Next Article in Journal
Changes of Physicochemical, Bioactive Compounds and Antioxidant Capacity during the Brewing Process of Zhenjiang Aromatic Vinegar
Next Article in Special Issue
The Early Years of 2,2′-Bipyridine—A Ligand in Its Own Lifetime
Previous Article in Journal
Numerical Investigation on the Effects of Dielectric Barrier on a Nanosecond Pulsed Surface Dielectric Barrier Discharge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hinged and Wide: A New P^P Ligand for Emissive [Cu(P^P)(N^N)][PF6] Complexes

by
Sarah Keller
1,2,
Matthias Bantle
1,
Alessandro Prescimone
1,
Edwin C. Constable
1 and
Catherine E. Housecroft
1,*
1
Department of Chemistry, University of Basel, BPR 1096, Mattenstrasse 24a, CH-4058 Basel, Switzerland
2
Chimie ParisTech, PSL University CNRS, Institute of Chemistry for Life and Health Sciences, Laboratory for Inorganic Chemical Biology, 11 rue Pierre et Marie Curie, F-75005 Paris, France
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(21), 3934; https://doi.org/10.3390/molecules24213934
Submission received: 24 October 2019 / Revised: 29 October 2019 / Accepted: 29 October 2019 / Published: 31 October 2019
(This article belongs to the Special Issue Bipyridines: Synthesis, Functionalization and Applications)

Abstract

:
Heteroleptic [Cu(BIPHEP)(N^N)][PF6] complexes (BIPHEP = 1,1′-biphenyl-2,2′-diylbis(diphenylphosphane)), in which N^N is 2,2′-bipyridine (bpy), 6-methyl-2,2′-bipyridine (6-Mebpy), 6-ethyl-2,2′-bipyridine (6-Etbpy), or 5,5′-dimethyl-2,2′-bipyridine (5,5′-Me2bpy), have been synthesized and characterized using multinuclear NMR spectroscopies and electrospray ionization mass spectrometry. The single crystal structures of [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O and [Cu(BIPHEP)(6-Etbpy)][PF6] confirm distorted tetrahedral {Cu(P^P)(N^N)} coordination environments. Each compound shows a quasi-reversible Cu+/Cu2+ process. In deaerated solution, the compounds are weak emitters. Powdered samples are yellow emitters (λemmax in the range 558–583 nm) and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] exhibits the highest photoluminescence quantum yield (PLQY = 14%). On cooling to 77 K (frozen 2-methyloxolane), the emission maxima are red-shifted and the excited state lifetimes increase from τ1/2 < 8 μs, to τ1/2 values of up to 53 μs, consistent with the compounds with N^N = 6-Mebpy, 6-Etbpy and 5,5′-Me2bpy exhibiting thermally activated delayed fluorescence (TADF).

Graphical Abstract

1. Introduction

Organic light-emitting diodes (OLEDs) and light-emitting electrochemical cells (LECs) contribute to modern solid-state lighting technologies, and significant advances in the performances of these devices have been made in recent years [1,2,3,4,5,6]. The discovery of the photophysical phenomenon of thermally activated delayed fluorescence (TADF), which allows both singlet and triplet excitons to be harvested for light emission, has injected huge excitement into the field [7,8,9]. In the TADF process, after excitation of a compound into the lowest-lying singlet excited state, fast intersystem crossing to the excited triplet state, at a slightly lower energy (<0.38 eV [9]), takes place. As the triplet state is long lived, and emission as phosphorescence is rather slow, the thermal energy at room temperature allows repopulation of the singlet excited state from which fluorescence occurs. The emission at room temperature is, therefore, a combination of phosphorescence from the triplet excited state and fluorescence from the singlet excited state, allowing, in theory, a 100% internal quantum efficiency. Although organic compounds have been widely used for TADF applications [10,11] metal complexes are particularly effective and can be easily tuned through ligand modification to possess desired excited state energies [9]. The development of LECs which contain ionic copper(I) complexes in the emissive layer is especially appealing, since it capitalizes on the use of an Earth-abundant metal and [Cu(P^P)(N^N)]+ complexes (N^N = diimine, P^P = bisphosphane) which exhibit TADF [9]. Application of these copper(I) coordination compounds follows early results from McMillin and co-workers, who were the first to report that copper(I) complexes comprising PPh3 or P^P ligands and 2,2′-bipyridine (bpy) or 1,10-phenanthroline (phen) exhibited low-lying metal-to-ligand charge transfer (MLCT) excited states [12,13].
To date, the P^P ligands most thoroughly explored for incorporation into emissive [Cu(P^P)(N^N)]+ complexes have been 4,5-bis(diphenylphosphanyl)-9,9-dimethyl-9H-xanthene (xantphos, IUPAC PIN (9,9-dimethyl-9H-xanthene-4,5-diyl)bis(diphenylphosphane)) and bis(2-(diphenylphosphanyl)phenyl)ether (POP, IUPAC PIN oxydi(2,1-phenylene)]bis(diphenylphosphane)) (Scheme 1). Both are classed as wide bite–angle bisphosphanes [14], and the P–Cu–P bite angles in the solid-state structures of [Cu(POP)(bpy)][PF6].CHCl3, [Cu(POP)(bpy)][PF6].Et2O and [Cu(xantphos)(bpy)][PF6] are 115.00(3)° [15], 111.97(3)° [16] and 113.816(14) [17], respectively. A wide variety of bpy-based ligands have been combined with POP and xantphos in systematic investigations of [Cu(P^P)(N^N)]+ emitters [3,17,18,19,20,21,22,23] and references cited therein]. In contrast, variations of the P^P ligand (other than POP and xantphos) have received less attention. Replacing the PPh2 units in xantphos with PMes2 units (Mes = mesityl) leads to a P–Cu–P angle in [Cu(xantphosMes2)(bpy)][PF6] (xantphosMes2 = 9,9-dimethyl-9H-xanthene-4,5-diyl)bis[di(2,4,6-trimethylphenyl)phosphane]) of 115.96(2)° [24]. We have observed that introducing PtBu2 in place of PPh2 groups in the xantphos derivative 9,9-dimethyl-9H-xanthene-4,5-diyl)bis[di(tert-butyl)phosphane] is detrimental, and gives rise to very weakly emitting [Cu(P^P)(N^N)][PF6] compounds, likely due to the vibrational quenching effects of the tert-butyl substituents [25]. Retaining aromatic substituents is therefore critical for efficient energy to light conversion. Two sterically demanding ligands that fulfil this requirement are 1,1′-biphenyl-2,2′-diylbis(diphenylphosphane) (BIPHEP) and 1,1′-binaphthalene-2,2′-diylbis(diphenylphosphane) (BINAP) (Scheme 1). A search of the Cambridge Structural Database (v. 5.40 with February 2019 updates [26]) for {Cu(BINAP)(N^N)}-containing compounds, in which N^N is constrained to a chelating ligand, reveals only six entries [27,28,29,30] and for these, the range of P–Cu–P bond angles is 99.25–105.13°. Both BINAP and BIPHEP are expected to exhibit atropisomerism, but, at room temperature, interconversion of (R)- and (S)-BINAP is slow, whereas the atropisomers (R)- and (S)-BIPHEP (Scheme 1) interconvert rapidly in solution [31]. Coordination of BIPHEP to a metal centre can lead to chiral metal coordination compounds. The preparation and photophysical properties of [Cu(BINAP)(N^N)]+ complexes have been reported [28], but, even though BIPHEP is commercially available, no [Cu(BIPHEP)(N^N)]+ complexes appear to have been described. We were, therefore, inspired to investigate the structural and solution properties of such complexes. In this work, we describe the preparation and characterization of [Cu(BIPHEP)(bpy)][PF6], along with three derivatives, in which the bpy metal-binding domain bears alkyl substituents in the 5- or 6-positions.

2. Results and Discussion

2.1. Synthesis and Characterization of [Cu(BIPHEP)(N^N][PF6] Complexes

The [Cu(BIPHEP)(N^N)][PF6] complexes, in which N^N is bpy, 6-Mebpy, 6-Etbpy or 5,5′-Me2bpy, were prepared by adding CH2Cl2 solutions containing both BIPHEP and the respective 2,2′-bipyridine ligand to CH2Cl2 solutions of [Cu(MeCN)4][PF6]. This led to an immediate colour change from colourless to yellow, indicating the formation of the heteroleptic complexes without ligand redistribution; formation of a homoleptic [Cu(N^N)2][PF6] complex would have been accompanied by the development of a distinctive red colour. Attempts to prepare the analogous complex [Cu(BIPHEP)(6,6′-Me2bpy)][PF6] resulted in a mixture of homoleptic complexes as well as the desired heteroleptic compound, and since equilibration between hetero- and homoleptic complexes persisted in the solution, this reaction was not investigated further. The compounds [Cu(BIPHEP)(bpy)][PF6], [Cu(BIPHEP)(6-Mebpy)][PF6], [Cu(BIPHEP)(6-Etbpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] were characterized by elemental analysis, mass spectrometry, and multinuclear NMR spectroscopies.
In the ESI-MS spectrum of each compound, the base peak was assigned to the [M–PF6]+ ion (Figures S1–S4, see Supporting Information). The room temperature 1H and 13C{1H} NMR spectra were recorded in acetone-d6 solution and signals were assigned (see Materials and Methods section) using 2D-methods. Aromatic ring and atom labelling are shown in Scheme 2. COSY, HMQC and HMBC spectra are displayed in Figures S5–S16 (see Supporting Information). Figure 1 shows that the signals arising from the coordinated bpy, 6-Mebpy, 6-Etbpy and 5,5′-Me2bpy ligands are well resolved and sharp. The non-equivalent pyridine rings in [Cu(BIPHEP)(6-Mebpy)][PF6] and [Cu(BIPHEP)(6-Etbpy)][PF6] (Figure 1b,c) were readily distinguished, starting with the characteristic signal for proton HA6 in the COSY spectrum and using the NOESY signal between the 6-alkyl group and proton HB5. In [Cu(BIPHEP)(bpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6], the 1H and 13C{1H} NMR spectra at room temperature reveal that the rings (rings C in Scheme 1) in the biphenyl unit of BIPHEP are equivalent. Consistent with this is the observation of a single signal in the 31P{1H} NMR spectrum for the cation, at δ −1.2 ppm and δ −0.6 ppm, respectively, for [Cu(BIPHEP)(bpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]. In [Cu(BIPHEP)(6-Mebpy)][PF6] and [Cu(BIPHEP)(6-Etbpy)][PF6], the rings of the biphenyl unit are non-equivalent, and this is illustrated by comparing Figure 1a,d with Figure 1b,c, with a splitting of the signal for the HC3 protons. This is consistent with the modelled structure [32] shown in Figure 2; the biphenyl rings remain non-equivalent, even with partial rotation about the trans-annular C–C bond of the biphenyl unit. At 298 K, the solution 31P{1H} NMR spectrum of [Cu(BIPHEP)(6-Mebpy)][PF6] (Figure 3) and [Cu(BIPHEP)(6-Etbpy)][PF6] (Figure S17 in the Supporting Information) exhibits two signals of similar chemical shifts, in addition to the signal for the [PF6] ion.
We have previously discussed in detail the dynamic processes observed for complexes of the types [Cu(POP)(6-Rbpy)]+ and [Cu(xantphos)(6-Rbpy)]+, where 6-Rbpy is a 6-substituted 2,2′-bipyridine [19,33]. In the 1H-NMR spectra in Figure 1, it is noticeable that the regions associated with the phenyl rings of the PPh2 units (rings labelled D) are broader for compounds containing asymmetrical 6-Mebpy and 6-Etbpy than for those with bpy or 5,5′-Me2bpy. The 13C{1H} NMR spectra in Figure 4 are instructive. In [Cu(BIPHEP)(bpy)][PF6], all signals are well-resolved and have been assigned with the aid of HMQC and HMBC spectra. The two sets of CD1/D1’, CD2/D2’, CD3/D3’ and CD4/D4’ signals arise from pairs facing towards or away from the bpy alkyl substituent (Figure 2), and the spectra are consistent with unhindered rotation of the PPh2 phenyl rings. Introducing the 6-alkyl substituent results in a broadening of all 13C{1H} NMR resonances arising from rings C and D (Figure 4b), consistent with increased barriers to dynamic behaviour. An inspection of the changes in line shapes of the signals for CD2 and CD2’, and for CD3 and CD3 (highlighted in Figure 4 by the red hashed lines) suggests that the energy barrier to the rotation of one set of phenyls (presumably the rings closest to the alkyl substituent) is higher than that for the second set.

2.2. Single Crystal X-ray Diffraction

X-ray quality crystals of [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O, [Cu(BIPHEP)(6-Etbpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2, were obtained by diffusion of Et2O into CH2Cl2 solutions of the respective compound. The latter two compounds crystallize in the monoclinic space groups P21/c and C2/c, respectively, while [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2 and [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O crystallize in the triclinic space group P−1. The solid-state structures of the [Cu(BIPHEP)(N^N)]+ cations are shown in Figure 5 and Figure 6, and selected structural parameters are summarized in Table 1. The twist in the backbone of the BIPHEP ligand renders each [Cu(BIPHEP)(N^N)]+ cation chiral (Figure 7), and both enantiomers are present in the unit cell.
The copper atoms in the complexes exhibit distorted tetrahedral geometries, with τ4 parameters [34] of 0.81 for [Cu(BIPHEP)(bpy)]+, 0.85 for [Cu(BIPHEP)(6-Mebpy)]+, 0.78 for [Cu(BIPHEP)(6-Etbpy)]+, and 0.80 for [Cu(BIPHEP)(5,5′-Me2bpy)]+. For Td symmetry, τ4 = 1.00, and the values observed here indicate a distortion towards a trigonal pyramidal geometry. Figure 7 displays the four [Cu(BIPHEP)(N^N)]+ cations, each viewed along the P...P vector; the same enantiomer is shown for each complex. The Cu–P distances are typical (2.2248(6) to 2.2570(7) Å) and the same is true for the Cu–N distances (range = 2.0321(18) to 2.086(2) Å, Table 1). The P–Cu–P angles lie in the range 101.56(4)°–106.24(2)°. For the analogous POP and xantphos complexes with eight-membered rings, the bite angle of the coordinated phosphanes is around 113° for the complexes with 6-Mebpy or 6-Etbpy [18,19], and in the range 111.97(3)°–115.00(3)° for [Cu(POP)(bpy)][PF6] and [Cu(xantphos)(bpy)][PF6] [15,16,17]. As expected, the chelating 2,2′-bipyridine ligands exhibit similar N–Cu–N angles, in the range 79.75(14)°–80.65(10)°. The two rings of the bpy ligand are twisted in all the complex cations, with N–C–C–N torsion angles between 6.4(5)° in [Cu(BIPHEP)(6-Etbpy)]+, and −12.1(4)° in [Cu(BIPHEP)(6-Mebpy)]+.
There are no trends in the structural parameters as a function of the number or steric demands of the alkyl substituents attached to the bpy unit, and it appears that packing effects are dominant. In the complexes containing [Cu(BIPHEP)(bpy)][PF6]+, [Cu(BIPHEP)(6-Mebpy)]+ and [Cu(BIPHEP)(5,5′-Me2bpy)]+, intracation π-stacking interactions occur between a phenyl ring of a PPh2 group and one ring of the BIPHEP backbone (Figure 5 and Figure 6b, and Figure S18 in the Supporting Information) although the angles between the ring planes of 25.7°, 15.1° and 18.0° mean that the interactions are not optimal. Short C–H...F cation...anion contacts feature in all of the structures, and the accommodation of solvent molecules in [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2 also contributes to packing effects.

2.3. Electrochemical Behaviour

The electrochemical behaviour of the [Cu(BIPHEP)(N^N)][PF6] compounds was investigated using cyclic voltammetry (CV), and cyclic voltammograms for CH2Cl2 solutions are shown in Figures S19–S22 (see Supporting Information). Each compound exhibits a reversible or quasi-reversible oxidation process (Table 2) assigned to a Cu+/Cu2+ oxidation, but no ligand-based reduction processes were observed within the solvent-accessible window. For comparison, Table 2 also reports the oxidation potentials of [Cu(P^P)(bpy)][PF6] and [Cu(P^P)(6-Mebpy)][PF6] (P^P = POP and xantphos) [17,18], measured under the same conditions as the BIPHEP-containing compounds. For the BIPHEP complexes, the oxidation potentials increase on going from the complex with the unsubstituted bpy (+0.59 V), to 6-Mebpy (+0.72 V) and 6-Etbpy (+0.73 V). This follows the trend in the corresponding xantphos-containing compounds (Table 2), although, with the more flexible POP ligand, the introduction of the 6-methyl substituent has little effect. The trends for the BIPHEP-containing compounds are consistent with the introduction of the 6-alkyl substituent, hindering the flattening of the tetrahedral copper(I) towards square planar copper(II), thereby moving the oxidation to a higher potential. The Cu+/Cu2+ oxidation occurs at similar potentials in [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] and [Cu(BIPHEP)(bpy)][PF6] (Table 2), indicating that the presence of methyl groups in these positions has little stabilizing effect on the copper(I) oxidation state. The data in Table 2 reveal that metal oxidation in [Cu(BIPHEP)(bpy)][PF6] occurs at significantly lower potential than in [Cu(POP)(bpy)][PF6] and [Cu(xantphos)(bpy)][PF6], suggesting that the flattening of the copper coordination sphere that accompanies oxidation from Cu+ to Cu2+ occurs more easily. This should also be reflected in the excited state lifetimes, which are discussed in the next section. We note that there is no clear correlation between the Cu+/Cu2+ oxidation potential in solution and the τ4 value (see Section 2.2), which gives a measure of the distortion away from a tetrahedral geometry (for which τ4 = 1.00) in the solid state.

2.4. Photophysical Properties

The solution absorption spectra of the [Cu(BIPHEP)(N^N)][PF6] compounds in CH2Cl2 are displayed in Figure 8 and absorption maxima are given in Table 3. The spectra for all four compounds are similar, with intense, high-energy (230–330 nm) absorptions arising from spin-allowed, ligand-based π*←π transitions. The broad absorption, with λmax close to 400 nm, is assigned to the MLCT band, and values of λmax and εmax vary little across the series of compounds.
When excited into the MLCT band at 390 nm, deaerated solutions of [Cu(BIPHEP)(bpy)][PF6] did not show a detectable emission, while [Cu(BIPHEP)(6-Mebpy)][PF6], [Cu(BIPHEP)(6-Etbpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] exhibited weak emissions, with values of λemmax = 645, 636 and 641 nm, respectively (Figure S23 in Supporting Information). Since these emissions were so weak, we focused on solid-state and frozen-glass emission behaviour. Emission maxima, photoluminescence quantum yields (PLQY) and excited state lifetimes (τ1/2) are presented in Table 4. Powdered samples are yellow emitters with values of λemmax, in the range 558–583 nm (Figure 9a), and are similar to those observed for [Cu(P^P)(bpy)][PF6], [Cu(P^P)(6-Mebpy)][PF6] and [Cu(P^P)(6-Etbpy)][PF6], in which P^P is POP or xantphos [17,18,19]. A blue shift in the emission maximum, upon going from liquid solution to a solid state, is consistent with previous observations of these families of compounds [19,22]. The highest PLQY for the solid-state compounds (Table 4) is 14% for [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] and this also shows the longest excited state lifetime (8 μs). This may indicate better protection of the copper centre against quenchers (e.g., O2 and H2O) [22], although substitution in the 5,5′-positions, in the case of [Cu(POP)(5,5′-(CF3)2bpy)][PF6] and [Cu(xantphos)(5,5′-(CF3)2bpy)][PF6], proved less beneficial than a 6,6′-substitution pattern [17]. As noted earlier, we were not able to isolate [Cu(BIPHEP)(6,6′-Me2bpy)][PF6] and are not, therefore, able to compare the effects of going from 5,5′-Me2bpy to 6,6′-Me2bpy in the case of the BIPHEP co-ligand.
In order to investigate whether the [Cu(BIPHEP)(N^N)][PF6] compounds were TADF emitters, their emission and excited state lifetimes were recorded at low temperature and compared to their values at room temperature. Since the repopulation of the singlet excited state by reverse intersystem crossing is suppressed at low temperatures (ΔE > kT), the excited state lifetime should be considerably longer, since phosphorescence from the triplet excited state should be the dominant process. The normalized emission spectra of the [Cu(BIPHEP)(N^N)][PF6] complexes in frozen Me-THF at 77 K are depicted in Figure 9b, and the emission maxima and excited state lifetimes are summarized in Table 4. A comparison of the values of λemmax in the powders at room temperature with those at 77 K reveals that the emission maxima are red-shifted for all the complexes, consistent with the triplet excited state being lower in energy than the singlet excited state. A significant extension of the excited state lifetimes is observed (Table 4), and, for the compounds in which the bpy bears alkyl substituents, values of τ1/2 are ca. 50 μs. These data are consistent with [Cu(BIPHEP)(N^N)][PF6] (N^N = 6-Mebpy, 6-Etbpy and 5,5′-Me2bpy) being TADF emitters. The increased excited state lifetime on going from 298 to 77 K is particularly noticeable for [Cu(BIPHEP)(6-Etbpy)][PF6], where τ1/2 increases from 1 μs at room temperature to 53 μs at 77 K. At the same time, the temperature-induced red-shift in this complex is the smallest in the series, and at 77 K, it has an emission maximum at the shortest wavelength compared to others in the series. This raises the question as to whether other effects such as non-radiative decay processes occur in [Cu(BIPHEP)(6-Etbpy)][PF6].
Since [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] exhibited the highest PLQY and longest excited state lifetime in the solid state, we decided to conduct preliminary tests in LECs. Simple bilayer devices were fabricated by spin-coating a solution of [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] on indium tin oxide-coated glass substrates, covered with a film of poly(3,4-ethylenedioxythiophene) polystyrene sulfonate. Devices were finished with the thermal evaporation of an aluminium cathode. LECs exhibited yellow electroluminescence with λemmax = 586 nm, red-shifted with respect to the solid-state photoluminescence. However, even at a current density of 100 A m−2, the maximum luminance was as low as 12 cd m−2, with a maximum external quantum efficiency (EQE) of 0.03%. The LEC also showed only a short lifetime, with t1/2 = 0.1 hours. In comparison, LECs containing [Cu(xantphos)(6-Mebpy)][PF6], [Cu(xantphos)(6-Etbpy)][PF6] and [Cu(POP)(6-Etbpy)][PF6] in the emissive layers exhibit values of t1/2 greater than 15, 40 and 80 h, respectively [18,19].

3. Materials and Methods

3.1. General

1H and 13C{1H} and 31P-NMR spectra were recorded on a Bruker Avance III-500 spectrometer (Bruker BioSpin AG, Fällanden, Switzerland) at 298 K. 1H and 13C-NMR chemical shifts were referenced, with respect to residual solvent peaks (δ TMS = 0), and 31P-NMR chemical shifts, with respect to δ(85% aqueous H3PO4) = 0 ppm. Solution absorption and emission spectra were measured using a Shimadzu UV-2600 spectrophotometer and a Shimadzu RF-5301PC spectrofluorometer, respectively (Shimadzu Schweiz GmbH, 4153 Aesch, Switzerland). A Shimadzu LCMS-2020 instrument (Shimadzu Schweiz GmbH, 4153 Aesch, Switzerland) was used to record electrospray ionization (ESI) mass spectra; samples were introduced as MeCN solutions with added formic acid.
Quantum yields (CH2Cl2 solutions or powder samples) were measured with a Hamamatsu absolute photoluminescence quantum yield spectrometer C11347 Quantaurus-QY (Hamamatsu Photonics, 4500 Solothurn, Switzerland). Emission lifetimes and powder emission spectra were recorded using a Hamamatsu Compact Fluorescence lifetime Spectrometer C11367 Quantaurus-Tau (Hamamatsu Photonics), with an LED light source with λexc = 365 nm. Low temperature emission and lifetime measurements were made using an Edinburgh Instruments LP920-KS instrument (Edinburgh Instruments, Livingston, UK), in frozen 2-methyloxolane (Me-THF); a 410 nm excitation was obtained from pulsed third-harmonic radiation from a Quantel Brilliant b Nd:YAG laser (Quantal Laser, Lumibird, 22300 Lannion, France), equipped with a Rainbow optical parameter oscillator. The laser pulse duration and pulse frequency were ~10 ns and 10 Hz, respectively, with a typical pulse energy of 7 mJ. An iCCD camera (Andor, Oxford Instruments GmbH, Wiesbaden, Germany) was used for detection of the spectra. Single-wavelength kinetics were recorded using a photomultiplier tube.
6-Methyl-2,2′-bipyridine (6-Mebpy) and 6-ethyl-2,2′-bipyridine (6-Etbpy) [35] and [Cu(MeCN)4][PF6] [36] were prepared following the literature procedures. BIPHEP was bought from Alfa Aesar (Alfa Aesar GmbH, 76185 Karlsruhe, Germany), and 2,2′-bipyridine (bpy), 5,5′-dimethyl-2,2′-bipyridine (5,5′-Me2bpy) and 6,6′-dimethyl-2,2′-bipyridine (6,6′-Me2bpy) from Sigma-Aldrich (Merck, 9470 Buchs, Switzerland). All chemicals were used as received.

3.2. [Cu(BIPHEP)(bpy)][PF6]

A colourless solution of bpy (39 mg, 0.25 mmol) and BIPHEP (131 mg, 0.25 mmol) in CH2Cl2 (20 mL) was added to a colourless solution of [Cu(MeCN)4][PF6] (93 mg, 0.25 mmol) in CH2Cl2 (20 mL), turning the solution yellow. After stirring at room temperature for 2 h, the solvent was removed in vacuo. The crude product was redissolved in CH2Cl2 and [Cu(BIPHEP)(bpy)][PF6] (63 mg, 0.071 mmol, 28%), precipitated as a yellow powder by layering with Et2O. 1H-NMR (500 MHz, acetone-d6, 298 K): δ/ppm 8.76 (dt, J = 8.2, 1.0 Hz, 2H, HA3), 8.42 (dd, J = 5.2, 1.4 Hz, 2H, HA6), 8.29 (td, J = 7.9, 1.6 Hz, 2H, HA4), 7.67 (m, 2H, HA5), 7.42–7.37 (m, 2H, HD4/D4), 7.34–7.26 (overlapping m, 16H, HD2+D2’+D3/D3’+D4/D4’+C5), 7.26–7.17 (overlapping m, 6H, HD3/D3’+C4), 7.10 (m, 2H, HC6), 6.81 (m, 2H, HC3). 13C{1H} NMR (126 MHz, acetone-d6, 298 K): δ/ppm 152.9 (t, JPC = 2 Hz, CA2), 151.4 (CA6), 146.5 (t, JPC = 9 Hz, CC1), 140.2 (CA4), 135.7 (t, JPC = 3.5 Hz, CC6), 135.0 (t, JPC = 8 Hz, CD2/D2), 134.5 (t, JPC = 8 Hz, CD2/D2), 132.9 (t, JPC = 18 Hz, CD1/D1), 132.2 (t, JPC = 14 Hz, CC2), 131.6 (t, JPC = 18 Hz, CD1/D1), 131.6 (CC3), 131.5 (t, JPC = 1 Hz, CD4/D4), 131.2 (t, JPC = 1 Hz, CD4/D4), 130.6 (CC5), 130.0 (t, JPC = 5 Hz, CD3/D3), 129.4 (t, JPC = 5 Hz, CD3/D3), 128.6 (t, JPC = 3 Hz, CC4), 127.2 (CA5), 123.8 (CA3). 31P{1H} NMR (202 MHz, acetone-d6, 298 K): δ/ppm −1.2 (broad, FWHM = 270 Hz, BIPHEP), −142.5 (septet, JPF = 700 Hz, [PF6]). ESI MS: m/z 741.10 [M–PF6]+ (base peak, calc. 741.16). Found C 61.53, H 4.37, N 3.22; C46H36CuF6N2P3∙H2O requires C 61.03, H 4.23, N 3.09.

3.3. [Cu(BIPHEP)(6-Mebpy)][PF6]

A colourless solution of 6-Mebpy (43 mg, 0.25 mmol) and BIPHEP (131 mg, 0.25 mmol) in CH2Cl2 (20 mL) was added to a colourless solution of [Cu(MeCN)4][PF6] (93 mg, 0.25 mmol) in CH2Cl2 (20 mL), turning the solution yellow. After stirring at room temperature for 2 h, the solvent was removed under reduced pressure. The crude product was redissolved in CH2Cl2 and layered with Et2O, to precipitate [Cu(BIPHEP)(6-Mebpy)][PF6] (202 mg, 0.23 mmol, 92%) as a yellow powder. 1H-NMR (500 MHz, acetone-d6, 298 K): δ/ppm 8.72 (dt, J = 8.3, 1.0 Hz, 1H, HA3), 8.62 (d, J = 7.9 Hz, 1H, HB3), 8.29–8.21 (overlapping m, 2H, HB4+A4), 8.04 (dt, J = 5.2, 0.8 Hz, 1H, HA6), 7.73 (dd, J = 7.8, 0.9 Hz, 1H, HB5), 7.54 (m, 1H, HA5), 7.46–7.10 (overlapping m, 26H, HD2+D2’+D3+D3’+D4+D4’+C4+C4’+C5+C5’+C6+C6’), 6.99 (m, 1H, HC3/C3), 6.88 (m, 1H, HC3/C3), 2.58 (s, 3H, HMe). 13C{1H} NMR (126 MHz, acetone-d6, 298 K): δ/ppm 160.5 (CB6), 153.5 (CA2/B2), 152.7 (CA2/B2), 151.0 (CA6), 145.7 (br, CC1+C1), 140.6 (CA4/B4), 140.2 (CA4/B4), 136.2 (br, CC6/C6), 135.2 (br, overlapping, CD2/D2’), 134.2 (CD2/D2’), 134.1 (CD2/D2’), 129.45 (CD3/D3’), 129.4 (CD3/D3’), 128.8 (br, CC4+C4’), 127.6 (CB5), 127.0 (CA5), 123.9 (CA3), 121.4 (CB3); signals for CC3, CC5, CC1, CC2, CD1 were poorly resolved. 31P{1H} NMR (202 MHz, acetone-d6, 298 K): δ/ppm −1.3 (broad, FWHM = 428 Hz, BIPHEP) −3.7 (broad, FWHM = 401 Hz, BIPHEP), −144.2 (septet, JPF = 700 Hz, [PF6]). ESI MS: m/z 755.16 [M–PF6]+ (base peak, calc. 755.18). Found C 61.89, H 4.83, N 3.38; C47H38CuF6N2P3∙H2O requires C 61.41, H 4.39, N 3.05.

3.4. [Cu(BIPHEP)(6-Etbpy)][PF6]

A solution of 6-Etbpy (46 mg, 0.25 mmol) and BIPHEP (131 mg, 0.25 mmol) in CH2Cl2 (20 mL) was added to a solution of [Cu(MeCN)4][PF6] (93 mg, 0.25 mmol) in CH2Cl2 (20 mL), causing an immediate change from colourless to yellow. After stirring at room temperature for 2 h, the solvent was removed in vacuo. The crude material was redissolved in CH2Cl2. Layering with Et2O precipitated [Cu(BIPHEP)(6-Etbpy)][PF6] (208 mg, 0.23 mmol, 92%) as a yellow powder. 1H-NMR (500 MHz, acetone-d6, 298 K): δ/ppm 8.72 (d, J = 8.2 Hz, 1H, HA3), 8.62 (d, J = 7.8 Hz, 1H, HB3), 8.31 (t, J = 7.9 Hz, 1H, HB4), 8.24 (td, J = 7.6, 1.6 Hz, 1H, HA4), 7.91 (m, 1H, HA6), 7.79 (d, J = 7.7 Hz, 1H, HB5), 7.51 (m, 1H, HA5), 7.42–7.12 (overlapping m, 25H, HD4+D4’+C5+C5’+D2+D2’+D3+D3’+C4+C4’+C6/C6’), 7.08 (m, 1H, CC6/C6’), 7.00 (m, 1H, HC3), 6.88 (m, 1H, HC3), 3.03 (q, J = 7.8 Hz, 2H, HEt), 1.08 (t, J = 7.6 Hz, 3H, HEt). 13C{1H} (126 MHz, acetone-d6, 298 K): 164.8 (CB6), 153.6 (CA2/B2), 152.4 (CA2/B2), 151.0 (CA6), 145.8 (CC1/C1’), 145.6 (CC1/C1’), 140.8 (CB4), 140.2 (CA4), 136.1 (CC6+C6’), 135.3 (overlapping m, CD2/D2’), 134.2 (CD2/D2’), 134.0 (CD2/D2’), 132.7–131.6 (overlapping m, CD1+D1’+C2+C2’), 131.5 (br, CC3+C3’), 131.0 (CD4/D4’), 130.9 (CD4/D4’), 130.8 (CD4/D4’), 130.5 (CD4/D4’), 130.0 and 129.8 (br, overlapping m, CD3/D3’+C5+C5’), 129.4 (br, overlapping m, CD3/D3’), 128.7 (br, CC4/C4’), 128.5 (br, CC4/C4’), 126.9 (CA5), 125.4 (CB5), 123.9 (CA3), 121.6 (CB3), 35.5 (CEt), 13.2 (CEt); see text discussion. 31P{1H} NMR (162 MHz, acetone-d6, 300 K) δ/ppm −1.3 (broad, FWHM = 320 Hz, BIPHEP), −4.0 (broad, FWHM = 320 Hz, BIPHEP), −144.2 (septet, JPF = 708 Hz, [PF6]); see text discussion. ESI MS: m/z 769.28 [M–PF6]+ (base peak, calc. 769.20). Found C 62.40, H 4.98, N 3.22; C48H40CuF6N2P3∙0.5H2O requires C 62.37, H 4.47, N 3.03.

3.5. [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]

A solution of 5,5′-Me2bpy (46 mg, 0.25 mmol) and BIPHEP (131 mg, 0.25 mmol) in CH2Cl2 (20 mL) was added to a solution of [Cu(MeCN)4][PF6] (93 mg, 0.25 mmol) in CH2Cl2 (20 mL). The yellow reaction mixture was stirred at room temperature for 2 h, and then solvent was removed under reduced pressure. The crude product was redissolved in CH2Cl2 and layering with Et2O led to the precipitation of yellow [Cu(BIPHEP)(5,5′-Me2bpy)][PF6] (215 mg, 0.23 mmol, 92%). 1H-NMR (500 MHz, acetone-d6, 298 K): δ/ppm 1H-NMR (500 MHz, acetone-d6, 298 K) δ /ppm 8.55 (d, J = 8.2 Hz, 2H, HB3), 8.04 (m, 2H, HB4), 7.96 (m, 2H, HB6), 7.44–7.39 (m, 2H, HD4/D4’), 7.38–7.29 (overlapping m, 12H, HD2/D2’+D3/D3’+D4/D4’+C5), 7.27–7.19 (m, 10H, HD2/D2’+D3/D3’+C4), 7.15 (m, 2H, HC6), 6.76 (m, 2H, HC3), 2.30 (s, 6H, HMe). 13C{1H} NMR (126 MHz, acetone-d6, 298 K) δ /ppm 151.5 (CB6), 150.5 (t, J = 2 Hz, CB2), 145.7 (t, J = 10 Hz, CC1), 140.2 (CB4), 137.3 (CB5), 135.8 (t, J = 4 Hz, CC6), 135.2 (t, J = 9 Hz, CD2/D2’), 134.6 (t, J = 8 Hz, CD2/D2’), 132.8 (t, J = 18 Hz, CD1/D1’), 132.2 (t, J = 14 Hz, CC2), 131.7 (t, J = 3 Hz, CC3), 131.6 (CD4/D4’), 131.3 (t, J = 18 Hz, CD1/D1’), 131.1 (CD4/D4’), 130.6 (CC5), 130.0 (t, J = 5 Hz, CD3/D3’), 129.3 (t, J = 5 Hz, CD3/D3’), 128.6 (t, J = 3 Hz, CC4), 122.8 (CB3), 18.1 (CMe). 31P{1H} NMR (162 MHz, acetone-d6, 300 K) δ/ppm −0.6 (broad, FWHM = 350 Hz, BIPHEP), −144.2 (septet, JPF = 708 Hz, [PF6]). ESI MS: m/z 769.28 [M–PF6]+ (base peak, calc. 769.20). Found C 58.24, H 4.30, N 3.02; C48H40CuF6N2P3∙CH2Cl2 requires C 58.84, H 4.23, N 2.80.

3.6. Crystallography

Single crystal X-ray diffraction data were collected on a Bruker Kappa Apex2 diffractometer with data reduction, solution and refinement, using the programs APEX [37] and CRYSTALS [38]. Structural analysis used the program Mercury v. 4.1.2 [39,40].

3.7. [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2

C47H38Cl2CuF6N2P3, M = 972.19, yellow plate, monoclinic, space group P21/c, a = 14.3304(6), b = 15.4136(6), c = 20.6025(7) Å, β = 105.844(2)°, U = 4377.8(3) Å3, Z = 4, Dc = 1.475 Mg m−3, μ(Cu-Kα) = 3.397 mm−1, T = 123 K. Total 46,656 reflections, 8189 unique, Rint = 0.071. Refinement of 6109 reflections (550 parameters) with I >2σ(I) converged at final R1 = 0.0505 (R1 all data = 0.0733), wR2 = 0.1151 (wR2 all data = 0.1308), gof = 0.9710. CCDC 1958102.

3.8. [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O

C51H49CuF6N2O1.5P3, M = 984.42, yellow block, monoclinic, space group C2/c, a = 29.5271(19), b = 16.7361(11), c = 20.5516(14) Å, β = 111.249(2)°, U = 9465.5(11) Å3, Z = 8, Dc = 1.381 Mg m−3, μ(Cu-Kα) = 2.157 mm−1, T = 123 K. Total 31,108 reflections, 8537 unique, Rint = 0.023. Refinement of 8430 reflections (584 parameters) with I >2σ(I) converged at final R1 = 0.0540 (R1 all data = 0.0544), wR2 = 0.1197 (wR2 all data = 0.1197), gof = 1.0146. CCDC 1958100.

3.9. [Cu(BIPHEP)(6-Etbpy)][PF6]

C48H40CuF6N2P3, M = 915.31, yellow block, triclinic, space group P−1, a = 10.5819(11), b = 13.4147(14), c = 15.0134(16) Å, α = 93.766(3) β = 100.826(3), γ = 92.405(3)°, U = 2085.5(4) Å3, Z = 2, Dc = 1.457 Mg m−3, μ(Cu-Kα) = 2.377 mm−1, T = 123 K. Total 29,365 reflections, 7594 unique, Rint = 0.032. Refinement of 6965 reflections (541 parameters) with I >2σ(I) converged at final R1 = 0.0412 (R1 all data = 0.0502), wR2 = 0.0822 (wR2 all data = 0.1385), gof = 1.3893. CCDC 1958101.

3.10. [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2

C49H42Cl2CuF6N2P3, M = 1000.25, yellow block, triclinic, space group P−1, a = 10.9543(6), b = 10.9953(6), c = 18.8408(11) Å, α = 94.9723(17), β = 92.9633(18), γ = 96.7453(17)°, U = 2240.7(2) Å3, Z = 2, Dc = 1.482 Mg m−3, μ(Cu-Kα) = 3.334 mm−1, T = 123 K. Total 29,278 reflections, 7850 unique, Rint = 0.030. Refinement of 7808 reflections (568 parameters) with I >2σ(I) converged at final R1 = 0.0453 (R1 all data = 0.0453), wR2 = 0.0461 (wR2 all data = 0.0461), gof = 1.1280. CCDC 1958099.

4. Conclusions

We have prepared and characterized the compounds [Cu(BIPHEP)(bpy)][PF6], [Cu(BIPHEP)(6-Mebpy)][PF6], [Cu(BIPHEP)(6-Etbpy)][PF6] and [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]. 1H, 13C{1H} and 31P{1H} NMR spectra, including the use of 2D methods, reveal dynamic processes involving the BIPHEP ligand. The single crystal structures of [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2, [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O and [Cu(BIPHEP)(6-Etbpy)][PF6] were determined and confirmed distorted tetrahedral {Cu(P^P)(N^N)} coordination environments. Each compound shows a quasi-reversible Cu+/Cu2+ process, which moves to a higher potential upon the introduction of 6-methyl or 6-ethyl substituent into the bpy unit. The [Cu(BIPHEP)(N^N)][PF6] compounds are weak emitters in the deaerated solution, whereas powdered samples are yellow emitters at room temperature, with values of λemmax between 558 and 583 nm, and PLQY values up to 14%. Evidence that the compounds with N^N = 6-Mebpy, 6-Etbpy and 5,5′-Me2bpy exhibit TADF comes from the fact that their excited state lifetimes increase from τ1/2 < 8 μs to τ1/2 values of up to 53 μs on cooling from room temperature to 77 K, and that the emission maxima are red-shifted.

Supplementary Materials

The following are available online. Crystallographic data (cifs) for the four structures. Figures S1–S4: mass spectra; Figures S5–S16: 2D NMR spectra; Figure S17: 31P{1H} NMR spectrum of [Cu(BIPHEP)(6-Etbpy)][PF6]; Figure S18: structural diagrams to illustrate π-stacking interactions in [Cu(BIPHEP)(bpy)]+, [Cu(BIPHEP)(6-Mebpy)]+ and [Cu(BIPHEP)(5,5′-Me2bpy)]+; Figures S19–S22: cyclic voltammograms; Figure S23: solution emission spectra.

Author Contributions

Project conceptualization, administration, supervision and funding acquisition, E.C.C. and C.E.H.; investigation, S.K.; M.B.; crystallography, A.P.; manuscript writing, C.E.H. and S.K.; manuscript editing, E.C.C.

Funding

This research was partially funded by the Swiss National Science Foundation, grant number 200020_162631.

Acknowledgments

We gratefully acknowledge the support of the University of Basel. We thank Oliver S. Wenger (University of Basel) for allowing access to instruments to measure low-temperature excited-state lifetimes. We thank Marco Meyer and Fabian Brunner (University of Basel) for recording the ESI-MS spectra and for assisting with electrochemical measurements, and Isaak Nohara (University of Basel) for verifying the absorption and solid-state emission spectra. We also acknowledge the contributions of Michele Sessolo and Ana M. Igual (University of Valencia) for LEC data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bizzarri, C.; Spuling, E.; Knoll, D.M.; Volz, D.; Bräse, S. Sustainable metal complexes for organic light-emitting diodes (OLEDs). Coord. Chem. Rev. 2018, 373, 49–82. [Google Scholar] [CrossRef]
  2. Volz, D.; Wallesch, M.; Fléchon, C.; Danz, M.; Verma, A.; Navarro, J.M.; Zink, D.M.; Bräse, S.; Baumann, T. From iridium and platinum to copper and carbon: new avenues for more sustainability in organic light-emitting diodes. Green Chem. 2015, 17, 1988–2011. [Google Scholar] [CrossRef]
  3. Fresta, E.; Costa, R.D. Beyond traditional light-emitting electrochemical cells—A review of new device designs and emitters. J. Mater. Chem. C 2017, 5, 5643–5675. [Google Scholar] [CrossRef]
  4. Elie, M.; Gaillard, S.; Renaud, J.L. Light-Emitting Electrochemical Cells: Concepts, Advances and Challenges; Costa, R.D., Ed.; Springer International Publishing: New York, NY, USA, 2017. [Google Scholar] [CrossRef]
  5. Salehi, A.; Fu, X.; Shin, D.-H.; So, F. Recent Advances in OLED Optical Design. Adv. Funct. Mater. 2019, 29, 1808803. [Google Scholar] [CrossRef]
  6. Luo, D.; Chen, Q.; Liu, B.; Qiu, Y. Emergence of Flexible White Organic Light-Emitting Diodes. Polymers 2019, 11, 384. [Google Scholar] [CrossRef]
  7. Wong, M.Y.; Zysman-Colman, E. Purely Organic Thermally Activated Delayed Fluorescence Materials for Organic Light-Emitting Diodes. Adv. Mater. 2017, 29, 1605444. [Google Scholar] [CrossRef]
  8. Yerson, H. (Ed.) Highly Eficient OLEDs: Materials Based on Thermally Activated Delayed Fluorescence; John Wiley & Sons: Weinheim, Germany, 2019; ISBN 978-3-527-33900-6. [Google Scholar]
  9. Czerwieniec, R.; Leitl, M.J.; Homeier, H.H.H.; Yerson, H. Cu(I) complexes—Thermally activated delayed fluorescence. Photophysical approach and material design. Coord. Chem. Rev. 2016, 325, 2–28. [Google Scholar] [CrossRef]
  10. Bergmann, L.; Zink, D.M.; Bauman, T.; Volz, D.; Bräse, S. Metal-Oragnic and Organic TADF Materials: Status, Challenges and Characterization. Top. Curr. Chem. 2016, 374, 22. [Google Scholar] [CrossRef]
  11. Yang, Z.; Mao, Z.; Xie, Z.; Zhang, Y.; Liu, S.; Zhao, J.; Xu, J.; Chi, Z.; Aldred, M.P. Recent Advances in Organic Thermally Activated Delayed Fluorescence Materials. Chem. Soc. Rev. 2017, 46, 915–1016. [Google Scholar] [CrossRef]
  12. Buckner, M.T.; McMillin, D.R. Photoluminescence from copper(I) complexes with low-lying metal-to-ligand charge transfer excited states. J. Chem. Soc. Chem. Commun. 1978, 759–761. [Google Scholar] [CrossRef]
  13. Rader, R.A.; McMillin, D.R.; Buckner, M.T.; Matthews, T.G.; Casadonte, D.J.; Lengel, R.K.; Whittaker, S.B.; Darmon, L.M.; Lytle, F.E. Photostudies of 2,2′-bipyridine bis(triphenylphosphine)copper(1+), 1,10-phenanthroline bis(triphenylphosphine)copper(1+), and 2,9-dimethyl-1,10-phenanthroline bis(triphenylphosphine)copper(1+) in solution and in rigid, low-temperature glasses. Simultaneous multiple emissions from intraligand and charge-transfer states. J. Am. Chem. Soc. 1981, 103, 5906–5912. [Google Scholar] [CrossRef]
  14. Kramer, P.C.J.; Van Leeuwen, P.W.N.M.; Reek, J.N.H. Wide Bite Angle Diphosphines: Xantphos Ligands in Transition Metal Complexes and Catalysis. Acc. Chem. Rev. 2001, 34, 895–904. [Google Scholar] [CrossRef]
  15. Costa, R.D.; Tordera, D.; Ortí, E.; Bolink, H.J.; Schönle, J.; Graber, S.; Housecroft, C.E.; Constable, E.C.; Zampese, J.A. Copper(I) complexes for sustainable light-emitting electrochemical cells. J. Mater. Chem. C 2011, 21, 16108–16118. [Google Scholar] [CrossRef]
  16. Yang, L.; Feng, J.-K.; Ren, A.-M.; Zhang, M.; Ma, Y.-G.; Liu, X.-D. Structures, Electronic States and Electroluminescent Properties of a Series of CuI Complexes. Eur. J. Inorg. Chem. 2005, 2005, 1867–1879. [Google Scholar] [CrossRef]
  17. Keller, S.; Brunner, F.; Junquera-Hernández, J.M.; Pertegás, A.; La-Placa, M.-G.; Prescimone, A.; Constable, E.C.; Bolink, H.J.; Ortí, E.; Housecroft, C.E. CF3 Substitution of [Cu(P^P)(bpy)][PF6] complexes: Effects on Photophysical Properties and Light-emitting Electrochemical Cell Performance. ChemPlusChem 2018, 83, 217–229. [Google Scholar] [CrossRef]
  18. Keller, S.; Constable, E.C.; Housecroft, C.E.; Neuburger, M.; Prescimone, A.; Longo, G.; Pertegás, A.; Sessolo, M.; Bolink, H.J. [Cu(bpy)(P^P)]+ containing light-emitting electrochemical cells: improving performance through simple substitution. Dalton Trans. 2014, 43, 16593–16596. [Google Scholar] [CrossRef]
  19. Keller, S.; Pertegás, A.; Longo, G.; Martínez, L.; Cerdá, J.; Junquera-Hernández, J.M.; Prescimone, A.; Constable, E.C.; Housecroft, C.E.; Ortí, E.; et al. Shine bright or live long: substituent effects in [Cu(N^N)(P^P)]+-based light-emitting electrochemical cells where N^N is a 6-substituted 2,2′-bipyridine. J. Mater. Chem. C 2016, 4, 3857–3871. [Google Scholar] [CrossRef]
  20. Fresta, E.; Volpi, G.; Milanesio, M.; Garino, C.; Barolo, C.; Costa, R.D. Novel Ligand and Device Designs for Stable Light-Emitting Electrochemical Cells Based on Heteroleptic Copper(I) Complexes. Inorg. Chem. 2018, 57, 10469–10479. [Google Scholar] [CrossRef]
  21. Brunner, F.; Martínez-Sarti, L.; Keller, S.; Pertegás, A.; Prescimone, A.; Constable, E.C.; Bolink, H.J.; Housecroft, C.E. Peripheral halo-functionalization in [Cu(N^N)(P^P)]+ emitters: influence on the performances of light-emitting electrochemical cells. Dalton Trans. 2016, 45, 15180–15192. [Google Scholar] [CrossRef]
  22. Alkan-Zambada, M.; Keller, S.; Martínez-Sarti, L.; Prescimone, A.; Junquera-Hernández, J.M.; Constable, E.C.; Bolink, H.J.; Sessolo, M.; Ortí, E.; Housecroft, C.E. [Cu(P^P)(N^N)][PF6] compounds with bis(phosphane) and 6-alkoxy, 6-alkylthio, 6-phenyloxy and 6-phenylthio-substituted 2,2′-bipyridine ligands for light-emitting electrochemical cells. J. Mater. Chem. C 2018, 6, 8460–8471. [Google Scholar] [CrossRef]
  23. Keller, S.; Prescimone, A.; Bolink, H.J.; Sessolo, M.; Longo, G.; Martínez-Sarti, L.; Junquera-Hernández, J.-M.; Constable, E.C.; Ortí, E.; Housecroft, C.E. Luminescent copper(I) complexes with bisphosphane and halogen-substituted 2,2′-bipyridine ligands. Dalton Trans. 2018, 47, 14263–14276. [Google Scholar] [CrossRef] [PubMed]
  24. Brunner, F.; Babaei, A.; Pertegás, A.; Junquera-Hernández, J.M.; Prescimone, A.; Constable, E.C.; Bolink, H.J.; Sessolo, M.; Ortí, E.; Housecroft, C.E. Phosphane tuning in heteroleptic [Cu(N^N)(P^P)]+ complexes for light-emitting electrochemical cells. Dalton Trans. 2019, 48, 446–460. [Google Scholar] [CrossRef] [PubMed]
  25. Keller, S.; Prescimone, A.; Constable, E.C.; Housecroft, C.E. Copper(I) and silver(I) complexes of 9,9-dimethyl-4,5-bis(di-tert-butylphosphino)xanthene: photophysical properties and structural rigidity under pressure. Photochem. Photobiol Sci. 2018, 17, 375–385. [Google Scholar] [CrossRef] [PubMed]
  26. Groom, C.R.; Bruno, I.J.; Lightfoot, M.P.; Ward, S.C. The Cambridge Structural Database. Acta Cryst. 2016, B72, 171–179. [Google Scholar] [CrossRef]
  27. Shi, Y.; Liu, X.; Shan, Y.; Zhang, X.; Kong, W.; Lu, Y.; Tan, Z.; Li, X.-L. Naked-eye repeatable off–on–off and on–off–on switching luminescence of copper(I)-1H-imidazo[4,5-f][1,10]phenanthroline complexes with reversible acid–base responses. Dalton Trans. 2019, 48, 2430–2441. [Google Scholar] [CrossRef]
  28. Saito, K.; Tsukuda, T.; Tsubomura, T. Synthesis and Luminescence of New Cu(I) Complexes Containing a Binap Ligand and a Diimine Ligand. Bull. Chem. Soc. Jpn. 2006, 79, 437–441. [Google Scholar] [CrossRef]
  29. Feng, X.-Y.; Xin, X.-L.; Guo, Y.-M.; Chen, L.-L.; Liang, Y.-Y.; Xu, M.; Li, X.-L. Synthesis, structure and solid luminescence of copper(I)–bromodiimine–diphosphine complexes. Polyhedron 2015, 101, 23–28. [Google Scholar] [CrossRef]
  30. Tan, C.-H.; Ma, X.; Zhu, Q.-L.; Huang, Y.-H.; Fu, R.-B.; Hu, S.-M.; Sheng, T.-L.; Wu, X.-T. Syntheses, crystal structures and properties of dinuclear copper(I) complexes. J. Mol. Struct. 2012, 1007, 26–30. [Google Scholar] [CrossRef]
  31. Oczipka, P.; Müller, D.; Leitner, W.; Franciò, G. Enantiodivergent asymmetric catalysis with the tropos BIPHEP ligand and a proline derivative as chiral selector. Chem. Sci. 2016, 7, 678–683. [Google Scholar] [CrossRef]
  32. Spartan ’16, v. 2.0.9, Wavefunction Inc.: Irvine, CA, USA, 2017.
  33. Brunner, F.; Graber, S.; Baumgartner, Y.; Häussinger, D.; Prescimone, A.; Constable, E.C.; Housecroft, C.E. The effects of introducing sterically demanding aryl substituents in [Cu(N^N)(P^P)]+ complexes. Dalton Trans. 2017, 46, 6379–6391. [Google Scholar] [CrossRef]
  34. Yang, L.; Powell, D.R.; Houser, R.P. Structural variation in copper(I) complexes with pyridylmethylamide ligands: structural analysis with a new four-coordinate geometry index, τ4. Dalton Trans. 2007, 955–964. [Google Scholar] [CrossRef] [PubMed]
  35. Shaul, M.; Cohen, Y. Novel Phenanthroline-Containing Trinuclear Double-Stranded Helicates: Self-Recognition between Helicates with Phenanthroline and Bipyridine Binding Sites. J. Org. Chem. 1999, 64, 9358–9364. [Google Scholar] [CrossRef]
  36. Kubas, G.J. Tetrakis(Acetonitrile)Copper(I) Hexafluorophosphate. Inorg. Synth. 1979, 19, 90–92. [Google Scholar] [CrossRef]
  37. APEX2, version 2 User Manual, M86-E01078, Bruker Analytical X-ray Systems Inc.: Madison, WI, USA, 2006.
  38. Betteridge, P.W.; Carruthers, J.R.; Cooper, R.I.; Prout, K.; Watkin, D.J. CRYSTALS version 12: Software for guided crystal structure analysis. J. Appl. Cryst. 2003, 36, 1487. [Google Scholar] [CrossRef]
  39. Macrae, C.F.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Shields, G.P.; Taylor, R.; Towler, M.; van de Streek, J. Mercury: visualization and analysis of crystal structures. J. Appl. Cryst. 2006, 39, 453–457. [Google Scholar] [CrossRef]
  40. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0 - new features for the visualization and investigation of crystal structures. J. Appl. Cryst. 2008, 41, 466–470. [Google Scholar] [CrossRef]
Sample Availability: Limited samples of the N^N ligands are available from the authors.
Scheme 1. Structures of the bis(phosphane) ligands POP, xantphos, BINAP and BIPHEP.
Scheme 1. Structures of the bis(phosphane) ligands POP, xantphos, BINAP and BIPHEP.
Molecules 24 03934 sch001
Scheme 2. The structure of the ligands and numbering scheme for NMR spectroscopic assignments. The phenyl rings in BIPHEP are labelled D.
Scheme 2. The structure of the ligands and numbering scheme for NMR spectroscopic assignments. The phenyl rings in BIPHEP are labelled D.
Molecules 24 03934 sch002
Figure 1. 1H-NMR spectra (aromatic regions, 500 MHz, acetone-d6, 298 K) of (a) [Cu(BIPHEP)(bpy)][PF6], (b) [Cu(BIPHEP)(6-Mebpy)][PF6], (c) [Cu(BIPHEP)(6-Etbpy)][PF6] and (d) [Cu(BIPHEP)(5,5′-Me2bpy)][PF6].
Figure 1. 1H-NMR spectra (aromatic regions, 500 MHz, acetone-d6, 298 K) of (a) [Cu(BIPHEP)(bpy)][PF6], (b) [Cu(BIPHEP)(6-Mebpy)][PF6], (c) [Cu(BIPHEP)(6-Etbpy)][PF6] and (d) [Cu(BIPHEP)(5,5′-Me2bpy)][PF6].
Molecules 24 03934 g001
Figure 2. Modelled structure (not energy minimized) of [Cu(BIPHEP)(6-Rbpy)]+, in which the alkyl group R is represented by the red sphere and the trans-annular bond in the biphenyl unit is shown in purple. Partial rotation about the latter renders the two D’ rings equivalent and the two D rings equivalent, but the pair C and C’, and the pair D and D’, remain non-equivalent.
Figure 2. Modelled structure (not energy minimized) of [Cu(BIPHEP)(6-Rbpy)]+, in which the alkyl group R is represented by the red sphere and the trans-annular bond in the biphenyl unit is shown in purple. Partial rotation about the latter renders the two D’ rings equivalent and the two D rings equivalent, but the pair C and C’, and the pair D and D’, remain non-equivalent.
Molecules 24 03934 g002
Figure 3. The signal arising from the [Cu(BIPHEP)(6-Mebpy)]+ cation in the 31P{1H} NMR spectrum (202 MHz, acetone-d6, 298 K) of [Cu(BIPHEP)(6-Mebpy)][PF6].
Figure 3. The signal arising from the [Cu(BIPHEP)(6-Mebpy)]+ cation in the 31P{1H} NMR spectrum (202 MHz, acetone-d6, 298 K) of [Cu(BIPHEP)(6-Mebpy)][PF6].
Molecules 24 03934 g003
Figure 4. 13C{1H} NMR spectra (126 MHz, acetone-d6, 298 K) of (a) [Cu(BIPHEP)(bpy)][PF6] and (b) [Cu(BIPHEP)(6-Etbpy)][PF6], with expansions of the regions δ 137–128 ppm (signals from CC6 to CC4), shown in the insets at the top-left. Scale: δ/ppm.
Figure 4. 13C{1H} NMR spectra (126 MHz, acetone-d6, 298 K) of (a) [Cu(BIPHEP)(bpy)][PF6] and (b) [Cu(BIPHEP)(6-Etbpy)][PF6], with expansions of the regions δ 137–128 ppm (signals from CC6 to CC4), shown in the insets at the top-left. Scale: δ/ppm.
Molecules 24 03934 g004
Figure 5. Structures of (a) the [Cu(BIPHEP)(bpy)]+ cation in [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2 and (b) the [Cu(BIPHEP)(6-Mebpy)]+ cation in [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O. Ellipsoids are plotted at 50% probability in (a) and 40% in (b); H atoms are omitted for clarity. Each cation exhibits π-stacking between two phenyl rings in BIPHEP backbone. In (a) the angle between the ring planes = 25.7°, centroid...centroid distance = 3.85 Å, and in (b) the angle between the ring planes = 15.1°, centroid...centroid distance = 3.64 Å.
Figure 5. Structures of (a) the [Cu(BIPHEP)(bpy)]+ cation in [Cu(BIPHEP)(bpy)][PF6]∙CH2Cl2 and (b) the [Cu(BIPHEP)(6-Mebpy)]+ cation in [Cu(BIPHEP)(6-Mebpy)][PF6]∙Et2O∙0.5H2O. Ellipsoids are plotted at 50% probability in (a) and 40% in (b); H atoms are omitted for clarity. Each cation exhibits π-stacking between two phenyl rings in BIPHEP backbone. In (a) the angle between the ring planes = 25.7°, centroid...centroid distance = 3.85 Å, and in (b) the angle between the ring planes = 15.1°, centroid...centroid distance = 3.64 Å.
Molecules 24 03934 g005
Figure 6. Structures of (a) the [Cu(BIPHEP)(6-Etbpy)]+ cation in [Cu(BIPHEP)(6-Etbpy)][PF6] and (b) the [Cu(BIPHEP)(5,5′-Me2bpy)]+ cation in [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2. Ellipsoids are plotted at 40% probability in (a) and 50% in (b); H atoms are omitted for clarity. For the π-stacking interaction shown in (b), the angle between the ring planes is 18.0° and the centroid...centroid distance is 3.70 Å.
Figure 6. Structures of (a) the [Cu(BIPHEP)(6-Etbpy)]+ cation in [Cu(BIPHEP)(6-Etbpy)][PF6] and (b) the [Cu(BIPHEP)(5,5′-Me2bpy)]+ cation in [Cu(BIPHEP)(5,5′-Me2bpy)][PF6]∙CH2Cl2. Ellipsoids are plotted at 40% probability in (a) and 50% in (b); H atoms are omitted for clarity. For the π-stacking interaction shown in (b), the angle between the ring planes is 18.0° and the centroid...centroid distance is 3.70 Å.
Molecules 24 03934 g006
Figure 7. Comparisons of the distorted tetrahedral coordination spheres (viewed along the P...P vector) of copper(I) in (a) [Cu(BIPHEP)(bpy)]+4 = 0.81), (b) [Cu(BIPHEP)(6-Mebpy)]+4 = 0.85), (c) [Cu(BIPHEP)(6-Etbpy)]+4 = 0.78), and (d) [Cu(BIPHEP)(5,5′-Me2bpy)]+4 = 0.80). The origin of the chirality is the twisted backbone of BIPHEP (shown in purple).
Figure 7. Comparisons of the distorted tetrahedral coordination spheres (viewed along the P...P vector) of copper(I) in (a) [Cu(BIPHEP)(bpy)]+4 = 0.81), (b) [Cu(BIPHEP)(6-Mebpy)]+4 = 0.85), (c) [Cu(BIPHEP)(6-Etbpy)]+4 = 0.78), and (d) [Cu(BIPHEP)(5,5′-Me2bpy)]+4 = 0.80). The origin of the chirality is the twisted backbone of BIPHEP (shown in purple).
Molecules 24 03934 g007
Figure 8. Solution absorption spectra of the [Cu(BIPHEP)(N^N)][PF6] complexes (CH2Cl2, 3.5–5.3 × 10−5 mol dm−3).
Figure 8. Solution absorption spectra of the [Cu(BIPHEP)(N^N)][PF6] complexes (CH2Cl2, 3.5–5.3 × 10−5 mol dm−3).
Molecules 24 03934 g008
Figure 9. (a) Normalized emission spectra of the powdered [Cu(BIPHEP)(N^N)][PF6] compounds, λexc = 365 nm. The curves for [Cu(BIPHEP)(bpy)][PF6] and [Cu(BIPHEP)(6-Mebpy)][PF6] are virtually superimposed. (b) Normalized emission spectra of the [Cu(BIPHEP)(N^N)][PF6] complexes in frozen Me-THF at 77 K. λexc = 410 nm.
Figure 9. (a) Normalized emission spectra of the powdered [Cu(BIPHEP)(N^N)][PF6] compounds, λexc = 365 nm. The curves for [Cu(BIPHEP)(bpy)][PF6] and [Cu(BIPHEP)(6-Mebpy)][PF6] are virtually superimposed. (b) Normalized emission spectra of the [Cu(BIPHEP)(N^N)][PF6] complexes in frozen Me-THF at 77 K. λexc = 410 nm.
Molecules 24 03934 g009
Table 1. Selected structural parameters for the [Cu(BIPHEP)(N^N][PF6] compounds.
Table 1. Selected structural parameters for the [Cu(BIPHEP)(N^N][PF6] compounds.
Cation in [Cu(BIPHEP)(N^N)][PF6]Cu–P Distance/ÅCu–N distance/ÅP–Cu–P Angle/DegN–Cu–N Angle/DegN–C–C–N Torsion Angle/Deg
[Cu(BIPHEP)(bpy)]+Cu1–P1 = 2.2415(10);
Cu1–P2 = 2.2292(9)
Cu1–N1 = 2.071(3);
Cu1–N2 = 2.040(3)
P2–Cu1–P1 = 104.70(4)N2–Cu1–N1 = 80.15(11)−11.7(5)
[Cu(BIPHEP)(6-Mebpy)]+Cu1–P1 = 2.2570(7);
Cu1–P2 = 2.2549(7)
Cu1–N1 = 2.086(2);
Cu1–N2 = 2.049(2)
P2–Cu1–P1 = 105.01(2)N1–Cu1–N2 = 80.65(10)−12.1(4)
[Cu(BIPHEP)(6-Etbpy)]+Cu1–P1 = 2.2410(11);
Cu1–P2 = 2.2322(10)
Cu1–N1 = 2.072(4);
Cu1–N2 = 2.063(4)
P2–Cu1–P1 = 101.56(4)N2–Cu1–N1 = 79.75(14)6.4(5)
[Cu(BIPHEP)(5,5′-Me2bpy)]+Cu1–P16 = 2.2248(6);
Cu1–P41 = 2.2515(6)
Cu1–N2 = 2.0321(18);
Cu1–N15 = 2.0751(18)
P16–Cu1–P41 = 106.24(2)N2–Cu1–N15 = 80.51(7)−10.3(3)
Table 2. Cyclic voltammetric data for [Cu(BIPHEP)(N^N)][PF6] complexes referenced to internal Fc/Fc+ = 0.0 V; CH2Cl2 (dried) solutions, with [nBu4N][PF6] as the supporting electrolyte and a scan rate of 0.1 V s−1. Processes are quasi-reversible.
Table 2. Cyclic voltammetric data for [Cu(BIPHEP)(N^N)][PF6] complexes referenced to internal Fc/Fc+ = 0.0 V; CH2Cl2 (dried) solutions, with [nBu4N][PF6] as the supporting electrolyte and a scan rate of 0.1 V s−1. Processes are quasi-reversible.
Cation in [Cu(BIPHEP)(N^N][PF6] E 1 / 2 o x /VEpcEpa/mV
[Cu(BIPHEP)(bpy)]++0.59100
[Cu(BIPHEP)(6-Mebpy)]++0.72129
[Cu(BIPHEP)(6-Etbpy)]++0.73133
[Cu(BIPHEP)(5,5′-Me2bpy)]++0.57197
[Cu(POP)(bpy)]++0.72 a110 a
[Cu(POP)(6-Mebpy)]++0.69 b-
[Cu(xantphos)(bpy)]++0.76 a110 a
[Cu(xantphos)(6-Mebpy)]++0.90 a100 a
a Data taken from reference [17]. b Data taken from reference [18].
Table 3. Solution absorption maxima for the [Cu(BIPHEP)(N^N)][PF6] complexes (CH2Cl2 with solution concentrations in the range 3.5–5.3 × 10−5 mol dm−3).
Table 3. Solution absorption maxima for the [Cu(BIPHEP)(N^N)][PF6] complexes (CH2Cl2 with solution concentrations in the range 3.5–5.3 × 10−5 mol dm−3).
Cation in [Cu(BIPHEP)(N^N)][PF6]λmax/nm (ε/dm3 mol−1 cm−1)
Ligand-Based π*←πMLCT
[Cu(BIPHEP)(bpy)]+244 sh (32,000), 291 (19,400), 309 sh (12,400)407 (3500)
[Cu(BIPHEP)(6-Mebpy)]+292 (19,400), 301 (18,500), 316 sh (14,500) 400 (3400)
[Cu(BIPHEP)(6-Etbpy)]+292 (22,500), 301 (22,100), 315 sh (15,800)400 (4000)
[Cu(BIPHEP)(5,5′-Me2bpy)]+248 sh (36,200), 301 (21,300), 319 sh (15,200) 390 (4400)
Table 4. Emission maxima, photoluminescence quantum yields (PLQY) and lifetimes (τ1/2) for [Cu(BIPHEP)(N^N)][PF6] complexes.
Table 4. Emission maxima, photoluminescence quantum yields (PLQY) and lifetimes (τ1/2) for [Cu(BIPHEP)(N^N)][PF6] complexes.
Cation in [Cu(BIPHEP)(N^N)][PF6]Powder aFrozen Glass (Me-THF, 77 K) b
λemmax/nm cPLQY/%τ1/2/μsλemmax/nm dτ1/2/μs
[Cu(BIPHEP)(bpy)]+566436157
[Cu(BIPHEP)(6-Mebpy)]+5687359545
[Cu(BIPHEP)(6-Etbpy)]+5833159053
[Cu(BIPHEP)(5,5′-Me2bpy)]+55814860049
aλexc = 365 nm; b λexc = 410 nm. c λemmax estimated error ± 2 nm. d λemmax estimated error ± 5 nm.

Share and Cite

MDPI and ACS Style

Keller, S.; Bantle, M.; Prescimone, A.; Constable, E.C.; Housecroft, C.E. Hinged and Wide: A New P^P Ligand for Emissive [Cu(P^P)(N^N)][PF6] Complexes. Molecules 2019, 24, 3934. https://doi.org/10.3390/molecules24213934

AMA Style

Keller S, Bantle M, Prescimone A, Constable EC, Housecroft CE. Hinged and Wide: A New P^P Ligand for Emissive [Cu(P^P)(N^N)][PF6] Complexes. Molecules. 2019; 24(21):3934. https://doi.org/10.3390/molecules24213934

Chicago/Turabian Style

Keller, Sarah, Matthias Bantle, Alessandro Prescimone, Edwin C. Constable, and Catherine E. Housecroft. 2019. "Hinged and Wide: A New P^P Ligand for Emissive [Cu(P^P)(N^N)][PF6] Complexes" Molecules 24, no. 21: 3934. https://doi.org/10.3390/molecules24213934

Article Metrics

Back to TopTop