% can be achieved for EuIII and TbIII , respectively;
[Tm(L45)2 ](DNEt3 ) [226]. Finally, green and red emission from those for SmIII and DyIII lie around 10–15%, while the quantum
ErIII is usually produced by upconversion (largest yield reported: yields reported for other visible-emitting ions reach only 1% for
12% for Gd2 O2 S:ErIII (10%) under 700 W/m2 excitation) [227], or PrIII and are much lower for the other ions.
42 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

Table 10
Selected overall quantum yields of NIR-emitting LnIII complexes at room temperature; see Schemes 1, 8 and 9 for formulae of the ligands.
L
Complex State/solvent ex /nm QLn /%a Ref.

[Pr(tta)3 (H2 O)2 ] DMSO 410 ∼0.18 [192]


[Pr(hfa)3 (pyz)2 ] CHCl3 340 3.1 × 10−3 [187]
[Pr(L32)3 ](NHEt3 )3 H2 O 322 4 × 10−4 (2) [191]

[CrNdCr(L46)3 ]9+ Solid 751 2.7(1) [233]


NdIII ⊂(ZnII MC)b Solid 370 1.13 [234]
[Nd(L47)3 (H2 O)] THF-d8 350–395 0.52 [235]
K[Nd(L48)4 ] CH3 CN 383 0.45 [236]
[Nd(L49a)3 ] Solid 350 0.33 [237]
[K(L50)3 AlNd]+ Solid 355 0.29(1) [238]
[Nd(L32)3 ](NHOct3 )3 Solid 322 0.21(1) [191]
[Nd(L51)3 ] MeCN/MeOH 395 0.17 [239]
[Nd2 L52] H2 O 250–360 0.085c [240]

[Ho(L32)3 ](NHOct3 )3 Solid 322 1 × 10−2 [191]


[Ho(hfa)3 (pyz)2 ] Solid 340 2 × 10−3 [187]
K[Ho(L53a)4 ] DMSO 340 2.3(2) × 10−3 [225]

[Er(L54)3 ] Zn(L55)2 film 405 7d [34]


[Er(L56)3 ](NO3 )3 MeCN 290 0.19(3) [214]
[Er3 (L57a)9 ] DMSO 380 0.039(4) [241]
ErIII ⊂(ZnII MC)b Solid 370 0.036(1) [234]
[Er(L49b)3 ] Solid 350–400 0.033(5) [242]
[K(L50)3 AlEr]+ Solid 355 0.028(4) [238]
K[Er(L48)4 ] CH3 CN 387 0.021 [236]
[Er(H2 L58)]3− H2 O 260–350 4.0 × 10−5 [243]

[Tm(L59)3 ] Solid 340 0.5 [128]


[Tm(L32)3 ](NHOct3 )3 Solid 322 6 × 10−2 (4) [191]
K[Tm(L48)4 ] CH3 CN 380 5.9(2) × 10−3 [236]
K[Tm(L53a)4 ] DMSO 340 3.8(2) × 10−3 [225]

[Yb(tta-d)3 (DMSO-d6 )2 ] DMSO-d6 300–350 6.1 [244]


Na[Yb(L60)4 ] Solid 450 3.7 [57]
K[Yb(L48)4 ] CH3 CN 375 3.8 [236]
[CrYbCr(L46)3 ]9+ Solid 751 3.0(3) [233]
[Yb(L61)(CoCp)]e H2 O 430 2.5 [245]
[Yb(L57b)2 (HL57b)2 Cl] DMSO 392 1.4 [246]
[Yb(L51)3 ] MeCN/MeOH 395 0.9 [239]
[Yb(H2 L58)]3− H2 O 260–350 0.37 [243]
a
Whenever available, experimental errors are given between parentheses.
b
ZnII metallacrown: (12-MCZn -4)2 (24-MCZn -8).
c
Partial quantum yield for 4 F3/2 →4 I9/2 and 4 F3/2 →4 I11/2 transitions.
d Er
Intrinsic quantum yield QEr .
e
CoCp = (C5 H5 )Co(PO(OMe)2 )3.

5.2. Near-infrared emitting ions efficiency in the triple-stranded trinuclear helicate


[CrNdCr(L46)3 ]9+ in which energy is transferred from the two
In coordination chemistry, studied NIR-emitting ions are often d-transition metal ions onto NdIII [233], pointing to the advantage
limited to the trio NdIII (0.86, 1.06, and 1.35 ␮m), ErIII (1.54 ␮m), of such an excitation mechanism. On the other hand, several
and YbIII (∼1 ␮m) because of their practical applications. However, NdIII complexes are reported having very large intrinsic quantum
ions such as PrIII (0.9–1.1 and 1.5 ␮m), SmIII (0.9–1.2 ␮m), DyIII yields, up to 47% for [Nd(SePh)3 ] [178]. All the other NIR-emitting
(0.9–1.4 ␮m), HoIII (0.98–1.5 and 1.9 ␮m), and TmIII (0.78 and ions have quantum yields usually lower than 0.1%, with a few
1.45 ␮m) also present useful emission in the NIR range; in purely exceptions such as [Tm(L59)3 ].
inorganic matrices, the corresponding quantum yields can be Ligands containing 8-hydroxyquinolinate moieties perform
very high, as exemplified for TmIII in germinate glasses for laser very well in sensitizing luminescence of NIR-emitting ions. Thus,
applications, displaying 71% NIR quantum yield for the 3 F4 (TmIII ) quantum yields for solid state samples of [Nd(L49a)3 ] [237]
emission at 1.9 ␮m [229]. In this section, we concentrate on coordi- and [Er(L49b)3 ] [242] reach 0.33% and 0.033%, respectively,
nation compounds and have to recognize that quantum yield data while for heterodinuclear YbIII NaI chelates with benzoxazole-
for the latter five LnIII ion are not common, particularly in view substituted 8-hydroxyquinolinates QYb L is as high as 3.7% [57].
that intrinsic instead of overall quantum yields are often reported The same sensitizing moiety has been incorporated into the
by use of Eq. (6), since very small QLn L are difficult to measure.
ditopic ligand H2 L50 which self-assemble into trimetallic heli-
Extensive reviews have been published on LnIII NIR lumines- cates [K(L50)3 AlLn]+ with overall quantum yields of 0.29, 0.028,
cence [230–232] so that only a few selected systems are listed in and 1.17% for Ln = NdIII , ErIII and YbIII [238]. Complexes with
Table 10. porphyrins derivatized with aromatic substituents have low-
YbIII complexes have the highest values of overall quantum lying triplet states (≈12 000 cm−1 for tetraphenylporphyrin) so
yield because this ion has the largest energy gap. When deuteration that they appear to be ideal for sensitizing NIR-luminescence.
L reaches 6.1% for [Yb(tta-d) (DMSO-d ) ]; as a compari-
is used, QYb 3 6 2 Unfortunately, quantum yields remain usually low (<1% for
L
son, QYb drops to 2.1% for the non-deuterated analogue and to 0.12% YbIII ), except for the complex with fluorinated porphyrin H2 L62
for [Yb(tta)3 (H2 O)2 ] in toluene [245]. The best NdIII complexes have capped with a CoIII cyclopentadienyl derivative displaying an
quantum yield in the range 1–3% in the solid state and 0.1–0.5% in amazingly large quantum yield in water (2.5%) [245]. Another
solution; one notes here the particularly good sensitization large quantum yield (∼3%) has been reported at 77 K for
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 43

N N OH
L46 O
N N
N
N Ln
N N N N N
O
N N
N HL51
N N K+ X
X
OH
O X
F F O
O O
F F F F X
N Al X
HL47 OH
X = H HL53a
F F Cl HL53b
O
F O O F O
N H
HL48 O
3 R N R
Br [K(L50)3AlLn]+ P P HL54
R = CONEt2 HL49a R O O R
F
F F
O F F F
N R=
Br N R
F
OH N F
S F
HL49b
F OH F F F
HL55
X
O
O
N N
L56
X N
O O
O O
NH HN N OH
X = H HL57a
N N Cl HL57b
O O
HO N HO N
N N N N OH
N O N O O
HO HO P
HO HO O
N
O O OH N
H6L52 HL59
HL60

SO3-
F
F F

N F F
N
O OH
F F
N NH N
F OR
O OH
N HN
F F R=
N
N
O
F F 4 O

(H4L58) 4-
SO3 -
H2L61 F F I
2
F Et2N O NEt2

Scheme 9. Ligands for highly luminescent NIR-emitting LnIII complexes.

[Yb(acac)(TFPP)] in methylcyclohexane glass, where TFPP is 5.3. Dual emissive ions


tetrakis(5,10,15,20-pentafluorophenyl)porpholactone [247]. Alter-
native classes of ligands able to efficiently sensitize NIR Most practical applications of LnIII luminescence are based
luminescence are azulene derivatives HL48 [236] and tropolone either on visible or NIR emission. However several ions emit both
HL53 [225] which both sensitize a whole series of NIR-emitting LnIII in the visible and NIR ranges, which may be of interest if dual
ions. properties are sought for, either in marking, i.e. the non-visible
44 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

emission can be used to hide some information, or in biosciences. References


Recent works have called attention on SmIII [197] and DyIII [208]
[1] D. Casanova, D. Giaume, E. Beaurepaire, T. Gacoin, J.P. Boilot, A. Alexandrou,
and dual emission of SmIII in a dendrimer has helped improving
Appl. Phys. Lett. 89 (2006), Art. Nr. 253103.
the reliability of live cell imaging by reducing artefacts [248]. [2] J. Zhao, D. Jin, E.P. Schartner, Y. Lu, Y. Liu, A.V. Zvyagin, L. Zhang, J.M. Dawes,
Simultaneous data on the quantum yields of visible and NIR P. Xi, J.A. Piper, et al., Nat. Nanotechnol. 8 (2013) 729–734.
emission under similar excitation conditions are very scarce for [3] S.W. Wu, G. Han, D.J. Milliron, S. Aloni, V. Altoe, D.V. Talapin, B.E. Cohen, P.J.
Schuck, Proc. Natl. Acad. Sci. U.S.A. 106 (2009) 10917–10921.
these ions, except for two recent studies. In the first one, solid state [4] J. Zhang, K. Ray, Y. Fu, J.R. Lakowicz, Chem. Commun. 50 (2014)
samples of [Ln(hfa)3 (pyz)2 ] have the following ratios between 9383–9386.
the quantum yields for visible and NIR emission: 419 for PrIII , [5] Y. Zhang, L. Zhang, R. Deng, J. Tian, Y. Zong, D. Jin, X. Liu, J. Am. Chem. Soc. 136
(2014) 4893–4896.
579 for DyIII , and 35 for TmIII [187]. The second study deals with [6] Y.Q. Lu, J.B. Zhao, R. Zhang, Y.J. Liu, D.M. Liu, E.M. Goldys, X.S. Yang, P. Xi, A.
[Ln(L32)3 ]3+ and the following ratios can be calculated from the Sunna, J. Lu, et al., Nat. Photonics 8 (2014) 33–37.
reported data for aqueous solutions: 15 for PrIII , 12 for SmIII , 65 [7] D. Casanova, C. Bouzigues, T.L. Nguyen, R.O. Ramodiharilafy, L. Bouzhir-Sima,
T. Gacoin, J.P. Boilot, P.L. Tharaux, A. Alexandrou, Nat. Nanotechnol. 4 (2009)
for DyIII , 11 for HoIII , and 40 for TmIII [191]; in this investigation, 581–585.
quantum yields are reported with 15–50% uncertainties, so that the [8] E.N. Harvey, A History of Luminescence. From the Earliest Times Until 1900,
calculated ratios bear large errors, but they point to the NIR emis- American Philosophical Society, Philadelphia, 1957.
[9] J.-C.G. Bünzli, S.V. Eliseeva, in: O.S. Wolfbeis, M. Hof (Eds.), Springer
sion being always far less intense than the visible one (Scheme 9). Series on Fluorescence. Lanthanide Luminescence: Photophysical, Analyti-
cal and Biological Aspects, vol. 7, Springer Verlag, Berlin, 2011, pp. 1–45
(Chapter 7).
6. Conclusions [10] J.-C.G. Bünzli, Kirk-Othmer Encyclopedia of Chemical Technology, Wiley
Online Library, 2013, pp. 1–43, http://dx.doi.org/10.1002/0471238961.
1201142019010215.a01.pub3.
Expertise in the design of adequate ligands for both coordina- [11] J.-C.G. Bünzli, S.V. Eliseeva, Chem. Sci. 4 (2013) 1939–1949.
tion and luminescence sensitization of trivalent lanthanide ions [12] G. Urbain, C.R. Acad. Sci. Paris 142 (1906) 205–207.
has seen a tremendous development during the past two decades. [13] S. Shionoya, W.M. Yen, Phosphor Handbook, CRC Press Inc., Boca Raton, FL,
1999.
Advances in instrumentation has helped researchers to refine their [14] J.H. Van Vleck, J. Phys. Chem. 41 (1937) 67–80.
analysis by collecting more quantitative data, lifetimes, quantum [15] S.I. Weissman, J. Chem. Phys. 10 (1942) 214–217.
yields, sensitization efficiencies, time-resolved spectra which, in [16] B.R. Judd, Phys. Rev. 127 (1962) 750.
[17] G.S. Ofelt, J. Chem. Phys. 37 (1962) 511–520.
turn, contributes to point out latent bottlenecks and to avoid them
[18] B.M. Walsh, in: B. Di Bartolo, O. Forte (Eds.), Advances in Spectroscopy for
by smart ligand design. Potential users of visible luminescence can Lasers and Sensing, Springer Verlag, Berlin, 2006, pp. 403–433.
simply rely on these studies to produce the targeted luminescent [19] F. Auzel, C.R. Acad. Sci. Paris Ser. B 262 (1966) 1016.
[20] F. Auzel, Chem. Rev. 104 (2004) 139–173.
molecular edifices. As pointed out in the previous sections, ligand-
[21] D.E. Cooper, A. D’Andrea, G.W. Faris, B. MacQueen, W.H. Wright, in: J.M. Van
excited quantum yields in excess of 50% for EuIII and 70% for TbIII Emon (Ed.), Immunoassay and Other Bioanalytical Techniques, CRC Press,
are routinely obtained while values for SmIII and DyIII are smaller, Taylor & Francis Group, Boca Raton, 2007, pp. 217–247 (Chapter 9).
around 10–15%, but remain sizable. The situation is somewhat [22] Q. Liu, W. Feng, F.Y. Li, Coord. Chem. Rev. 273 (2014) 100–110.
[23] B. Liu, C. Li, D. Yang, Z. Hou, P. Ma, Z. Cheng, H. Lian, S. Huang, J. Lin, Eur. J.
different for NIR-emitting ions. Minimization of non-radiative de- Inorg. Chem. 2014 (2014) 1906–1913.
activation processes is well understood but not yet mastered at the [24] J.-C.G. Bünzli, A.-S. Chauvin, in: J.-C.G. Bünzli, V.K. Pecharsky (Eds.), Handbook
experimental level and only a handful of compounds have quan- on the Physics and Chemistry of Rare Earths, Elsevier Science B.V., Amsterdam,
2014, pp. 169–281 (Vol. 44, Chapter 261).
tum yields larger than 3–5% and none >10%, despite smart designs [25] R.T. Wegh, H. Donker, A. Meijerink, Phys. Rev. B 56 (1997) 13841–13848.
[34]. Therefore this aspect remains an open field much in need of [26] E. Soini, H. Kojola, Clin. Chem. 29 (1983) 65–68.
clever contributions. Inorganic–organic hybrid systems might offer [27] I. Hemmilä, Applications of Fluorescence in Immunoassays, Wiley Inter-
science, New York, 1991.
a solution. [28] G. Mathis, Clin. Chem. 39 (1993) 1953–1959.
Another facet which still needs improvement is achieving [29] J.-C.G. Bünzli, Chem. Rev. 110 (2010) 2729–2755.
reliable predictions of photophysical properties by theoretical cal- [30] J.-C.G. Bünzli, in: A. de Bettencourt-Dias (Ed.), Luminescence in Lanthanide
Coordination Compounds and Nanomaterials, Wiley-Blackwell, Oxford, 2014
culations. Much has been improved, but much remains to be done
(Chapter 4, 125–197).
in view of the intricate problem generated by the modelling of the [31] R.J. Mears, L. Reekie, S.B. Poole, D.N. Payne, Electron. Lett. 22 (1986)
numerous energy migration paths in a given compound. This is 159–160.
[32] M.J.F. Digonnet, H.J. Shaw, IEEE J. Quantum Electron 18 (1982) 746–754.
important because too many authors rely on too simplistic and
[33] S.V. Eliseeva, J.-C.G. Bünzli, New J. Chem. 35 (2011) 1165–1176.
incorrect models to interpret their data and to tailor their lumi- [34] H.Q. Ye, Z. Li, Y. Peng, C.C. Wang, T.Y. Li, Y.X. Zheng, A. Sapelkin, G. Adamopou-
nescent systems; relative energy of electronic states is not the only los, I. Hernandez, P.B. Wyatt, et al., Nat. Mater. 13 (2014) 382–386.
characteristic to take into account, energy migration rates are of [35] J.D.B. Bradley, M. Pollnau, Laser Photon. Rev. 5 (2011) 368–403.
[36] D. Geskus, S. Aravazhi, S.M. Garcia-Blanco, M. Pollnau, Adv. Mater. 24 (2012)
paramount importance and should be considered in parallel to the OP19–OP22.
energy scheme. [37] B.G. Wybourne, Spectroscopic Properties of Rare Earths, Wiley Interscience,
Altogether, given the tremendous interest in lanthanide lumi- New York, 1965.
[38] G.H. Dieke, Spectra and Energy Levels of Rare Earth Ions in Crystals, Inter-
nescence and the burgeoning of related applications, I am confident science Publishers, New York, 1968.
that more detailed attention will be given to all aspects pertain- [39] P.S. Peijzel, A. Meijerink, R.T. Wegh, M.F. Reid, G.W. Burdick, J. Solid State
ing to the design of LnIII luminescent edifices and this, allied with Chem. 178 (2005) 448–453.
[40] C. Görller-Walrand, K. Binnemans, in: K.A. Gschneidner Jr., L. Eyring (Eds.),
more quantitative experimental data gathered on the energy trans- Handbook on the Physics and Chemistry of Rare Earths, vol. 25, Elsevier Sci-
fer processes and radiationless de-activations, will contribute to ence B.V., Amsterdam, 1998, pp. 101–264 (Chapter 167).
maintain lanthanide luminescence as indispensable tool in many [41] C. Görller-Walrand, K. Binnemans, in: K.A. Gschneidner Jr., L. Eyring (Eds.),
Handbook on the Physics and Chemistry of Rare Earths, Elsevier Science B.V.,
fields of materials and biological sciences.
Amsterdam, 1996, pp. 121–283 (Chapter 155).
[42] K. Okada, Y. Kaizu, H. Kobayashi, J. Chem. Phys. 75 (1981) 1577–1578.
[43] Y. Kaizu, K. Miyakawa, K. Okada, H. Kobayashi, M. Sumitani, K. Yoshihara, J.
Acknowledgments Am. Chem. Soc. 107 (1985) 2622–2626.
[44] P. Dorenbos, J. Phys.: Condens. Matter 15 (2003) 2645–2665.
[45] P.F. Smet, I. Moreels, Z. Hens, D. Poelman, Material 3 (2010) 2834–2883.
The author acknowledges financial support from EPFL and [46] P. Dorenbos, J. Lumin. 91 (2000) 91–106.
FJIRSM (through the CAS/SAFEA International Partnership Program [47] P. Dorenbos, J. Phys.: Condens. Matter 15 (2003) 575–594.
for Creative Research Teams). [48] J.-C.G. Bünzli, F. Ihringer, Inorg. Chim. Acta 246 (1996) 195–205.
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 45

[49] L.E. Roy, D. Ortiz-Accosta, E.R. Batista, B.L. Scott, M.W. Blair, I. May, R.E. Del [96] N.M. Shavaleev, S.V. Eliseeva, R. Scopelliti, J.-C.G. Bünzli, Inorg. Chem. 49
Sesto, R.L. Martin, Chem. Commun. 46 (2010) 1848–1850. (2010) 3927–3936.
[50] A. Vogler, H. Kunkely, Inorg. Chim. Acta 359 (2006) 4130–4138. [97] N.M. Shavaleev, S.V. Eliseeva, R. Scopelliti, J.-C.G. Bünzli, Chem. Eur. J. 15
[51] J.-C.G. Bünzli, S.V. Eliseeva, in: V.W.-W. Yam (Ed.), Comprehensive Inorganic (2009) 10790–10802.
Chemistry II, vol. 8, Elsevier B.V., Amsterdam, 2013, pp. 339–398 (Chapter [98] N.M. Shavaleev, R. Scopelliti, F. Gumy, J.-C.G. Bünzli, Inorg. Chem. 48 (2009)
8.08). 5611–5613.
[52] J.L. Prather, Atomic Energy Levels in Crystals, National Bureau of Standards [99] A.P.S. Samuel, J. Xu, K.N. Raymond, Inorg. Chem. 48 (2009) 687–698.
Monograph 19, Washington, DC, 1961. [100] M. Latva, H. Takalo, V.M. Mukkala, C. Matachescu, J.-C. Rodriguez-Ubis, J.
[53] P.A. Tanner, Chem. Soc. Rev. 42 (2013) 5090–5101. Kankare, J. Lumin. 75 (1997) 149–169.
[54] C. Görller-Walrand, L. Fluyt, A. Ceulemans, W.T. Carnall, J. Chem. Phys. 95 [101] J. Bruno, W.D. Horrocks, R.J. Zauhar, Biochemistry 31 (1992) 7016–7026.
(1991) 3099–3106. [102] V.F. Plyusnin, A.S. Kupryakov, V.P. Grivin, A.H. Shelton, I.V. Sazanovich,
[55] M.H.V. Werts, R.T.F. Jukes, J.W. Verhoeven, Phys. Chem. Chem. Phys. 4 (2002) A.J.H.M. Meijer, J.A. Weinstein, M.D. Ward, Photochem. Photobiol. Sci. 12
1542–1548. (2013) 1666–1679.
[56] A. Aebischer, F. Gumy, J.-C.G. Bünzli, Phys. Chem. Chem. Phys. 11 (2009) [103] C. Yang, L.M. Fu, Y. Wang, J.P. Zhang, W.T. Wong, X.C. Ai, Y.F. Qiao, B.S. Zou,
1346–1353. L.L. Gui, Angew. Chem. Int. Ed. 43 (2004) 5010–5013.
[57] N.M. Shavaleev, R. Scopelliti, F. Gumy, J.-C.G. Bünzli, Inorg. Chem. 48 (2009) [104] L.D. Carlos, W.M. Faustino, O.L. Malta, J. Braz. Chem. Soc. 19 (2008) 299–301.
7937–7946. [105] N.S. Baek, Y.H. Kim, S.G. Roh, B.K. Kwak, H.K. Kim, Adv. Funct. Mater. 16 (2006)
[58] K. Van den Eeckhout, P.F. Smet, D. Poelman, Material 3 (2010) 1873–1882.
2536–2566. [106] P. Guerriero, P.A. Vigato, J.-C.G. Bünzli, E. Moret, J. Chem. Soc. Dalton Trans.
[59] C.D.S. Brites, P.P. Lima, N.J.O. Silva, A. Millan, V.S. Amaral, F. Palacio, L.D. Carlos, (1990) 647–655.
New J. Chem. 35 (2011) 1177–1183. [107] R.C. Howell, K.V.N. Spence, I.A. Kahwa, A.J.P. White, D.J. Williams, J. Chem. Soc.
[60] D.O. Olawale, T. Dickens, W.G. Sullivan, O.I. Okoli, J.O. Sobanjo, B. Wang, J. Dalton Trans. (1996) 961–968.
Lumin. 131 (2011) 1407–1418. [108] R. Rodriguez-Cortinas, F. Avecilla, C. Platas-Iglesias, D. Imbert, J.-C.G. Bünzli,
[61] M.P. Hehlen, M. Sheik-Bahae, R.I. Epstein, in: J.-C.G. Bünzli, V.K. Pecharsky A. de Blas, T. Rodriguez-Blas, Inorg. Chem. 41 (2002) 5336–5349.
(Eds.), Handbook on the Physics and Chemistry of Rare Earths, Elsevier Science [109] M.A. Katkova, A.V. Borisov, G.K. Fukin, E.V. Baranov, A.S. Averyushkin, A.G.
B.V., Amsterdam, 2014, pp. 179–260 (Vol. 45, Chapter 265). Vitukhnovsky, M.N. Bochkarev, Inorg. Chim. Acta 359 (2006) 4289–4296.
[62] M. Lovric, D. Suter, A. Ferrier, P. Goldner, Phys. Rev. Lett. 111 (2013), Art. Nr. [110] V. Vicinelli, P. Ceroni, M. Maestri, V. Balzani, M. Gorka, F. Vögtle, J. Am. Chem.
020503. Soc. 124 (2002) 6461–6468.
[63] J.-C.G. Bünzli, A.-S. Chauvin, H.K. Kim, E. Deiters, S.V. Eliseeva, Coord. Chem. [111] G.A. Hebbink, S.I. Klink, L. Grave, P.G.B.O. Alink, F.C.J.M. Van Veggel,
Rev. 254 (2010) 2623–2633. ChemPhysChem 3 (2002) 1014–1018.
[64] M. Kleinerman, J. Chem. Phys. 51 (1969) 2370–2381. [112] J.H. Ryu, Y.K. Eom, J.-C.G. Bünzli, H.K. Kim, New J. Chem. 36 (2012) 723–731.
[65] L.N. Puntus, K.A. Lyssenko, I. Pekareva, J.-C.G. Bünzli, J. Phys. Chem. B 113 [113] V.F. Zolin, L.N. Puntus, V.I. Tsaryuk, V.A. Kudryashova, J. Legendziewicz, P.
(2009) 9265–9277. Gawryszewska, R. Szostak, J. Alloys Compd. 380 (2004) 279–284.
[66] L.M. Fu, X.C. Ai, M.Y. Li, X.F. Wen, R. Hao, Y.S. Wu, Y. Wang, J.P. Zhang, J. Phys. [114] S.G. Roh, N.S. Baek, Y.H. Kim, H.K. Kim, Bull. Korean Chem. Soc. 28 (2007)
Chem. A 114 (2010) 4494–4500. 1249–1255.
[67] F.F. Chen, Z.Q. Chen, Z.Q. Bian, C.H. Huang, Coord. Chem. Rev. 254 (2010) [115] L.N. Puntus, A.-S. Chauvin, S. Varbanov, J.-C.G. Bünzli, Eur. J. Inorg. Chem.
991–1010. (2007) 2315–2326.
[68] S. Faulkner, L.S. Natrajan, W.S. Perry, D. Sykes, Dalton Trans. (2009) [116] D.B. Nie, Z.Q. Chen, Z.Q. Bian, J.Q. Zhou, Z.W. Liu, F.F. Chen, Y.L. Zhao, C.H.
3890–3899. Huang, New J. Chem. 31 (2007) 1639–1646.
[69] G.F. de Sá, O.L. Malta, C.D. Donega, A.M. Simas, R.L. Longo, P.A. Santa-Cruz, E.F. [117] M. Räsänen, H. Takalo, J. Rosenberg, J. Mäkelä, K. Haapakka, J. Kankare, J.
da Silva, Coord. Chem. Rev. 196 (2000) 165–195. Lumin. 146 (2014) 211–217.
[70] J.D.L. Dutra, T.D. Bispo, R.O. Freire, J. Comput. Chem. 35 (2014) 772–775. [118] L.N. Puntus, K.A. Lyssenko, M.Y. Antipin, J.-C.G. Bünzli, Inorg. Chem. 47 (2008)
[71] F.R. Gonçalves e Silva, O.L. Malta, C. Reinhard, H.U. Güdel, C. Piguet, J.E. Moser, 11095–11107.
J.-C.G. Bünzli, J. Phys. Chem. A 106 (2002) 1670–1677. [119] S. Sivakumar, M.L.P. Reddy, A.H. Cowley, K.V. Vasudevan, Dalton Trans. 39
[72] G.B. Rocha, R.O. Freire, N.B. Da Costa, G.F. de Sá, A.M. Simas, Inorg. Chem. 43 (2010) 776–786.
(2004) 2346–2354. [120] L.N. Puntus, E.V. Sergeeva, D.Y. Antonov, I.V. Anan’ev, K.A. Lyssenko, I.S.
[73] R.O. Freire, A.M. Simas, J. Chem. Theor. Comput. 6 (2010) 2019–2023. Pekareva, F. Kajzar, I. Rau, Proc. SPIE 8189 (2012), Art. Nr. 81890Y.
[74] J.D.L. Dutra, M.A.M. Filho, G.B. Rocha, R.O. Freire, A.M. Simas, J.J.P. Stewart, J. [121] F. Pointillart, T. Cauchy, O. Maury, Y. Le Gal, S. Golhen, O. Cador, L. Ouahab,
Chem. Theor. Comput. 9 (2013) 3333–3341. Chem. Eur. J. 16 (2010) 11926–11941.
[75] M.A.M. Filho, J.D.L. Dutra, G.B. Rocha, R.O. Freire, A.M. Simas, RSC Adv. 3 (2013) [122] M. Shi, C.R. Ding, J.W. Dong, H.Z. Wang, Y.P. Tian, Z.J. Hu, Phys. Chem. Chem.
16747–16755. Phys. 11 (2009) 5119–5123.
[76] S. Biju, R.O. Freire, Y.K. Eom, R. Scopelliti, J.-C.G. Bünzli, H.K. Kim, Inorg. Chem. [123] F. Pointillart, A. Bourdolle, T. Cauchy, O. Maury, Y. Le Gal, S. Golhen, O. Cador,
53 (2014) 8407–8417. L. Ouahab, Inorg. Chem. 51 (2012) 978–984.
[77] F. Gutierrez, C. Tedeschi, L. Maron, J.P. Daudey, J. Azema, P. Tisnes, C. Picard, [124] A. D’Aleo, A. Bourdolle, S. Brustlein, T. Fauquier, A. Grichine, A. Duperray, P.L.
R. Poteau, J. Mol. Struct. Theochem. 756 (2005) 151–162. Baldeck, C. Andraud, S. Brasselet, O. Maury, Angew. Chem. Int. Ed. 51 (2012)
[78] F. Gutierrez, C. Tedeschi, L. Maron, J.P. Daudey, R. Poteau, J.L. Azéma, P. Tisnes, 6622–6625.
C. Picard, Dalton Trans. (2004) 1334–1347. [125] V. Tsaryuk, K. Zhuravlev, V. Kudryashova, V. Zolin, J. Legendziewicz, I.
[79] O.L. Malta, J. Non-Cryst. Solids 354 (2008) 4770–4776. Pekareva, P. Gawryszewska, J. Photochem. Photobiol. A: Chem. 197 (2008)
[80] S. Sato, M. Wada, Bull. Chem. Soc. Jpn. 43 (1970) 1955–1962. 190–196.
[81] M.H. Ha-Thi, J.A. Delaire, V. Michelet, I. Leray, J. Phys. Chem. A 114 (2010) [126] W.M. Faustino, O.L. Malta, G.F. de Sá, Chem. Phys. Lett. 429 (2006) 595–599.
3264–3269. [127] A.F. Shestakov, N.S. Emelyanova, J. Mol. Struct. Theochem. 954 (2010)
[82] J. Andres, A.S. Chauvin, Phys. Chem. Chem. Phys. 15 (2013) 15981–15994. 124–129.
[83] Y. Haas, G. Stein, Chem. Phys. Lett. 8 (1971) 366–368. [128] S. Shuvaev, V. Utochnikova, L. Marciniak, A. Freidzon, I. Sinev, R. Van Deun,
[84] J.-C.G. Bünzli, J.-R. Yersin, Helv. Chim. Acta 65 (1982) 2498–2506. R.O. Freire, Y. Zubavichus, W. Grünert, N. Kuzmina, Dalton Trans. 43 (2014)
[85] E.E.S. Teotonio, G.M. Fett, H.F. Brito, W.M. Faustino, G.F. de Sá, M.C. Felinto, 3121–3136.
R.H.A. Santos, J. Lumin. 128 (2008) 190–198. [129] P. Mukherjee, C.M. Shade, A.M. Yingling, D.N. Lamont, D.H. Waldeck, S. Petoud,
[86] O.L. Malta, H.F. Brito, J.F.S. Menezes, F.R. Gonçalves e Silva, C.D. Donega, S. J. Phys. Chem. A 115 (2010) 4031–4041.
Alves, Chem. Phys. Lett. 282 (1998) 233–238. [130] A.P. de Silva, T.S. Moody, G.D. Wright, Analyst 134 (2009) 2385–2393.
[87] W.M. Faustino, S.A. Junior, L.C. Thompson, G.F. de Sá, O.L. Malta, A.M. Simas, [131] T. Terai, K. Kikuchi, S. Iwasawa, T. Kawabe, Y. Hirata, Y. Urano, T. Nagano, J.
Int. J. Quantum Chem. 103 (2005) 572–579. Am. Chem. Soc. 128 (2006) 6938–6946.
[88] P.P. Lima, R.A.S. Ferreira, R.O. Freire, F.A.A. Paz, L.S. Fu, S. Alves, L.D. Carlos, O.L. [132] T. Terai, K. Kikuchi, Y. Urano, H. Kojima, T. Nagano, Chem. Commun. 48 (2012)
Malta, ChemPhysChem 7 (2006) 735–746. 2234–2236.
[89] P.P. Lima, S.S. Nobre, R.O. Freire, S.A. Junior, R.A. SaFerreira, U. Pischel, O.L. [133] Y. Chen, W. Guo, Z. Ye, G. Wang, J. Yuan, Chem. Commun. 47 (2011)
Malta, L.D. Carlos, J. Phys. Chem. C 111 (2007) 17627–17634. 6266–6268.
[90] A.P. Souza, L.C.V. Rodrigues, H.F. Brito, S. Alves, O.L. Malta, J. Lumin. 130 (2010) [134] Y. Xiao, Z. Ye, G. Wang, J. Yuan, Inorg. Chem. 51 (2012)
181–189. 2940–2946.
[91] L.D. Carlos, J.A. Fernandes, R.A.S. Ferreira, O.L. Malta, I.S. Goncalves, P. Ribeiro- [135] Z.C. Dai, L. Tian, Z.Q. Ye, B. Song, R. Zhang, J.L. Yuan, Anal. Chem. 85 (2013)
Claro, Chem. Phys. Lett. 413 (2005) 22–24. 11658–11664.
[92] W.M. Faustino, O.L. Malta, G.F. de Sá, J. Chem. Phys. 122 (2005), Art. Nr. 054109. [136] Z.C. Dai, L. Tian, Y.N. Xiao, Z.Q. Ye, R. Zhang, J.L. Yuan, J. Mater. Chem. B 1 (2013)
[93] F.R. Gonçalves e Silva, R.L. Longo, O.L. Malta, C. Piguet, J.-C.G. Bünzli, Phys. 924–927.
Chem. Chem. Phys. 2 (2000) 5400–5403. [137] T. Inada, Y. Funasaka, K. Kikuchi, Y. Takahashi, H. Ikeda, J. Phys. Chem. A 110
[94] C. Piguet, J.-C.G. Bünzli, in: K.A. Gschneidner Jr., J.-C.G. Bünzli, V.K. Pecharsky (2006) 2595–2600.
(Eds.), Handbook on the Physics and Chemistry of Rare Earths, Elsevier Science [138] T. Lazarides, M.A.H. Alamiry, H. Adams, S.J.A. Pope, S. Faulkner, M.D. Ward,
B.V., Amsterdam, 2010, pp. 301–553 (Vol. 40, Chapter 247). Dalton Trans. (2007) 1484–1491.
[95] A.Y. Freidzon, A.V. Scherbinin, A.A. Bagaturyants, M.V. Alfimov, J. Phys. Chem. [139] J.L. Kropp, M.W. Windsor, J. Chem. Phys. 39 (1963) 2769–2770.
A 115 (2011) 4565–4573. [140] Y. Haas, G. Stein, J. Phys. Chem. 75 (1971) 3677.
46 J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47

[141] A. Beeby, I.M. Clarkson, R.S. Dickins, S. Faulkner, D. Parker, L. Royle, A.S. [189] S. Meshkova, Z. Topilova, V. Matiichuk, N. Pokhodylo, I. Kovalevskaya, I.
de Sousa, J.A.G. Williams, M. Woods, J. Chem. Soc., Perkin Trans. 2 (1999) Rakipov, P. Doga, Russ. J. Coord. Chem. 37 (2011) 309–315.
493–503. [190] A. Brandner, T. Kitahara, N. Beare, C. Lin, M.T. Berry, P.S. May, Inorg. Chem. 50
[142] V.L. Ermolaev, E.B. Sveshnikova, Russ. Chem. Bull. Int. Ed. 63 (1994) 905–922. (2011) 6509–6520.
[143] E. Deiters, B. Song, A.-S. Chauvin, C. Vandevyver, J.-C.G. Bünzli, Chem. Eur. J. [191] N. Wartenberg, O. Raccurt, E. Bourgeat-Lami, D. Imbert, M. Mazzanti, Chem.
15 (2009) 885–900. Eur. J. 19 (2013) 3477–3482.
[144] R.M. Supkowski, W.d. Horrocks Jr., Inorg. Chim. Acta 340 (2002) 44–48. [192] A.I. Voloshin, N.M. Shavaleev, V.P. Kazakov, J. Lumin. 93 (2001) 199–204.
[145] R.M. Supkowski, W.d. Horrocks Jr., Inorg. Chem. 38 (1999) 5616–5619. [193] X.H. Wei, L.Y. Yang, S.Y. Liao, M. Zhang, J.L. Tian, P.Y. Du, W. Gu, X. Liu, Dalton
[146] W.d. Horrocks Jr., D.R. Sudnick, J. Am. Chem. Soc. 101 (1979) 334–335. Trans. 43 (2014) 5793–5800.
[147] T. Kimura, Y. Kato, J. Alloys Compd. 275–277 (1998) 806–810. [194] J. Shi, Y. Hou, W. Chu, X. Shi, H. Gu, B. Wang, Z. Sun, Inorg. Chem. 52 (2013)
[148] A. Beeby, B.P. Burton-Pye, S. Faulkner, G.R. Motson, J.C. Jeffery, J.A. McCleverty, 5013–5022.
M.D. Ward, J. Chem. Soc. Dalton Trans. (2002) 1923–1928. [195] J. Cuan, B. Yan, RSC Adv. 4 (2014) 1735–1743.
[149] S. Faulkner, A. Beeby, M.-C. Carrié, A. Dadabhoy, A.M. Kenwright, P.G. Sammes, [196] J.M. Stanley, C.K. Chan, X.P. Yang, R.A. Jones, B.J. Holliday, Polyhedron 29
Inorg. Chem. Commun. 4 (2001) 187–190. (2010) 2511–2515.
[150] I. Hemmilä, V.M. Mukkala, Crit. Rev. Clin. Lab. Sci. 38 (2001) 441–519. [197] S. Biju, Y.K. Eom, J.-C.G. Bünzli, H.K. Kim, J. Mater. Chem. C 1 (2013) 6935–6944.
[151] P. Wang, J.P. Ma, Y.B. Dong, Chem. Eur. J. 15 (2009) 10432–10445. [198] M. Bredol, U. Kynast, C. Ronda, Adv. Mater. 3 (1991) 361–367.
[152] C. Daiguebonne, N. Kerbellec, O. Guillou, J.-C.G. Bünzli, F. Gumy, L. Catala, T. [199] G. Zucchi, V. Murugesan, D. Tondelier, D. Aldakov, T. Jeon, F. Yang, P. Thuéry,
Mallah, N. Audebrand, Y. Gérault, K. Bernot, et al., Inorg. Chem. 47 (2008) M. Ephritikhine, B. Geffroy, Inorg. Chem. 50 (2011) 4851–4856.
3700–3708. [200] O. Moudam, B.C. Rowan, M.A.H. Alamiry, P. Richardson, B.S. Richards, A.C.
[153] W.H. Zhu, Z.M. Wang, S. Gao, Inorg. Chem. 46 (2007) 1337–1342. Jones, N. Robertson, Chem. Commun. 2009 (2009) 6649–6651.
[154] L. Ma, O.R. Evans, B.M. Foxman, W.B. Lin, Inorg. Chem. 38 (1999) 5837–5840. [201] L.D. Carlos, C.D. Donega, R.Q. Albuquerque, S. Alves, J.F.S. Menezes, O.L. Malta,
[155] Q. Zhu, T. Sheng, R. Fu, S. Hu, J. Chen, S. Xiang, C. Shen, X. Wu, Cryst. Growth Mol. Phys. 101 (2003) 1037–1045.
Des. 9 (2009) 5128–5134. [202] M. Soulié, F. Latzko, E. Bourrier, V. Placide, S.J. Butler, R. Pal, J.W. Walton, P.L.
[156] A.R. Ramya, M.L.P. Reddy, A.H. Cowley, K.V. Vasudevan, Inorg. Chem. 49 (2010) Baldeck, B. Le Guennic, C. Andraud, et al., Chem. Eur. J. 20 (2014) 8636–8646.
2407–2415. [203] E. Brunet, O. Juanes, R. Sedano, J.-C. Rodriguez-Ubis, Photochem. Photobiol.
[157] H.B. Zhang, L.J. Zhou, J. Wei, Z.H. Li, P. Lin, S.W. Du, J. Mater. Chem. 22 (2012) Sci. 1 (2002) 613–618.
21210–21217. [204] P.R. Matthes, J. Nitsch, A. Kuzmanoski, C. Feldmann, A. Steffen, T.B. Marder, K.
[158] S. Biju, M.L.P. Reddy, A.H. Cowley, K.V. Vasudevan, J. Mater. Chem. 19 (2009) Muller-Buschbaum, Chem. Eur. J. 19 (2013) 17369–17378.
5179–5187. [205] G. Zucchi, O. Maury, P. Thuery, F. Gumy, J.-C.G. Bünzli, M. Ephritikhine, Chem.
[159] M.L. Sun, J. Zhang, Q.P. Lin, P.X. Yin, Y.G. Yao, Inorg. Chem. 49 (2010) Eur. J. 15 (2009) 9686–9696.
9257–9264. [206] S. Petoud, S.M. Cohen, J.-C.G. Bünzli, K.N. Raymond, J. Am. Chem. Soc. 125
[160] M.V. Lucky, S. Sivakumar, M.L.P. Reddy, A.K. Paul, S. Natarajan, Cryst. Growth (2003) 13324–13325.
Des. 11 (2011) 857–864. [207] J.G. Kang, T.J. Kim, H.J. Kang, Y. Park, M.K. Nah, J. Lumin. 128 (2008) 1867–1872.
[161] A. Monguzzi, R. Tubino, F. Meinardi, A.O. Biroli, M. Pizzotti, F. Demartin, F. [208] S. Biju, N. Gopakumar, J.-C.G. Bünzli, R. Scopelliti, H.K. Kim, M.L.P. Reddy, Inorg.
Quochi, F. Cordella, M.A. Loi, Chem. Mater. 21 (2009) 128–135. Chem. 52 (2013) 8750–8758.
[162] C. Piguet, C.F.G.C. Geraldes, in: K.A. Gschneidner Jr., J.-C.G. Bünzli, V.K. [209] Y.H. Luo, F.X. Yue, X.Y. Yu, X. Chen, H. Zhang, Cryst. Eng. Commun. 15 (2013)
Pecharsky (Eds.), Handbook on the Physics and Chemistry and Rare Earths, 6340–6348.
vol. 33, Elsevier Science B.V., Amsterdam., 2003, pp. 353–463 (Vol. 33, Chapter [210] R. Shunmugam, G.N. Tew, Polym. Adv. Technol. 18 (2007) 940–945.
215). [211] J.G. Kang, H.J. Kang, J.S. Jung, S.S. Yun, C.H. Kim, Bull. Korean Chem. Soc. 25
[163] Y. Hasegawa, Y. Kimura, K. Murakoshi, Y. Wada, J.-H. Kim, N. Nakashima, T. (2004) 852–858.
Yamanaka, S. Yanagida, J. Phys. Chem. 100 (1996) 10201–10205. [212] A.-S. Chauvin, F. Gumy, D. Imbert, J.-C.G. Bünzli, Spectrosc. Lett. 37 (2004) 517
[164] Y. Hasegawa, Y. Wada, S. Yanagida, J. Photochem. Photobiol. C: Photochem. [erratum: 40 (2007) 193–532].
Rev. 5 (2004) 183–202. [213] J.-C.G. Bünzli, A.-S. Chauvin, C.D.B. Vandevyver, B. Song, S. Comby, Ann. N.Y.
[165] L. Winkless, R.H.C. Tan, Y. Zheng, M. Motevalli, P.B. Wyatt, W.P. Gillin, Appl. Acad. Sci. 1130 (2008) 97–105.
Phys. Lett. 89 (2006), Art. 111115. [214] A. de Bettencourt-Dias, P.S. Barber, S. Bauer, J. Am. Chem. Soc. 134 (2012)
[166] R.H.C. Tan, M. Motevalli, I. Abrahams, P.B. Wyatt, W.P. Gillin, J. Phys. Chem. B 6987–6994.
110 (2006) 24476–24479. [215] D.B.A. Raj, B. Francis, M.L.P. Reddy, R.R. Butorac, V.M. Lynch, A.H. Cowley,
[167] F. Quochi, R. Orru, F. Cordella, A. Mura, G. Bongiovanni, F. Artizzu, P. Deplano, Inorg. Chem. 49 (2010) 9055–9063.
M.L. Mercuri, L. Pilia, A. Serpe, J. Appl. Phys. 99 (2006) 053520-1–053520-4. [216] J. Wu, Z. Ye, G. Wang, D. Jin, J. Yuan, Y. Guang, J. Piper, J. Mater. Chem. 19
[168] I. Hernandez, R.H.C. Tan, J.M. Pearson, P.B. Wyatt, W.P. Gillin, J. Phys. Chem. B (2009) 1258–1264.
113 (2009) 7474–7481. [217] Y.Q. Wu, M. Shi, L.Z. Zhao, W. Feng, F.Y. Li, C.H. Huang, Biomaterials 35 (2014)
[169] Y. Zheng, J. Pearson, R.H.C. Tan, W.P. Gillin, P.B. Wyatt, J. Mater. Sci. Mater. 5830–5839.
Electron. 20 (2009) 430–434. [218] N. Wartenberg, O. Raccurt, D. Imbert, M. Mazzanti, E. Bourgeat-Lami, J. Mater.
[170] R. Pizzoferrato, R. Francini, S. Pietrantoni, R. Paolesse, F. Mandoj, A. Monguzzi, Chem. C 1 (2013) 2061–2068.
F. Meinardi, J. Phys. Chem. A 114 (2010) 4163–4168. [219] V. Divya, M.L.P. Reddy, J. Mater. Chem. C 1 (2013) 160–170.
[171] A. Mech, A. Monguzzi, F. Cucinotta, F. Meinardi, J. Mezyk, L. De Cola, R. Tubino, [220] M.T. Alonso, E. Brunet, O. Juanes, J.-C. Rodriguez-Ubis, J. Photochem. Photobiol.
Phys. Chem. Chem. Phys. 13 (2011) 5605–5609. A: Chem. 147 (2002) 113–125.
[172] A. Mech, A. Monguzzi, F. Meinardi, J. Mezyk, G. Macchi, R. Tubino, J. Am. Chem. [221] H. Hakala, P. Liitti, K. Puukka, J. Peuralahti, K. Loman, J. Karvinen, P. Ollikka,
Soc. 132 (2010) 4574–4576. A. Ylikoski, V.M. Mukkala, J. Hovinen, Inorg. Chem. Commun. 5 (2002)
[173] C. Bischof, J. Wahsner, J. Scholten, S. Trosien, M. Seitz, J. Am. Chem. Soc. 132 1059–1062.
(2010) 14334–14335. [222] H. Hakala, P. Liitti, J. Peuralahti, J. Karvinen, V.M. Mukkala, J. Hovinen, J. Lumin.
[174] C. Doffek, N. Alzakhem, M. Molon, M. Seitz, Inorg. Chem. 51 (2012) 4539–4545. 113 (2005) 17–26.
[175] K. Norton, G.A. Kumar, J.L. Dilks, T.J. Emge, R.E. Riman, M.G. Brik, J.G. Brennan, [223] S. Quici, M. Cavazzini, G. Marzanni, G. Accorsi, N. Armaroli, B. Ventura, F.
Inorg. Chem. 48 (2009) 3573–3580. Barigelletti, Inorg. Chem. 44 (2005) 529–537.
[176] G.A. Kumar, R.E. Riman, L.A.D. Torres, S. Banerjee, A.D. Romanelli, T.J. Emge, [224] G. Zucchi, A.-C. Ferrand, R. Scopelliti, J.-C.G. Bünzli, Inorg. Chem. 41 (2002)
J.G. Brennan, Chem. Mater. 19 (2007) 2937–2946. 2459–2465.
[177] B.F. Moore, G.A. Kumar, M.C. Tan, J. Kohl, R.E. Riman, M.G. Brik, T.J. Emge, J.G. [225] J. Zhang, P.D. Badger, S.J. Greib, S. Petoud, Angew. Chem. Int. Ed. 44 (2005)
Brennan, J. Am. Chem. Soc. 133 (2011) 373–378. 2508–2512.
[178] A. Kornienko, B.F. Moore, G.A. Kumar, M.C. Tan, R.E. Roman, M.G. Brik, T.J. [226] J. Wahsner, M. Seitz, Inorg. Chem. 52 (2013) 13301–13303.
Emge, J.G. Brennan, Inorg. Chem. 50 (2011) 9184–9190. [227] R. Martin-Rodriguez, S. Fischer, A. Ivaturi, B. Froehlich, K.W. Krämer, J.C. Gold-
[179] G.I. Abutalybov, A.A. Mamedov, Glass Phys. Chem. 33 (2007) 663–666. schmidt, B.S. Richards, A. Meijerink, Chem. Mater. 25 (2013) 1912–1921.
[180] F. Quochi, M. Saba, H. Artizzu, M.L. Mercuri, P. Deplano, A. Mura, G. Bongio- [228] P.S. Peijzel, A. Meijerink, Chem. Phys. Lett. 401 (2005) 241–245.
vanni, J. Phys. Chem. Lett. 1 (2010) 2733–2737. [229] R. Xu, L. Xu, L. Hu, J. Zhang, J. Phys. Chem. A 115 (2011) 14163–14167.
[181] H. Suzuki, Y. Nishida, S. Hoshino, Mol. Cryst. Liq. Cryst. 406 (2003) 27–37. [230] S. Comby, J.-C.G. Bünzli, in: K.A. Gschneidner Jr., J.-C.G. Bünzli, V.K. Pecharsky
[182] L. Aboshyan-Sorgho, C. Besnard, P. Pattison, K.R. Kittilsved, A. Aebischer, (Eds.), Handbook on the Physics and Chemistry of Rare Earths, Elsevier Science
J.-C.G. Bünzli, A. Hauser, C. Piguet, Angew. Chem. Int. Ed. 50 (2011) B.V., Amsterdam, 2007, pp. 217–470 (Vol. 37, Chapter 235).
4108–4112. [231] J.-C.G. Bünzli, S.V. Eliseeva, J. Rare Earths 28 (2010) 824–842.
[183] E.B. Sveshnikova, V.L. Ermolaev, Opt. Spectrosc. 111 (2011) 34–50. [232] S.V. Eliseeva, J.-C.G. Bünzli, Chem. Soc. Rev. 39 (2010) 189–227.
[184] C. Doffek, N. Alzakhem, C. Bischof, J. Wahsner, T. Güden-Silber, J. Lügger, C. [233] L. Aboshyan-Sorgho, H. Nozary, A. Aebischer, J.-C.G. Bünzli, P.-Y. Morgantini,
Platas-Iglesias, M. Seitz, J. Am. Chem. Soc. 134 (2012) 16413–16423. K.R. Kittilsved, A. Hauser, S.V. Eliseeva, S. Petoud, C. Piguet, J. Am. Chem. Soc.
[185] J. Scholten, G.A. Rosser, J. Wahsner, N. Alzakhem, C. Bischof, F. Stog, A. Beeby, 134 (2012) 12675–12684.
M. Seitz, J. Am. Chem. Soc. 134 (2012) 13915–13917. [234] E.R. Trivedi, S.V. Eliseeva, J. Jankolovits, M.M. Olmstead, S. Petoud, V.L. Peco-
[186] C. Doffek, J. Wahsner, E. Kreidt, M. Seitz, Inorg. Chem. 53 (2014) 3263–3265. raro, J. Am. Chem. Soc. 136 (2014) 1526–1534.
[187] Z. Ahmed, K. Iftikhar, J. Phys. Chem. A 117 (2013) 11183–11201. [235] M. Iwamuro, Y. Wada, T. Kitamura, N. Nakashima, S. Yanagida, Phys. Chem.
[188] M.D. Regulacio, M.H. Pablico, J.A. Vasquez, P.N. Myers, S. Gentry, M. Prushan, Chem. Phys. 2 (2000) 2291–2296.
S.W. Tam-Chang, S.L. Stoll, Inorg. Chem. 47 (2008) 1512–1523. [236] J. Zhang, S. Petoud, Chem. Eur. J. 14 (2008) 1264–1272.
J.-C.G. Bünzli / Coordination Chemistry Reviews 293–294 (2015) 19–47 47

[237] N.M. Shavaleev, R. Scopelliti, F. Gumy, J.-C.G. Bünzli, Inorg. Chem. 48 (2009) [243] S. Comby, D. Imbert, A.-S. Chauvin, J.-C.G. Bünzli, Inorg. Chem. 45 (2006)
2908–2918. 732–743.
[238] M. Albrecht, O. Osetska, J.-C.G. Bünzli, F. Gumy, R. Fröhlich, Chem. Eur. J. 15 [244] M.P. Tsvirko, S.B. Meshkova, V.Y. Venchikov, Z.M. Topilova, D.V. Bol’shoi, Opt.
(2009) 8791–8799. Spectrosc. (Engl. Transl.) 90 (2001) 669–673.
[239] H.B. Wei, G. Yu, Z.F. Zhao, Z.W. Liu, Z.Q. Bian, C.H. Huang, Dalton Trans. 42 [245] T. Zhang, X. Zhu, C.C.W. Cheng, W.M. Kwok, H.L. Tam, J. Hao, D.W.J. Kwong,
(2013) 8951–8960. W.K. Wong, K.L. Wong, J. Am. Chem. Soc. 133 (2011) 20120–20122.
[240] Y.V. Korovin, N.V. Rusakova, Y.A. Popkov, V.P. Dotsenko, J. Appl. Spectrosc. [246] F. Artizzu, F. Quochi, M. Saba, L. Marchio, D. Espa, A. Serpe, A. Mura, M.L.
(Engl. Transl.) 69 (2002) 841–844. Mercuri, G. Bongiovanni, P. Deplano, ChemplusChem 77 (2012) 240–248.
[241] F. Quochi, F. Artizzu, M. Saba, F. Cordella, M.L. Mercuri, P. Deplano, M.A. Loi, [247] G.E. Khalil, E.K. Thompson, M. Gouterman, J.B. Callis, L.R. Dalton, N.J. Turro, S.
A. Mura, G. Bongiovanni, J. Phys. Chem. Lett. 1 (2009) 141–144. Jockusch, Chem. Phys. Lett. 435 (2007) 45–49.
[242] M. Albrecht, O. Osetska, J. Klankermayer, R. Fröhlich, F. Gumy, J.-C.G. Bünzli, [248] A. Foucault-Collet, C.M. Shade, I. Nazarenko, S. Petoud, S.V. Eliseeva, Angew.
Chem. Commun. (2007) 1834–1836. Chem. Int. Ed. 53 (2014) 2927–2930.

You might also like