Вы находитесь на странице: 1из 23

Accepted Manuscript

New Luminescent Copper(I) Complexes With Extended π-Conjugation

Kévin Soulis, Christophe Gourlaouen, Chantal Daniel, Alessia Quatela, Fabrice


Odobel, Errol Blart, Yann Pellegrin

PII: S0277-5387(17)30746-5
DOI: https://doi.org/10.1016/j.poly.2017.11.026
Reference: POLY 12928

To appear in: Polyhedron

Received Date: 30 September 2017


Accepted Date: 18 November 2017

Please cite this article as: K. Soulis, C. Gourlaouen, C. Daniel, A. Quatela, F. Odobel, E. Blart, Y. Pellegrin, New
Luminescent Copper(I) Complexes With Extended π-Conjugation, Polyhedron (2017), doi: https://doi.org/10.1016/
j.poly.2017.11.026

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
New Luminescent Copper(I) Complexes With Extended π-
Conjugation

Kévin Soulis,a Christophe Gourlaouen,b Chantal Daniel,b Alessia Quatela,c Fabrice Odobel,a
Errol Blart,a Yann Pellegrina,*

a
Université UNAM, CNRS, Chimie Et Interdisciplinarité: Synthèse, Analyse, Modélisation (CEISAM), UMR 6230,
2, rue de la Houssinière – BP 92208, 44322 Nantes Cedex 3, France. E-mail: yann.pellegrin@univ-nantes.fr; Tel:
+33 (0)2 76 64 51 74
b
Laboratoire de Chimie Quantique Institut de Chimie UMR 7177 CNRS-Université de Strasbourg, 4, Rue Blaise
Pascal CS 90032, F-67081 Strasbourg Cedex, France.
c
HORIBA France SAS, Avenue de la Vauve - Passage Jobin Yvon CS 45002 - 91120 Palaiseau – France

Abstract

While copper(I)-bis(diimine) complexes [CuI(L)2]+ are considered as potent substitutes for


[RuII(bpy)3]2+, they exhibit low molar extinction coefficients with respect to ruthenium parent
analogues. One interesting possibility to improve the light collection ability of [CuI(L)2]+
consists in increasing the length of the Cu-L dipole. In order to achieve this goal, we propose
in this contribution to fuse aromatic rings onto the 2,9-di-nbutyl-1,10-phenanthroline core and
examine how the properties of the corresponding copper(I) complexes are impacted.
Electrochemical, absorption and emission properties are assessed; rewardingly, the
envisioned approach was successful since extinction coefficients above 10000 M-1.cm-1 were
measured. All copper(I) complexes remain photoluminescent, with emission maxima greatly
varying from 725 to 815 nm, strongly affected by the molecular structures. A rationale to
explain the variations of the emission quantum yields is proposed.

Keywords

Copper(I) complexes; phenanthroline; photoluminescence; extinction coefficients


Introduction

Copper(I) complexes [CuI(L)2]+ where L is a chelating diimine ligand such as 2,2’-bipyridine


(bpy) or 1,10-phenanthroline (phen) have attracted considerable attention since McMillins’
breakthrough in 1980,1 revealing that those molecular species can feature near IR (NIR)
photoluminescence if the ligands L are sterically challenged in α of the chelating nitrogen
atoms. There has been since an extensive work on the impact of bulky substituents around
the copper(I) coordination sphere, leading to an overall understanding of the photo-induced
processes within such copper complexes.2-17 Briefly, [CuI(L)2]+ complexes undergo an
internal charge transfer upon MLCT excitation following the equation below:

[CuI(L)2]+ + hν → [CuII(L)(L•-)]+

When the complex is excited, the copper(I) ion is transiently oxidized into copper(II) while an
electron from the 3dCu orbitals is transferred onto a ligand centred π* orbital. Given that
copper(I) and copper(II) ions display different preferred geometries (tetrahedral and square
planar respectively), the excitation of those complexes in their metal-to-ligand-charge-
transfer (MLCT) states entails a dramatic flattening of the structure which is responsible for
quick, radiationless deactivation of the excited state.5,11,18 This deactivation is due to the
strong stabilization of the lowest excited states during flattening favouring non radiative
decay by virtue of the gap law. Additionally, the distortion of the complex frees space above
the copper(II) ion allowing the latter to be attacked by any kind of nucleophile to reach the
stable five-coordinate square planar pyramidal geometry; solvent molecules play an
important role in this process.6,19,20

A collection of copper(I)-bis(diimine) luminescent complexes, both homoleptic ([Cu(L)2]+) and


heteroleptic ([Cu(L)(L’)]+)21,22 exhibiting interesting properties in their excited states have thus
been isolated and studied in various domains (dye sensitized solar cells,23-31 photo-induced
water splitting,32 organic photochemistry,32-37 supramolecular photo-induced electron
transfer).38-46 Those complexes are particularly well-suited for photo-reductive mechanisms
since their photo-reductive power is higher than the champion photosensitizer [RuII(bpy)3]2+.47

However, bearing in mind that the ultimate aim is to replace the latter by copper(I)-
bis(diimine) complexes, there remain a few issues to address. In particular, [CuI(L)2]+
complexes are weak emitters, kinetically labile and endowed with small extinction
coefficients. Great progresses have been made with copper complexes [CuI(L)2]+ where L is
tert-butyl,7,13 sec-butyl4,9 or iso-propyl10 2,9-substituted phen. These complexes, however,
remain weak absorbers with extinction coefficients ranging between 3100 and 7500 L.mol-
1
.cm-1.7 Incidentally, grafting methyl groups in positions 3, 4, 7 and 8 on the already 2,9-
disubstituted phen core was very beneficial from all points of view, since the emission lifetime
was greatly improved and the extinction coefficient too (ca. 104 L.mol-1.cm-1).9,10 The
inductive effects of the methyl groups tend to increase the dipole length and this materializes
into a more intense MLCT band.6,48

A similar approach consists in using chromogenic diimine ligands:29,49 by increasing the


extent of π conjugation over the phen core and tethering electron donating groups, the dipole
length is increased and additional ligand centred transitions can overlay with the MLCT band.
Very high extinction coefficients have been obtained using this approach, but loss of the
MLCT excited state properties were experienced in a few cases because the absorption
features of the ligand override those of the copper-diimine chromophore.

To sum up, increasing the extent of π delocalization is a good strategy to improve the
extinction coefficients of the MLCT and we propose in this contribution to apply the latter by
fusing aromatic rings to the phenanthroline ligand. Several examples can be found in the
literature where concatenation of extra aromatic rings lead to improved ε, the epitome of all
being famous dipyrido[3,2-a:2’,3’-c]phenazine ligand, best known as dppz.50,51 The latter
ligand results from the fusion of the phenazine electron acceptor with a bipyridine cavity. This
ligand and complexes thereof have been abundantly studied owing to their intriguing
electronic properties52-55 and tested in many different fields such as cytotoxic agents
design,56-58 photo-induced electron transfer platforms,59,60 catalysis61-64 and of course DNA
probes.65 Although the phenazine spacer is known to be electronically decoupled from the
dipyrido coordination sphere,52,66 the extinction coefficient of e.g. [RuII(bpy)2(dppz)]2+ is
significantly higher than parent [RuII(bpy)3]2+,60 probably thanks to the increased dipole
length. Fusion of one additional benzene onto dppz yielded yet again improved extinction
coefficients, to a lesser extent though.67 Importantly, the ligand centred absorption features of
[RuII(bpy)2(dppz)]2+ stand below the Ru→L MLCT and do not interfere with the MLCT
transition. Thus, dppz seems a good candidate to improve copper(I) complexes light
harvesting capacity, since [CuI(L)2]+ complexes exhibit absorption profiles which are
surprisingly similar to those of [RuII(bpy)3]2+. The latter was incidentally employed in the field
of diphosphine-diimine64 and diimineA-diimineB,68 heteroleptic copper(I) complexes and
afforded indeed better absorbing complexes than their associated references. Besides, a
little studied ligand, yet closely related to dppz and showing increased π conjugation too is
ligand np (Figure 1), consisting of an anthracene spacer fused to a bipyridine cavity.69-71
Although np and dppz are very look-alike, their electronic properties are quite different. In
particular, the anthracene moiety is a far less potent electron acceptor than phenazine, and
this translates into very different electrochemical behaviours. Importantly, [RuII(bpy)2(np)]2+
features an improved MLCT extinction coefficient compared to [RuII(bpy)3]2+.69
Dppz and np ligands are thus our primary targets for this contribution. They however cannot
be used as such and it is mandatory to tether bulky substituents in α of the chelating nitrogen
atoms of dppz and np to design room temperature, solution phase luminescent copper(I)
complexes. Interestingly, Guo et al. have published a theoretical article where heteroleptic
copper(I) complexes bearing methyl-substituted dppz and np have been reported for future
use in dye sensitized solar cells.70 Critically, the complexes were praised for their enhanced
MLCT oscillator strengths, justifying the interest in preparing these species. We thus
embarked in the synthesis of ligands L1 and L3 (Figure 1), consisting respectively of dppz
and np molecules encumbered with two nbutyl chains on both sides of the diimine chelate.
On our way, we isolated ligand L2, displaying a potentially high dipole length too, and
decided to explore the ground and excited states properties of the corresponding copper(I)
complex.

In this article, we evidence the fusing of aromatic rings onto phenanthroline as a good
approach to increase the extinction coefficients of related copper(I) complexes without
sacrificing to too large an extent the luminescence quantum yields.

Figure 1. Structures of the three ligands L1, L2 and L3 and their associated homoleptic copper(I)
complexes C1, C2 and C3. Structure of model complex C4.

Experimental section

Chemicals were purchased from Sigma-Aldrich or Alfa Aesar and used as received. Thin-
layer chromatography (TLC) was performed on aluminium sheets precoated with Merck 5735
Kieselgel 60F254. Column chromatography was carried out either with Merck 5735 Kieselgel
60F (0.040-0.063 mm mesh).1H spectra were recorded on an AVANCE 300 UltraShield
BRUKER. Chemical shifts for 1H NMR spectra are referenced relative to residual protium in
the deuterated solvent (CDCl3  = 7.26 ppm). NMR spectra were recorded at room
temperature, chemical shifts are written in ppm and coupling constants in Hz. High-resolution
mass (HR-MS) spectra were obtained either by electrospray ionization coupled with high
resolution ion trap orbitrap (LTQ-Orbitrap, ThermoFisher Scientific,) or by MALDI-TOF-TOF
(Autoflex III, Bruker), working in ion-positive or ion-negative mode. Electrochemical
measurements were made under an argon atmosphere in CH2Cl2 with 0.1 M Bu4NPF6. Cyclic
voltammetry experiments were performed by using an Autolab PGSTAT 302N
potentiostat/galvanostat. A standard three-electrode electrochemical cell was used.
Potentials were referenced to a saturated calomel electrode (SCE) as internal reference. All
potentials are quoted relative to SCE. The working electrode was a glassy carbon disk and
the auxiliary electrode was a Pt wire. In all the experiments the scan rate was 100 mV .s-1.
UV-visible absorption spectra were recorded on a UV-2401PC Shimadzu, using 1 cm path
length cells. Emission spectra were recorded on a Fluoromax-4 Horiba Jobin Yvon
spectrofluorimeter (1 cm quartz cells). Luminescence decays were recorded with a
DELTAFLEX time correlated single photon counting system (HORIBA) on degassed
dichloromethane solutions.

All calculations were performed with ADF 2013 package72 at the DFT level of theory with
B3LYP functional.73 The atoms were described by the all electrons slater-type TZP basis
set.74 Scalar relativistic corrections were included through the ZORA Hamiltonian.75 Solvent
corrections for dichloromethane were introduced using PCM model.76 The structures were
fully optimized and the absorption spectra were computed by means of TD-DFT77 on these
structures. For the spectra, only the non-equilibrium response of the solvent were included
and Tamm-Dancoff78 corrections introduced. In all the calculations, the butyl chains were
replaced by methyl.

2,9-di-nbutyl-1,10-phenanthroline2, 2,9-di-nbutyl-1,10-phenanthroline-5,6-dione79 (1) and


complex C480 were synthesized according to previously published literature procedures.

Ligand 3,6-di-nbutyl-dipyrido[3,2-a:2’,3’-c]phenazine L1: 0.12 mmol of 1 (40 mg) and


0.12 mmol of 1,2-diaminobenzene 2 (13.4 mg) were dissolved in ethanol (20 mL) and heated
to reflux for 8 hours. The solvent was then evaporated and dissolved in dichloromethane (10
mL). The latter solution was washed three times with aqueous 1M HCl (100 mL x 2), then
pure water. The organic phase was subsequently dried on sodium sulphate, filtered and
evaporated to dryness, affording a pale yellow powder.
Yield: 38 mg (80%). 1H NMR (300 MHz, CDCl3, 25°C): δ = 9.56 (2H, d, J = 8.1 Hz), 8.34 (2H,
dd, J = 3.3 Hz, J = 6.6 Hz), 7.90, (2H, dd, J = 3.3 Hz, J = 6.6 Hz), 7.68 (2H, d, J = 8.4 Hz),
3.28 (4H, m), 1.96 (4H, m), 1.55 (4H, m), 1.03 (6H, t, J = 7.5 Hz) ppm. HRMS (MALDI) m/z:
[M+H]+ calculated for C26H27N4+: 395.2236; found: 395.2227. Δ = 2.3 ppm.

Ligand 3,6-di-nbutyl-naphto[2,3-f][4,5]phenanthroline-9,14-dicarbonitrile L2: 1 (0.98


mmol, 315 mg) and 1,2-diacetonitrilebenzene (1.13 mmol, 176 mg) were dissolved in
acetonitrile (25 mL) and heated to reflux. DBU (1,8-Diazabicyclo[5.4.0]undec-7-ene, 1 mmol,
156 mg) was then added to the solution which turned dark green at once. After 4 hours of
reflux, the medium was let to return to room temperature and water was added. The mixture
was extracted twice with dichloromethane. The organic phase was then dried on sodium
sulphate, filtered and evaporated to dryness. The crude was chromatographied on silica gel
(eluent: dichloromethane/triethylamine 98/2 (v/v)). The first fraction contained the desired
compound and 1,2-diacetonitrilebenzene (respectively 43% and 57% molar, estimated by
NMR). The latter mixture was dissolved in dichloromethane, degassed by argon bubbling,
and a degassed solution of Cu(CH3CN)4PF6 (0.21 mmol, 77 mg) was syringed herein. After
30 minutes, the deep red solution is evaporated to dryness and subjected to chromatography
column on silica gel (eluent: gradient from pure dichloromethane to a mixture of
dichloromethane and methanol (98/2 v/v)). The collected orange fraction was then dissolved
in acetonitrile (5 mL) and a saturated aqueous solution of KCN (5 mL) was added. The
solution changes from dark red to light yellow within a few seconds. The latter was extracted
twice with dichloromethane, dried on sodium sulphate, filtered and evaporated to dryness.
The resulting yellow powder was the desired compound.
Yield: 189 mg (44% on the overall process). 1H NMR (300 MHz, CDCl3, 25°C): δ = 9.69 (2H,
d, J = 8.7 Hz), 8.63 (2H, dd, J = 3.3 Hz, J = 6.6 Hz), 7.92, (2H, dd, J = 3.3 Hz, J = 6.6 Hz),
7.60 (2H, d, J = 8.7 Hz), 3.21 (4H, m), 1.94 (4H, m), 1.53 (4H, m), 1.03 (6H, t, 7.5 Hz) ppm.

Ligand 3,6-di-nbutyl-naphto[2,3-f][4,5]phenanthroline L3: L2 (0.34 mmol, 150 mg) was


suspended in a mixture of ethanol (7 mL) and water (1 mL). KOH (54 mmol, 3 g) was added.
The brown solution was set in an autoclave and heated under autogeneous pressure at
200°C for two days. When back at room temperature, the resulting slurry is poured into water
(10 mL) and extracted with dichloromethane (20 mL, three times). The organic phase was
washed with water, dried over sodium sulphate, filtered and evaporated under reduced
pressure. The crude was then subjected to chromatography on silica gel (eluent: gradient of
dichloromethane 100% to a mixture of dichloromethane and methanol 95/5 (v/v)) and a
yellow powder was collected.
Yield: 20 mg (15%). 1H NMR (300 MHz, CDCl3, 25°C): δ = 8.99 (2H, s), 8.94 (2H, d, J = 8.4
Hz), 8.09 (2H, dd, J = 3.3 Hz, J = 6.3 Hz), 7.59 (4H, m), 3.19 (4H, m), 1.93 (4H, m), 1.52 (4H,
m), 1.02 (6H, t, J = 7.2 Hz). HRMS (MALDI) m/z: [M+H]+ calculated for C28H29N2+: 393.2331;
found: 393.2346. Δ = 3.8 ppm.

Complex [Cu(L1)2]+(PF6-) C1: ligand L1 (0.076 mmol, 30 mg) was dissolved in


dichloromethane (5 mL) and thoroughly degasssed by argon bubbling in an ultrasonic bath.
A degassed dichloromethane solution of Cu(CH3CN)4PF6 (0.038 mmol, 14.2 mg) was
syringed into the former; the solution turned red and was left at room temperature for half an
hour. The solvent was then removed by rotary evaporation under reduced pressure. The
crude was filtered through a silica plug in dichloromethane and evaporation of the filtrate
afforded C1.
Yield: 34 mg (90%). 1H NMR (300 MHz, CDCl3, 25°C): δ = 9.87 (4H, d, J = 8.1 Hz), 8.48 (4H,
dd, J = 3.3 Hz, J = 6.3 Hz), 8.05, (4H, dd, J = 3.6 Hz, J = 6.6 Hz), 8.02 (4H, d, J = 8.4 Hz),
2.86 (8H, m), 1.47 (8H, m), 0.83 (8H, m), 0.33 (12H, t, J = 7.2 Hz) ppm. HRMS (MALDI) m/z:
[M]+ calculated for C52H52N8Cu+: 851.3611; found: 851.3618. Δ = 0.8 ppm.

Complex [Cu(L2)2]+(PF6-) C2: complex C2 was isolated during the synthesis procedure of
ligand L2.
1
H NMR (300 MHz, CDCl3, 25°C): δ = 10.18 (4H, d, J = 8.7 Hz), 8.80 (4H, dd, J = 3.0 Hz, J =
6.6 Hz), 8.11 (4H, dd, J = 3.3 Hz, J = 6.6 Hz), 7.98 (4H, d, J = 8.7 Hz), 2.85 (8H, m), 1.47
(8H, m), 0.84 (8H, m), 0.40 (12H, t, J = 7.2 Hz). HRMS (MALDI) m/z: [M]+ calculated for
C60H52N8Cu+: 947.3611; found: 947.3592. Δ = 2.0 ppm.

Complex [Cu(L3)2]+(PF6-) C3: ligand L3 (0,025 mmol, 10 mg) was dissolved in


dichloromethane (5 mL) and thoroughly degassed by argon bubbling in an ultrasonic bath. A
degassed dichloromethane solution of Cu(CH3CN)4PF6 (0.013 mmol, 4.7 mg) was syringed
into the former; the solution turned red and was left at room temperature for half an hour. The
solvent was then removed by rotary evaporation under reduced pressure. The crude was
subjected to preparative thin layer chromatography, using as eluent the mixture
acetonitrile/H2O/saturated aqueous KNO3 (100/8/2 in volumes). The bright orange stripe was
collected and the silica was washed several times with a mixture of dichloromethane and
methanol (95/5 v/v) until it was colourless. The orange organic phase was evaporated under
reduced pressure affording a pure batch of C3.
Yield: 14.5 mg (58%). 1H NMR (300 MHz, CDCl3, 25°C): δ = 9.36 (4H, d, J = 8.7 Hz), 9.26
(4H, s), 8.23, (4H, m), 7.90 (4H, d, J = 8.4 Hz), 7.69 (4H, m), 2.77 (8H, m), 1.41 (8H, m),
0.75 (8H, m), 0.30 (12H, t, J = 7.2 Hz) ppm. HRMS (MALDI) m/z: [M]+ calculated for
C56H56N4Cu+: 847.3801; found: 847.3785. Δ = 1.9 ppm.

Results and discussion

1. Synthesis

Scheme 1. Synthetic routes for ligands L1, L2 and L3.

The syntheses of all ligands is depicted in scheme 1, and rely on the same starting material,
namely 2,9-di-nbutyl-1,10-phenanthroline-5,6-dione 1, obtained in good yields by oxidation of
2,9-di-nbutyl-1,10-phenanthroline.79 Following the well-established dppz synthesis protocol,
condensation of 1 onto 1,2-phenylenediamine 2 in refluxing ethanol afforded ligand L1. The
presence of floppy butyl alkyl chains endows L1 with a higher solubility than dppz; hence,
although the latter precipitated as a pure compound in the reaction mixture, a purification
step by chromatography was necessary to get a batch of pure L1. The synthesis of L3
proved to be more challenging. The related np ligand (namely L2 with no alkyl chains, see
figure 1) was synthesized twice before with quite low yields following two different strategies.
Our own protocol is inspired by both, and is presented in scheme 1. First, condensation of 1
with 1,2-diacetonitrilebenzene 3 in presence of 1,8-Diazabicyclo[5.4.0]undec-7-ene (DBU)
yielded L2. Once again, because of the higher solubility provided by the nbutyl chains, L2 did
not precipitate in the reaction medium. All our attempts to isolate L2 by precipitation
invariably lead to mixtures of L2 and 3, a mixture which proved to be very difficult to purify by
chromatography column. Taking advantage of the coordinating ability of L2 vs. 3, we
prepared complex C2 by adding the appropriate amount of Cu(CH3CN)4PF6 to the latter
mixture. C2 was then easily separated from 3 by chromatography. C2 was finally treated in a
mixture of CH3CN, water and potassium cyanide; the red solution quickly turned opaque as a
bright yellow precipitate of pure L2 formed. A batch of pure C2 was retrieved for further
analysis. The latter was recovered by extraction with dichloromethane, with an overall yield
of 44% over the whole process. The low yield could partly be explained by the fact that the
starting dione 1 is not stable in basic medium, all the more if there are traces of water
(formation of fluorenone).68 The synthesis of L3 consists in a one-pot hydrolysis of both nitrile
groups on L2 followed by decarboxylation in a strongly alkaline medium at high temperature.
Only small amounts of ligand L3 were obtained, despite all our efforts to increase the yield of
the reaction. Increasing the duration of the reaction did not improve the yield.

On the other hand, the synthesis of complexes C1 and C3 was straightforward.


Cu(CH3CN)4PF6 (1 eqv.) was added to dichloromethane solutions of L1 and L3 (2 eqv.)
under argon at room temperature. The solutions immediately turned orange-red,
unmistakable sign that coordination of copper(I) by two diimine ligands has taken place. As
mentioned above, C2 was incidentally obtained during the purification process of L2. High
resolution mass and NMR spectra of all ligands and complexes are all in agreement with the
proposed structures.

Electrochemistry

Cyclic voltammetry was performed on dichloromethane solutions of each complex in order to


probe their electronic properties in the ground state. Voltammograms are given in figure 2
and relevant potentials are reported in table 1. In all cases, a pseudo-reversible oxidation
wave is observed around 1 V vs. SCE, which can be unambiguously assigned to the
copper(II)/copper(I) redox couple. The difference between the anodic peak potential Epa and
the cathodic peak potential Epc is a good indicator of the reversibility of the electrochemical
process and should be close to 60 mV for a one-electron exchange process; it is much larger
in this case and for copper(I)-diimine complexes in general owing to the flattening of the
coordination sphere occurring when tetrahedral copper(I) is oxidized to square planar
copper(II).15,81 The butyl chains prevent this distortion to some extent, thus justifying the
Cu(II)/Cu(I) half wave potential is higher for sterically challenged copper(I) complexes than
for “naked” archetypal [CuI(phen)2]+.17 Accordingly, the steric bulk imposed by bulky
substituents in α of the chelating nitrogen atoms has a great influence on the Cu(II)/Cu(I) half
wave potential;4,12 however, since C1, C2, C3 and model C4 feature the same coordination
cage, we can safely assume that the relative oxidation and reduction potentials of all
complexes are mainly governed by electronic effects. The latter likely accounts for the more
positive half wave potentials measured for all complexes compared to reference C4. The
electron withdrawing effects brought by the phenazine (for C1), the dicarbonitrileanthracene
(for C2) and the anthracene (for C3) spacers deplete the electron density on the metal
centre, thus shifting the oxidation potential to more anodic values. C1 and C2 feature more
positive oxidation potentials than C3 because the anthracene unit is less π-accepting than
phenazine or dicarbonitrileanthracene. This is in contrast with observations made on
[RuII(bpy)2(dppz)]2+, [RuII(bpy)2(np-CN)]2+ and [RuII(bpy)2(np)]2+ where the presence of
phenazine, dicarbonitrileanthracene or anthracene moieties did not influence the oxidation
potentials of the complexes.69

In the case of C1 and C2, a rather prominent wave at ca. 0.8 V vs. SCE, absent when the
potential sweeping is stopped before the CuII/CuI oxidation wave, can be attributed to the
partial destruction of the complexes upon oxidation, likely the loss of one ligand upon
oxidation. The strong electron withdrawing character of the phenazine and
dicarbonitrileanthracene fragments may deplete the electron density on the nitrogen atoms,
therefore affecting the σ-donation ability of the latter. This phenomenon is not observed in
the case of C3 (nor C4),4 suggesting anthracene has a milder effect on the electronic
properties of the complex.

5 µA

E (V) / SCE

-1.6 -1.1 -0.6 -0.1 0.4 0.9 1.4

Figure 2. Cyclic voltammograms of dichloromethane solutions of C1, C2 and C3 (from bottom to top)
in presence of tetra-nbutylammonium hexafluorophosphate (0.1 M) at room temperature. Potentials
-1
are referenced vs. SCE, platinum disc used as working electrode. Sweep rate: 100 mV.s .

Copper(I)-diimine complexes usually display very negative reduction potentials, well below
the limit of electroactivity window of dichloromethane (ca. -1.5 V vs. SCE in our conditions),
and corresponding to one-electron ligand centred reductions. Yet, complexes C1 and C2
feature well defined cathodic waves and by comparison with previously published work on
related ruthenium complexes,60,69,82 those reduction waves were assigned to the addition of
an extra electron on the phenazine and dicarbonitrileanthracene spacers, respectively.
Logically, the more easily reduced complex C2 is the more difficult to oxidize too.
Conversely, complex C3 does not exhibit any electrochemical event when sweeping towards
negative potentials. The concatenation of two phenyl rings onto phenanthroline does not
increase to a sufficient extent the π accepting power of ligand L3 to promote the reduction of
the latter above -1.5 V vs. SCE. This is in line with the observations from Albano et al. made
on corresponding ruthenium complexes.69

complex E / V vs. SCE (ΔEp / mV) λabs / nm (ε / L.mol-1.cm-1) λem / nm (Φem x 10-4 / %) τ / ns
C1 1.11 (140) -1.19 (180) 456 (10900) 725 (3.6) 53
C2 1.16 (110) -0.88a 456 (10460) 815 (n.d.) n.d.
C3 0.99 (120) n.d. 470 (8600) 730 (5.8) 67
C4 0.92 (150) n.d. 457 (7000)4 725 (9)4 1504
Table 1. Relevant data extracted from electrochemistry and spectroscopic measurements.
a
Irreversible wave.

Electronic absorption spectroscopy

The absorption spectra of complexes C1, C2 and C3 were recorded in dichloromethane


(figure 3), the associated data are gathered in table 1. All three new complexes absorb in the
visible domain, owing to the famous broad MLCT spanning from 400 to 650 nm. Additional
bands take place between 350 and 420 nm; these are monitored on the spectra of the free
ligands (see supporting information, figure S1) and were thus assigned to ligand centred π-
π* transitions. Other intense ligand centred transitions take place in the UV; the latter are
red-shifted for the complexes with respect to the free ligands, a common feature confirming
the coordination of the ligands to the copper(I) ion. The MLCT of complexes C1 and C2 are
look-alike and exhibit a maximum absorbance wavelength very similar to model C4’s.
However, the transitions are substantially broader (see supporting information, figure S2) and
the absorption onsets appear thus red-shifted with respect to C4, showing that the phenazine
and anthracenedicarbonitrile moieties exert a stabilizing effect on the MLCT. The MLCT
absorption maximum of complex C3 is obviously red-shifted compared to C4 on the one
hand, and C1 and C2 on the other. The transition appears thinner than C1 and C2’s and the
absorption onset follows the order: C2>C1~C3 where “>” means “exhibits a longer absorption
onset wavelength than”. Thus, it seems difficult to relate the electronic effects monitored by
electrochemistry and those estimated by UV-Vis spectroscopy. The difference between
“optical” and “redox” orbitals is here again exemplified; this is all the more true since the
MLCT for [CuI(L)2]+ complexes is not a S0 to S1 transition, the latter being symmetry
forbidden.83

140000
extinction coefficient (L.mol-1.cm-1)

120000

100000

80000

60000

40000

20000

0
250 350 450 550 650 750
wavelength / nm

Figure 3. UV-Vis absorption spectra of complexes C1 (solid black line), C2 (solid grey line) and C3
(dotted black line) in dichloromethane.

Rewardingly, the extinction coefficients of all three complexes are higher than those of
common copper(I) complexes such as C4. This is likely a result of the increased dipole
moment along the z-axis of the molecule. Accordingly, L1 and L2 probably exhibit higher
dipolar moments than L3 because of the strong electronic effects imposed by the intracyclic
nitrogen atoms (for L1) and the nitrile groups (for L2). This translates into a larger extinction
coefficient for C1 and C2 compared to C3.

Calculations

The structures of the three complexes were fully optimized. The general structures of the
three complexes are very similar with one major difference. The complex C1 and C3 are
perfectly symmetric with a D2d symmetry. The complex C2 has a lower symmetry: the
anthracenedicarbonitrile part is no longer planar. The presence of the cyano groups on the
anthracene part induces an out of plane twist of the ligand (see supporting information, Error!
Reference source not found.). However, the symmetric structure is very close in energy, the
D2d equivalent of C2 is only 5 kcal.mol-1 less stable than distorded structure. Such a small
difference suggests that the ligands are oscillating between their two possible distorted
structures and probably the complex is dynamically planar; this is confirmed by the NMR
data showing only one species on the spectra.

All the structures were then optimized without symmetry and the absorption spectra were
computed on these structures for the three complexes (see supporting information, figure S4
and table S1). The theoretical results are in good agreement with the experimental ones. A
first set of transitions is present between 450 and 500 nm in the three complexes. This is
followed by a second set of transitions in the region of 325-400 nm. Then the most intense
band appears. We retrieve the qualitative experimental order. The lowest energy peak is for
C2 at 311 nm (294 and 318 nm exp.) followed by that of C3 at 293 nm (exp: 294 nm). The
most energetic transition is that of C1 at 282 nm (exp: 280 nm). However, we do not retrieve
the splitting of the peak of C2, which may arise from the dynamical behavior of the ligand.

The spectrum of C3 is the most simple and will be described first. The first absorbing band is
located at 484 nm and corresponds to a MLCT from the copper(I) ion to the phenantroline
part of the ligand (Error! Reference source not found.c). Two lower energy states are present
(534 and 516 nm) but have negligible intensities and have the same nature. A massif is
present between 325 to 400 nm. The peak at 375 nm is the combination of four electronic
transitions, two of them being MLCT from a deeper d orbital of copper again towards the
phenantroline part of the ligand. The other two are charge transfer inside the ligands from the
external part towards the central one (Error! Reference source not found.c). The third peak at
346 nm is generated by a single transition and is a π-π* transition delocalized on the whole
ligand. Finally, the very intense peak at 293 nm is generated by a single transition
corresponding to a π-π* excitation delocalized on the whole ligand. In C1, the replacement of
one C-H group by a N atom in the ligand leads to a qualitatively similar spectra with small
differences. The first intense peak is slightly blue-shifted at 470 nm but has exactly the same
nature as the peak at 484 nm in C3. However, a weak transition appears in C1 at 539 nm
due to the presence of the nitrogen. This new peak corresponds to a MLCT from copper to
the phenazine part of the ligand (Error! Reference source not found.a). A similar transition was
already present in C3 but was strongly blue shifted and had no absorption intensity. The
second intense band in C1 appears at 339 nm and is a π-π* transition delocalized on the
whole ligand similar to the weak peaks at 346 nm of C3 but much more intense. The
presence of the nitrogen in the ligand C1 instead of C-H in C3 modifies the orbital
distribution. The LUMO and LUMO+1, which were localized on the phenantroline part in C3
(as in most [CuI(L)2]+ model complexes), are now localized on the phenazine part of the
ligand in C1, the LUMO+2 and LUMO+3 corresponding to the phenantroline part. This is
corroborated by the electrochemical measurements, and explains the presence of new peaks
at low energy in C1. The presence of the CN group in C2 again change the orbital
distribution, making the anthracenedicarbonitrile part of the ligand much more electrophilic.
Indeed, the LUMO and LUMO+1 are localized on the anthracenedicarbonitrile part of the
ligand in C2 as in C1 but they are also delocalized on these CN groups. This is the reason of
the existence of low energy transitions of MLCT type (Error! Reference source not found.) at
624 nm. Then we retrieve the intense MLCT band at 473 nm present in C1 and in C3. New
bands appear around 405 nm which correspond to π-π* excitations localized on the
anthracenedicarbonitrile.

Transition / nm
complex Nature of the transition
(computed)

C1 539

C2 624

484

C3

375
293

Figure 4. Nature of the transitions determined by electron density differences between the excited
state and the ground state in C1, C2 and C3. In red, area electronically depleted upon transitions and
in green area electronically enriched upon transition.

The changes of the ligand in the complexes affect the nature of the lowest unoccupied orbital
and therefore the nature and energy of the lowest transitions. Compared to the reference C3,
there are additional LUMOs at lower energy, sources of new MLCT in C1 and C2, which are
not present in C3. In all case the main transitions are similar and weakly affected by the
modification introduced on the ligand, though it may be less true for C2. These extra MLCT
present in C1 and C2 are at the origin of the broadening of the MLCT band.

Luminescence properties

The steady state emission spectra of dichloromethane solutions of complexes C1, C2 and
C3 with matching absorbance at the excitation wavelength are given in figure 5 and related
data are reported in table 1. All complexes are luminescent but show different behaviours.
The emission spectra of C1, C3 and reference compound C4 are very comparable as
regards their shape and emission maxima, despite their very different molecular structures. A
similar observation has been made on parent ruthenium complexes.69 A very slight red-shift
(ca. 5 nm) of the emission maximum wavelength for C3 compared to C4 could be grounded
in the extended π electronic system of ligand L3 compared to plain phenanthroline. As
regards C1, the dppz ligand is reputed for its peculiar electronic structure, as mentioned
above; it has been indeed well demonstrated that the phenazine spacer and the chelating
pyridyl rings were not electronically coupled. Thus, the emission maxima wavelengths of
solutions of the famous complexes [RuII(bpy)2(dppz)]2+ and [RuII(bpy)3]2+ are virtually
identical. A similar behaviour can be observed here in the case of C1.
On the other hand, the maximum emission wavelength of complex C2 is substantially red-
shifted, which likely reflects the strong π-accepting nature of the anthracenedicarbonitrile
moiety. Intriguingly, the luminescence behaviour of C2 is markedly different from C1 and C3
whereas their absorption spectra are quite similar. Nevertheless, in the case of C2 the
spectrum shape is broader and the onset of absorption occurs at longer wavelength than C1
or C3 (see figure S2 and TD-DFT calculations reported in table S1). Moreover, one must
bear in mind that the observed MLCT band is not a S0 to S1 transition, the latter being
symmetry-forbidden for tetrahedral copper(I) complexes.5,83 As such, we only have few
experimental data on the energy of the S1 state from the absorption spectra of C1, C2 and
C3 and drawing conclusions from the latter could be irrelevant. Relationships between the
absorption and emission energies are besides difficult to establish in the particular case of
copper(I)-bis(diimine) complexes.17 Additionally, one cannot exclude a contribution from a
ligand centred emission as proposed by Albano et al. for the related ruthenium complex.69

450000

400000

350000
luminescence intensity (a.u.)

300000

250000

200000

150000

100000

50000

0
500 550 600 650 700 750 800 850
wavelength / nm

Figure 5. Emission spectra of complexes C1 (black solid line), C2 (grey solid line) and C3 (dotted
black line) in degassed dichloromethane. Solutions are absorption matched at the wavelength of
excitation (456 nm for C1 and C2, 470 nm for C3).

The strong variations of the emission quantum yields among the three complexes are
remarkable features on figure 5. The position of C2’s emission band is too far in the red part
of the visible to allow measuring a trustworthy quantum yield with our spectrometer.
Nevertheless, one can deduct from the absorption matched emission spectra of the three
complexes the following trend: Φ(C3) > Φ(C1) > Φ(C2) where Φ is the emission quantum
yield. A lower luminescence quantum yield for C2 is expectable given the strongly red-shifted
emission band by virtue of the gap law. On the other hand, the latter cannot account for the
fact that both C1 and C3 have lower emission quantum yields than reference C4. Partial
quenching of the luminescence by photo-induced intramolecular electron transfer can be
ruled out although copper(I)-bis(diimine) complexes are justly considered as very potent
photo-reductants. In our case, a direct photo-induced electron transfer from copper(I) to L1,
L2 or L3 upon excitation in the MLCT is endergonic by several hundreds of meV (see
supplemental information, table S2 and comments).

Interestingly though, the emission quantum yields and the extinction coefficients for C1, C3
and C4 are progressing in opposite directions: the higher ε, the lower the quantum yield. It is
well established now that copper(I) complexes such as C4 exhibit a thermally activated
delayed fluorescence mechanism (TADF).10,84-86 As such, the emission band of C4 is mainly
composed of a S1 to S0 transition. Yet in our approach we aimed at increasing the extinction
coefficients of copper complexes by enlarging the moment of the copper-ligand dipole. S0 to
Sn transitions are indeed favoured but in return S1 to S0 de-excitation transition too, leading to
an overall decrease of the emission quantum yield.6,12,17 Time resolved emission decays,
measured by time-correlated single photon counting after excitation at 440 nm show that
both complexes C1 and C3 have indeed much shorter emission lifetimes than complex C4
(ca. 50 ns vs. 150 ns respectively).4 However, the emission lifetimes of C1 and C3 are very
similar (unfortunately we could not manage to record a suitable fluorescence decay for
complex C2 because of the appearance of an intense parasitic signal during acquisition,
likely grounded in partial dissociation of the complex). The decays are given in figure S5, and
are in agreement with the usual behaviour of parent copper(I) complexes such as C4: a first
short component can be assigned to the various deactivation pathways following excited
state formation (flattening, internal conversion, intersystem crossing, prompt fluorescence).
Longer lived phases correspond to the radiative decay and were estimated at 53 and 47 ns
for complexes C1 and C3 respectively. Another phenomenon should therefore take place
explaining the weaker emission quantum yield of C1 compared to C3. Typically, since the
radiative constants are roughly similar, then the non-radiative constants are to be
incriminated, since Φ = kr/(kr + knr), where Φ, kr and knr are the emission quantum yield, the
radiative and the non-radiative constants, respectively. Knowing that lower lying MLCT
excited states have been highlighted in the case of C1 (and C2), their participation to the
non-radiative de-excitation pathway is plausible.82
Conclusion

In order to increase the notably low extinction coefficients of the MLCT of [CuI(L)2]+
complexes (where L is a diimine ligand), we designed three ligands derived from the
phenanthroline, aiming at increasing the length of the molecular dipole. The strategy was to
fuse aromatic rings onto the phenanthroline core. The consequences on the electronic
structure of the complexes were rationalized. In particular, all complexes were luminescent in
the near infrared domain, with variable emission quantum yields. It seemed that enlarging the
length of the dipole has two main effects: on one hand, it indeed increases the extinction
coefficient (above 10000 M-1.cm-1) but on the other it increases the non-radiative deactivation
pathways. In both cases, this is due to the increased probability of the transitions between
singlet states. The emission quantum yields are still modest, but great improvements could
be made by increasing the steric bulk around the copper(I) ion first at the level of the
substituents in α of nitrogen atoms (e.g. with isopropyl groups) and second by decorating the
phenanthroline moiety with methyl groups. This would improve not only the luminescence
intensity of the corresponding complexes, but the extinction coefficient too as a consequence
of enlarged dipole length.6,9,10,48

Acknowledgements

The authors wish to thank the HORIBA-Jobin Yvon Facility for giving access to a Deltaflex
TC-SPC device, Pr. Chantal Daniel for fruitful discussions, the ANR (PERCO program
n°ANR-16-CE07-0012-01), the HPC of Strasbourg for computing facilities and the CNRS for
funding.

Bibliography

(1) Blaskie, M. W.; McMillin, D. R. Inorg. Chem. 1980, 19, 3519.


(2) Everly, R. M.; Ziessel, R.; Suffert, J.; McMillin, D. R. Inorg. Chem. 1991, 30, 559.
(3) Ichinaga, A. K.; Kirchhoff, J. R.; McMillin, D. R.; Dietrich-Buchecker, C. O.; Marnot, P. A.;
Sauvage, J. P. Inorg. Chem. 1987, 26, 4290.
(4) Eggleston, M. K.; McMillin, D. R.; Koenig, K. S.; Pallenberg, A. J. Inorg. Chem. 1997, 36, 172.
(5) Iwamura, M.; Takeuchi, S.; Tahara, T. Accounts of Chemical Research 2015, 48, 782.
(6) Capano, G.; Rothlisberger, U.; Tavernelli, I.; Penfold, T. J. The Journal of Physical Chemistry A
2015, 119, 7026.
(7) Green, O.; Gandhi, B. A.; Burstyn, J. N. Inorg. Chem. 2009, 48, 5704.
(8) Huang, J.; Mara, M. W.; Stickrath, A. B.; Kokhan, O.; Harpham, M. R.; Haldrup, K.; Shelby, M.
L.; Zhang, X.; Ruppert, R.; Sauvage, J. P.; Chen, L. X. Dalton Transactions 2014, 43, 17615.
(9) McCusker, C. E.; Castellano, F. N. Inorg. Chem. 2013, 52, 8114.
(10) Garakyaraghi, S.; Crapps, P. D.; McCusker, C. E.; Castellano, F. N. Inorg. Chem. 2016, 55,
10628.
(11) Shaw, G. B.; Grant, C. D.; Shirota, H.; Castner, E. W.; Meyer, G. J.; Chen, L. X. J. Am. Chem. Soc.
2007, 129, 2147.
(12) Miller, M. T.; Gantzel, P. K.; Karpishin, T. B. Inorg. chem. 1999, 38, 3414.
(13) Gothard, N. A.; Mara, M. W.; Huang, J.; Szarko, J. M.; Rolczynski, B.; Lockard, J. V.; Chen, L. X.
Journal of Physical Chemistry A 2012, 116, 1984.
(14) Mara, M. W.; Jackson, N. E.; Huang, J.; Stickrath, A. B.; Zhang, X.; Gothard, N. A.; Ratner, M.
A.; Chen, L. X. The Journal of Physical Chemistry B 2013, 117, 1921.
(15) Accorsi, G.; Armaroli, N.; Duhayon, C.; Saquet, A.; Delavaux-Nicot, B.; Welter, R.; Moudam,
O.; Holler, M.; Nierengarten, J.-F. European Journal of Inorganic Chemistry 2010, 2010, 164.
(16) Lavie-Cambot, A.; Cantuel, M.; Leydet, Y.; Jonusauskas, G.; Bassani, D. M.; McClenaghan, N.
D. Coord. Chem. Rev. 2008, 252, 2572.
(17) Scaltrito, D. V.; Thompson, D. W.; O'Callaghan, J. A.; Meyer, G. J. Coord. Chem. Rev. 2000,
208, 243.
(18) Iwamura, M.; Watanabe, H.; Ishii, K.; Takeuchi, S.; Tahara, T. J. Am. Chem. Soc. 2011, 133,
7728.
(19) Agena, A.; Iuchi, S.; Higashi, M. Chemical Physics Letters 2017, 679, 60.
(20) Smolentsev, G.; Soldatov, A. V.; Chen, L. X. The Journal of Physical Chemistry A 2008, 112,
5363.
(21) Schmittel, M.; Ganz, A. Chem. Commun. 1997, 999.
(22) Miller, M. T.; Gantzel, P. K.; Karpishin, T. B. J. Am. Chem. Soc. 1999, 121, 4292.
(23) Alonso-Vante, N.; Nierengarten, J.-F.; Sauvage, J.-P. J. Chem. Soc., Dalton Trans. 1994, 1649.
(24) Hewat, T. E.; Yellowlees, L. J.; Robertson, N. Dalton Transactions 2014, 43, 4127.
(25) Bessho, T.; Constable, E. C.; Graetzel, M.; Hernandez Redondo, A.; Housecroft, C. E.; Kylberg,
W.; Nazeeruddin, M. K.; Neuburger, M.; Schaffner, S. Chemical Communications 2008, 3717.
(26) Sakaki, S.; Kuroki, T.; Hamada, T. J. Chem. Soc., Dalton Trans. 2002, 840.
(27) Housecroft, C. E.; Constable, E. C. Chem. Soc. Rev. 2015, 44, 8386.
(28) Sandroni, M.; Favereau, L.; Planchat, A.; Akdas-Kilig, H.; Szuwarski, N.; Pellegrin, Y.; Blart, E.;
Le Bozec, H.; Boujtita, M.; Odobel, F. Journal of Materials Chemistry A 2014, 2, 9944.
(29) Wills, K. A.; Mandujano-Ramirez, H. J.; Merino, G.; Oskam, G.; Cowper, P.; Jones, M. D.;
Cameron, P. J.; Lewis, S. E. Dyes Pigm. 2016, 134, 419.
(30) Magni, M.; Biagini, P.; Colombo, A.; Dragonetti, C.; Roberto, D.; Valore, A. Coord. Chem. Rev.
2016, 322, 69.
(31) Malzner, F. J.; Willgert, M.; Constable, E. C.; Housecroft, C. E. J. Mater. Chem. A 2017, 5,
13717.
(32) Khnayzer, R. S.; McCusker, C. E.; Olaiya, B. S.; Castellano, F. N. J. Am. Chem. Soc. 2013, 135,
14068.
(33) Paria, S.; Reiser, O. ChemCatChem 2014, 6, 2477.
(34) Reiser, O. Acc. Chem. Res. 2016, 49, 1990.
(35) Kern, J.-M.; Sauvage, J.-P. Journal of the Chemical Society, Chemical Communications 1987,
546.
(36) Pirtsch, M.; Paria, S.; Matsuno, T.; Isobe, H.; Reiser, O. Chemistry-a European Journal 2012,
18, 7336.
(37) Baralle, A.; Fensterbank, L.; Goddard, J.-P.; Ollivier, C. Chemistry – A European Journal 2013,
19, 10809.
(38) Flamigni, L.; Heitz, V.; Sauvage, J.-P. Struct. Bonding (Berlin, Ger.) 2006, 121, 217.
(39) Mara, M. W.; Bowman, D. N.; Buyukcakir, O.; Shelby, M. L.; Haldrup, K.; Huang, J.; Harpham,
M. R.; Stickrath, A. B.; Zhang, X.; Stoddart, J. F.; Coskun, A.; Jakubikova, E.; Chen, L. X. J. Am. Chem.
Soc. 2015, 137, 9670.
(40) Huang, J.; Buyukcakir, O.; Mara, M. W.; Coskun, A.; Dimitrijevic, N. M.; Barin, G.; Kokhan, O.;
Stickrath, A. B.; Ruppert, R.; Tiede, D. M.; Stoddart, J. F.; Sauvage, J.-P.; Chen, L. X. Angewandte
Chemie International Edition 2012, 51, 12711.
(41) Sandroni, M.; Maufroy, A.; Rebarz, M.; Pellegrin, Y.; Blart, E.; Ruckebusch, C.; Poizat, O.;
Sliwa, M.; Odobel, F. The Journal of Physical Chemistry C 2014, 118, 28388.
(42) Barnsley, J. E.; Scottwell, S. O.; Elliott, A. B. S.; Gordon, K. C.; Crowley, J. D. Inorg. Chem. 2016,
55, 8184.
(43) Armaroli, N.; Accorsi, G.; Gisselbrecht, J.-P.; Gross, M.; Eckert, J.-F.; Nierengarten, J.-F. New J.
Chem. 2003, 27, 1470.
(44) Listorti, A.; Accorsi, G.; Rio, Y.; Armaroli, N.; Moudam, O.; Gégout, A.; Delavaux-Nicot, B.;
Holler, M.; Nierengarten, J.-F. Inorg. Chem. 2008, 47, 6254.
(45) Megiatto, J. D.; Li, K.; Schuster, D. I.; Palkar, A.; Herranz, M. Á.; Echegoyen, L.; Abwandner, S.;
de Miguel, G.; Guldi, D. M. The Journal of Physical Chemistry B 2010, 114, 14408.
(46) Lazorski, M. S.; Gest, R. H.; Elliott, C. M. J. Am. Chem. Soc. 2012, 134, 17466.
(47) Armaroli, N. Chem. Soc. Rev 2001, 30, 113.
(48) Phifer, C. C.; McMillin, D. R. Inorg. Chem. 1986, 25, 1329.
(49) Sandroni, M.; Kayanuma, M.; Rebarz, M.; Akdas-Kilig, H.; Pellegrin, Y.; Blart, E.; Le, B. H.;
Daniel, C.; Odobel, F. Dalton Trans. 2013, 42, 14628.
(50) Ackermann, M. N.; Interrante, L. V. Inorg. Chem. 1984, 23, 3904.
(51) Chambron, J. C.; Sauvage, J. P.; Amouyal, E.; Koffi, P. Nouv. J. Chim. 1985, 9, 527.
(52) van der Salm, H.; Elliott, A. B. S.; Gordon, K. C. Coord. Chem. Rev. 2015, 282, 33.
(53) Horvath, R.; Gordon, K. C. Inorganica Chimica Acta 2011, 374, 10.
(54) Berger, S.; Fiedler, J.; Reinhardt, R.; Kaim, W. Inorg. Chem. 2004, 43, 1530.
(55) Fees, J.; Ketterle, M.; Klein, A.; Fiedler, J.; Kaim, W. J. Chem. Soc., Dalton Trans. 1999, 2595.
(56) Sasmal, P. K.; Saha, S.; Majumdar, R.; De, S.; Dighe, R. R.; Chakravarty, A. R. Dalton Trans.
2010, 39, 2147.
(57) Ma, T.; Xu, J.; Wang, Y.; Yu, H.; Yang, Y.; Liu, Y.; Ding, W.; Zhu, W.; Chen, R.; Ge, Z.; Tan, Y.; Jia,
L.; Zhu, T. Journal of Inorganic Biochemistry 2015, 144, 38.
(58) Scalese, G.; Correia, I.; Benítez, J.; Rostán, S.; Marques, F.; Mendes, F.; Matos, A. P.; Costa
Pessoa, J.; Gambino, D. Journal of Inorganic Biochemistry 2017, 166, 162.
(59) Wouters, K. L.; Tacconi, N. R.; Konduri, R.; Lezna, R. O.; MacDonnell, F. M. Photosynth. Res.
2006, 87, 41.
(60) Amouyal, E.; Homsi, A.; Chambron, J. C.; Sauvage, J. P. J. Chem. Soc., Dalton Trans. 1990,
1841.
(61) Troian-Gautier, L.; Moucheron, C. Molecules 2014, 19, 5028.
(62) White, T. A.; Witt, S. E.; Li, Z.; Dunbar, K. R.; Turro, C. Inorg. Chem. 2015, 54, 10042.
(63) White, T. A.; Maji, S.; Ott, S. Dalton Trans. 2014, 43, 15028.
(64) Heberle, M.; Tschierlei, S.; Rockstroh, N.; Ringenberg, M.; Frey, W.; Junge, H.; Beller, M.;
Lochbrunner, S.; Karnahl, M. Chem. - Eur. J. 2017, 23, 312.
(65) Friedman, A. E.; Chambron, J. C.; Sauvage, J. P.; Turro, N. J.; Barton, J. K. J. Am. Chem. Soc.
1990, 112, 4960.
(66) Brennaman, M. K.; Meyer, T. J.; Papanikolas, J. M. The Journal of Physical Chemistry A 2004,
108, 9938.
(67) Hartshorn, R. M.; Barton, J. K. J. Am. Chem. Soc. 1992, 114, 5919.
(68) Pellegrin, Y.; Sandroni, M.; Blart, E.; Planchat, A.; Evain, M.; Bera, N. C.; Kayanuma, M.; Sliwa,
M.; Rebarz, M.; Poizat, O.; Daniel, C.; Odobel, F. Inorg. Chem. 2011, 50, 11309.
(69) Albano, G.; Belser, P.; De Cola, L.; Gandolfi, M. T. Chemical Communications 1999, 1171.
(70) Wei, S.; Shao, Y.; Shi, X.; Lu, X.; Li, K.; Zhao, Z.; Guo, C.; Zhu, H.; Guo, W. Org. Electron. 2016,
29, 142.
(71) Lu, W.; Vicic, D. A.; Barton, J. K. Inorg. Chem. 2005, 44, 7970.
(72) te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca Guerra, C.; van Gisbergen, S. J. A.;
Snijders, J. G.; Ziegler, T. Journal of Computational Chemistry 2001, 22, 931.
(73) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. The Journal of Physical Chemistry
1994, 98, 11623.
(74) Van Lenthe, E.; Baerends, E. J. Journal of Computational Chemistry 2003, 24, 1142.
(75) Lenthe, E. v.; Ehlers, A.; Baerends, E.-J. The Journal of Chemical Physics 1999, 110, 8943.
(76) Pye, C. C.; Ziegler, T. Theoretical Chemistry Accounts 1999, 101, 396.
(77) van Gisbergen, S. J. A.; Snijders, J. G.; Baerends, E. J. Computer Physics Communications 1999,
118, 119.
(78) Hirata, S.; Head-Gordon, M. Chemical Physics Letters 1999, 314, 291.
(79) Ott, S.; Faust, R. Synthesis 2005, 2005, 3135.
(80) Pallenberg, A. J.; Koenig, K. S.; Barnhart, D. M. Inorg. Chem. 1995, 34, 2833.
(81) Federlin, P.; Kern, J. M.; Rastegar, A.; Dietrich-Buchecker, C.; Marnot, P. A.; Sauvage, J. P. New
J. Chem. 1990, 14, 9.
(82) Véry, T.; Ambrosek, D.; Otsuka, M.; Gourlaouen, C.; Assfeld, X.; Monari, A.; Daniel, C.
Chemistry – A European Journal 2014, 20, 12901.
(83) Parker, W. L.; Crosby, G. A. The Journal of Physical Chemistry 1989, 93, 5692.
(84) Kirchhoff, J. R.; Gamache, R. E.; Blaskie, M. W.; Del Paggio, A. A.; Lengel, R. K.; McMillin, D. R.
Inorg. Chem. 1983, 22, 2380.
(85) Everly, R. M.; McMillin, D. R. The Journal of Physical Chemistry 1991, 95, 9071.
(86) Siddique, Z. A.; Yamamoto, Y.; Ohno, T.; Nozaki, K. Inorg. Chem. 2003, 42, 6366.
Synopsis : increasing the π-conjugation of diimine ligands allowed to increase the dipole length of the
copper(I) complexes yielding photoluminescent copper(I) complexes with substantially improved
light harvesting efficiencies.

Вам также может понравиться