Open Access Article
Zubair Ahmed*a,
Rian E. Adernea,
Jiang Kaib,
Jackson A. L. C. Resendec,
Helmut I. Padilla-Chavarríab and
Marco Cremona
*a
aDepartment of Physics, Pontifícia Universidade Católica do Rio de Janeiro, PUC-Rio, Rio de Janeiro, RJ 22453-970, Brazil. E-mail: cremona@fis.puc-rio.br; zubairchem011@gmail.com
bDepartment of Chemistry, Pontifícia Universidade Católica do Rio de Janeiro, PUC-Rio, Rio de Janeiro, RJ 22453-970, Brazil
cInstituto de Ciências Extas e da Terra, Centro Universitário do Araguaia, Universidade Federal do Mato Grosso, 78600-000, Barra do Garças, MT, Brazil
First published on 27th March 2017
An in situ reaction of two optoelectronically active organic ligands (anionic thenoyltrifluoroacetylacetone, tta−, and neutral triphenylphosphine oxide, tppo) with erbium(III) ion in the presence of a base yielded a new erbium complex, [Er(tta)3(tppo)]. The solid and solution structure of the complex was established by X-ray crystallography, NMR, ESI-MS, FTIR, TGA and Raman spectroscopy. They indicate that the Er(III) ion is coordinated to seven oxygen atoms of three tta ligands and one tppo ligand in a monocapped octahedral geometry. The Judd–Ofelt parameters of the complex were determined by a least squares fitting and dealt with its chemical structure. On UV excitation through ligand mediation, the complex shows the characteristic near-infrared emission of the corresponding Er(III) ion at 1534 nm. Furthermore, a near infra-red organic light emitting diode (NIR-OLED) was fabricated with structure: ITO/[N,N′-bis(naphthalen-2-yl)-N,N′-bis(phenyl)benzidine]/[Er(tta)3(tppo)]/[2,2′,2′′-(1,3,5-benzinetriyl)-tris(1-phenyl-1-H-benzimidazole)]/LiF/Al. This device, with maximum applied voltage of 23 V, shows a total quenching of visible emission and electroluminescence in the C-band region (1534 nm) which is suitable for third communication window applications in fiber optics. Finally, an organic diode was fabricated to determine charge carrier mobility of the complex using a steady-state method.
Among the sensitizers of Ln ions, β-diketones are the most versatile ligands which is owing to (i) their strong π–π* electronic transitions in the UV region, (ii) their strong coordination ability and (iii) the possibility of functionalization. Generally, their combination with Ln ion and O-/N-donor synergic/ancillary ligands leads to formation of fully saturated complexes leaving no space, in the vicinity of Ln ion, for high-energy O–H/C–H oscillators of solvent molecules.5,6 This, therefore, reinforces the thermal and chemical properties of their Ln complexes which are very critical for optoelectronic applications. The bidentate ancillary ligands in such complexes deliver symmetric coordination geometries (8- and 10-coordinate) which are relatively less supportive of radiative transitions in the emission process.8 It points toward the use of monodentate ligands that produce asymmetric Ln coordination geometries which are expected to give strong emission efficiency.
The O-donor triphenylphosphine oxide is a very versatile ligand because it possesses tunable excited energy states and produces strong Ln–ligand bonds. Recently, the Ln complexes of this ligand, particularly those of europium and terbium, have been exploited as regards visible light photoluminescence.5a,9 However, its Ln complexes for NIR photoluminescence (PL) and electroluminescence (EL) applications, particularly with erbium, are rarely reported. For instance, two monometallic (eight- and nine-coordinate) and one heterometallic (eight-coordinate) complexes of Er with tppo and β-diketones have been reported.10,11 The authors used perfluorinated β-diketones to avoid high energy C–H oscillators in the vicinity of the Er ion in order to get efficient emission. In contrast, we have carried out a reaction in which fluorinated β-diketone, tppo and ErCl3 in an appropriate ratio have been used to get an asymmetric seven-coordinate Er complex that is anticipated to be highly luminescent. The PL and EL of erbium are very interesting because it emits at 1.54 μm that corresponds to the amplification range of erbium-doped fiber amplifiers. The development of new erbium(III) complexes and optimization of their 1.54 μm NIR emission open up new possibilities in the fabrication of efficient photonic integrated circuits.12
In this article, a new erbium(III) complex, tris(thenoyltrifluoroacetylacetonate)mono(triphenylphosphine oxide), or [Er(tta)3(tppo)], was synthesized by an in situ method and characterized thoroughly. Single crystal X-ray diffraction and 2D-NMR spectroscopy indicate that the complex is a seven-coordinate structure arranged in a distorted monocapped octahedral geometry. It is a highly asymmetric structure that suggests efficient energy transfer from coordinated ligands to Er ion. The steady-state NIR PL spectra of the pure complex and its PMMA-doped sample were studied and are presented. The Judd–Ofelt parameters (Ωt) of the complex were calculated from spectroscopic measurements that give an insight into local structure and bonding around the Er ion. Furthermore, a NIR-OLED based on the present complex was fabricated with structure: ITO/[N,N′-bis(naphthalen-2-yl)-N,N′-bis(phenyl)benzidine] (30 nm)/[Er(tta)3(tppo)] (40 nm)/[2,2′,2′′-(1,3,5-benzinetriyl)-tris(1-phenyl-1-H-benzimidazole)] (30 nm)/LiF (0.1 nm)/Al (100 nm). Finally, an organic diode was fabricated to determine charge carrier mobility in this novel complex by using a Mott–Gurney model.
The elemental analyses were performed using Flash EA equipment (Thermo Electron, Model 1112). The infrared spectra were recorded with a PerkinElmer Spectrum 2 spectrophotometer in the range 4000–400 cm−1. The NMR spectra of the ligands and the complex in CDCl3 were recorded with a Bruker AVANCE II 400 NMR spectrometer, Department of Chemistry, PUC-Rio. Cyclic voltammetry (CV) was conducted with a COMPACTSTAT.e using the conventional three-electrode configuration consisting of a graphite working electrode, a platinum wire auxiliary electrode, and an Ag/AgCl reference electrode. The cyclic voltammograms were obtained in CH2Cl2 using tetrabutylammonium hexafluorophosphate (NBu4PF6) (0.1 M) as the supporting electrolyte at scan rate of 0.05 V s−1. Ferrocene/ferricenium (Fc/Fc+) was used as internal reference during the measurement. Raman spectroscopy was performed using a micro-Raman microscope (Horiba, model XploRA), at an excitation wavelength of 638 nm. The laser excitation beam was focused onto crystalline samples with an intensity of 25 W mm−2 by a 10× objective lens. The scan was performed using an acquisition time of 5 s with 20 accumulations with a range between 50 and 1900 cm−1 and a spectral resolution of 4 cm−1. Mass spectrometry measurement was obtained using a mass spectrometer (Thermo Scientific, model Orbitrap XL) with electrospray ionization (ESI-MS) in positive mode. The operating parameters of the ionization source were: sprayer voltage: 3 kV, capillary voltage: 42 V, capillary temperature: 275 °C, desolvation temperature: 350 °C, sheath gas flow (nitrogen): 9 (arbitrary unit). The electronic spectrum of the complex was recorded with a HP Hewlett Packard 8452 A diode array spectrophotometer, with the sample contained in a 1 cm3 stoppered quartz cell of 1 cm path length in the concentration range between 1 × 10−5 and 1 × 10−3 M. PL spectra were measured using a Model Quanta Master spectrometer with a 150 W xenon lamp as the excitation source and InGaAs photomultiplier tube as detector. The current–luminance–voltage properties were measured by using a Keithley source measurement unit (Keithley 2400) and Newport Powermeter Model 1936-C.
| Empirical formula | C42H27ErF9O7PS3 |
| Formula weight | 1109.04 |
| Temperature | 293(2) K |
| Wavelength | 0.71073 Å |
| Crystal system | Trigonal |
| Space group | P31c |
| Unit cell dimensions | a = 14.9084(4) Å, α = 90° |
| b = 14.9084(4) Å, β = 90° | |
| c = 13.0146(3) Å, γ = 120° | |
| Volume | 2505.09(15) Å3 |
| Z | 2 |
| Density (calculated) | 1.470 Mg m−3 |
| Absorption coefficient | 1.908 mm−1 |
| F(000) | 1094 |
| Crystal size | 0.37 × 0.18 × 0.13 mm3 |
| Theta range for data collection | 2.222 to 26.394° |
| Index ranges | −18 ≤ h ≤ 18, −18 ≤ k ≤ 18, −16 ≤ l ≤ 16 |
| Reflections collected | 25 604 |
| Independent reflections [R(int) = 0.0548] | 3437 [R(int) = 0.0548] |
| Completeness to theta = 25.242° | 99.9% |
| Refinement method | Full-matrix least-squares on F2 |
| Data/restraints/parameters | 3437/37/218 |
| Goodness-of-fit on F2 | 1.060 |
| Final R indices [I > 2sigma(I)] | R1 = 0.0301, wR2 = 0.0705 |
| R Indices (all data) | R1 = 0.0425, wR2 = 0.0752 |
| Absolute structure parameter | −0.005(7) |
| Largest diff. peak and hole/e Å−3 | 0.617 and −0.371 |
| Bond length/Å | Bond angle/° | ||
|---|---|---|---|
| a Symmetry transformation coordinates; (#1): −x + y − 1, −x − 1, z; (#2): − y − 1, x − y, z′′. | |||
| Er(1)–O1 | 2.231(7) | O(1)–Er(1)–O(1A)#1 | 80.88(11) |
| Er(1)–O(1A)#1 | 2.284(4) | O(1)–Er(1)–O(1A) | 80.88(11) |
| Er(1)–O(1A) | 2.284(4) | O(1A)#1–Er(1)–O(1A) | 117.54(6) |
| Er(1)–O(1A)#2 | 2.284(4) | O(1)–Er(1)–O(1A)#2 | 80.88(11) |
| Er(1)–O(2A)#2 | 2.284(4) | O(1A)#1–Er(1)–O(1A)#2 | 117.54(6) |
| Er(1)–O2A | 2.284(4) | O(1A)–Er(1)–O(1A)#2 | 117.54(6) |
| Er(1)–O(2A)#2 | 2.284(4) | O(1)–Er(1)–O(2A)#2 | 132.74(13) |
| O(1A)#1–Er(1)–O(2A)#2 | 146.36(17) | ||
| O(1A)–Er(1)–O(2A)#2 | 76.14(16) | ||
| O(1A)#2–Er(1)–O(2A)#2 | 74.32(17) | ||
| O(1)–Er(1)–O(2A) | 132.74(13) | ||
| O(1A)#1–Er(1)–O(2A) | 76.14(16) | ||
| O(1A)–Er(1)–O(2A) | 74.32(17) | ||
| O(1A)#2–Er(1)–O(2A) | 146.36(17) | ||
| O(2A)#2–Er(1)–O(2A) | 79.0(2) | ||
| O(1)–Er(1)–O(2A)#1 | 132.74(13) | ||
| O(1A)#1–Er(1)–O(2A)#1 | 74.32(17) | ||
| O(1A)–Er(1)–O(2A)#1 | 146.36(17) | ||
| O(1A)#2–Er(1)–O(2A)#1 | 76.14(16) | ||
| O(2A)#2–Er(1)–O(2A)#1 | 79.0(2) | ||
| O(2A)–Er(1)–O(2A)#1 | 79.0(2) | ||
All the different layers of the device were sequentially deposited in high-vacuum environment by thermal evaporation onto ITO substrates with a sheet resistance of 15 Ω square−1. The ITO substrate was initially cleaned by ultrasonification using a detergent solution followed by toluene degreasing and then cleaned again by ultrasonification with pure isopropyl alcohol. Finally, the substrate was dried using nitrogen gas. The base pressure was 6.6 × 10−5 Pa, whereas during the evaporation process the pressure was ∼10−4 Pa. The deposition rates for organic compounds were in the range of 0.3 Å s−1 to 2 Å s−1. The layer thickness was controlled in situ through a quartz crystal monitor and confirmed with a profilometer measurement. The fabricated device had an active area of around 2 mm2 and operated in forward bias voltage with ITO as the positive electrode and Al as the negative one.
:
3
:
1. The reaction mixture was stirred for 7 h at room temperature. During stirring a white precipitate was formed, and it was repeatedly filtered off. The filtrate was concentrated and left to slowly evaporate at room temperature. After three days, pink-coloured crystals appeared which were washed with ethanol and chloroform, and then recrystallized from an ethanol/hexane solution.
[Er(tta)3(tppo)], pink colour. Yield (82%). Elemental analysis: calcd for ErC42H27F9O7S3P1: C, 45.44; H, 2.54; N, 6.24%. Found: C, 45.82; H, 2.63; N, 6.32%. 1H NMR (400 MHz, CDCl3, 300 K; δ, ppm): 17.83 (br, 6H, tppo), 10.57 (br, 3H, tppo), 8.84 (d, 6H, tppo), 7.41 (d, 3H, tta), 6.11 (s, 3H, tta), 3.78 (br, 3H, tta) and −16.12 (s, 3H, tta, C–H). FTIR data (cm−1): 3062 (w), 1602 (vs), 1541 (s), 1509 (w), 1440 (s), 1413 (s), 1357 (s), 1310 (vs), 1249 (v), 1165 (s), 1064 (w), 1020 (w), 789 (w), 724 (s), 541 (s). Melting point 137 °C.
:
3
:
3
:
1 (Scheme 1). The reaction gave the complex in good yield (82%). The complex is resistant to air and moisture, and highly soluble in common organic solvents.
O vibrations, and the bands at 600 and 671 cm−1 can be assigned to P–C6H5 vibrations.20 These bands in the spectrum of the complex have low intensity and are slightly shifted to higher frequencies relative to free tppo. This perturbation in P
O bond vibrations indicates the complexation of tppo ligand to Er ion. Moreover, the multiple peaks between 1650 and 1500 cm−1 (O
C–C
C(str) + CF3(str)) could be attributed to the coordinated β-diketonate (tta).21
Ostr).9 While in the spectrum of the complex, this stretching mode appears at lower frequency with very much less intensity. The shift and relatively low intensity of the (P
Ostr) band suggest the interaction/coordination of tppo ligand to the Er ion. Moreover, the vibration bands appearing in the region of 700–400 cm−1 correspond to v(Er–Otta) and v(Er–Otppo) bands.22 This is also indicative of evidence about the existence of coordination bonds between erbium and tta, and erbium and tppo.
![]() | ||
| Fig. 4 (A) Molecular structure of the complex, (B) coordination geometry around Er and (C) molecular packing showing H⋯F interactions. Thermal ellipsoids are drawn at the 30% probability level. | ||
![]() | ||
| Fig. 5 1H NMR spectra of the complex (a), tppo (b), and Htta (c) and 19F NMR spectrum of the complex (d). Insets: high resolution. | ||
The NMR spectrum of the Er complex is very interesting because the Er(III) ion is paramagnetic and induces sizable downfield/upfield shifts of protons of coordinated diamagnetic ligands. In the COSY spectrum, the complex shows correlations between protons 2 and 3, and the signals 6 and 7. Besides this, no correlation was found for other signals. The protons 4 of tta, being closer to metal ion, show highest upfield shift and appear at −16.01 ppm (δ). In contrast, protons 5 of tppo, being in close proximity to metal ion, show the highest downfield shift and appear at 18.02 ppm (δ). This downfield/upfield shift of protons of the coordinated ligands in opposite directions substantiates that these shifts are dipolar in nature. On the other hand, the spectrum of the complex displays 7 signals which are assigned to 27 protons; 4 signals due to tta equivalent to 12 protons and 3 signals due to tppo equivalent to 15 protons, confirming the tta/tppo ratio of 3
:
1. It substantiates the coordination of three tta and one tppo to Er(III) ion making the complex a seven-coordinate structure. Moreover, the spectrum covers a wide range of chemical shifts (−17 to +18 ppm (δ)) which are in good agreement with those reported for other ternary Ln(III) β-diketone complexes.5b,5c,6a
The emission spectrum of the complex was recorded in chloroform at room temperature. Upon the direct excitation of the ligands at 380 nm, the spectrum shows a peak at 1534 nm that covers a large spectral range from 1440 to 1645 nm. It is attributed to a typical 4I13/2 → 4I15/2 transition of Er(III) ion (Fig. 7). On the other hand, the singlet/triplet states of tta9,28 and tppo29 ligands are 25
164/18
954 and 36
376/18
954 cm−1, respectively, with energy differences ΔE (S1−T1) of 6982 and 17
422 cm−1. The singlet/triplet states of the tta and tppo ligands were also calculated from emission spectra of the present Er complex recorded in the visible region at low and room temperatures (Fig. S3 in ESI†). The observed values (tta, 501 nm; tppo, 519 nm) are in good agreement with those reported in the literature.9,27,28 The ΔE (S1−T1) between these ligands suggests that the energy gap is more appropriate for the tta ligand since a ΔE (S1−T1) of 5000 cm−1 is generally required for efficient intersystem crossing relaxation.30 To understand the energy transfer processes in the complex, a mechanism is proposed (Fig. 8) which suggests that the tta ligand is first excited to its S1 level and then energy transfer occurs to its T1 level via intersystem crossing (ISC). At the same time, some of energy transfer takes place from S1 of tta to T1 of tppo since they lie very close to each other. Thereafter, the T1 states of tta and tppo transfer the energy to upper levels of Er(III) ion via resonant energy transfer (RET).31 The populated levels (4S3/2, 4F9/2, 4I9/2 and 4I11/2) of Er(III) ion relax the energy to lower level, 4I13/2, which upon efficient nonradiative decay gives rise to the sensitized emission at 1534 nm.
![]() | ||
| Fig. 7 NIR emission spectra of the complex in pure form and PMMA-doped sample, under excitation (λmax = 380 nm). Inset: photostability curve of the complex. | ||
![]() | ||
| Fig. 8 Proposed energy level diagram showing Er ion excited state and the singlet/triplet states of tta and tppo sensitizers, and the energy transfer processes. | ||
It is well established that polymers doped with Ln β-diketonate complexes possess enhanced electrical, optical and thermal properties since such complexes are well processable and compatible with the polymer matrixes.31 In this study, the present complex was doped into PMMA matrix and its emission spectrum was recorded and compared (Fig. 7). The spectrum of doped sample shows an enhancement in the emission intensity of 4I13/2 → 4I15/2 transition as compared to that observed for the pure complex. Furthermore, the full width at half-maximum (fwhm) of the 4I13/2 → 4I15/2 transition of doped sample is 67 nm which is higher than that observed for pure complex (52 nm). The larger fwhm for the doped sample is in agreement with other reported Er(III)-doped materials.32 It suggests that they can be used in a wide gain bandwidth for optical-amplification applications.
| EHOMO = − (1.4 ± 0.1) × (qVCV) − (4.6 ± 0.08) eV. |
The ELUMO was determined by subtracting the singlet energy gap (Eg) from the EHOMO level. The determined HOMO and LUMO levels for the complex are −2.66 and −5.92 eV.
The complex shows good film-forming properties such as good volatility and transparency, light weight and easy thermal evaporation. These characteristics prompted us to use the complex in OLED fabrication. Subsequently, a triple-layered device was fabricated with the structure: ITO/β-NPB (30 nm)/[Er(tta)3(tppo)] (40 nm)/TPBi (30 nm)/LiF (0.1 nm)/Al (100 nm) (Fig. 9). The device shows EL emission at 1534 nm, measured at 2, 2.5 and 3 mA driving currents corresponding to 4I13/2 → 4I15/2 transition of the Er(III) ion (Fig. 10).
![]() | ||
| Fig. 9 Energy level diagram of device with the structure: ITO/β-NPB (30 nm)/[Er(tta)3(tppo)] (40 nm)/TPBi (30 nm)/LiF (0.1 nm)/Al (100 nm). | ||
The EL emission spectrum of the device matches very well with the PL spectrum, except for a broadening at higher wavelength. This could be related to an increase in the temperature of the material, which in turn leads to the electronic population redistribution in the excited 4I13/2 multiplet. Moreover, the device did not show any ligand-associated EL emission in the visible region indicating an efficient charge transfer from the organic ligands to the Er(III) ion, i.e. exciton-harvesting processes by the ligands.27 The threshold voltage (Von) of the device was 14 V and it gave a maximum irradiance of 0.069 μW cm−2 with a current density (J) of 129 mA cm−2 at 23 V (Fig. 11). The device was resistive up to an applied voltage of 24 V. Beyond this voltage the device started to degrade. The present device shows higher EL efficiency as compared to those reported for [Er(tfac)3(bath)]-, [Er(tfac)3(5NO2phen)]- and [Er(tfac)3(bipy)]-based devices with Von values of 4, 11 and 6 V respectively.27 However, the Von value obtained for the present device is slightly higher than those for the above reported Er complex-based devices. This is owing to the use of LiF/Al electrode instead of Ca electrode. Moreover, Von of the present complex-based is lower than that reported for an ErQ-based device35 (tfac = 1,1,1-trifluoro-2,4-pentanedione, bath = bathophenanthroline, 5NO2phen = 5-nitro-1,10-phenanthroline, bipy = 2,2′-bipyridine and Q = 8-hydroxyquinoline).
![]() | ||
| Fig. 11 Irradiance–current density curve of the complex. Inset: current density–voltage (J–V) curve. | ||
![]() | (1) |
This equation is analytically derived from the Poisson and continuity eqn (2) and (3). This derivation assumes that the mobility μ is constant within the device and the device boundaries at x = 0 and x = L, E = 0 and
, respectively.
![]() | (2) |
| J = p(x)eμE(x) | (3) |
The μ0 value of the complex was determined by using a single layer device which is fabricated with the following structure: ITO/[Er complex]/Al. The J–V curve of the complex gave μ0 of 6.44 × 10−9 cm2 V−1 s−1 (<5 V). It is comparable to that reported for the closely related complex [Er(tfnb)3bipy] (8.8 × 10−9 cm2 V−1 s−1) (tfnb = 4,4,4-trifluoro-1-(2-naphthyl)-1,3-butanedione; bipy = 2,2′-bipyridine).38 This is a sizable carrier mobility that could be due to the presence of electron-transporting tppo ligand. The μ0 value of the present complex represents total carrier mobility, which is due to the movement of both electrons as well as holes. The hole injection barrier between ITO (workfunction, ω = 4.7 eV) and the complex (ω = 5.98 eV) is very small, i.e. 1.28 eV. Similarly, the electron injection barrier between Al (ω = 4.2 eV) and the complex is 1.78 eV. This suggests an equal probability of hole and election injection into the device.
![]() | (4) |
The spectrum of [Er(tta)3(tppo)] shows eight transitions as given in Table 3. Among them, the most intense transitions are 4G11/2 (Er-VIII) and 2H11/2 (Er-IV) and are classified as hypersensitive transitions.40 Usually, the Ω2 value is closely related to the hypersensitive transitions, i.e. the larger the Pos of the hypersensitive transition is, the greater the Ω2 value. The Pos values of these hypersensitive transitions are 17- and 8-fold higher than that of Er(III) aqua-ion. Moreover, Pos of these transitions of the present complex in chloroform are higher than those reported for [Er(acac)3(H2O)] (17.61)7 and [Er(acac)3phen] (22.61)7 in methanol. The higher Pos suggests a highly asymmetric coordination environment around the Er ion in the present complex as compared to the above mentioned Er complexes.
| S′L′J′ | Transitions Er3+ (←4I15/2) | Spectral ranges (cm−1) | Er3+ aqua-ion (P × 106) | Oscillator strength (P × 106) | Ωt=2,4,6 (× 10−20 cm2) |
|---|---|---|---|---|---|
| a Present complex.b [Er(acac)3(H2O)]7.c [Er(acac)3phen]7. | |||||
| 4I11/2 | Er-I | 9809–10 657 |
0.19 | 0.91a, (—)b, [0.17]c | Ω2 = 34.31a, Ω2 = 25.71b, Ω2 = 32.28c |
| 4F9/2 | Er-II | 14 657–15 707 |
1.94 | 1.93a, (1.17)b, [1.07]c | |
| 4S3/2 | Er-III | 17 958–18 515 |
0.41 | 0.61a, (0.25)b, [0.35]c | Ω4 = 1.45a, Ω4 = 0.95b, Ω4 = 0.91c |
| 2H11/2 | Er-IV | 18 603–19 711 |
2.91 | 24.12a, (17.61)b, [22.67]c | |
| 4F7/2 | Er-V | 19 923–21 080 |
2.22 | 2.10a, (1.07)b, [1.54]c | |
| 4F3/2, F5/2 | Er-VI | 21 765–24 941 |
1.10 | 0.50a, (0.43)b, [0.36]c | Ω6 = 1.85a, Ω6 = 0.75b, Ω6 = 0.81c |
| 2H9/2 | Er–VII | 24 100–24 941 |
0.51 | 0.81a, (0.29)b, [0.45]c | |
| 4G11/2 | Er-VIII | 25 610–26 802 |
5.91 | 102.36a, (63.12)b, [74.98]c | |
| Ω2α(2t + 1)∑|Asp|2(2s + 1)−1 | (5) |
The determined Ω2 value for the present complex, [Er(tta)3(tppo)], in chloroform is 34.31 which is higher than those reported for [Er(acac)3·H2O] (25.71) and [Er(acac)3phen] (32.28).7 The difference between average bond lengths of Er–O(tta) and Er–O(tppo) in the present complex is 0.053 Å, which is lower than those reported for Er–O(acac) and Er–N(phen) in [Er(acac)3phen] (0.284 Å), and Er–O(acac) and Er–O(H2O) in [Er(acac)3·H2O] (0.1 Å). This result suggests that the present complex must have high symmetry since a smaller difference in bond length leads to higher symmetry of the field charge density. However, the present complex is a seven-coordinate complex and possesses a monocapped octahedral geometry which is more asymmetric as compared to eight-coordinate [Er(acac)3phen] having symmetrical square antiprismatic geometry. On the other hand, [Er(acac)3·H2O] is also a seven-coordinate complex with a monocapped trigonal prism geometry. Nevertheless, the present complex seems more asymmetric since it possesses a tta ligand having –CF3 group (electron withdrawing) on one terminal and thiophene ring (electron donating) on another terminal, i.e. asymmetric substitution. This changes the electron density at keto–enol oxygen atoms that leads to different electron populations on the carbonyl oxygen atoms in the ligand. Moreover, the oxygen atom of tppo has different electron density from keto–enol oxygen atoms since it is attached to electronegative phosphorus atom with a P
O bond. This suggests that there is a different field charge distribution around the Er(III) ion in the present complex. Consequently, it can be inferred that Asp of [Er(tta)3(tppo)] should be higher than those of [Er(acac)3·H2O] and [Er(acac)3phen], and therefore the Ω2 value of [Er(tta)3(tppo)] is larger.
The determined Ω6 value of the present complex [Er(tta)3(tppo)] (1.85) is higher than those reported for [Er(acac)3·H2O] (0.75) and [Er(acac)3phen] (0.81). In such systems, the Er(III) ion is coordinated to the ligands by σ-bonds which are formed between the filled 2p orbitals of the ligands and the empty 6s orbitals of Er(III) ion. The overlap of these orbitals leads to σ-electron donation from the ligands to Er(III) ion. It results in an increase of 6s electron density or decrease of 5d electron density and, therefore, the Ω6 value decreases. In the present complex, the Er–O(tppo) bond is less covalent as compared Er–N(phen) bonds in [Er(acac)3phen], since oxygen atom is more electronegative than nitrogen atom. While in comparison to Er–O(water) in [Er(acac)3H2O], the Er–O(tppo) bond is also assumed as less covalent because tppo has strong electron-attracting atoms/groups like phosphorus and phenyl rings on oxygen as compared to oxygen of water having only two hydrogen atoms. This result implies that σ-electron donation is less from tppo ligand to 6s orbital of Er ion in [Er(tta)3(tppo)] and, therefore, the Ω6 value is higher.
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 1487439. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra27473k |
| This journal is © The Royal Society of Chemistry 2017 |