Next Article in Journal
Dental Implants with a Calcium Ions-Modified Surface and Platelet Concentrates for the Rehabilitation of Medically Compromised Patients: A Retrospective Study with 5-Year Follow-Up
Next Article in Special Issue
Luminescence and Electrochemical Activity of New Unsymmetrical 3-Imino-1,8-naphthalimide Derivatives
Previous Article in Journal
Off-Axis Flexural Properties of Multiaxis 3D Basalt Fiber Preform/Cementitious Concretes: Experimental Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Acceptor–Donor–Acceptor Systems Based on Bis-(Imino-1,8-Naphthalimide)

by
Sonia Kotowicz
1,*,†,
Mateusz Korzec
1,*,†,
Agnieszka Katarzyna Pająk
1,
Sylwia Golba
2,
Jan Grzegorz Małecki
1,
Mariola Siwy
3,
Justyna Grzelak
4,
Sebastian Maćkowski
4 and
Ewa Schab-Balcerzak
1,3
1
Institute of Chemistry, University of Silesia, 9 Szkolna Str., 40-006 Katowice, Poland
2
Institute of Materials Engineering, University of Silesia, 75 Pulku Piechoty Str., 41-500 Chorzow, Poland
3
Centre of Polymer and Carbon Materials, Polish Academy of Sciences, 34 M. Curie-Sklodowska Str., 41-819 Zabrze, Poland
4
Faculty of Physics, Institute of Physics, Astronomy and Informatics, Nicolaus Copernicus University, 5 Grudziadzka Str., 87-100 Torun, Poland
*
Authors to whom correspondence should be addressed.
Contributed equally to this work: Sonia Kotowicz, Mateusz Korzec.
Materials 2021, 14(11), 2714; https://doi.org/10.3390/ma14112714
Submission received: 2 May 2021 / Revised: 18 May 2021 / Accepted: 18 May 2021 / Published: 21 May 2021

Abstract

:
In this paper, six novel symmetrical bis-(imino-1,8-naphthalimides) differing in core and N-substituent structure were synthesized, and their thermal (TGA, DSC), optical (UV-Vis, PL), electrochemical (DPV, CV) properties were evaluated. The compounds were stable to 280 °C and could be transferred into amorphous materials. Electrochemical investigations showed their ability to occur reductions and oxidations processes. They exhibited deep LUMO levels of about −3.22 eV and HOMO levels above −5.80 eV. The optical investigations were carried out in the solutions (polar and non-polar) and in films and blends with PVK:PBD. Bis-(imino-1,8-naphthalimides) absorbed electromagnetic radiation in the range of 243–415 nm and emitted light from blue to yellow. Their capacity for light emission under voltage was preliminarily tested in devices with an active layer consisting of a neat compound and a blend with PVK:PBD. The diodes emitted green or red light.

Graphical Abstract

1. Introduction

The discovery of organic compounds with semiconductor properties has opened up new possibilities for the development of electronics. A significant advantage of organic semiconductors is the practically unlimited possibility of modifying their physicochemical properties by appropriately designing their chemical structure. The imide rings as n-type semiconductors are very popular as electron acceptors with the possibility of functionalizing [1,2,3,4]. Due to their structure, imide rings can be divided into five- and six-membered rings obtained from phthalic or naphthalene/perylene anhydrides. Many publications associated with imide rings in organic electronics include symmetrical structures (phthalic diimides, naphthalene, perylene) [5,6,7,8,9,10]. The imide molecules’ properties can be changed by modifying the substituent type attached to the nitrogen atom or, more broadly, by modifying the perylene or naphthalene ring (4-C or 3-C) [9]. The introduction of structures with electron-donating properties to the naphthalene ring allows the obtainment of materials with a low band gap, which is beneficial in optoelectronic applications [11]. Compounds containing imide rings are characterized by many valuable properties, i.e., high thermal, chemical/oxidation resistance and the ability to self-organize [9]. Imides obtained from naphthalene anhydride as derivatives of 1,8-naphthalimide are widely used in medicine and biology, as compounds with high antitumor activity, cell dyes for bioimaging [12,13,14] and as ion detectors [15]. Most compounds containing 1,8-naphthalimide show absorption and emission in the visible spectrum region and good photo- and thermal-stability [16,17,18,19,20,21,22]. 1,8-Naphthalimide derivatives in OLEDs and OPVs have been studied as electron-transporting materials [23,24], emitters [25,26,27,28,29,30,31], emission polymers additionally transporting electrons [32], as a host in the guest-host structure (acting as a matrix for the blue emitter) [9], and as green dopant [33]. 1,8-Naphthalimides showed high electron mobility [34], and further the presence of a donor substituent allows to obtain compounds with dominant transport of holes and higher emission intensity [35,36,37,38] and high electron affinity [39]. Asymmetrical 1,8-naphthalimides structures, constructed from a single imide ring, are often tested as compounds substituted in the 4-C position for organic electronics [1,39,40]. In the papers [8,9,41,42,43,44] the thermal, optical, and electrochemical properties of azomethinediimides, as potential candidates for optoelectronic were presented. To the best of our knowledge, only few papers present naphthalimide derivatives with imine bond substituted in the 4-C or 3-C position investigated towards applications in organic electronics [1,45,46].
In the Supplementary Information in Figure S1 (see ESI) the structures of bisnaphthalimide derivatives are presented, dividing them according to the type of linker between 1,8-naphthalimide unit and the core of the compound. The described bisnaphthalimide derivatives include symmetric donor-acceptor structures with linkers, such as, vinyl [47], ethynyl [36], azo- [39,48], imino- [45], C-S [49] or C-O [50] bonds as well as without linkers by directly C–C binding with aryl [38,51]. Moreover, the bisnaphthalimides described so far in their core contained the following derivatives: phenyl, carbazole, triphenylamine, thiophene or bitiophene (Figure S1) [51]. To bisnaftaloimisdes are also include structures through the imide part [52,53,54] as well as through a bridge, e.g., amide [55,56]. Should be noted that the current research focuses on the synthesis and study of derivatives substituted in the 4-C position of the naphthalimide ring [51] while there are few reports on derivatives substituted in the 3-C position of the naphthalimide [45], as was mentioned above. In Table S1 (in ESI) collected data from optical and electrochemical measurements for the selected compounds shown in Figure S1 based on the available literature.
Herein, the synthesis of new six symmetrical bis-(imino-1,8-naphthalimides) formed the A–π–D–π–A system and its characterization are reported. The study is focused on impact of donor and N-substituent structure on their thermal, photophysical, and electrochemical properties. The research was supplemented with the Density Functional Theory (DFT) calculations. Additionally, the preliminary attempts for applications of obtained bis-(imino-1,8-naphthalimides) in OLEDs are presented.

2. Materials, Methods and Synthesis

Related information to measurements, films and device preparations as well as DFT calculations are given in the ESI. Describe of the synthesis is also given in the ESI.

2.1. Materials

Dialdehydes, TFA, Pd/C, 3-nitro-1,8-naphthalic anhydride, hydrazine, hexylamine, benzylamine, 4-methylbenzylamine, 2-phenethylamine, triethylamine (99%), PVK, PBD, Bu4NPF6, NaOH, PPh3, Pd(PPh3)2Cl2 were purchased from Sigma Aldrich (Merck). Trimethylsilylacetyle (98%) and 4-bromobenzaldehyde (99%) were purchased from Acros Organic. Solvents were purchased from Sigma Aldrich (Merck) and Avantor S.A. PEDOT:PSS and glass with ITO surface were purchased from OSSILA.

2.2. Structural Characterization

5,5′-(triphenylamine-4,4′-diimine)-bis(2-(2-hylhexyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (1a)
Yellow solid; Yield= 75%, 1H NMR (δ, 400 MHz, CDCl3, ppm): 8.63 (s, 2H), 8.55–8.52 (m, 4H), 8.18 (d, J = 8.1 Hz, 2H), 7.95 (d, J = 1.8 Hz, 2H), 7.90 (d, J = 8.6 Hz, 4H), 7.74 (t, J = 7.8 Hz, 2H), 7.42 (t, J = 7.6 Hz, 2H), 7.29–7.22 (m, 7H), 4.25–4.16 (m, 4H), 1.83–1.70 (m, 4H), 1.51–1.29 (m, 12H), 0.91 (t, J = 7.0 Hz, 6H). 13C NMR (δ, 101 MHz, CDCl3, ppm): 164.13, 164.03, 161.15, 150.66, 150.39, 146.15, 133.54, 132.68, 130.49, 130.11, 129.92, 127.38, 127.20, 126.5, 126.39, 125.43, 125.17, 124.62, 123.77, 123.26, 122.76, 40.58, 31.57, 28.09, 26.81, 22.58, 14.07. Anal. Calcd for C56H51N5O4 (858.04 g/mol): C (78.39%), H (5.99%), N (8.16%), Found: C (78.08%), H (5.91%), N (7.96%).
5,5′-(triphenylamine-4,4′-diimine)-bis(2-(2-benzyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione (2a)
Light brown solid; Yield= 45%, 1H NMR (δ, 400 MHz, CDCl3, ppm): 8.62 (s, 2H), 8.58–8.53 (m, 4H), 8.20 (d, J = 8.1 Hz, 2H), 7.98 (d, J = 1.8 Hz, 2H), 7.90 (d, J = 8.6 Hz, 4H), 7.75 (t, J = 7.8 Hz, 2H), 7.59 (d, J = 7.2 Hz, 4H), 7.42 (t, J = 7.8 Hz, 2H), 7.34 (t, J = 7.4 Hz, 4H), 7.29–7.25 (m, 9H), 5.41 (s, 4H). 13C NMR (δ, 101 MHz, CDCl3, ppm): insufficient concentration. Anal. Calcd for C58H39N5O4 (869.07 g/mol): C (80.07%), H (4.52%), N (8.05%), Found: C (79.78%), H (4.47%), N (8.07%).
5,5′-(triphenylamine-4,4′-diimine)-bis(2-(2-metylobenzyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (3a)
Light brown solid; Yield= 69%,1H NMR (δ, 400 MHz, CDCl3ppm): 8.62 (s, 2H), 8.58–8.53 (m, 4H), 8.20 (d, J = 8.3 Hz, 2H), 7.99 (d, J = 1.7 Hz, 2H), 7.90 (d, J = 8.6 Hz, 4H), 7.76 (m, 2H), 7.49 (d, J = 7.9 Hz, 4H), 7.41 (t, J = 7.8 Hz, 2H), 7.31–7.23 (m, 7H), 7.14 (d, J = 7.8 Hz, 4H), 5.39 (s, 4H), 2.33 (s, 6H). 13C NMR (δ, 101 MHz, CDCl3, ppm): 164.28, 164.13, 161.21, 150.69, 150.42, 146.15, 137.16, 134.34, 133.75, 132.72, 130.46, 130.38, 129.92, 129.11, 129.00, 127.41, 126.83, 126.51, 126.47, 125.42, 125.33, 124.98, 123.76, 123.27, 122.74, 43.39, 21.12. Anal. CalcdforC60H43N5O4 (898.01 g/mol): C (80.25%), H (4.83%), N (7.80%), Found: C (79.93%), H (4.67%), N (7.94%).
5,5′-(thiophene-2,5-diimine)-bis(2-(2-phenethyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (4b)
Light brown solid; Yield= 69%,1H NMR (δ, 400 MHz, CDCl3ppm): 8.87 (s, 2H), 8.65–8.55 (m, 4H), 8.25 (d, J = 8.2 Hz, 2H), 8.09 (m, 2H), 7.80 (t, J = 7.7 Hz, 2H), 7.67 (s, 2H), 7.41 (m, 4H), 7.35 (t, J = 7.4 Hz, 4H), 7.30–7.23 (m, 2H), 4.43 (m, 4H), 3.13–2.97 (m, 4H). 13C NMR (δ, 101 MHz, CDCl3, ppm): 163.92, 163.81, 154.14, 149.13, 146.62, 138.77, 133.91, 133.14, 132.65, 130.64, 129.03, 128.52, 127.65, 126.83, 126.49, 126.06, 124.42, 123.86, 122.74, 41.86, 34.31.Anal. Calcd for C46H32N4O4S (736.84 g/mol): C (74.98%), H (4.38%), N (7.60%), Found: C (79.72%), H (4.43%), N (7.56%).
5,5′-(biphenyl-4,4′-diimine)-bis(2-(2-hylhexyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (1c)
Yellow solid; Yield= 65%, 1H NMR (δ, 400 MHz, CDCl3, ppm): 8.74 (s, 2H), 8.60–8.52 (m, 4H), 8.22 (d, J = 7.9 Hz, 2H), 8.10 (d, J = 8.2 Hz, 4H), 8.02 (d, J = 1.6 Hz, 2H), 7.84 (d, J = 8.2 Hz, 4H), 7.77 (t, J = 7.7 Hz, 2H), 4.28–4.14 (m, 4H), 1.85–1.70 (m, 4H), 1.54–1.29 (m, 12H), 0.92 (t, J = 6.8 Hz, 6H). 13C NMR (δ, 101 MHz, CDCl3, ppm): 164.02, 161.53, 150.32, 143.46, 135.39, 133.62, 132.61, 130.35, 129.76, 127.65, 127.46, 126.54, 125.01, 124.78, 123.83, 122.77, 40.60, 31.56, 28.09, 26.80, 22.57, 14.06. Anal. Calcd for C50H46N4O4 (766.92 g/mol): C (78.30%), H (6.05%), N (7.31%), Found: C (77.90%), H (5.99%), N (7.34%).
5,5′-(ethyne-1,2-diyl-4,4′-diimine)-bis(2-(2-hylhexyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (1d)
Yellow solid; Yield= 55%. 1H NMR (δ, 400 MHz, CDCl3, ppm): 8.70 (s, 2H), 8.57 (d, J = 8.3 Hz, 4H), 8.23 (d, J = 8.0 Hz, 2H), 8.03–7.96 (m, 6H), 7.79 (t, J = 7.7 Hz, 2H), 7.71 (d, J = 8.1 Hz, 4H), 4.28–4.10 (m, 4H), 1.82–1.67 (m, 4H), 1.50–1.25 (m, 12H), 0.91 (m, 6H). 13C NMR (δ, 101 MHz, CDCl3, ppm): 164.43, 160.89, 149.94, 145.09, 136.27, 133.70, 133.09, 131.61, 129.57, 129.10, 127.56, 127.23, 125.01, 124.78, 122.84, 121.99, 114.04, 40.54, 31.57, 28.08, 26.72, 22.57, 14.05. Anal. Calcd for C52H46N4O4 (790.95 g/mol)C (78.96%), H (5.86%), N (7.08%), Found: C (78.86%) H (5.76%), N (6.93%).

3. Result and Discussion

3.1. Synthesis and Structural Characterization

The bis-(imino-1,8-naphthalimides) were obtained in a three-step reaction (Figure 1b). In the first stage, the condensation of commercially available 3-nitro-1,8-naphthalic anhydride with various amines, such as (1) hexylamine, (2) benzylamine, (3) 4-methylbenzylamine, (4) 2-phenethylamine in the ethanol was performed. Next, the nitro group was reduced to the amine by using 10% Pd/C as catalyst and hydrazine in ethanol [46,57]. Then, a reaction was carried out between the synthesized amines and commercially available dialdehydes such as: (a) 4,4′-diformyltriphenylamine, (b) 2,5-diformylthiophene, (c) 4,4′-biphenyldicarbaldehyde, as well as (d) dialdehyde obtained in the Sonogashira reaction. The obtained targeted compounds are presented in Figure 1.
The chemical structure and purity of synthesized final compounds were confirmed based on 1H NMR, 13C NMR (Figures S2 and S3 in ESI), and elemental analysis. Moreover, for molecule 1a the correlation spectra COSY and HMQC have performed for more detailed analysis of its structure (Figure S4 in ESI). In HMQC spectrum seventeen carbon atoms correlated with hydrogen atoms are seen (Figure S4). Two signals of isolated carbons in the imide part in the range of 160 ÷ 164 ppm are observed (Table 1). Thus, based on the analysis of the 2D spectra, signals for the imine bond in the other compounds were assigned (Table 1).
The chemical shift for the carbons in the imine bond depends on the core structure (a, b, c, and d presented in Figure 1). Compounds containing an aryl core (c and d) have a carbon assigned to imine (–N=CH–) about 150 ppm, with thiophene core at 154 ppm (b) and with triphenylamine (TPA) core at 161 ppm (a). In comparison, the signals of the imine in the range of 8.6–8.9 ppm in the 1HNMR spectrum were seen (Table 1). Experimental content of nitrogen, carbon, and hydrogen atoms were found to be consistent with the theoretical value.

3.2. Thermal Characterization

The thermal stability, phase transition temperatures (Tm, Tc), and glass transition temperatures (Tg) of bis-(imino-1,8-naphthalimide) derivatives were determined by the thermogravimetric analysis (TGA) and the differential scanning calorimetry (DSC), respectively. Organic materials dedicated to electronic applications should show a sufficiently high temperature of the beginning of thermal decomposition (T5), and melting (Tm) or glass transition temperatures (Tg), which allows for trouble-free integration in devices [46,58,59,60]. The collected data from thermal investigations are presented in Table 2. The exemplary DSC thermograms are given in Figure 2. (for other compounds in Figure S6).
The presented bis-(imino-1,8-naphthalimide) derivatives exhibited temperatures of 5% weight loss above 280 °C. Considering the impact of core structure comparing 1a, 1c and 1d the highest T5 = 426 °C was recorded for the compound with triphenylamine core and hexyl chain (1a). The compounds with triphenylamine (1a–3a) showed different T5 dependent on the structure attached to the nitrogen in the imide ring. The T5 was grown from 426 °C (1a) with hexyl chain to 431 °C (2a) with benzyl and 446 °C (3a) with 4-methylbenzyl (1a < 2a < 3a). This same behavior, growing T5 dependent on the substituent attached to the nitrogen atom in imide unit, was noticed in our previous publication [61]. The temperature of 5% weight loss was lowered by the presence of the ethynyl bond (1d) compared to compound 1c, moreover the two steps of decomposition of 1d was noticed. The first step of decomposition of 1d at the 275 °C can be assigned to degradation of ethynyl linkage [62].
The melting temperature (Tm) as the endothermic peak was registered during the first heating scan in the range of 173–259 °C, which indicates that bis-(imino-1,8-naphthalimide) derivatives were obtained as crystalline compounds [46]. During the second heating scan (after rapid cooling) the glass transition temperature (Tg) was recorded in the range 74–140 °C (Table 2). In the case of compounds with thiophene (4b) and biphenyl (1c) structure upon further heating above Tg, the “cold crystallization temperature” (Tc, as the exothermic peak) and Tm were seen (Figure 2b). The presence of the glass transition temperature confirmed the ability of investigated molecules to transform from crystalline into the amorphous state. Thus, bis-(imino-1,8-naphthalimide) derivatives are molecular glasses [45,63]. The lack of tendency for crystallization in the second heating scan was confirmed for 1a–3a and 1d, which means that the presented molecules are stable molecular glasses. The Tg was strongly depended on the chemical structure where the highest temperature was recorded for the compounds with TPA core and benzyl ring (2a) and the lowest for 1c with biphenyl and hexyl chain (Tg = 74 °C).
Symmetrical 1,8-naphthalimide with TPA core and hexyl chain (1a) had higher glass transition temperature (Tg = 86 °C) than its unsymmetrical analogue—3-(4-(diphenylamine)-N-benzo)-N-hexyl-1,8-naphthalimide (Tg = 51 °C) described in our former work [46]. This same tendency was noticed for T5 and Tm. In the case of symmetrical 1,8-naphthalimide with thiophene core and phenylethyl structure (4b) the higher Tm and Tg were registered compare with symmetrical imide with thiophene core and hexyl chain—5,5′-(thiophene-2,5-diylbis(methan-1-yl-1-yli-dene))bis(azan-1-yl-1-ylidene)bis(2-hexyl-1H-benzo[de]isoquinoline-1,3(2H)-dione), (Tm = 239 °C; Tg= 84 °C) described in our former publication [45].

3.3. Electrochemical Investigations

Electrochemical investigations were performed by cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in 0.1M Bu4NPF6 electrolyte in dichloromethane (DCM) (10−3 mol/dm3). Based on the CV and DPV voltammograms the onset potentials of reduction, and oxidation (Ered(onset) and Eox(onset)) were determined and the electron affinities (EA) and ionization potentials (IP) were estimated. The cyclic voltammograms of the 1a, 1c, 1d compounds are presented in Figure 3. whereas, for others are shown in Figure S7. The electrochemical data are collected in Table 3.
The ionization potentials (IP) were obtained between −5.80 and −5.46 eV, and the electron affinities (EA) between −3.91 and −3.22 eV with the energy band gap (Eg) below 2.39 eV (above 1.66 eV). The obtained derivatives show the irreversible oxidation processes as well as irreversible reduction processes, except for 2a (with triphenylamine and benzyl), 3a (with triphenylamine and 4-methylbenzyl) and 4b (with thiophene and phenylethyl), where the quasi-reversible first reduction process was seen (ΔE = 80 mV for 3a and ΔE = 110 mV for 2a, ΔE = 110 mV for 4b) (see Figure S7). The reduction process is related to acceptor moieties namely to the azomethine linkage (–N=CH–) and imide ring where the reduction process can occur. The first reduction step was seen between −1.46 and −1.99 V, dependent on the structures of the compounds, and indicates the reduction of –N=CH– [45].
The reduction process C=O → C=O in the imide ring was seen as two peaks (Ered2 and Ered3). The properly separated and defined peaks were registered by the DPV method except for compound 1c, where the one broad peak was shown (also during the oxidation process). Presence of the ethyne-1,2-diylbenzene structure strongly affected on the Ered1 position which is seen comparing the compound 1c (Ered1CV = −1.92 V) and 1d (Ered1CV = −1.66 V). The synthesized bis-(imino-1,8-naphthalimide) derivatives were characterized by the multi-step processes of oxidation, except for 1c. The first oxidation process, seen about 0.65 V, can be assigned to the oxidation of the imine bond and may vary depending on the donor substituent to the –N=CH– bond. The other Eox peaks are connected with donor moieties attached to the imine bond and the nitrogen in the imide ring as was reported in our previous work [46].
The unsymmetrical analog of 1a, reported in our former work [46], exhibited both, oxidation and reduction, irreversible processes. The presence of the second 1,8-naphthalimide group lowered the Ered and Eg. However, as was mentioned earlier, the quasi-reversible first reduction process for compounds with triphenylamine core (2a and 3a) was observed, which means that presence of benzyl ring (2a) and 4-methylbenzyl (3a) strongly affected the first reduction process. Differences were also notice for IP, EA and Eg values, between 5,5′-(thiophene-2,5-diylbis(methan-1-yl-1-yli-dene))bis(azan-1-yl-1-ylidene)bis(2-hexyl-1H-benzo[de]isoquinoline-1,3(2H)-dione) and 5,5′-(thiophene-2,5-diylbis(methan-1-yl-1-yli-dene))bis(azan-1-yl-1-ylidene)bis(2-(3,5-dimethylphenyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione), described in our former publication [45] and compound 4b, presented in this work. Ionization potentials and electron affinities levels are higher for 4b molecule (with phenylethyl substituent) than for other compounds.
It can be noticed the value of the energy gap in the tested compounds (1a–3a, 4b, 1c, 1d) are relatively lower (1.58–2.36 eV) compared to the compounds 1–20, collected in Table S1 (Eg > 2.10 eV). Furthermore, based on the cyclic voltammetry studies in CH2Cl2, it can be concluded that the presence of an azo- or imino-linker (2.10–2.68 eV) seems to be important for lowering energy gap in contrast to the derivatives with C–C bond (e.g., -C≡C-, Eg = 2.55–2.94 eV). This dependence may be due to better electrochemical activity of azo- and imine- derivatives in contrast to other analyzed compounds (Figure S1, Table S1). In this case, the indication of the linker seems to be justified, because the above-mentioned tendency can be described independently of the core and the substituent in the imides part for analyzed compounds (Figure S1, compare molecules 3–6 with 9–14).

3.4. Theoretical Calculations

The theoretical calculations were performed with the density functional theory (DFT) carried out using the Gaussian09 program on B3PW91/6-311g++ level. The molecular geometry of the singlet ground and first excited states of the compounds were optimized in the gas phase and its electronic structures and electronic transitions were calculated with use of the Polarizable Continuum Model (PCM) in dichloromethane for comparison of HOMO, LUMO energies with electrochemical data and in chloroform. The optimized geometries of the compounds are depicted in Figure S8. in the ESI. In the ground state molecules of the compounds display deviation from planarity within the central part (TPA (a series), thiophene (4b), biphenyl (1c) and ethyne-1,2-diylbenzene (1d)) and 1,8-naphthalimide (Table S2). Comparing the energies of HOMOs and LUMOs determined on the basis of electrochemical data (Table S6) with theoretically calculated values it can be noticed that the calculated HOMO energies of a series (TPA core) compounds correspond with the experimental values of IP determined from CV measurements. The discrepancies between the experimental and the calculated energy values of the HOMO level for 1c, 1d and 4b are greater. The calculated LUMOs energies were overestimated but the calculated values of the HOMO and LUMO energies were used only for consistency with geometry optimization. For a more detailed description of the molecular orbitals the contribution of a molecule parts, i.e., core, –N=CH–, 1,8-naphthalimide and substituent R to a molecular orbital was calculated. The obtained DOS diagrams are presented in Figure S10. in the ESI and composition of selected molecular orbitals are gathered in Table S3. Calculations show that LUMO in the compounds of 1a–3a and 1c is localized on the 1,8-naphthalimide part. In 1d and 4b LUMO comprises the central molecule part with the azomethine–1,8-naphthalimide fragment. HOMO is localized on the triphenylamine core of the 1a–3a compounds. In the case of 1c, 1d and 4b molecules HOMO comprises the central part (biphenyl (1c), ethyne-1,2-diylbenzene (1d), thiophene (4b)) including 1,8-naphthalimide linked by imine bond. The aromatic substituents on the 1,8-naphtalenediimide nitrogen play role at the HOMO-3, HOMO–4 levels, in the case of 2a, 3a, and especially in 4b compound where HOMO-2 is localized on the phenylethyl substituent (Table S3). The transition corresponding to the excitation wavelength (340 nm vide infra) has, in all compounds, a mixed intra molecular charge transfer/locally excited (ICT/LE) nature (Table S4). However, in the case of 2a, the excitation is related to the charge transfer from 1,8-naphthalimide to triphenylamine core, and in 3a from TPA to the 1,8-naphthalimide fragment with imine linker. The R substituents take part in the excitation process in the case of compound 4b for which the transition is related to π → π* transition.

3.5. Photophysical Properties

The electronic absorption (UV-Vis) and photoluminescence (PL) spectra were recorded in non-polar and polar solvents, such as chloroform and N-methyl-2-pyrrolidone (NMP) (c= 10−5 mol/dm3), and in the solid-state as films and blends with PVK:PBD matrix on the glass substrates. The UV-Vis spectra are presented in Figure 4. Data from absorption and emission measurements are collected in Table 4.
Bis-(imino-1,8-naphthalimide) in chloroform and NMP solutions were absorbed electromagnetic radiation with λmax in the range 243–415 nm (5.10–2.99 eV). The absorption bands in the higher energy (λmax = 243–274 nm) can be attributed to π → π* electron transitions in the aromatic rings, while the λmax above 328 nm to π →π* naphthalimide [46,64]. The λmax in the lower energies, above 360 nm, belong to the charge transfer (CT) between the core (at the 3-C position) and naphthalimide [64]. The shift of the maximum of the absorption band in CHCl3 and NMP towards lower energies was observed for bis-(imino-1,8-naphthalimides) with a triphenylamine (TPA) core (1a–3a) compared to other compounds (Figure 4). The shift of λmax towards lower energies indicates a better degree of conjugation for 1a–3a, confirmed by NMR measurements.
The same position of λmax in chloroform and N-methyl-2-pyrrolidone for the compound 1d with ethyne-1,2-diylbenzene substituent was observed, suggesting that the difference in dipole moments between the excited and the ground state are minimal [40]. The bathochromic shift of λmax of molecule with a triple bond (-C≡C-), (1d) compared to compound with the biphenyl core (1c) in solutions were noticed (∆λmax = 35 nm) (Table 4). No differences in λmax position between molecules with TPA core in chloroform solution were seen, however, dissimilarities were recorded in emission spectra. The hypsochromic shift of λmax of molecules with benzyl (2a, λmax = 387 nm) and 4-methylbenzyl (3a, λmax = 397 nm) attached to the nitrogen in the imide ring compared to 1a with hexyl chain (λmax = 413 nm) registered in NMP was seen.
Bis-(imino-1,8-naphthalimide) derivatives absorbed radiation in films with λmax in the range 345–413 nm. The hypsochromic shift in films was noticed for molecules with TPA core depending on the substituent in the imide unit (hexyl chain 1a, λmax= 413nm > phenyl ring 2a, λmax = 402 nm > 4-methylphenyl 3a, λmax = 397 nm). Thus, comparing film with chloroform solution, the hypsochromic shift of ∆λmax= 13 nm (2a) and ∆λmax = 18 nm (3a) was seen. Differences in the λmax for other molecules were slight (∆λmax = 2–7 nm).
The second naphthalimide group, present in 1a molecule, allowed to obtain bathochromically shifted absorption spectrum (by approximately 31 nm in CHCl3 and 36 nm in the film) comparing to molecule presented in our former work [46] 3-(4-(diphenylamine)-N-benzo)-N-hexyl-1,8-naphthalimide, due to difference in the degree conjugation. The position of the maximum absorption bands for 5,5′-(thiophene-2,5-diylbis(methan-1-yl-1-yli-dene))bis(azan-1-yl-1-ylidene)bis(2-hexyl-1H-benzo[de]isoquinoline-1,3(2H)-dione) and 5,5′-(thiophene-2,5-diylbis(methan-1-yl-1-yli-dene))bis(azan-1-yl-1-ylidene)bis(2-(3,5-dimethylphenyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (published in former work [45]) and 5,5′-(thiophene-2,5-diimine)-bis(2-(2-phenethyl)-1H-benzo[de]isoquinoline-1,3(2H)-dione) (4b) are very similar, which allows concluding that interactions between thiophene core-imine bond and imide unit are more crucial compare to substituent in the imide ring.
Bis-(imino-1,8-naphthalimides) showed the ability to emission of light in solutions. In PL spectra one band with maximum (λem) in the range of 450–544 nm from blue to green light was observed (Figure 5c). The emission data are summarized in Table 4. The naphthalimide 4b with thiophene core and 1c with biphenyl core showed the λem in the blue spectral region in both solutions. The red shift of λem in NMP was seen for 1a and 1d compared to CHCl3 (blue spectral region in CHCl3, green spectral region in NMP, see Figure S12) under external UV-light λex= 366 nm, but for 2a and 3a, the opposite behavior was noticed. The various excitation wavelengths (λex) did not affect the maximum PL band (λem) in accordance with the Kasha’s rule.
The photoluminescence spectrum of the molecule with triphenylamine core and hexyl chain (1a) was hypsochromically shifted about 32 nm in CHCl3 solution compared to PL spectrum of compounds with benzyl (2a) and 4-methylphenyl (3a) (Table 4). An opposite behavior was detected in the NMP solution. Furthermore, in the solid-state in the films and blends, this behavior was not seen (λem at about 540 and 500 nm). The quantum yield (ΦPL) in the films of 1a–3a was about 4%, but compounds in chloroform showed higher ΦPL 25, 26 and 17% for 1a, 2a, 3a, respectively. In the S1 state the planarity distortion of the 1,8–naphthalimide–X–1,8–naphthalimide part is reduced, which has impact on the delocalization and conjugation of π-electrons in the molecules. Changing the molecule’s geometry in the excited state is easier in the solution than in film which may explain the very low quantum yields of the film emissions. The most bathochromically shifted spectrum was recorded for the molecule with thiophene core and phenylethyl (4b) in film (λem = 577 nm, the yellow spectral region; Figure 5d) with the lowest ΦPL (also in CHCl3, ΦPL = 1%). The PL maxima for all the molecules in the films were bathochromically shifted compared to the chloroform solution, except for 1c (Figure S14).
In the next step of the research the PL spectra of the poly(N-vinylcarbazole) (PVK):2-tert-butylphenyl-5-biphenyl-1,3,4-oxadiazole (PBD) (1:1) blends were registered, where the energy transfer can take place (Förster transfer or exchange (Dexter) mechanism) [65,66]. The energy transfer process in the host-guest structure may occur, consisting of the energy transfer from the host matrix (PVK:PBD) to the guest molecule in the ground state [59]. The above mechanisms are processes of non-radiative energy transfer and may coexist in the case of minimal distances between the guest and the host. The Förster Energy Transfer (FRET) results from dipole-dipole interactions and the Dexter energy transfer requires overlapping of electron clouds [59]. Effective energy transfer occurs when the emission intensity of the guest increases and the host decreases in the host’s presence [67,68]. Furthermore, the effective FRET energy transfer is possible when host’s emission is overlapping with the absorption spectrum of the guest [65,68]. The PL spectrum of the matrix (PVK:PBD) overlapping with the absorption spectrum in the case of the molecules with TPA (1a–3a) and thiophene (4b) core, for the other molecules partial overlap was seen (Figure 4c). In PL spectra of the blends, one (for 4b) or two bands were found (Figure S14). The band with λem about 390 nm came from the emission of the PVK:PBD matrix and the other originate from the bis-(imino-1,8-naphthalimides) emission (see Table 4, Figure 5a,b). The presence of the second emission band (at about 390nm) in the PL spectra of PVK:PBD blends indicates no complete energy transfer. It is worth mentioning that for compound 1a, with TPA and hexyl chain, the highest value of the quantum yield in the PVK:PBD blends was obtained (ΦPL = 11.3%).
The unsymmetrical analogue of 1a with TPA core, presented in our previous work [46] emitted light with the same λem position in the solid-state but in solutions, the red shift (in CHCl3,∆λem = 64 nm) and blue shift (in NMP, ∆λem = 32 nm) was observed. The unsymmetrical 3-(4-(diphenylamine)-N-benzo)-N-hexyl-1,8-naphthalimideexhibited lower ΦPL and shorter lifetime (τ) compared to 1a [46]. The PL spectrum in CHCl3 of the molecule 4b is bathochromically shifted compared to the compound with the 3,5-dimethylphenyl substituent (∆λem = 18 nm) presented in publication [46]. Moreover, the compound 2a showed hypsochromic shift of the UV-Vis and PL spectra (∆λmax = 16 nmin CHCl3 and 33 nm in film; ∆λem = 47 nm in CHCl3 and 7 nm in film) compared to the synthesized 1,8-naphthalimides with TPA at 4-C position (C–C linker) and benzyl ring in the imide unit (N-C bond) presented by Arunchai et al. [40]. Differences were also noticed in ΦPL and τ, where 5,5′-(triphenylamine-4,4′-diimine)-bis(2-(2-benzyl)-1H-benzo[de]isoquinoline-1,3(2H)-dionehad the longest life-time but lowest ΦPL in the chloroform solution compared to naphthalimide [65].
The most similar values of the λmax, λem and quantum yield are represented by bisimides with carbon–carbon aryls linkers (at 4-C position of thiophene (18), EDOT(19) and bithiophene(20), Table S1) compared to the compounds presented in this work. Moreover, the collected molecules in the Table S1 (ESI) and bis-(imino-1,8-naphthalimides) emitted light from blue to green spectral region.

3.6. Electroluminescence Study

The synthesized bis-(imino-1,8-naphthalimides) were used as an active layer or as components of the active layer in the organic electroluminescence diodes (OLEDs). The following configuration of devices ITO/PEDOT: PSS/molecules/Al and ITO/PEDOT: PSS/PVK: PBD: molecules 2 wt.%/Al were applied. PVK is often used in organic electronics to build an active layer due to the high mobility of holes [59,69]. PVK was mixed with PBD, which exhibited high mobility of electrons, creating a two-component matrix and finally a host-guest active layer [66,70,71]. Data from electroluminescence (EL) measurements are collected in Table 5 and the EL spectra of chosen fabricated devices are presented in Figure 6 (and Figure S16).
Only two devices, which the bis-(imino-1,8-naphthalimides) acts as an active layer, with the molecules 4b and 1d, did not show electroluminescence, for other diodes the red light was registered (λEL ≈ 675 nm). A lack of emission induced by external voltage of devices with compound 4b (thiophene core and phenylethyl substituent) can be explained by low ΦPL and short PL lifetime (τ) in the solid state. The presence of 2 wt.% compounds dispersed molecularly in the matrix as active layer in diode allowed for registered emission of light with the maximum EL band (λEL) at the green visible spectral region (Figure 6).
The highest electroluminescence intensity was registered for device with triphenylamine core and hexyl chain (1a) dispersed in PVK:PBD matrix (2 wt.% content of 1a) (Table 5). Due to the high EL intensity the device with PVK:PBD contains 1 and 15 wt.% of 1a were prepared (Figure 6). It was showed that 15 wt.% content of the 1a molecule in the PVK:PBD matrix allowed to obtain the diode with the highest electroluminescence intensity, however the external voltage was high. It should be emphasized that the EL spectra were registered under external voltage about 10 V. The guest-host devices with 1a emitted green light with the shift of λEL position together within crease of its content in the PVK:PBD matrix (λEL = 530 nm < 540 nm < 555 nm).
The energy levels of the HOMO and LUMO orbital’s of bis-(imino-1,8-naphthalimides) are respectively higher and lower than a binary matrix energy levels (Figure 6), which may indicate the dominance of the trapping mechanism [68,72]. However, in the recombination process in OLEDs, the processes of energy transfer (FRET or/and Dexter) and the charge trapping mechanism may coexist [73]. Comparing EL spectra with PL spectra the λEL shift towards longer wavelengths was observed (Table 4 and Table 5). This behavior can be attributed to electroplex formation, both in the bis-(imino-1,8-naphthalimides), which can also explain shoulder in the case of device with an active layer containing 1a molecule, as well as in the guest-host structures [49,74]. At this stage of research, the diode parameters (luminance, luminous efficacy) were not measured.
The device based on the unsymmetrical 3-(4-(diphenylamine)-N-benzo)-N-hexyl-1,8-naphthalimide (described in [46], dispersed in PVK:PBD with 2 wt.% content) emitted light with λEL = 538 nm, which means, that the presence of an additional 1,8-naphthalimide group did not have an effect on the λEL position (compared with 1a). Larger differences were observed in the case of diodes with 15 wt.% content of molecules in the matrix (∆λEL = 14 nm).

4. Conclusions

Six symmetrical bis-(imino-1,8-naphthalimides) with triphenylamine (TPA), thiophene, biphenyl, and ethyne-1,2-diylbenzenecoreinthe 3-C position of naphthalene structure and hexylamine (1a,1c,1d), benzylamine (2a), 4-methylbenzylamine (3a), and 2-phenethylamine (4b) substituents in the imide ring were synthesized. These molecules showed temperature of 5% weight loss above 280 °C and were obtained as crystalline compounds with ability to transform into stable molecular materials. The glass transition temperature was strongly dependent on the chemical structure. Bis-(imino-1,8-naphthalimides) were electrochemically active and showed the energy band gap below 2.39 eV. Based on the DFT calculations, HOMO is localized on the central part of the compounds—core, and in the case of 1c, 1d, and 4b is mixed (core–imine bond–naphthalimide). LUMO is localized on the 1,8-naphthalimide part, and for 1d and 4b LUMO is also mixed. The presented molecules emitted blue, green and yellow light in the investigated media with the higher quantum yields in solutions (except 1d and 4b). The bis-(imino-1,8-naphthalimides) were acting as the active layer and as a component of the active layer in the preliminary EL studies. The highest EL intensity was registered for the device with the 15 wt.% of 1a (triphenylamine core and hexyl chain) in PVK:PBD matrix. Molecule 1a, in our opinion, is promising for optoelectronic applications and can be further investigated and modified as a component of the binary blends or green/red light emitter.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/ma14112714/s1, Figure S1: Structure of the compounds divided into the linker in the donor-acceptor system (4-C or 3-C position) (Publications: 37, 39, 40, 46, 48, 49, 50, 52), Table S1: Photophysical and electrochemical properties of 1–20 derivatives (Figure S1.) based on the literature data, Figure S2: 1HNMRof the investigated compounds (400 MHz, CDCl3), Figure S3: 13C NMR of the investigated compounds (400 MHz, CDCl3), Figure S4: HMQC and COSY correlation spectra for 1a (400 MHz, CDCl3), Figure S5: The investigated compounds under (a) 1a, (c) 1c,(d) 1d day light and (b) 1a under UV-light, Figure S6: DSC thermograms of (a) 2a, (b) 3a and (c) 4b, Figure S7: The (a) reduction and (b) oxidation processes of 2a, 3a, 4b (Pt, v= 0.1 V/s, 0.1 mol/dm3 Bu4NPF6 in CH2Cl2 with 10−3 mol/dm3 of compounds), Figure S8: Optimized geometries of the compounds (a) 1a, (b) 2a, (c) 3a, (d) 1c, (e) 1d, (f) 4b, Figure S9: Contours of HOMO and LUMO of the compounds (a) 1a, (b) 2a, (c) 3a, (d) 1c, (e) 1d, (f) 4b, Figure S10: Density-of-states diagrams of ground state of the molecules, Figure S11: Experimental and calculated UV-Vis spectra of bis-(imino-1,8-naphthalimide) derivatives, Table S2: Calculated dipole moments (in CHCl3) and geometrical parameters of the molecules in ground and first singlet excited state, Table S3: Composition of the selected MO in ground state of the compounds, Table S4: The calculated electronic transitions corresponding to excitation resulting most intense luminescence in CHCl3 solution, Figure S12: (a) 1a and (b) 1d under day light and under UV-light (λ = 366 nm) in NMP, Table S5: PL life-time (τ) measurements data, Figure S13. PL life-time (τ) curves of (a) 2a, (b) 3a and (c) 4b in chloroform solution, Figure S14. Emission spectra of (a) 2a; (b) 3a; (c) 4b and (d) 1c with PVK:PBD., Figure S15. Emission spectra of 1a and PVK:PBD, Table S6. EA, IP, HOMO, LUMO, Eg and Egopt, Figure S16. Electroluminescence (EL) spectra of the working devices under an applied voltage. Above the graphs the device structures are shown. Presented are maxima of EL intensity.

Author Contributions

Conceptualization, S.K. and M.K.; formal analysis, S.K., M.K., S.M. and E.S.-B..; investigation, S.K., M.K., A.K.P., S.G., M.S. and J.G.; writing—original draft preparation, S.K., M.K., A.K.P., S.G. and J.G.M.; writing—review & editing, S.K., M.K., S.M. and E.S.-B.; visualization, S.K., M.K., J.G.M., M.S. and J.G.; supervision, E.S.-B. All authors have read and agreed to the published version of the manuscript.

Funding

The research was co-financed by The National Centre for Research and Development (NCBiR) under grant LIDER XI no. LIDER/39/0137/L-11/19/NCBR/2020 and by the funds granted under the Research Excellence Initiative of the University of Silesia in Katowice.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

The GAUSSIAN-09 calculations were carried out in the Wrocław Centre for Networking and Supercomputing, WCSS, Wrocław, Poland, http://www.wcss.wroc.pl; accessed date: 11 March 2021 (grant number 18). Authors thank to K. Bednarczyk for TGA and H. Janeczek for DSC measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jin, R.F.; Chang, Y.F. A theoretical study on photophysical properties of triphenylamine-cored molecules with naphthalimide arms and different π-conjugated bridges as organic solar cell materials. Phys. Chem. Chem. Phys. 2015, 17, 2094–2103. [Google Scholar] [CrossRef]
  2. Grabchev, I.; Staneva, D. Photophysical properties of new polymerizable 1,8-Naphthalimides and their copolymers with methylmethacrylate. Z. Nat. 2003, 58, 558–562. [Google Scholar] [CrossRef]
  3. Refat, M.S.; Al Didamony, H.; Abou El-Nour, K.M.; Grabchev, I.; El-Zayat, L. Synthesis and characterizations of charge-transfer complexes of 1,8-naphthalimides with different acceptors. Bulg. Chem. Commun. 2010, 42, 279–299. [Google Scholar]
  4. Bag, S.S.; Pradhan, M.K.; Kundu, R.; Jana, S. Highly solvatochromic fluorescent naphthalimides: Design, synthesis, photophysical properties and fluorescence switch-on sensing of ct-DNA. Bioorg. Med. Chem. Lett. 2013, 23, 96–101. [Google Scholar] [CrossRef] [PubMed]
  5. Bijak, K.; Janeczek, H.; Grucela-Zajac, M.; Schab-Balcerzak, E. New room-temperature thermotropic perylene-based bisimides: Synthesis, liquid crystalline, light-emitting and electrochemical properties. Opt. Mater. 2013, 35, 1042–1050. [Google Scholar] [CrossRef]
  6. Schab-Balcerzak, E.; Iwan, A.; Krompiec, M.; Siwy, M.; Tapa, D.; Sikora, A.; Palewicz, M. New thermotropic azomethine-naphthalene diimides for optoelectronic applications. Synth. Met. 2010, 160, 2208–2218. [Google Scholar] [CrossRef]
  7. Bijak, K.; Grucela-Zajac, M.; Janeczek, H.; Wiacek, M.; Schab-Balcerzak, E. New azomethine-phthalic diimides:Synthesis and thermal, optical and electrochemical characterization. Synth. Met. 2013, 175, 146–154. [Google Scholar] [CrossRef]
  8. Grucela-Zajac, M.; Bijak, K.; Kula, S.; Filapek, M.; Wiacek, M.; Janeczek, H.; Skorka, L.; Gasiorowski, J.; Hingerl, K.; Sariciftci, N.S.; et al. (Photo)physical properties of new molecular glasses end-capped with thiophene rings composed of diimide and imine units. J. Phys. Chem. C 2014, 118, 13070–13086. [Google Scholar] [CrossRef]
  9. Grucela-Zajac, M.; Bijak, K.; Zaleckas, E.; Grigalevicius, S.; Wiacek, M.; Janeczek, H.; Schab-Balcerzak, E. Electronic and thermal properties of compounds bearing diimide, azomethine and triphenylamine units. Opt. Mater. 2014, 37, 543–551. [Google Scholar] [CrossRef]
  10. Rybakiewicz, R.; Djurado, D.; Cybulski, H.; Dobrzynska, E.; Kulszewicz-Bajer, I.; Boudinet, D.; Verilhac, J.M.; Zagorska, M.; Pron, A. Arylene bisimides with trirylamine N-substituents as new solution processable organic semiconductors: Synthesis, spectroscopic, electrochemical and electronic properties. Synth. Met. 2011, 161, 1600–1610. [Google Scholar] [CrossRef]
  11. Sakai, N.; Mareda, J.; Vauthey, E.; Matile, S. Core-substituted naphthalenediimides. Chem. Commun. 2010, 46, 4225–4237. [Google Scholar] [CrossRef] [Green Version]
  12. Xiao, H.; Chen, M.; Shi, G.; Wang, L.; Yin, H.; Mei, C. A novel fluorescent molecule based on 1,8-naphthalimide: Synthesis, spectral properties, and application in cell imaging. Res. Chem. Intermed. 2010, 36, 1021–1026. [Google Scholar] [CrossRef] [Green Version]
  13. Saito, G.; Velluto, D.; Resmini, M. Synthesis of 1,8-naphthalimide-based probes with fluorescent switch triggered by flufenamic acid. R. Soc. Opern Sci. 2018, 5, 172137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Duke, R.M.; Veale, E.B.; Pfeffer, F.M.; Kruger, P.E.; Gunnlaugsson, T. Colorimetric and fluorescent anion sensors: An overview of recent developments in the use of 1,8-naphthalimide-based chemosensors. Chem. Soc. Rev. 2010, 39, 3936–3953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Chang, I.J.; Choi, M.G.; Jeong, Y.A.; Lee, S.H.; Chang, S.K. Colorimetric determination of Cu2+ in simulated wastewater using naphthalimide-based Schiff base. Tetrahedron Lett. 2017, 58, 474–477. [Google Scholar] [CrossRef]
  16. Ulla, H.; Kiran, M.R.; Garudachari, B.; Satyanarayan, M.; Umesh, G.; Isloor, A. Blue emitting halogen-phenoxy substituted 1,8-naphthalimides for potential organic light emitting diode applications. Opt. Mater. 2014, 37, 311–321. [Google Scholar] [CrossRef]
  17. Yordanova, S.; Grabchev, I.; Stoyanov, S.; Milusheva, V.; Petkov, I. Synthesis and functional characteristics of two new yellow-green fluorescent PAMAM dendrimers periphery modified with 1,8-naphthalimides. Inorg. Chim. Acta 2014, 409, 89–95. [Google Scholar] [CrossRef]
  18. Bojinov, V.B.; Simeonov, D.B. Synthesis of highly photostable blue-emitting 1,8-naphthalimides and their acrylonitrile copolymers. Polym. Degrad. Stab. 2010, 95, 43–52. [Google Scholar] [CrossRef]
  19. Prezhdo, O.V.; Uspenskii, B.V.; Prezhdo, V.V.; Boszczyk, W.; Distanov, V.B. Synthesis and spectral-luminescent characteristics of N-sustituted 1,8-naphthalimides. Dye. Pigment. 2007, 72, 42–46. [Google Scholar] [CrossRef]
  20. Cao, X.; Meng, L. Large red-shifted fluorescent emission via intermolecular π-π stacking in 4-ethynyl-1,8-naphthalimide-based supramolecular assemblies. Langmuir 2014, 30, 11753–11760. [Google Scholar] [CrossRef]
  21. Sonalin, S.; Sakthivel, K.; Nagarajan, S. Functionalization of 1,8-naphthalimides-an approach towards air-stable n-type organic semiconductors. Mater. Today Proc. 2018, 5, 16592–16597. [Google Scholar] [CrossRef]
  22. Zhu, M.; Miao, J.; Hu, Z.; Chen, Y.; Liu, M.; Murtaza, I.; Meng, H. A novel A-D-A small molecule with 1,8-naphthalimide as a potential non-fullerene acceptor for solution processable solar cells. Dye. Pigment. 2017, 142, 39–50. [Google Scholar] [CrossRef]
  23. Wang, S.; Zeng, P.; Liu, Y.; Yu, G.; Sun, X.; Niu, H.; Zhu, D. Luminescent properties of a novel naphthalimide-fluorene molecule. Synth. Met. 2005, 150, 33–38. [Google Scholar] [CrossRef]
  24. Do, T.T.; Takeda, Y.; Manzhos, S.; Bell, J.; Tokito, S.; Sonar, P. Naphthalimide and capped anthraquinone based solution-processable n-channel organic semiconductors: Effect of alkyl chain engineering on charge transport. J. Mater. Chem. C 2018, 6, 3774–3786. [Google Scholar] [CrossRef] [Green Version]
  25. Bojinov, V.; Grabchev, I. Synthesis of new polymerizable 1,8-naphthalimide dyes containing a 2-hydroxyphenylbenzotriazole fragment. Dye. Pigment. 2003, 59, 277–283. [Google Scholar] [CrossRef]
  26. Pluczyk, S.; Laba, K.; Schab-Balcerzak, E.; Bijak, K.; Kotowicz, S.; Lapkowski, M. Electrochemical and spectroelectrochemical properties of new polymers with diimide subunits. J. Electroanal. Chem. 2017, 795, 90–96. [Google Scholar] [CrossRef]
  27. Kolosov, D.; Adamovich, V.; Djurovich, P.; Thompson, M.E.; Adachi, C. 1,8-Naphthalimides in phosphorescent organic LEDs: The interplay between dopant, exciplex, and host emission. J. Am. Chem. Soc. 2002, 124, 9945–9954. [Google Scholar] [CrossRef] [PubMed]
  28. Kucheryavy, P.; Li, G.; Vyas, S.; Hadad, C.; Glusac, K.D. Electronic properties of 4-substituted naphthalimides. J. Phys. Chem. A 2009, 113, 6453–6461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Sun, F.; Jin, R. Optical and charge transport properties of N-butyl-1,8-naphthalimide derivatives as organic light-emitting materials: A theoretical study. J. Lumin. 2014, 149, 125–132. [Google Scholar] [CrossRef]
  30. Jin, R.; Tang, S. Theoretical study on optical and electronic properties of bipolar molecules with 1,8-naphthalimide and triphenylamine moieties as organic light-emitting materials. J. Mol. Graph. Model. 2013, 42, 120–128. [Google Scholar] [CrossRef]
  31. Chai, W.; Jin, R. Theoretical investigations into optical and charge transfer properties of donor-acceptor 1,8-naphthalimide derivatives as possible organic light-emitting materials. J. Mol. Struct. 2016, 1103, 177–182. [Google Scholar] [CrossRef]
  32. Liu, J.; Li, Y.; Wang, Y.; Sun, H.; Lu, Z.; Wu, H.; Peng, J.; Huang, Y. Synthesis and luminescent properties of blue sextuple-hydrogen-bond self-assembly molecular duplexes bearing 4-phenoxy-1,8-naphthalimide moieties. Opt. Mater. 2012, 34, 1535–1542. [Google Scholar] [CrossRef]
  33. Liu, J.; Tu, G.; Zhou, Q.; Cheng, Y.; Geng, Y.; Wang, L.; Ma, D.; Jing, X.; Wang, F. Highly efficient green light emitting polyfluorene incorporated with 4-diphenylamino-1,8-naphthalimides as green dopant. J. Mater. Chem. 2006, 16, 1431–1438. [Google Scholar] [CrossRef]
  34. Zhang, J.; Zhang, X.; Xiao, H.; Li, G.; Liu, Y.; Li, C.; Huang, H.; Chen, X.; Bo, Z. 1,8-naphthalimide-based planar small molecular acceptor for organic solar cells. ACS Appl. Mater. Interfaces 2016, 8, 5475–5483. [Google Scholar] [CrossRef]
  35. Gautam, P.; Sharma, R.; Misra, R.; Keshtov, M.L.; Kuklin, S.A.; Sharma, G.D. Donor-acceptor-acceptor (D-A-A) type 1,8-naphthalimides as non-fullerene small molecule acceptors for bulk heterojunction solar cells. Chem. Sci. 2017, 8, 2017–2024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Saini, A.; Thomas, K.R.J. Bis-naphthalimides bridged by electron acceptors: Optical and self-assembly characteristics. RSC Adv. 2016, 6, 71638–71651. [Google Scholar] [CrossRef]
  37. Zhang, J.; Xiao, H.; Zhang, X.; Wu, Y.; Li, G.; Li, C.; Chen, X.; Ma, W.; Bo, Z. 1,8-Naphthalimide-based nonfullerene acceptors for wide optical band gap polymer solar cells with an ultrathin active layer thickness of 35nm. J. Mater. Chem. C 2016, 4, 5656–5663. [Google Scholar] [CrossRef]
  38. Zhengneng, J.; Najun, L.; Chuanfeng, W.; Huajiang, J.; Jianmei, L.; Qizhong, Z. Synthesis and fluorescence property of some novel 1,8-naphthalimide derivatives containing a thiophene ring at the C-4 position. Dye. Pigment. 2013, 96, 204–210. [Google Scholar] [CrossRef]
  39. Gan, J.A.; Song, Q.L.; Hou, X.Y.; Chen, K.; Tian, H. 1,8-naphthalimides for non-doping OLEDs: The tunable emission color from blue, green to red. J. Photochem. Photobiol. A Chem. 2004, 162, 399–406. [Google Scholar] [CrossRef] [Green Version]
  40. Arunchai, R.; Sudyoadsuk, T.; Prachumrak, N.; Namuangruk, S.; Promarak, V.; Sukwattanasinitt, M.; Rashatasakhon, P. Synthesis and characterization of new triphenylamine-1,8-naphthalimides for organic light-emitting diode applications. New J. Chem. 2015, 39, 2807–2814. [Google Scholar] [CrossRef]
  41. Schab-Balcerzak, E.; Grucela, M.; Malecki, G.; Kotowicz, S.; Siwy, M.; Janeczek, H.; Golba, S.; Praski, A. Azomethine diimides and-capped with anthracene moieties: Experimental and theoretical investigations. J. Mol. Struct. 2017, 1128, 462–470. [Google Scholar] [CrossRef]
  42. Bijak, K.; Filapek, M.; Wiacek, M.; Janeczek, H.; Grucela, M.; Smolarek, K.; Mackowski, S.; Schab-Balcerzak, E. Preparation and characterization of new aliphatic-tailed five- and six-membered azomethine-diimides. Mater. Chem. Phys. 2016, 171, 97–108. [Google Scholar] [CrossRef]
  43. Nowak, E.M.; Sanetra, J.; Grucela, M.; Schab-Balcerzak, E. Azomethine naphthalene diimides as component of active layers in bulk heterojunction solar cells. Mater. Lett. 2015, 157, 93–98. [Google Scholar] [CrossRef]
  44. Iwan, A.; Boharewicz, B.; Tazbir, I.; Sikora, A.; Schab-Balcerzak, E.; Grucela-Zajac, M.; Skórka, Ł. Structural and electrical properties of mixture based on P3HT:PCBM and low band gap naphthalene diimide-imines. Synth. Met. 2014, 189, 183–192. [Google Scholar] [CrossRef]
  45. Schab-Balcerzak, E.; Siwy, M.; Filapek, M.; Kula, S.; Malecki, G.; Laba, K.; Lapkowski, M.; Janeczek, H.; Domanski, M. New core-substituted with electron-donating group 1,8-naphthalimides towards optoelectronic applications. J. Lumin. 2015, 166, 22–39. [Google Scholar] [CrossRef]
  46. Kotowicz, S.; Korzec, M.; Siwy, M.; Golba, S.; Malecki, J.G.; Janeczek, H.; Mackowski, S.; Bednarczyk, K.; Libera, M.; Schab-Balcerzak, E. Novel 1,8-naphthalimides substituted at 3-C position: Synthesis and evaluation of thermal, electrochemical and luminescent properties. Dye. Pigment. 2018, 158, 65–78. [Google Scholar] [CrossRef]
  47. Mikroyannidis, J.A.; Ye, S.; Liu, Y. Electroluminesent divinylene- and trivinylene-molecules with terminal naphthalimide or phthalimide segments. Synth. Met. 2009, 159, 492–500. [Google Scholar] [CrossRef]
  48. Gudeika, D.; Lygaitis, R.; Mimaitė, V.; Grazulevicius, J.; Jankauskas, V.; Lapkowski, M.; Data, P. Hydrazones containing electron-accepting and electron-donating moieties. Dye. Pigment. 2011, 91, 13–19. [Google Scholar] [CrossRef]
  49. Steidl, L.; Jhaveri, S.J.; Ayothi, R.; Sha, J.; McMullen, J.D.; Ng, S.Y.C.; Zipfel, W.R.; Zentel, R.; Ober, C.K. Non-ionic photo-acid generators for applications in two-photon lithography. J. Mater. Chem. 2009, 19, 505–513. [Google Scholar] [CrossRef]
  50. Luo, Y.; Wang, Y.; Chen, S.; Wang, N.; Qi, Y.; Zhang, X.; Yang, M.; Huang, Y.; Li, M.; Yu, J.; et al. Facile Access to Twisted Intramolecular Charge-Transfer Fluorogens Bearing Highly Pretwisted Donor–Acceptor Systems Together with Readily Fine-Tuned Charge-Transfer Characters. Small 2017, 13, 1604113. [Google Scholar] [CrossRef]
  51. Gudeika, D. A review of investigation on 4-substituted 1,8-naphthalimide derivatives. Synth. Met. 2020, 262, 116328. [Google Scholar] [CrossRef]
  52. Wang, Y.-F.; Lua, H.Y.; Chen, C.; Li, M.; Chen, C.-F. 1,8-Naphthalimide-based circularly polarized TADF enantiomers as the emitters for efficient orange-red OLEDs. Org. Electron. 2019, 70, 71–77. [Google Scholar] [CrossRef]
  53. Kumar, S.; Muhammad, S.; Koh, J.; Khalid, M.; Ayub, K. A combined experimental and computational study of 2,2’-(diazene-1,2-diylbis(4,1-phenylene))bis(6-(butylamino)-1H-benzo[de]isoquinoline-1,3(2H)-dione): Synthesis, optical and nonlinear optical properties. Optik 2019, 182, 162952. [Google Scholar] [CrossRef]
  54. Staneva, D.; Grabchev, I.; Soumillion, J.P.; Bojinov, V. A new fluorosensor based on bis-1,8-naphthalimide for metal cations and protons. J. Photochem. Photobiol. A Chem. 2007, 189, 192–197. [Google Scholar] [CrossRef]
  55. Esteban-Gómez, D.; Fabbrizzi, L.; Licchelli, M.; Sacchi, D. A two-channel chemosensor for the optical detection of carboxylic acids, including cholic acid. J. Mater. Chem. 2005, 15, 2670–2675. [Google Scholar] [CrossRef]
  56. Korzec, M.; Kotowicz, S.; Łaba, K.; Łapkowski, M.; Małecki, J.G.; Smolarek, K.; Mackowski, S.; Schab-Balcerzak, E. Naphthalene Diimides Prepared by a Straightforward Method and Their Characterization for Organic Electronics. Eur. J. Org. Chem. 2018, 15, 1756–1760. [Google Scholar] [CrossRef]
  57. Korzec, M.; Malarz, K.; Mrozek-Wilczkiewicz, A.; Rzycka-Korzec, R.; Schab-Balcerzak, E.; Polański, J. Live cell imaging by 3-imino-(2-phenol)-1,8-naphthalimides: The effect of ex vivo hydrolysis. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 238, 118442. [Google Scholar] [CrossRef]
  58. McEwan, J.A.; Clulow, A.J.; Nelson, A.; Yepuri, N.R.; Burn, P.L.; Gentle, I.R. Dependence of Organic Interlayer Diffusion on Glass-Transition Temperature in OLEDs. ACS Appl. Mater. Interfaces 2017, 9, 14153–14161. [Google Scholar] [CrossRef]
  59. Skórka, Ł.; Kurzep, P.; Wiosna-Sałyga, G.; Łuszczyńska, B.; Wielgus, I.; Wróbel, Z.; Ulański, J.; Kulszewicz-Bajer, I. New diarylaminophenyl derivatives of carbazole: Effect of substituent position on their redox, spectroscopic and electroluminescent properties. Synth. Met. 2017, 228, 1–8. [Google Scholar] [CrossRef]
  60. Sęk, D.; Kotowicz, S.; Kula, S.; Siwy, M.; Szłapa-Kula, A.; Małecki, J.G.; Maćkowski, S.; Schab-Balcerzak, E. Thermal, spectroscopic, electrochemical, and electroluminescent characterization of malononitrile derivatives with triphenylamine structure. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2019, 210, 136–147. [Google Scholar] [CrossRef]
  61. Korzec, M.; Kotowicz, S.; Rzycka-Korzec, R.; Schab-Balcerzak, E.; Małecki, J.G.; Czichy, M.; Łapkowski, M. Novel β-ketoenamines versus azomethines for organic electronics: Characterization of optical and electrochemical properties supported by theoretical studies. J. Mater. Sci. 2020, 55, 3812–3832. [Google Scholar] [CrossRef]
  62. Kim, K.-W.; Kim, G.-H.; Park, C.-E.; Choi, J.-H. 1,8-Naphthalimide Derivatives Containing Ethynyl Linkage and Blue Light Emitting Properties. Bull. Korean Chem. Soc. 2017, 38, 956–959. [Google Scholar] [CrossRef]
  63. Kula, S.; Szlapa-Kula, A.; Kotowicz, S.; Filapek, M.; Bujak, K.; Siwy, M.; Janeczek, H.; Maćkowski, S.; Schab-Balcerzak, E. Phenanthro[9,10-d]imidazole with thiophene rings toward OLEDs application. Dye. Pigment. 2018, 159, 646–654. [Google Scholar] [CrossRef]
  64. Adaci, M.; Murata, Y.; Nakamura, S. Spectral Similarity and Difference of Naphthalenetetracarboxylic Dianhydride, Perylenetetracarboxylic Dianhydride, and Their Derivatives. J. Phys. Chem. 1995, 99, 14240–14246. [Google Scholar] [CrossRef]
  65. Kotowicz, S.; Siwy, M.; Filapek, M.; Malecki, J.G.; Smolarek, K.; Grzelak, J.; Mackowski, S.; Slodek, A.; Schab-Balcerzak, E. New donor-acceptor-donor molecules based on quinoline acceptor unit with Schiff base bridge: Synthesis and characterization. J. Lumin. 2017, 183, 458–469. [Google Scholar] [CrossRef]
  66. Glowacki, I.; Szamel, Z. The nature of trapping sites and recombination centres in PVK and PVK–PBD electroluminescent matrices seen by spectrally resolved thermoluminescence. J. Phys. D Appl. Phys. 2010, 43, 295101. [Google Scholar] [CrossRef]
  67. Bujak, P.; Kulszewicz-Bajer, I.; Zagorska, M.; Maurel, V.; Wielgus, I.; Pron, A. Polymers for electronics and spintronics. Chem. Soc. Rev. 2013, 42, 8895–8999. [Google Scholar] [CrossRef]
  68. Hussain, S.A. An Introduction to Fluorescence Resonance Energy Transfer (FRET). arXiv 2009, arXiv:0908.1815. [Google Scholar]
  69. Grykien, R.; Luszczynska, B.; Glowacki, I.; Kurach, E.; Rybakiewicz, R.; Kotwica, K.; Zagorska, M.; Pron, A.; Tassini, P.; Maglione, M.G.; et al. Photo- and electroluminescent properties of bithiophene disubstituted 1,3,4-thiadiazoles and their application as active components in organic light emitting diodes. Opt. Mater. 2014, 37, 193–199. [Google Scholar] [CrossRef]
  70. Lee, D.Y.; Lee, M.H.; Lee, C.J.; Park, S.K. Driving Characteristics of Poly(N-vinylcarbazole) and 2-(4-biphenylyl)-5-(4-tert-butylphenyl)-1,3-4-oxadiazole-Based Polymer Light Emitting Diodes. Electron. Mater. Lett. 2013, 9, 663–668. [Google Scholar] [CrossRef]
  71. Yang, X.; Neher, D.; Hertel, D.; Daubler, T.K. Highly Efficient Single-Layer Polymer Electrophosphorescent Devices. Adv. Mater. 2004, 16, 161–166. [Google Scholar] [CrossRef]
  72. Kotwica, K.; Bujak, P.; Wamil, D.; Pieczonka, A.; Wiosna-Salyga, G.; Gunka, P.A.; Jaroch, T.; Nowakowski, R.; Luszczynska, B.; Witkowska, E.; et al. Structural, Spectroscopic, Electrochemical, and Electroluminescent Properties of Tetraalkoxydinaphthophenazines: New Solution-Processable Nonlinear Azaacenes. J. Phys. Chem. C 2015, 119, 10700–10708. [Google Scholar] [CrossRef]
  73. Yersin, H. Highly Efficient OLEDs with Phosphorescent Materials; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008; ISBN 978-527-40594-7. [Google Scholar]
  74. Wen, L.; Li, F.; Xie, J.; Wu, C.; Zheng, Y.; Chen, D.; Xu, S.; Guo, T.; Qu, B.; Chen, Z.; et al. Electroplex emission at PVK/Bphen interface for application in white organic light-emitting diodes. J. Lumin. 2011, 131, 2252–2254. [Google Scholar] [CrossRef]
Figure 1. (a) Synthesis of dialdehyde: (I) 4-bromobenzaldehyde, trimethylsilylacetylene, Et3N, PdCl2, PPh3,Cu, 65 °C, 3 h, (II) MeOH, K2CO3, 2 h, rt., (III) 4-bromobenzaldehyde, 4-ethynylbenzaldehyde, Pd(PPh3)2Cl2, PPh3, CuI, Et3N, 70 °C, 15 h. (b) Synthesis of the compounds: (IV) 3-nitro-1,8-naphthalic anhydride, amine (1, 2, 3 or 4), EtOH, 2 h in reflux, (V) 10% Pd/C, EtOH, hydrazine, 60 °C, N2, 6 h, (VI)EtOH, CF3COOH, dialdehyde (a, b, c, d).
Figure 1. (a) Synthesis of dialdehyde: (I) 4-bromobenzaldehyde, trimethylsilylacetylene, Et3N, PdCl2, PPh3,Cu, 65 °C, 3 h, (II) MeOH, K2CO3, 2 h, rt., (III) 4-bromobenzaldehyde, 4-ethynylbenzaldehyde, Pd(PPh3)2Cl2, PPh3, CuI, Et3N, 70 °C, 15 h. (b) Synthesis of the compounds: (IV) 3-nitro-1,8-naphthalic anhydride, amine (1, 2, 3 or 4), EtOH, 2 h in reflux, (V) 10% Pd/C, EtOH, hydrazine, 60 °C, N2, 6 h, (VI)EtOH, CF3COOH, dialdehyde (a, b, c, d).
Materials 14 02714 g001
Figure 2. The DSC thermograms of (a) 1a and (b) 1c registered in I and II heating scan.
Figure 2. The DSC thermograms of (a) 1a and (b) 1c registered in I and II heating scan.
Materials 14 02714 g002
Figure 3. The voltammograms of the (a) reduction and (b) oxidation processes of the 1a, 1c, 1d (Pt, v = 0.1V/s, 0.1 mol/dm3 Bu4NPF6 in CH2Cl2 with 10−3 mol/dm3 of compounds).
Figure 3. The voltammograms of the (a) reduction and (b) oxidation processes of the 1a, 1c, 1d (Pt, v = 0.1V/s, 0.1 mol/dm3 Bu4NPF6 in CH2Cl2 with 10−3 mol/dm3 of compounds).
Materials 14 02714 g003
Figure 4. Absorption spectra of bis-(imino-1,8-naphthalimide) derivatives in (a) N-methyl-2-pyrrolidone; (b) chloroform, and (c) film together with PL spectrum of PVK:PBD.
Figure 4. Absorption spectra of bis-(imino-1,8-naphthalimide) derivatives in (a) N-methyl-2-pyrrolidone; (b) chloroform, and (c) film together with PL spectrum of PVK:PBD.
Materials 14 02714 g004
Figure 5. Emission spectra (PL) (a) of 1a; (b) 1d; (c) in chloroform solution and in (d) the films of bis-(imino-1,8-naphthalimide) derivatives.
Figure 5. Emission spectra (PL) (a) of 1a; (b) 1d; (c) in chloroform solution and in (d) the films of bis-(imino-1,8-naphthalimide) derivatives.
Materials 14 02714 g005
Figure 6. Diagram of HOMO and LUMO energy levels of diode components and EL spectra of the working devices under an applied voltage with their photo.
Figure 6. Diagram of HOMO and LUMO energy levels of diode components and EL spectra of the working devices under an applied voltage with their photo.
Materials 14 02714 g006aMaterials 14 02714 g006b
Table 1. Selected NMR chemical shifts (ppm) and multiplicities for compounds.
Table 1. Selected NMR chemical shifts (ppm) and multiplicities for compounds.
CodeImide
13C (ppm)
Imine
1H (ppm)13C (ppm)
1a164.1, 164.08.6 (s, 2H)161.2
2a-8.6 (s, 2H)-
3a164.3, 165.18.6 (s, 2H)161.2
4b163.9, 164.88.9 (s, 2H)154.1
1c164.0, 161.58.7 (s, 2H)150.3
1d164.4, 160.98.7 (s, 2H)149.9
Table 2. Thermal stability and transition temperatures of investigated 1,8-naphthalimide derivatives.
Table 2. Thermal stability and transition temperatures of investigated 1,8-naphthalimide derivatives.
CompoundTGADSC
I Heating ScanII Heating Scan
T5 aTmax bTm cTg eTc dTm c
(°C)(°C)(°C)(°C)(°C)(°C)
1a42646417386ndnd
2a431479259140ndnd
3a446488209138ndnd
4b428498255118170226
1c38743222074148225
1d285275, 479251127ndnd
a T5, T10—temperature based on 5 and 10% weight loss from TGA curves, b Temperature of the maximum decomposition rate from DTG curves. c Melting temperature, d Cold crystallization temperature, e Glass transition temperature, nd—not detected.
Table 3. Electrochemical data of the bis-(imino-1,8-naphthalimide) derivatives.
Table 3. Electrochemical data of the bis-(imino-1,8-naphthalimide) derivatives.
CompoundMethodEred 1Ered(onset)Eox 1Eox(onset)EAIPEg
[V][V][V][V][eV][eV][eV]
1aCV−1.72 a−1.230.73 a0.53−3.87−5.631.76
DPV−1.46−1.220.500.36−3.88−5.461.58
2aCV−1.82 b−1.730.69 a0.55−3.37−5.652.28
DPV−1.89−1.770.610.53−3.33−5.632.30
3aCV−1.96 b−1.830.67 a0.56−3.27−5.662.39
DPV−1.99−1.880.580.48−3.22−5.582.36
4bCV−1.75 b−1.530.76 a0.46−3.35−5.562.21
DPV−1.67−1.570.780.56−3.43−5.662.23
1cCV−1.92 a−1.380.76 a0.41−3.72−5.511.79
DPV−1.96−1.280.700.38−3.82−5.481.66
1dCV−1.66 a−1.270.90 a0.70−3.83−5.801.97
DPV−1.62−1.190.800.70−3.91−5.801.89
IP = −5,1-Eox(onset)·|e|, EA = −5,1-Ered(onset)·|e|, Eg = Eox(onset)-Ered(onset). Measurements inCH2Cl2 with concentration 10−3 mol/dm3 and electrolyte 0.1 mol/dm3Bu4NPF6. Pt as the working electrode. a Irreversible process. b Quasi-reversible process. Ered 1–the first reduction process, Eox 1–the first oxidation process. Ered(onset)–the onset potential of the first reduction process, Eox(onset)–the onset potential of the first oxidation process.
Table 4. Collected UV-Vis and PL properties of bis-(imino-1,8-naphthalimides).
Table 4. Collected UV-Vis and PL properties of bis-(imino-1,8-naphthalimides).
CompoundMediumUV-visPL
λmax (nm) (ε·104) aλem(nm)Stokes ShiftΦ (%)τ c (ns)X2
(cm−1) b
1aCHCl3 d345(2.5), 411(9.8)50163562516.91.045
NMP d342(2.5), 413(2.8)5445831---
FILM345, 41353755913.812.50.930
PVK:PBD e310 sh, 344 sh, 424388, 504-11.3--
2aCHCl3 d246(14.1), 260 sh, 345(8.1), 415(10.0)53353352619.11.017
NMP d345(3.5), 387(6.9),414sh5166460---
FILM348, 40254464934.016.21.061
PVK:PBD e310 sh, 344 sh, 429 sh384, 499-7.5--
3aCHCl3 d246(14.0), 260 sh, 345(10.1), 415(12.1)53353351421.01.063
NMP d345(3.8), 393(8.9), 414 sh5146460---
FILM345, 39754468073.715.31.098
PVK:PBD e310 sh, 344 sh, 429 sh384, 499-7.3--
4bCHCl3 d246 sh, 274(23.2), 345(5.8), 390(4.7)5156224122.00.935
NMP d345(10.1), 387(5.6)4503618---
FILM348, 39757742211.72.161.019
PVK:PBD e310 sh, 344 sh384-2.3--
1cCHCl3 d243(3.6), 340(9.1)5039445716.71.123
NMP d267(21.2), 340 (4.6)51710069---
FILM345435, 490 sh59974.811.01.056
PVK:PBD e310 sh, 344 sh399, 477-3.6--
1dCHCl3 d345(9.8), 376(8.8)4705319311.21.012
NMP d345(17.8), 376(8.8)5257548---
FILM360, 389 sh, 424 sh471 sh, 509-4.69.01.047
PVK:PBD e310 sh, 344 sh386, 498-8.4--
a ε—Absorption coefficient, (dm3·mol−1·cm−1). b Stokes shifts calculated according to the equation Δν = (1/λabs − 1/λem) × 107 (cm−1). c The average time from the multi-exponential decay profiles (Table S4). d c = 10−5 mol/dm3. sh—shoulder. e 2 wt.%—concentration of compound in the matrix PVK:PBD. Underlined data indicates the dominant band.
Table 5. The electroluminescence data of prepared diodes.
Table 5. The electroluminescence data of prepared diodes.
CompoundλELUELMaxEL Intensity
(nm)(V)(Counts)
1a670 a372207
530 b2416,214
540 c2614,228
555 d28195,060
2a675 a12608
535 c9267
3a675 a16489
532 c11226
1c675 a19125
525 c231837
1d529 c237091
a ITO/PEDOT:PSS/molecules/Al. b ITO/PEDOT:PSS/PVK:PBD:molecule 1 wt.%/Al. c ITO/PEDOT:PSS/PVK:PBD:molecules 2 wt.%/Al.d ITO/PEDOT:PSS/PVK:PBD:molecule 15 wt.%/Al. λEL—the maximum of EL band. UELMax—external voltage for maximal EL intensity.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kotowicz, S.; Korzec, M.; Pająk, A.K.; Golba, S.; Małecki, J.G.; Siwy, M.; Grzelak, J.; Maćkowski, S.; Schab-Balcerzak, E. New Acceptor–Donor–Acceptor Systems Based on Bis-(Imino-1,8-Naphthalimide). Materials 2021, 14, 2714. https://doi.org/10.3390/ma14112714

AMA Style

Kotowicz S, Korzec M, Pająk AK, Golba S, Małecki JG, Siwy M, Grzelak J, Maćkowski S, Schab-Balcerzak E. New Acceptor–Donor–Acceptor Systems Based on Bis-(Imino-1,8-Naphthalimide). Materials. 2021; 14(11):2714. https://doi.org/10.3390/ma14112714

Chicago/Turabian Style

Kotowicz, Sonia, Mateusz Korzec, Agnieszka Katarzyna Pająk, Sylwia Golba, Jan Grzegorz Małecki, Mariola Siwy, Justyna Grzelak, Sebastian Maćkowski, and Ewa Schab-Balcerzak. 2021. "New Acceptor–Donor–Acceptor Systems Based on Bis-(Imino-1,8-Naphthalimide)" Materials 14, no. 11: 2714. https://doi.org/10.3390/ma14112714

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop