M. Judith Percino*a,
Margarita Ceróna,
Perumal Venkatesana,
Enrique Pérez-Gutiérreza,
Pilar Santosa,
Paulina Ceballosa,
Armando E. Castilloa,
Paola Gordillo-Guerraa,
Karnambaram Anandhana,
Oracio Barbosa-Garcíab,
Wilson Bernalb and
Subbiah Thamotharanc
aUnidad de Polímeros y Electrónica Orgánica, ICUAP, Benemérita Universidad Autónoma de Puebla, Val 3-Ecocampus Valsequillo, Independencia O2 Sur 50, San Pedro Zacachimalpa, Pue., Mexico 7296. E-mail: judith.percino@correo.buap.mx
bCentro de Investigaciones en Óptica A. P. 1-948, 37150, León, Guanajuato, Mexico
cBiomolecular Crystallography Laboratory, Department of Bioinformatics, School of Chemical and Biotechnology, SASTRA Deemed University, Thanjavur 613 401, India
First published on 12th September 2019
2-(4-((2-Hydroxyethyl)(methyl)amino)benzylidene)malononitrile (HEMABM) was synthesized from 4-[hydroxymethyl(methyl)amino]benzaldehyde and propanedinitrile to obtain a low molecular weight fluorescent material with an efficient solid-state emission and electroluminescence properties comparable to the well-known poly(2-methoxy-5(2′-ethyl)hexoxyphenylenevinylene) (MEH-PPV). The HEMABM was used to prepare an organic light-emitting diode by a solution process. Despite the title compound being a small molecule, it showed optical properties and notable capacity to form a film with smooth morphology (10.81 nm) closer to that of polymer MEH-PPV (10.63 nm). The preparation of the device was by spin coating, the electrical properties such as threshold voltage were about 1.0 V for both HEMABM and MEH-PPV, and the luminance 1300 cd m−2 for HEMABM and 2600 cd m−2 for MEH-PPV. This low molecular weight compound was characterized by SCXRD, IR, NMR, and EI. Besides a quantitative analysis of the intermolecular interactions by PIXEL, density functional theory (DFT) calculations are reported.
Soluble organic semiconductors for OLEDs have received important attention since they offer a potential for the production by solution processing techniques such as spin-coating, ink-jet printing or drop casting.10,11 Within this, the conjugated organic polymers have been generally considered as a suitable material for these techniques because of their good film forming capacity.12,13 In recent research, the numerous organic electroluminescent (EL) materials including polymers has focused on solution processable of small molecules which possess advantages of having well-defined structures and being uniformly reproducible.14,15 The parameters for the manufacture conditions such as the dye device, quantum yield, solvent or vacuum deposition, size of the films, etc. are very important to design a new compound with adequate optical properties and to understand structure–properties relationships.16,17 In this manner, synthesis of molecules with a substituent that interacted with the solvents, it's able to form films might be a valuable approach to solution processed OLEDs.18,19 The extended conjugated structures of well-defined small molecules involving alkyl and alkoxyl chains15,20,21 may provide better solubility and higher glass thermal stability while maintaining high purity and reliable molecular structure–property correlations.22
On the other hand, the molecular orientation of organic semiconductors play an important role on small-molecule light-emitting diodes (OLEDs), i.e. the films morphology. Molecular orientation at the microscopic level can be reached in vacuum-deposited amorphous films or in spin coating for a better device performance.23,24 Thin films based on organic small-molecule consist of single units, and each one has specific characteristics, such as geometric and electronic structures, as well as absorption and emission spectra. Therefore, the properties as solids came from of single molecule before fabricating the thin films or crystals. This means that the properties are shifted or changed depending on intermolecular interactions they form. Also, it is interesting that although most chemical phenomena take place in solution, the solvent is usually thought of as a spectator, acting merely as a medium to hold the reactants and allow them to encounter each other by diffusion.25 It well known cases as solvated electrons26 and the charge transfer to solvent transitions27 of simple anions, the solvent creates the electronic states that are of importance, primarily because the electrons of interest are not otherwise bound without the stabilizing presence of the solvent. Moreover, in electron transfer and related reactions, reorganization of the surrounding solvent molecules is the driving force to move the electron from the donor to the acceptor, and thus determines the reaction rate.28–30 Dielectric effects from collective solvent motions can lead to solvatochromic shifts in the electronic absorption spectra of solutes,31,32 and solvent molecules can provide viscous drag that changes the dynamics and thus the branching ratios of unimolecular isomerization reactions.33
Recently, the thin-film assemblies, most of which are amorphous in nature, are easily broken up by trying to deposit electron transporting layer (ETL), because small molecules are typically attached to each other only by weak intermolecular forces such as van der Waals, H-bonding and π stacking interactions. Consequently, the highest reported efficiency of solution-processed small molecule OLEDs still relies on a vacuum-evaporated films, which is not practical for low-cost mass production of scalable devices. The solution process of small molecules encounters major difficulties to fabricate multilayers architectures because of redissolution of underlying layers. This problem can be avoided using orthogonal solvents,34,35 or by cross-linkable the organic functional materials36,37 or designing a compound that showing a highly efficient solution-processed small-molecule for OLEDs.
Alternatively, cyano-substituted pyridine derivatives have been reported as electron-transporting (ET) properties and have been used in OLED devices showing acceptable efficiency.38,39 Herein, malononitrile based HEMABM derivative (Fig. 1) was developed by introducing two electron-withdrawing cyano-group substituents on the HEMAB. Emphasis is in comparing with the commercial polymer MEH-PPV, because it exhibits a long-conjugated system contrary to the title molecule, but the optical properties are very similar.
In this study, it was prepared a small molecular weight compound with orange emission and it can be deposited as thin film by solution process. Their optical and luminance properties are compared with MEH-PPV. The electroluminescent behavior of the emitting material was studied in devices showing a wide spectral to orange. We focused on the study on the effect of the morphology development of solution processed OLEDs and to evaluate the manufacturing related to physical properties. The solvent study with 2-(4-((2-hydroxyethyl)(methyl)amino)benzylidene)malononitrile (HEMABM) was carried out before fabrication of OLED, for which we have chosen different polarities solvents such as MeOH, isopropyl alcohol (IPA), ethyl acetate (EtOAc), THF, chlorobenzene (Ph-Cl), CHCl3, 1,4-dioxane and toluene. It was investigated the morphological and photophysical properties of the solution-processed films by atomic force microscopy (AFM), and the electrical and electro-optical properties of solution-processed, as well as their current–voltage–luminance characteristics are reported. Finally, to identify the origin of optical properties, electronic structure calculations were performed. Further, the intermolecular interactions existing in the crystal structures and their energies were quantified for various dimers by PIXEL and DFT approaches in order to correlate with the intermolecular interaction that play role in solution and in the films.
Solvents | Absorbance (a.u.) | ||||
---|---|---|---|---|---|
Solution | [M1] | [M2] | [M3] | [M4] | [M5] |
a [M1] = 5.5 × 10−5 M; [M2] = 2.75 × 10−5 M; [M3] = 1.38 × 10−5 M; [M4] = 0.69 × 10−5 M; [M5] = 0.34 × 10−5 M. | |||||
MeOH | 1.686 | 0.851 | 0.434 | 0.210 | — |
IPA | 2.672 | 1.339 | 0.694 | 0.364 | 0.181 |
EtOAc | 3.370 | 1.711 | 0.888 | 0.465 | 0.240 |
THF | 3.198 | 1.614 | 0.832 | 0.433 | 0.216 |
Ph-Cl | 3.332 | 1.693 | 0.854 | 0.427 | 0.219 |
CHCl3 | 3.344 | 1.652 | 0.899 | 0.584 | 0.336 |
1,4-Dioxane | 2.865 | 1.424 | 0.731 | 0.359 | 0.179 |
Toluene | 4.616 | 2.302 | 1.166 | 0.600 | 0.304 |
Fig. 2 The absorption spectra of HEMABM in different solvents with two different concentrations of solution (a) 5.50 × 10−5 M and solution (b) 0.69 × 10−5 M. |
Fig. 3 The selected absorption spectrum of HEMABM in two different concentrations of solution (a) 5.50 × 10−5 M and solution (b) 0.34 × 10−5 M in different solvents. |
λabs (DFT) | eV | f | Major transition (%) | λexp |
---|---|---|---|---|
Gas | ||||
410 | 3.02 | 0.6718 | H → L(93) | — |
327 | 3.79 | 0.0921 | H−1 → L(51); H−3 → L(35) | — |
292 | 4.25 | 0.0594 | H → L+2(48); H−3 → L(31); H−1 → L(15) | — |
Methanol | ||||
445 | 2.78 | 0.8412 | H → L(95) | 430 |
352 | 3.52 | 0.1343 | H−1 → L(77); H−3 → L(17) | |
295 | 4.19 | 0.0876 | H−3 → L(70); H−1 → L(14) | 269 |
Isopropyl alcohol | ||||
447 | 2.78 | 0.8555 | H → L(96) | 431 |
352 | 3.53 | 0.1314 | H−1 → L(77); H−3 → L(18) | |
296 | 4.19 | 0.0872 | H−3 → L(70); H−1 → L(14) | 268 |
Ethyl acetate | ||||
443 | 2.80 | 0.8499 | H → L(96) | 428 |
347 | 3.57 | 0.1257 | H−1 → L(74); H−3 → L(21) | |
295 | 4.20 | 0.0914 | H−3 → L(67); H−1 → L(17) | 269 |
THF | ||||
445 | 2.78 | 0.8600 | H → L(96) | 430 |
349 | 4.61 | 0.1110 | H−1 → L(75); H−3 → L(20) | |
295 | 4.20 | 0.0895 | H−3 → L(68); H−1 → L(18) | 270 |
Chlorobenzene | ||||
449 | 2.76 | 0.8882 | H → L(97) | 433 |
348 | 3.57 | 0.1219 | H−1 → L(75); H−3 → L(20) | |
296 | 4.20 | 0.0873 | H−3 → L(68); H−1 → L(16) | — |
CHCl3 | ||||
445 | 2.79 | 0.8686 | H → L(100) | 431 |
346 | 3.58 | 0.1224 | H−1 → L(74); H−3 → L(22) | |
295 | 4.20 | 0.0900 | H−3 → L(67); H−1 → L(17) | 270 |
1,4-Dioxane | ||||
439 | 2.83 | 0.8574 | H → L(96) | 424 |
339 | 3.66 | 0.1169 | H−1 → L(63); H−3 → L(27) | |
294 | 4.22 | 0.0928 | H−3 → L(62); H−1 → L(21) | 271 |
Toluene | ||||
442 | 2.81 | 0.8769 | H → L(97) | 427 |
340 | 3.64 | 0.1168 | H−1 → L(70); H−3 → L(26) | |
294 | 4.22 | 0.0911 | H−3 → L(63); H−1 → L(20) | — |
Fig. 4 The important frontier molecular orbitals of HEMABM calculated by using PBEPBE/6-311++G (d, p) level of theory in gas phase and in 1,4-dioxane. The orbitals plotted with isovalue of 0.02 Å−3. |
From Table 2, the computed λmax absorption found at 410 nm in the gas phase and in the range of 439–449 nm in different solvents which belongs to the H → L transition. The second absorption is found at 327 nm in gas and 339–349 nm in solvents which belongs to H−1 → L and H−3 → L transitions. Additionally, another weak absorption is found in the range of 290–300 nm in gas and solvents which belongs to H−3 → L and H−1 → L transitions. The earlier intense (H → L) transition is might be due to the intermolecular charge transfer transition and the latter two absorptions belong to π → π* and n → π* transitions, respectively.
From the experimental, we found two absorption bands, the λmax in the range of 430–427 nm and a smaller absorption band emerged below 300 nm as mentioned earlier. Experimental λmax is may be due to CT and π → π* transitions and latter smaller absorption belongs to n → π* transitions, respectively. Because the absorption intensity for λmax (in experimental) is decreased as decreasing the concentrations of solution (Fig. 2) as described earlier. Further, to understand CT transition between the molecules, we carried out the NBO analysis for most stabilized dimer (D1) which formed by the strong O–H⋯N intermolecular interaction. Crystallography information and the intermolecular interactions in crystal structure of HEMABM is depicted in the next section. The NBO analysis suggested that the charge transfer occur between the lone pair electron of N2 atom and the antibonding orbital of O1–H1(O) bond and it is stabilized with 6 kcal mol−1.
The energy gap (ΔE(L–H)) is nearly the same in most of the solvents (2.02–2.07 eV) except the 1,4-dioxane and toluene (2.11 eV), see Table 3. It is interesting to note that the HOMO–LUMO energy level of HEMABM is very closer to the HOMO–LUMO of MEH-PPV (HOMO = 5.3 eV, LUMO = 3.0 eV).42 But, the ΔE(L–H) value for MEH-PPV is slightly higher than the HEMABM. This difference is mainly arising from the energy levels of LUMO orbital in HEMABM (Table 3). We observed that the ΔE(L–H) value is increased with decreasing the polarity of the solvent except CHCl3.
Solvents | HOMO | LUMO | ΔE(L–H) | Solvation energya |
---|---|---|---|---|
a Energy difference between the optimized structure of the gas phase and in respective solvents. | ||||
Gas | −5.53 | −3.34 | 2.19 | — |
MeOH | −5.29 | −3.27 | 2.02 | 13.24 |
IPA | −5.30 | −3.27 | 2.03 | 12.86 |
CHCl3 | −5.35 | −3.28 | 2.07 | 10.19 |
EtOAc | −5.33 | −3.28 | 2.05 | 10.91 |
THF | −5.33 | −3.28 | 2.05 | 11.45 |
Ph-Cl | −5.33 | −3.28 | 2.05 | 10.77 |
1,4-Dioxane | −5.41 | −3.30 | 2.11 | 6.64 |
Toluene | −5.40 | −3.29 | 2.11 | 7.07 |
Briefly, the least ΔE(L–H) value obtained in methanol (2.02 eV) and highest value obtained in toluene and 1,4-dioxane (2.11 eV). The computed solvation energy for HEMABM in different solvents are listed in Table 3 and we found that the solvation energy is increased with increasing the solvent polarity. The very low solvation energy is observed in 1,4-dioxane with 6.64 kcal mol−1 and this may be a reason for the formation of excellent morphology in the 1,4-dioxane mixture.
Solvent | rt | 85 °C |
---|---|---|
a rt = room temperature; δs = partially soluble; s = completely soluble. | ||
MeOH | s | S |
IPA | δs | S |
EtOAc | s | S |
THF | s | S |
Ph-Cl | δs | S |
CHCl3 | s | S |
1,4-Dioxane | s | S |
Toluene | δs | S |
To investigate the distinctive nature of small molecule HEMABM films from solution-processed, solutions were prepared in MeOH, IPA, EtOAc, THF, Ph-Cl, CHCl3, 1,4-dioxane and toluene to analyze its quality and morphology by AFM, Fig. 5. The HEMABM solution preheated at 85 °C for 20 min (exception in case of the MeOH, CHCl3 and THF) was deposited on glass substrate, Fig. 6(a). The glass substrate was previously coated with PEDOT:PSS polymer layer with around 40 nm and dried at 120 °C for 20 min to better HEMABM adhesion. The HEMABM films were analyzed by AFM in order to know the quality and morphology. The AFM results suggested that the good homogeneity surface with all the solvents and total coverage area of HEMABM films excepted to the toluene solution (Fig. 6(b)).
In the case of toluene mixture, a star shaped HEMABM crystals were grown in the glass substrate after the drying a couple of hours. The film is not homogeneous, it has gaps, agglomerates and crystals. Only some portion of the film was regular and its morphology has an average roughness of 1.0 nm. The film images at 100 μm showed some homogeneous area with small holes and its corresponding image on the left side of the Fig. 7 which showed the needle-shaped crystals formed on the film (Fig. 7). The morphology and roughness of each film with all solvents Table 5 and Fig. 8 are compared at 10–20 μm scale. We can see from the whole films the formation of agglomerates with a length of 10–20 μm. However, the most homogeneous films were obtained using MeOH, 1,4-dioxane and Ph-Cl, but on a smaller scale in homogeneous areas the morphology is comparable.
Fig. 8 Representative AFM image of the morphology for films of compound HEMABM at (a) 10 μm scale and (b) 20 μm scale. |
The Fig. 9 shows the absorption and photoluminescence of the films for each solvent used. The λmax absorption maximum is between 420 nm and 434 nm for MeOH → Ph-Cl, it was not observed a solvent effect on absorption wavelength at 421 nm, which corresponding to the transition observed for HEMABM in solution, only the band shape is broad, as it was expected for solid state.30,42
Fig. 9 HEMABM absorption and PL emissions spectra of the films (glass substrate) prepared with all solvents. |
From film thickness the HEMABM concentration was calculated for best comparison of the absorption and photoluminescence, Table 6. However, the spectra did not show the absorption below 300 nm as observed in the solution spectra (see Fig. 9), so the solution were deposited on quartz substrate. The spectra are shown in Fig. 10, and it shows that the absorption in the range 215–300 nm appeared as the transitions described earlier in solution. Also, it is interesting note that with the solvent 1,4-dioxane the band intensity is higher than with MeOH, which is an indication to the molecules interaction.
Solvent | Thickness (nm) | [HEMABM] × 10−5 M | Imax | Abs. 155 nm | Abs.a |
---|---|---|---|---|---|
a Normalized values [intensity absorbance/thickness]; I = intensity (Abs. at λ = 421 nm). | |||||
MeOH | 155 | 4.37 | 1.33 | 1.330 | 1 |
IPA | 232 | 3.08 | 1.48 | 0.989 | 0.743 |
CHCl3 | 250 | 2.05 | 1.32 | 0.818 | 0.615 |
EtOAc | 242 | 6.48 | 1.88 | 1.204 | 0.905 |
THF | 235 | 3.18 | 0.69 | 0.455 | 0.342 |
Ph-Cl | 187 | 1.11 | 0.6 | 0.497 | 0.374 |
1,4-Dioxane | 165 | 1.15 | 0.77 | 0.723 | 0.544 |
Toluene | 200 | 0.897 | 0.4 | 0.310 | 0.233 |
Fig. 12 shows representative electrical parameters for devices with HEMABM compounds as well as MEH-PPV. The threshold voltage for devices based on MEH-PPV was slightly lower. These low threshold voltages can be compared with those reported by Ha43 and Hewidy44 for OLEDs with MEH-PPV as emitting layer. The luminance for devices based on HEMABM was of about 1300 cd m−2 meanwhile for MEH-PPV was about 2600 cd m−2. Also, it is observed that the electroluminescence wavelength (EL) of OLEDs corresponded to the PL wavelength (λem 611) and 586 nm for HEMABM and MEH-PPV respectively, Fig. 13.
Fig. 13 Photoluminescence and electroluminescence emission of the film and OLED based on MEH-PPV and HEMABM. |
Fig. 14 shows the graphics with the measurement of the optical band gap for the compound HEMABM. The band gap of the MEH-PPV from different reports,44,45 is in the ranges between 2.00 and 2.10 eV. The values were estimated using onset wavelength. The HEMABM absorption value of λmax and λem calculated using optoelectronic module of Schrödinger Material Suite.46 gave the values at 390 nm and 735 nm respectively (Table S4†). The band gap is 2.92 eV (HOMO (eV) = −6.1399 and LUMO (eV) = −3.2087).
Fig. 15 The ORTEP molecular structure of HEMABM, with displacement ellipsoids drawn at the 50% probability level. The photo shows the orange color of the crystal. |
Single crystal X-ray diffraction analysis revealed that HEMABM crystals exhibited a structure with disordered over either two or three orientations. The fragment C1 > C10 (including N1 and N2) is disordered over two orientations, and the occupancy factor of the major component of the disorder refines to 0.598(13). The fragment N3 > O1 (including C11, C12 and C13) is disordered over three orientations, and the occupancy factors of the three components refine to 0.531(3), 0.2429(19) and 0.227(3) (Fig. 15).
The disorder may be due to intermolecular interaction between the –OH group of N-methylethanol moiety's and one of the –CN group of malononitrile moiety in the other molecule. This interaction is a key interaction for the formation of molecular packing which forming a macrocyclic ring with R22(22) motif (Fig. 16) and the intermolecular interaction O–H⋯N could be mainly was observed in the crystal packing of HEMABM, Table 7. These interactions affect the bond length corresponding to CN group. N(1)C(1) and –C(2)N(2) bond are of 1.138(7) Å and 1.156(7) Å respectively. (For N(1′)C(1′) and N(2′)C(2′) bonds of 1.161(11) Å and 1.136(10) Å respectively) Table 8. The values for N(1)C(1) and N(2′)C(2′) are within the reported values for Car–CN,47 but for the N(2)C(2) and N(1′)C(1′) distances are larger due to hydrogen bonds distances, a, due to the atoms involved in chain motif formation.47 This bond length variation is might be due to the charger transfer between the interacting functional groups and similar bond length variation is also observed in the phenyl ring affected by the donor group methylaminoethanol and the acceptor group malononitrile (Table 8), contributing to the disorder that exhibit the molecular structure and the dimers observed in the package,48 Fig. 15.
Fig. 16 Crystal packing diagram of HEMABM showing the intermolecular interaction between –CN and O–H groups.51 |
D–H⋯A | d(D–H) | d(D⋯A) | <DHA | Symm. op. 2 |
---|---|---|---|---|
O(1)⋯N(2) | O(1)–H(1) | N2⋯H1 | O(1)–H(1)⋯N2 | −x, −y, 1 − z |
3.022 | 0.8400 | 2.182 | 173.94 | |
2.5–3.2 | 1.5–2.2 | 170–180 |
Bond | Length (Å) |
---|---|
C(3)–C(4) | 1.373 (6) |
C(4)–C(5) | 1.430 (7) |
C(5)–C(6) | 1.414 (7) |
C(5)–C(10) | 1.418 (7) |
C(6)–C(7) | 1.371 (7) |
C(7)–C(8) | 1.415 (4) |
C(8)–N(3) | 1.402 (6) |
C(8)–C(9) | 1.423 (6) |
C(9)–C(10) | 1.372 (7) |
The arrangement of the molecules in the crystal is not completely face-to-face manner, Fig. 16. The dimer showed a weak interaction between aromatic rings; with two centroid–centroid distances: one of 3.834 Å (shift distance of 1.872 Å) and 3.917 Å (shift distance of 1.905 Å), which were calculated using OLEX 2.49 This π stacked dimer is shown in Fig. 17, and these parameters for the π⋯π interaction is found in the range of typical stacking interaction (<4.00 Å and offsets of 1.6–1.8 Å).50
Fig. 17 Perspective view of the molecular structures showing weak π⋯π interactions and O–H⋯N interactions.49 |
Fig. 18 Different interacting dimers (D1–D5) in the crystal structure of HEMABM along with the interaction energies. |
Dimer | Distance a | Ecoul | Epol | EDisp (% Disp) | Erep | Etot | ΔEcp | Symmetry | Possible interactions | Geometry (Å, °) | HS label | ||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
H⋯A | D⋯A | ∠D–H⋯A | |||||||||||
D1 | 5.076 | −18.0 | −7.0 | −13.5 (35) | 18.7 | −19.9 | −24.34 | −X, −y, 1 − z | O1–H1⋯N2 | 2.03 | 3.022 | 173 | 1 |
D2 | 4.185 | −6.4 | −1.9 | −17.1 (67) | 13.7 | −11.7 | −17.99 | 1 − X, −y, 1 − z | Cg1⋯Cg1 | 3.834(2) | |||
D3 | 9.863 | −5.2 | −1.7 | −3.6 (34) | 5.1 | −5.3 | −4.87 | 1/2 − x, −1/2 + y, 3/2 − z | C4–H4⋯O1 | 2.43 | 3.404 | 148 | 2 |
C10–H10⋯O1 | 2.28 | 3.276 | 152 | 3 | |||||||||
D4 | 10.304 | −3.0 | −0.8 | −2.5 (40) | 1.8 | −4.4 | −5.07 | 1/2 − x, 1/2 + y, 1/2 − z | C7–H7⋯N2 | 2.61 | 3.694 | 174 | |
D5 | 9.270 | −2.7 | −0.9 | −2.3 (39) | 2.5 | −2.9 | −3.41 | 1/2 + x, 1/2 − y, 1/2 + z | C11–H11A⋯N1 | 2.65 | 3.489 | 133 | 4 |
C9–H9⋯N1 | 2.58 | 3.502 | 142 | ||||||||||
Lattice energy | −20.9 | −8.5 | −33.0 (52) | 30.2 | −32.2 |
Another 2D molecular layers are formed by the combinations of interaction in dimers D3–D5 which propagates along the bc plane (Fig. 21(b)). It is worthy noted that all the dimers (D1–D5) are predominately stabilized with electrostatic contribution in the range of 60–65% except in D2. The dimer D2, mainly stabilized with dispersion contribution (67%) because it is a π stacking dimer. The total lattice energy for HEMABM is −32.2 kcal mol−1, and the dispersion (52%) and electrostatic (48%) is contributions are approximately equal in stabilizing the crystal structure.
Further, to qualitatively analyse the intermolecular interactions present in the title molecule, we performed HS analysis and 2D fingerprint plots. The HS diagram and shape index (SI) of HEMABM is shown in Fig. 19. In HS diagram shown bright red color spot for O1–H1⋯N2(D1); C4–H4⋯O1; C10–H10⋯O1(D3) and C9–H9⋯N1(D5) interactions which are labelled in Table 9. The SI diagram shows that the red and blue color triangles on the surface of phenyl ring is due to the π stacking interactions (D2) in HEMABM structure. The 2D decomposed FP plot show the relative contribution of various intermolecular contacts in HEMABM is shown in Fig. 20. The H⋯H contacts are predominated in the crystal (35.8%) and it was shown as a double spike in de = di ≅ 1.1 Å. The second significant intermolecular contacts are N⋯H with a contribution of 29.9% and it shown as a sharp spike at de = di ≅ 1.2 Å. The C⋯H and O⋯H contacts are contributing 14.3% and 6.9% of total surface. It noted that the contribution of C⋯C contacts is comparably higher (10.1%) than the O⋯H contacts (6.9%) in the crystal structure and these C⋯C contacts are concentrated around de = di ≅ 1.8 Å as a green dot (it highlighted in red color circle) which indicate the existence of π stacking interaction in HEMABM.
Fig. 21 (a) The crystal packing diagram of HEMABM; (b) part of crystal packing of HEMABM showing the molecular layered structure formed by the combinations of interactions in D3–D5. |
The current density versus voltage (J–V) and luminous efficiency versus voltage (L–V) curves were measured simultaneously using a power supply (Newark element I4, Keithley 2400) with an in-house-designed and calibrated detection system. The J–V curve is recorded by direct processing of data acquired from the used Keithley 2400 apparatus. Luminous density is estimated through the voltage delivered by a photodiode located at fixed distance from the OLED. Photodiode calibration was performed by using the luminance of commercial LEDs, at different wavelengths and considering all geometrical parameters involved in the detection system.
The OLEDs fabrication, it architecture was glass/ITO/PEDOT:PSS/emissive layer/cathode. The PEDOT:PSS layer (50 nm) was deposited by spin coating and dried 10 minutes at 120 °C and normal atmosphere. The emissive layer (about 80 nm) was deposited from a solution with dioxane as solvent for compound HEMABM and chlorobenzene for MEH-PPV; 20 mg ml−1 and 5 mg ml−1 respectively. The solutions and substrates were pre-heated at 85 °C, films were dried also at 85 °C in normal atmosphere. As cathode 100 nm of Al were thermal evaporated also the eutectic alloy of Bi:In:Sn (melting point 65 °C) was deposited at 100 °C in normal atmosphere by drop casting.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 1868089. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c9ra05425a |
This journal is © The Royal Society of Chemistry 2019 |