Mithun Dasa,
Biswa Nath Ghoshb,
Antonio Bauzác,
Kari Rissanenb,
Antonio Frontera*c and
Shouvik Chattopadhyay*a
aDepartment of Chemistry, Inorganic Section, Jadavpur University, Kolkata 700032, India. E-mail: shouvik.chem@gmail.com; Tel: +91-33-24572941
bDepartment of Chemistry, Nanoscience Center, University of Jyväskylä, P. O. Box 35, 40014 Jyväskylä, Finland
cDepartament de Química, Universitat de les Illes Balears, Crta. de Valldemossa km 7.5, 07122 Palma de Mallorca, Baleares, Spain. E-mail: toni.frontera@uib.es
First published on 7th August 2015
Two mononuclear cobalt(III) Schiff base complexes with azide [Co(L)(N3)(L′)] (1) and [Co(L)(N3)(L′′)] (2) {where HL = 1-((2-(diethylamino)ethylimino)methyl)naphthalene-2-ol, HL′ = 2-hydroxy-1-naphthaldehyde and HL′′ = acetylacetone} have been synthesized and characterized by elemental analysis, IR and UV-Vis spectroscopy and single crystal X-ray diffraction studies. Both complexes show mononuclear structures with azide as terminal coligand. Structural features have been examined in detail that reveal the formation of interesting supramolecular networks generated through non-covalent forces including hydrogen bonding, C–H⋯H–C and C–H/π interactions. These interactions have been studied energetically by means of theoretical DFT calculations. We have also analyzed the unexpected O⋯O interactions observed in one complex between the oxygen atoms of the coordinated aldehyde groups using several computational tools, including Bader's “atoms-in-molecules” (AIM) and natural bond orbital (NBO) analyses.
On the other hand, the formation of supramolecular assemblies is an appealing research topic now-a-days. The most commonly used approach is to employ hydrogen bonds for engineering the structures of such complexes.8 The crystalline architecture of the complexes could also be managed by several other well established non-covalent interactions, such as, π-stacking, cation–π, and C–H/π forces, which have also attracted considerable interest due to their relative strength, directionality, and ability to act synergically, thereby providing an organizing force for the association of molecules into distinct supramolecular assemblies.9 Moreover, other less recognized forces, such as, σ/π-hole, lone-pair/π and anion/π interactions, have also been utilized in sensors, probes, photonic devices, catalysis etc.10 In this context, it is not irrelevant to state that intra-molecular non-bonded interaction between sulfur and sulfur (or oxygen or nitrogen) atoms has also been observed in a large number of organosulfur complexes controlling the conformation of small and large molecules.11 In these molecules, the non-bonded S⋯S, S⋯O or S⋯N distances are significantly shorter than the sum of the corresponding van der Waals radii (3.62, 3.32 or 3.35 Å) in the crystalline structure.12
In the present work, we have synthesized two new cobalt(III) azide complexes with a tridentate N2O donor Schiff base and 2-hydroxynaphthyl-1-carboxaldehyde or acetylacetone. The structures have been confirmed by single crystal X-ray diffraction analysis. Structural features have been examined in detail that reveal the formation of interesting supramolecular networks generated through non-covalent forces including hydrogen bonding, C–H⋯H–C and C–H/π interactions.13 These interactions have been studied energetically by means of theoretical DFT calculations.14 Most interesting observation is the existence of O⋯O interactions in the supramolecular assembly of one complex. The O⋯O interactions between the oxygen atoms of the coordinated aldehyde groups have also been analyzed using several computational tools, including Bader's “atoms-in-molecules” (AIM)15 and natural bond orbital (NBO)16 analyses.
Elemental analysis was performed using a PerkinElmer 240C elemental analyzer. IR spectra in KBr (4500–500 cm−1) were recorded using a PerkinElmer Spectrum Two FT-IR spectrophotometer. 1H NMR spectrum was recorded on Bruker DRX-300 NMR Spectrometer at 300 MHz using DMSO-d6 as solvent. Electronic spectra in DMSO (800–300 nm) were recorded in PerkinElmer LAMBDA 35 UV/Vis spectrophotometer. Fluorescence spectra were obtained on SHIMADZU RF-5301PC Spectrofluorophotometer at room temperature. Lifetime measurements were recorded using Hamamatsu MCP photomultiplier (R3809) and were analyzed by using IBHDAS6 software. Intensity decay profiles were fitted to the sum of exponentials series
Yield: 340 mg (63%). Anal. calc. for C28H28CoN5O3 (541.48): C, 62.11; H, 5.21; N, 12.93%. Found: C, 62.3; H, 5.3; N, 12.7%. FT-IR (KBr, cm−1): 1618, 1602 (CO), 1583 (C
N), 2023 (N3). UV-Vis, λmax (nm) [εmax (L mol−1 cm−1)](acetonitrile): 219 (90
235), 268 (62
163) 319 (24
645), 404 (6446), 422 (6076). 1H NMR (DMSO-d6) (ppm) δ: 9.77 (s, 1H, –CH
O), 8.83 (s, 1H, –CH
N), 8.71 (d, J = 4.18 Hz, 1H, Ar–H), 8.37 (d, J = 4.11 Hz, 1H, Ar–H), 7.60 (m, 3H, Ar–H), 7.45 (m, 3H, Ar–H), 7.28 (dd, J = 7.34, 7.22 Hz, 1H, Ar–H), 7.15 (dd, J = 7.61, 7.53 Hz, 1H, Ar–H), 6.87 (dd, J = 9.07 Hz, 9.09 Hz, 1H, Ar–H), 6.68 (d, J = 9.05 Hz, 1H, Ar–H), 4.38 (m, 2H, –CH2CH3), 3.15 (m, 1H, –CH2CH2), 2.79 (m, 2H, –CH2CH3), 2.06 (m, 2H, –CH2CH2), 1.21 (m, 1H, –CH2CH2), 1.55 (t, 6.19 Hz, 3H, –CH2CH3), 0.76 (t, 6.14 Hz, 3H, –CH2CH3).
Yield: 290 mg (62%). Anal. calc. for C22H28CoN5O3 (469.42): C, 59.41; H, 5.58; N, 13.86%. Found: C, 59.6; H, 5.4; N, 13.7%. FT-IR (KBr, cm−1): 1621 (CO), 1577 (C
N), 2023 (N3). UV-Vis, λmax (nm) [εmax (L mol−1 cm−1)](acetonitrile): 224 (58
391), 266 (57
462), 408 (5601), 424 (5455). 1H NMR (DMSO-d6) (ppm) δ: 8.72 (s, 1H, –CH
N), 8.14 (d, J = 4.26 Hz, 1H, Ar–H), 7.65 (d, J = 3.97 Hz, 1H, Ar–H), 7.55 (d, J = 4.55 Hz, 1H, Ar–H), 7.44 (dd, J = 7.14, 7.56 Hz, 1H, Ar–H), 7.16 (dd, J = 7.65, 7.46 Hz, 1H, Ar–H), 7.08 (d, J = 4.57 Hz, 1H, Ar–H), 5.62 (s, 1H, –COCH–), 4.29 (m, 2H, –CH2CH3), 3.15 (m, 1H, –CH2CH2), 2.91 (m, 1H, –CH2CH2), 2.68 (m, 2H, –CH2CH3), 2.37 (m, 1H, –CH2CH2), 2.27 (s, 3H, –COCH3), 1.85 (m, 1H, –CH2CH2), 1.63 (s, 3H, –COCH3), 1.06 (t, 6.77 Hz, 3H, –CH2CH3), 0.81 (t, 6.99 Hz, 3H, –CH2CH3).
Complexes | 1 | 2 |
Formula | C28H28CoN5O3 | C22H28CoN5O3 |
Formula weight | 541.48 | 469.42 |
Crystal size (mm) | 0.12 × 0.18 × 0.24 | 0.14 × 0.21 × 0.32 |
Temperature (K) | 123 | 170 |
Crystal system | Tetragonal | Orthorhombic |
Space group | P43212 | P212121 |
a (Å) | 10.2252(2) | 9.55072(15) |
b (Å) | 10.2252(2) | 12.7917(2) |
c (Å) | 50.0545(10) | 18.0663(3) |
Z | 8 | 4 |
dcalc (g cm−3) | 1.375 | 1.413 |
μ (mm−1) | 5.459 | 0.811 |
F(000) | 2256 | 984 |
Total reflections | 9712 | 28![]() |
Unique reflections | 4839 | 5055 |
Observed data [I > 2σ(I)] | 4405 | 4429 |
R(int) | 0.051 | 0.056 |
R1, wR2 (all data) | 0.0442, 0.0940 | 0.0472, 0.0845 |
R1, wR2 [I > 2σ(I)] | 0.0403, 0.0927 | 0.0364, 0.0782 |
Largest diff. in peak and hole (e Å−3) | 0.33, −0.21 | 0.36, −0.25 |
We have also optimized the position of the hydrogen atoms of the crystal structures in order to compare the theoretically calculated (BP86-D3/def2-TZVP) positions to those obtained using isotopic thermal parameters (riding model). The positions of the hydrogen atoms are similar and consequently we have used the crystallographic ones. The geometries of the theoretically calculated complexes are included in the ESI.†
Complex | 1 | 2 |
---|---|---|
Co(1)–O(1) | 1.894(2) | 1.874(2) |
Co(1)–O(2) | 1.909(2) | 1.922(2) |
Co(1)–O(3) | 1.909(2) | 1.915(2) |
Co(1)–N(1) | 1.883(2) | 1.887(2) |
Co(1)–N(2) | 2.068(2) | 2.066(2) |
Co(1)–N(3) | 1.932(2) | 1.949(2) |
Complex | 1 | 2 |
---|---|---|
O(1)–Co(1)–O(2) | 89.75(9) | 87.60(7) |
O(1)–Co(1)–O(3) | 87.33(8) | 86.03(7) |
O(1)–Co(1)–N(1) | 93.56(8) | 93.49(8) |
O(1)–Co(1)–N(2) | 176.66(8) | 179.02(8) |
O(1)–Co(1)–N(3) | 88.74(1) | 90.68(8) |
O(2)–Co(1)–O(3) | 92.32(8) | 93.87(7) |
O(2)–Co(1)–N(1) | 87.52(8) | 88.61(8) |
O(2)–Co(1)–N(2) | 93.58(8) | 93.19(8) |
O(2)–Co(1)–N(3) | 178.11(9) | 178.28(8) |
O(3)–Co(1)–N(1) | 179.10(9) | 177.46(8) |
O(3)–Co(1)–N(2) | 92.85(9) | 94.51(8) |
O(3)–Co(1)–N(3) | 86.49(9) | 85.85(8) |
N(1)–Co(1)–N(2) | 86.28(9) | 85.94(8) |
N(1)–Co(1)–N(3) | 93.69(9) | 91.66(9) |
N(2)–Co(1)–N(3) | 87.95(1) | 88.53(8) |
Both the aliphatic and aromatic parts of the tridentate ligand in complex 1 establish a variety of C–H/π interactions in the solid state forming a supramolecular 1D chain (Fig. 2A). Interestingly, the hydrogen atoms of the diethylamino group interact to one face of the π-system of the naphthalene moiety and concurrently the aromatic hydrogen atoms of the same ligand interact to the opposite face facilitating the formation of the 1D chain. We have studied this interaction using a dimeric model (Fig. 2B) to keep the size of the system computationally approachable. We have computed the interaction energy of this dimer which is ΔE1 = −17.5 kcal mol−1, which is large because of the enhanced acidity of the interacting hydrogen atoms due to the coordination of the ligand to the cobalt(III) metal center that strengthens the C–H/π interactions. As a matter of fact, if the calculations are performed without the cobalt(III) transition metal (protonating the ligand to keep the model neutral), the interaction energy is reduced to −7.5 kcal mol−1 for the aliphatic and to −5.2 kcal mol−1 for the aromatic C–H/π interaction. To corroborate this explanation, we have computed the atomic charges of the interacting hydrogen atoms, which are 0.17 and 0.19e for the aliphatic and aromatic hydrogen atoms, respectively, when the ligand is coordinated to cobalt(III). The charges are reduced to 0.15e (for both hydrogen atoms) when the ligand is not coordinated, thus supporting the acidity enhancement of the interacting hydrogen atoms of the ligand upon complexation.
![]() | ||
Fig. 2 (A) X-ray fragment of complex 1. (B) Theoretical model used to estimate the C–H/π interactions. Distances in Å. |
In addition to the C–H/π bonding network, non-covalent C–H⋯H–C interactions are also observed in the solid state of complex 1 (see Fig. 3) in the formation of a self-assembled dimer where the azide coligand also participates establishing hydrogen bonding interactions. Recently a combined computational and CSD study has demonstrated the importance of C–H⋯H–C interactions weak interactions in the solid state.31 In fact the binding energy of the dimer of dodecahedrane,31 that is stabilized only by C–H⋯H–C interactions, is approximately 3 kcal mol−1. To further analyze the importance on this interaction, we have computed the binding energies in a series of theoretical dimers based on the X-ray structure (see Fig. 3). In Fig. 3B the dimer retrieved from the X-ray geometry is shown where, in addition to the C–H⋯H–C interactions, a hydrogen bonding network with the participation of the azide and aldehyde moieties is present. The interaction energy of this dimer is very large and negative (ΔE2 = −20.9 kcal mol−1) due to the hydrogen bonding interactions. We have also computed a theoretical model where the azide coligands have been replaced by hydrides (Fig. 3C) and consequently the hydrogen bonding interactions involving azide are not present. As a result the interaction energy is significantly reduced to ΔE3 = −12.7 kcal mol−1 that corresponds to the hydrogen bonding interactions established between the symmetrically related aldehyde groups [coordinated to cobalt(III)] and the C–H⋯H–C interactions. Moreover, we have computed another theoretical model where the azide coligands and the ethyl substituents of the amino group have been replaced by hydrides and hydrogen atoms, respectively (Fig. 3D) and consequently only the hydrogen bonding interaction involving coordinated aldehyde groups are evaluated. As a result the interaction energy is further reduced to ΔE4 = −6.9 kcal mol−1 that corresponds to both complementary CO⋯HC hydrogen bonding interactions. The contribution of the C–H⋯H–C interaction can be evaluated as ΔEC–H⋯H–C = ΔE4 − ΔE3 = −5.8 kcal mol−1. The large binding energy for this interaction can be rationalized by means of the large number of individual and additive C–H⋯H–C contacts. Similar results have been observed in long open chain alkanes, where the number of side-on contacts between neighbouring molecules increases resulting in large dimer dissociation energies, as high as 4.5 kcal mol−1 for n-hexane.31b
![]() | ||
Fig. 3 Theoretical model used to estimate the C–H⋯H–C and hydrogen bonding interactions. Distances in Å. |
It can be observed in Fig. 3 that the O⋯O distance between the oxygen atoms of the coordinated aldehyde groups is shorter than the sum of van der Waals radii (∑RvdW = 3.04 Å). We have investigated the physical nature of this counterintuitive “like–like” interaction between the oxygen atoms (red dashed line in Fig. 3). We have first computed the non-covalent interaction (NCI) plot of the dimer of 1 using the crystallographic coordinates to analyze the intermolecular non-covalent interactions. The NCI plot is a visualization index based on the electron density and its derivatives, and enables identification and visualization of non-covalent interactions efficiently. The isosurfaces correspond to both favorable and unfavorable interactions, as differentiated by the sign of the second density Hessian eigenvalue and defined by the isosurface color. NCI analysis allows an assessment of host–guest complementarity and the extent to which weak interactions stabilize a complex. The information provided by NCI plots is essentially qualitative, i.e. which molecular regions interact. The color scheme is a red-yellow-green-blue scale with red for ρcut+ (repulsive) and blue for ρcut− (attractive). Yellow and green surfaces correspond to weak repulsive and weak attractive interactions, respectively.32 In Fig. 4A we show the representation of the NCI plot computed for the dimer. It can be observed a region between the ethyl groups that confirm the existence of favorable C–H⋯H–C interactions. In addition, there is long isosurface that corresponds to the hydrogen bonding network in the region of the azide ligand. Finally there is also a green surface in the region where the aldehyde groups interact. In order to further analyze the aldehyde–aldehyde interaction we have optimized a model dimer using formaldehyde and the interaction energy is ΔE5 = −1.6 kcal mol−1 that is modest (Fig. 4B). Since this value is small, we have also computed the interaction energy at a higher level of theory, i.e. CCSD(T)/def2-TZVP. As a result, the interaction energy is similar (−1.3 kcal mol−1), giving reliability to the DFT-D3 method. Moreover, we have also used the symmetry adapted perturbation theory (SAPT)33a to provide quantitative information on the magnitude of the electrostatic, induction and dispersion interactions. For the formaldehyde dimer complex the DFT-SAPT method gives a total interaction energy of Et(SAPT) = −1.0 kcal mol−1, that is in reasonable agreement with the CCSD(T) interaction energy. The partitioning of the interaction energy into the individual contributions is as follows: electrostatic term is Eele = −2.4 kcal mol−1, exchange term is Eexc = +3.1 kcal mol−1, induction term is Eind = −0.2 and dispersion term is Edisp = −1.5 kcal mol−1. Therefore the aldehyde–aldehyde interaction is dominated by the electrostatic and dispersion terms. Strikingly the AIM analysis shows that the interaction is characterized by a bond critical point (red sphere) and a bond path that connects both oxygen atoms confirming the existence of chalcogen–chalcogen interaction. The distribution of critical points does not show a bond critical point connecting the hydrogen to the oxygen atoms. It should be mentioned that the optimized O⋯O distance (2.98 Å) is very similar to the distance observed in the X-ray structure (2.94 Å). We have also analyzed how the interaction energy is affected by the O⋯O distance (Table 4). The interaction energy is reduced in more than 50% upon increasing 0.5 Å the O⋯O distance. At this point, in order to investigate the O⋯O interaction from an orbital point of view we have performed Natural Bond Orbital (NBO)33b calculations in the formaldehyde dimer focusing our attention on the second order perturbation analysis that is very useful to study donor acceptor interactions. Interestingly, we have found that the lone pair (lp orbital) of one oxygen atom interacts with the C–O and C–H antibonding orbitals of the opposite formaldehyde anion and vice versa with a concomitant second order stabilization energy of E(2) = 0.18 kcal mol−1 for each interaction. In addition the energetic differences between the lp and σ* orbitals are only 0.50 and 0.75 a.u. for the C–H and C–O bonds, respectively. Therefore the orbital stabilization energy that can be attributed to the O⋯O is approximately 0.34 kcal mol−1 in the dimer. We have also computed the AIM analysis of the model dimer of complex 1 shown in Fig. 3C in order to investigate if the distribution of critical points is similar to the one found in the formaldehyde dimer. The distribution is shown in Fig. 5A. It can be observed that the distribution in this model of complex 1 also exhibits two bond critical points that characterize the hydrogen bonds in addition to the bond CP and bond path that characterize the O⋯O interaction. This results agrees with the larger binding energy obtained for the dimer of complex 1 compared to the formaldehyde dimer. The properties of the charge density at the bond CPs are summarized in Table 5. The values of the Laplacian are positive as is common in closed shell interactions. The kinetic energy and the total energy density are higher at the bond CP that connects the oxygen atoms than that at the bond CP that characterizes the hydrogen bonds. Finally, we also represent in Fig. 5B the distribution of critical points that characterizes the C–H⋯H–C interaction described above (see Fig. 3A). This interaction is characterized by the presence of three bond CPs and bond paths that inter-connect three hydrogen atoms of the ethyl groups.
O⋯O distance (Å) | ΔE (kcal mol−1) |
---|---|
2.98 | −1.6 |
3.06 | −1.5 |
3.21 | −1.0 |
3.46 | −0.7 |
Complex | ρ | ∇2ρ | H(r) | G(r) | V(r) | ε |
---|---|---|---|---|---|---|
CP1 | 0.0092 | 0.0387 | 0.0019 | 0.0078 | −0.0059 | 0.4139 |
CP2 | 0.0090 | 0.0420 | 0.0021 | 0.0084 | −0.0064 | 0.7379 |
CP3 | 0.0050 | 0.0206 | 0.0013 | 0.0039 | −0.0026 | 0.1496 |
CP4 | 0.0051 | 0.0206 | 0.0013 | 0.0039 | −0.0027 | 0.1395 |
CP5 | 0.0042 | 0.0155 | 0.0008 | 0.0031 | −0.0023 | 0.0594 |
CP6 | 0.0057 | 0.0154 | 0.0006 | 0.0032 | −0.0026 | 0.3015 |
CP7 | 0.0043 | 0.0128 | 0.0007 | 0.0025 | −0.0019 | 0.6784 |
For complex 2 we have focused our attention to the C–H/π interactions and the influence of the metal center of the strength of the interaction. The dimers retrieved from the X-ray structure are shown in Fig. 6. In one dimer the hydrogen atoms that participate in the interaction belong to the aliphatic part of the tridentate Schiff base ligand (Fig. 6A) and in the other dimer the hydrogen atoms belong to the acetylacetonate (4-oxopent-2-en-2-olate) coligand (Fig. 6B). The interaction energies of the dimers are ΔE6 = −11.1 kcal mol−1 for the Schiff base interaction and ΔE7 = −8.2 kcal mol−1 for the acetylacetonate coligand. We have used two theoretical models to evaluate the influence of the coordination to the transition metal upon the interaction energies. In both models we have eliminated the metal center and protonated the oxygen atom of the ligand that is coordinated to the cobalt(III) ion in the real system (Fig. 6C and D). As a result both interaction energies weaken to ΔE8 = −6.0 kcal mol−1 and ΔE9 = −6.1 kcal mol−1, confirming that the coordination of the ligand to the metal center increases the acidity of the interacting hydrogen atoms and enhances the C–H/π interactions. For the latter models we have also computed the interaction energy at a higher level of theory (RI-MP2) in order to validate the DFT method used in this study. The interaction energies are in reasonable agreement, giving reliability to the methodology used herein (Fig. 6).
![]() | ||
Fig. 6 Theoretical models based of the X-ray structure of complex 2 used to evaluate the C–H/π interactions. Distances are in Å. |
We have also confirmed the existence of the aforementioned C–H/π interactions using the AIM analysis of critical points and bond paths (Fig. 7). In the dimer where the interacting hydrogen atoms belong to the aliphatic part of the tridentate Schiff base ligand (Fig. 7A) the distribution of critical points show four bond CPs that connect several hydrogen atoms with the carbon atoms of the aromatic rings. The interaction is further characterized by the presence of several ring CPs that are generated as a consequence of the formation of several supramolecular rings. In the other dimer the hydrogen atoms belong to the acetylacetonate coligand (Fig. 7B) and the interaction is characterized by the presence of three bond CPs that connect three hydrogen atoms with three carbon atoms of the aromatic part of the ligand. The interaction is further characterized by the presence of two ring CPs. The properties of the density at two representative bond CPs are summarized in Table 5.
![]() | ||
Fig. 7 Distribution of bond and ring critical points (red and yellow spheres, respectively) and bond paths in two dimers of complex 2. |
Kinetic studies were performed to understand the extent of the catalytic efficiency. For this purpose, 1.0 × 10−5 M solutions of the complex was treated with substrate, maintaining pseudo first order conditions. For a particular complex substrate mixture, time scans at the maximum band of 2-aminophenoxazine-3-one were carried out for a period of 30 min, and the initial rate was determined by linear regression from the slope of the absorbance versus time, and each experiment was performed thrice and average values were noted. As shown in Fig. 9, the initial rates of the reaction versus concentration of the substrate plots show rate saturation kinetics. This observation indicates that an intermediate complex substrate adduct formed at a preequilibrium stage and that the irreversible substrate oxidation is the rate determining step of the catalytic cycle. This type of saturation rate dependency on the concentration of the substrate can be treated with the Michaelis–Menten model, which upon linearization gives a double reciprocal Lineweaver–Burk plot to analyze values of the parameters Vmax, KM, and Kcat. The observed and simulated initial rates versus substrate concentration plot and the Lineweaver–Burk plot for 2 are shown in Fig. 9 and 10, respectively. Analysis of the experimental data yielded Michaelis binding constant (KM) value of 4.39 × 10−3 and Vmax value of 8.09 × 10−5 M−1. The turnover number (Kcat) value is obtained by dividing the Vmax by the concentration of the complex used, and is found to be 8.32 s−1.
![]() | ||
Fig. 9 Linear Lineweaver–Burk plots for the oxidation of o-aminophenol catalyzed by complex 2. Symbols and solid lines represent experimental and simulated profiles, respectively. |
1H NMR spectra data for complexes 1 and 2 have been summarized in experimental section and the spectra have been depicted in Fig. S1† and 11 respectively. Imine protons (–CHN) of 1 and 2 appear as a singlet at 8.83 ppm and 8.72 ppm respectively whereas, aldehyde protons (–CH
O) of 1 appears at 9.77 ppm. A sharp singlet at 5.62 ppm and two sharp singlet at 2.27 ppm and 1.63 ppm appears in the spectrum of 2 suggesting the presence of acetylacetonate moiety (Fig. 11).
Complex 1 shows emission at 344 nm in the UV region, whereas complex 2 exhibits emission in the visible region at 431 nm upon irradiation with UV-light (300 nm) in acetonitrile at room temperature (Fig. 12). The exited state mean lifetimes of complexes 1 and 2 are 3.17 ns and 6.7 ns respectively (Fig. 13). This emission band can be attributed to intra ligand fluorescent 1(π → π*) emission of the coordinated ligand.37
Relative fluorescence quantum yields for two complexes were measured in acetonitrile using quinine sulfate (in 0.5 (M) H2SO4, Φ = 0.54) as the quantum yield standard.38 The fluorescence quantum yields of complexes 1 and 2 are 0.0046, 0.0041 respectively. If Φx and Φs are the quantum yields of a given fluorophore species ‘x’ and the standard ‘s’, respectively, then
Φx = Φs(Fx/Fs)(Ax/As) |
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1047381 and 1047382. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra13960k |
This journal is © The Royal Society of Chemistry 2015 |