Assembly, Optical Properties a

0 downloads 0 Views 10MB Size Report
This manuscript has been accepted after peer review and appears as an. Accepted Article online prior to editing, proofing, and formal publication of the final ...
A Journal of

Accepted Article Title: Phenothiazine Based Oligo(p-phenylenevinylene)s: Substituents Affected Self-Assembly, Optical Properties and Morphology Induced Transport Authors: C Arivazhagan, Sitakanta Satapathy, Arijit Jana, Partha Malakar, Edamana Prasad, and Sundargopal Ghosh This manuscript has been accepted after peer review and appears as an Accepted Article online prior to editing, proofing, and formal publication of the final Version of Record (VoR). This work is currently citable by using the Digital Object Identifier (DOI) given below. The VoR will be published online in Early View as soon as possible and may be different to this Accepted Article as a result of editing. Readers should obtain the VoR from the journal website shown below when it is published to ensure accuracy of information. The authors are responsible for the content of this Accepted Article. To be cited as: Chem. Eur. J. 10.1002/chem.201801810 Link to VoR: http://dx.doi.org/10.1002/chem.201801810

Supported by

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER Phenothiazine Based Oligo(p-phenylenevinylene)s: Substituents Affected Self-Assembly, Optical Properties and Morphology Induced Transport

Dedication ((optional))

Abstract: Designing intramolecular charge-transfer (ICT) based luminogenic ordered assemblies exhibiting significant electrical transport is a challenging task in the field of organic optoelectronics. In this context, a series of novel phenothiazine-based oligo(pphenylenevinylene)s (OPV16) derivatives were designed and their structureproperty relationship was investigated. Upon examining their photophysical properties, all the OPVs were found to exhibit significant intramolecular charge transfer (ICT) characteristics in organic solvents. While inspecting the self-assembly behaviour, the OPV with long alkyl chain on the central phenyl core (OPV4) underwent gelation in organic solvent mixtures through strong hydrophobic interactions of the long hexadecyl chains and   interactions from their aromatic counter parts. Computational studies revealed a lamellar packing of molecules in the assembly. Interestingly, the degree of ICT and the gelation abilities of OPVs were significantly influenced by the electronic nature of the substituents appended to the peripheral phenothiazines. Further, the AC impedance results revealed an increase in storage and electronic transport for the fluorescent thin films prepared by an increase in the content of OPV4 in PMMA.

Introduction Organic functional molecular systems with well-defined architectures and morphologies have gained enormous attraction due to their heavy potential usage in diverse fields, such as molecular electronics and light-energy conversion.[1,2] Organic -functional units are endowed with characteristic electronic and optical properties, attributed to their fairly delocalized electrons throughout the conjugated network. This enables researchers to establish intricate insights and shedding light onto their structureproperty relationship and selfassembly.[3,4] The non-covalent interactions, such as, hydrogen bonding,  and electrostatic forces play a decisive role in tuning the factors regulating the self-assembly and final

[a]

Dr. C. Arivazhagan, Mr. S. Satpathy, Mr. A. Jana, Dr. P. Malakar, Prof. Dr. E. Prasad, Prof. Dr. S. Ghosh Department of Chemistry Indian Institute of Technology Madras Chennai 600 036 (India) E-mail: [email protected] [email protected] Supporting information and the ORCID identification number(s) for the author(s) of this article can be found under http://doi.org/10.1002/xxxxx.

morphologies in such systems.[5] Ordered self-assembly of functional -systems has been an active area of research with potential applications in the field of flexible photonics and electronics.[6] This attributes to facilitation of long range intraand intermolecular delocalization of charge carriers within the well-defined ordered network. Controlling the different hierarchies of the aggregates, from molecules to devices is one of the key issues in soft-matter electronics.[7,8] In order to comprehend probable functions, both the chemical composition of the -conjugated molecules and the degree of alignment of their molecular chains must be controlled. The planar geometries facilitate the formation of stacked assemblies, while the interchain electronic coupling determines the performance of such -conjugated systems within electronic devices.[9] One of the approaches adopted to solve this problem is the formation of extended structures of self-assembled systems in appropriate solvent medium, which leads to gelation of a large volume of solvent mass within the molecular pockets.[10] A large number of findings reported on molecular gel systems are mostly serendipitous.[11] Nevertheless, the past few years have witnessed significant progress towards the understanding of supramolecular self-assembled architectures.[12] This expedite the design of a variety of low molecular weight gelators (LMWGs) from various research groups that form strong supramolecular gels with interesting properties useful for potential applications in petrochemical industry, medicine, tissue engineering, water purification, catalysis, sensors and advanced materials.[13-19] Among these self-assembling systems, the linear conjugated molecular systems have gained considerable attention due to their promising optoelectronic properties and useful in designing flexible or portable next generation organic electronic devices.[20] Some of the linearly -conjugated oligomeric systems with gelation abilities reported in the literature includes derivatives of thiophenes, phenylenevinylenes, phenylenethynylenes, cyanostilbenes and phenylenes.[21-25] On the other hand, the linear oligo(p-phenylenevinylene)s (OPVs) based systems are given special preferences over others, attributing to simple synthetic protocols and excellent electronic properties.[26] However, proposing a design paradigm for luminogenic OPV systems, exhibiting significant electronic storage and transport properties through controlled assembly is a rare phenomena and is still a field of ongoing research. In the following, we describe the synthesis, characterization and photophysical properties of a series of intramolecular charge-transfer (ICT) active OPV derivatives and their

This article is protected by copyright. All rights reserved.

Accepted Manuscript

C. Arivazhagan, Sitakanta Satapathy, Arijit Jana, Partha Malakar, Edamana Prasad* and Sundargopal Ghosh*[a]

10.1002/chem.201801810

Chemistry - A European Journal

organogelation abilities, tested in various organic solvents mixtures. Electrochemical AC impedance studies have also been carried out for the fluorescent thin films of OPV4 doped with PMMA.

Results and Discussion Synthesis and Characterization The general synthetic scheme for the preparation of OPV16 is shown in Scheme 1 and the detailed synthesis of intermediates is given in the experimental section. The starting materials, 10ethyl-phenothiazine-3-carbaldehyde, 7-bromo-10-ethyl-10Hphenothiazine-3-carbaldehyde, 1,4-bis(diethylphosphonatomethyl)-2,5dihexyloxybenzene, 1,4-bis(diethylphosphonatomethyl)-2,5-dihexadecyloxybenzene were prepared according to literature procedures.[27-29] 7-(dimesitylboranyl)-10-ethyl-10Hphenothiazine-3-carbaldehyde was newly synthesized and characterized by multinuclear NMR (1H, 13C and 11B) spectroscopic analyses. The target molecules OPV16 were synthesized by Horner–Wadsworth–Emmons reaction between aldehydes and phosphite esters. OPV16 were thoroughly characterized by 1H, 13C NMR spectroscopy and HRMS (see Supporting Information). The molecular structure of OPV2 was further characterized by single crystal X-ray diffraction analysis.

through the vacant p orbital of tricoordinate boron. These assignments are comparable with the related compounds reported for other dimesitylboryl phenothiazine-based compounds[30]. In emission studies, methylcyclohexane solutions of non-boron OPVs show the peak at 575 nm, and the boron containing OPVs show the peak at 545 nm, when excited at their longest absorption maxima (Figure 1, bottom). Furthermore, the emission properties of OPV16 were explored in different solvents with varying polarities (Table S1). Unlike absorption properties, the emission properties of these OPVs are notable and displayed significant solvatochromism with higher Stokes shifts. The boron containing OPVs showed higher emission quantum yields (f) in solvents of different polarity.

Optical Properties Methylcyclohexane (MCH) solutions of OPV16 showed two absorptions around 300-500 nm (Figures 1 and S1). The higher energy bands ( OPV5 > OPV6.

This article is protected by copyright. All rights reserved.

Accepted Manuscript

as it could be safely turned upside down (Figures 3a and b). The critical gel concentration (CGC) of OPV4 was determined in a wide range of solvent mixtures and the values are provided in Table S3. OPV4 forms gel in different type of organic solvent mixtures at a low CGC (0.8 wt %). The reversibility of the gel was found to occur through mechanical stimuli as the gel exhibit thixotropic behavior. For example, shaking of the vial containing gel for 2 min turns the gels into the free-flowing materials, which take 15-20 hour to reform the gel. On the other hand, OPV5 and OPV6 form only partial/weak or no gels in different solvents mixtures. This is presumably due to the presence of bulky substituents (dimesitylboryl and bromine moieties).

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER To describe and understand the crystal packing and its implications for supramolecular chemistry, the single crystal Xray structure analysis was undertaken. The solid-state X-ray structure of OPV2, shown in Figures 5 and S9, provides an insight into the possible types of intermolecular interactions present in the molecular unit cell. OPV2 shows dimeric - stacking units in the crystal structure (d-= 3.689 Å, slip angle 82.3°; d- is the distance between the geometrical centres of the aromatic systems). The packing order in its crystal structure implies that multiple arrangements are possible (Figure S9). This in turn results in non-gelating nature of these short alkyl chains OPVs.

assembly to form gelation. This can be inferred from the PXRD results in Figure S10, Supporting Information.

Figure 5. Crystal structure of OPV2 (Disordered groups removed for clarity).

Powder X-Ray Diffraction Analysis Powder X-ray diffraction experiment (PXRD) was performed in order to get the self-assembly pattern for the OPV4 molecular components during its gelation. Sharp peaks were observed in the small angle (2θ = 1-15 ) scattering region, implying high crystallinity and order with long range periodicity upon lamellar packing of the OPV molecules. The 2θ values are estimated as 2.65 (d100 = 33.2 Å), 5.54 (d200 = 15.9 Å), 6.95 (d300 = 12.7 Å), 9.99 (d400 = 8.84 Å) and 12.12 (d500 = 7.29 Å) with a reciprocal ratio of 1:2:3:4:5, signifying long range order for the selfassembled fibers. It is noteworthy to mention that although the system was not responding to single crystal stability upon several attempts, the long range lamellar packing due to significant molecular interactions resulted in crystallinity of the architecture in nanoscale region. In order to delve into the molecular interactions, responsible for such self-assembly mode, we performed the MD simulations for the OPV4 dimer upto 100 ns using optimized potential for liquid simulations (OPLS) force field parameters. The dimer exerted strong  interactions for the peripheral phenothiazine and central phenyl rings around 3.52 Å and 3.6 Å, respectively. This validates our experimental values of 3.4 Å (2θ= 25.7) (Figure 6) and 3.7 Å (2θ = 24.1), observed in the wide-angle diffraction region. Also, strong interactions of the hydrophobic long hexadecyl chains were reflected with minor interdigitation. Such single component assemblies with strong molecular interactions can manipulate several new functions. In the present case, the luminescence intensity in gel phase was augmented due to the well-defined packing mode of OPV’s in the assembly. The lameller packing of OPV4 molecules was retained even in other solvents, in which it underwent self-

Figure 6. a) X-ray diffraction profile for xerogel OPV4 (8 mg/mL), selfassembled from (THF/EtOH; 2/8). Dispersive stacking distances for OPV4 dimer, obtained after 100 ns of MD simulation, using OPLS force field parameters and b) Schematic represents the molecular packing paradigm leading to self-assembled crystalline nanofibers.

Electrochemical AC Impedance Spectroscopy (EIS studies) Thin films of thickness ~0.18 mm of OPV4 at various compositions were fabricated by blending them with polymer matrix, poly(methylmethacrylate) (PMMA). The high luminescence properties of OPV4 remained unperturbed in thin film state (Figure 7, top panel). The solid thin films were brittle in nature. The SEM images for the films made up of only PMMA revealed a highly transparent sheet type morphology (Figure S11). On the other hand, the blended PMMA films of OPV4 revealed reinforced spherical particles, suggesting a coassembly pattern (vide Figure 7b). The morphology of OPVs at a lessor doping concentration of 10 (Figure 7a) and 12.5 wt% (Figure 7b) in PMMA reveals clear, ordered and homogeneous spherical domains with distinct grain

This article is protected by copyright. All rights reserved.

Accepted Manuscript

Single Crystal X-Ray Analysis

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER

Figure 8. a) AC-impedance Nyquist plots (with Z-fit) for the thin films made up of 25 %, 16.6 %, 12.5 % and 10 % of OPV4 (wt % in PMMA) with PMMA (Inset shows the high frequency region) and b) the model equivalent circuit derived from the Nyquist plots for analysis purpose.

Figure 7. Top pictures: Thin films made up of different wt% of OPV4 dissolved in PMMA solution. Bottom pictures: a and b) SEM images of OPV4 at 12.5 wt% in PMMA and c, d) SEM images of OPV4 at 25 wt% in PMMA.

In general, PMMA is very less conductive in nature.[33] However, the presence of OPV4 resulted in the storage of electronic charge in the polymer matrix as well as their transportation under an applied AC potential. To understand the electronic carrier properties in such blended thin films, we have employed a non-destructive characterization technique known as electronic AC impedance spectroscopy (EIS). Figure 8a represents the Nyquist plots for all the films made by varying the concentration of OPV4 and maintaining the PMMA composition as constant. A model equivalent circuit is always indispensable in order to scrutinize the impedance results, which in the present case is manifested from the respective Nyquist plots shown, in Figure 8b. ‘R1’ and ‘R2’ in the circuit model denotes the resistances to ionic and electronic transport, respectively.[34] The resistance and capacitance values are estimated using the Z-fit program (EC lab, version 11.10) and are summarized in Table S5. The film samples are fitted with two resistances, R1 and R2, which are the resistances to ionic and electronic (bulk) charge transport, respectively.

A detailed analysis of the conductance, resistance and capacitance values suggest that the increase of OPV4 in the film by 25 % (by wt%) with respect to PMMA has amplified the electronic conductance {2.3 ( 0.3)  106 S} by four orders of magnitude as compared to that of pristine PMMA film {5.7 ( 0.2)  1010 S}. This is also accompanied by the simultaneous increase in the film’s charge storage abilities. The augmentation of carriers can be inferred from the increase in capacitance value for the film (18.6  0.004 F), where the charge storage ability is found to be 186 times higher as compared to pristine PMMA (0.1  0.003 F). The other probable factor that contributes here towards the increase in conductance and storage abilities for the films is the morphology, as inferred from the SEM studies (vide Figure 3). The SEM images of the thin films reveals strengthening of the assembly with gradually increasing the content of OPV4 in the polymer matrix. At the highest concentration, i.e., at 25% of OPV4, the film depicts almost a hierarchical assembly of thick fibers reinforcing the spherical domains. We believe that the formation of such reinforced assembly can potentially act as a channel in the gaps existing in between the spherical domains, amplifying the storage and transport abilities. The ionic conductance most likely predominates throughout the film with increase in concentration for OPV4, i.e. 25% of OPV4 in PMMA, giving rise to a very low resistance of 113.2  0.01  throughout the film. The decrease in electronic resistances also follow a similar trend (vide Table S5). This ionic conduction stems from the semi-infinite diffusion of ions at the electrodes in the high frequency region, giving rise to a straight line in the impedance spectrum and synergistically stimulates the electronic transport from the sample.

Conclusions In conclusion, we have synthesized highly fluorescent ICT active phenothiazine based oligo(p-phenylvinylene) derivatives, OPV16. The significant solvatochromism with higher Stokes shifts were also observed in the photoluminescence properties of these compounds. The solvatochromic shifts for non-boron compounds were higher than those containing the boron moieties due to the presence of sterically bulky substituents (BMes2). The emission quantum yields (f) for the boron containing compounds were relatively high. Among all the

This article is protected by copyright. All rights reserved.

Accepted Manuscript

boundaries. At a relatively greater concentration of 16.6 wt% (Figure 7c), the ordered spherical domains became closely packed at the surface of the transparent PMMA sheet. This indicates that the sheet morphology of PMMA acts as a support in order to hold these ordered spherical domains close to each other. At this stage, the grain boundaries disappeared. The assembly became stronger at 25 wt% of OPV4 in PMMA, revealing a dense network of closely packed eclipsed domains (Figure 7d) with almost no grain boundaries. The origin of such closely packed architectures is reasonable, which is already evident from the potential of OPV4 to exert gelation at higher concentration, as discussed in previous sections.

10.1002/chem.201801810

Chemistry - A European Journal

compounds prepared, the molecular components of OPV4 undergoes a monolayer lameller assembly revealing the formation of hiererchically ordered 1D fibers in a variety of organic solvent mixtures. Our theoretical calculations further substantiate the influnce of electronic nature of the hetero atoms in nullifying the gelation tendency for the compounds in the order of OPV4 > OPV5 > OPV6. The AC impedance studies for the fluorescent thin films of OPV4 with PMMA indicated an amplified charge storage ability with an increase in the content of OPV4. Such ICT and transport active luminogens can exert great potential in the field of organic optoelectronics, such as field effect transistors (FETs), light emitting diodes (LEDs) and nonlinear optics applications.

Experimental Section General Instrumentation The 1H, 13C{1H} NMR and 11B{1H} NMR were collected on a Bruker 400 MHz or 500 MHz spectrometers. Electrospray mass (ESI-MS) spectra were recorded on a Qtof Micro YA263 HRMS and Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) measurements were carried out on Bruker Ultra Flextreme instrument (MALDI-TOF) by using 2,5-dihydroxybenzoic acid as a matrix and a ground steel target plate. All photophysical measurements were made in standard quartz cuvettes (1 cm  1 cm). The UV/Vis spectroscopic studies were performed on a Jasco V-660 or V-750 Spectrophotometers. Fluorescence spectra were recorded using Jasco FP-6300 or Horiba Jobin Yvon Fluoromax-4 spectro fluorimeters. The scanning electron microscopic studies were carried out using a Hitachi S-4800. Single crystal X-ray data were collected using Bruker Kappa apex II CCD single crystal diffractometer. Powder-XRD patterns were recorded on a Bruker D8 Advance X-ray diffractometer using CuK radiation ( = 1.54178 Å). Materials Starting materials 10-ethyl-phenothiazine-3-carbaldehyde, 7-bromo-10ethyl-10H-phenothiazine-3-carbaldehyde, 1,4-bis(diethylphosphonato methyl)-2,5-dihexyloxybenzene, 1,4-bis(diethylphosphonatomethyl)-2,5dihexadecyloxybenzene were prepared according to literature procedures.[27-29] 3-bromo-7-(5,5-dimethyl-1,3-dioxan-2-yl)-10-ethyl-10H-phenothiazine 7-bromo-10-ethyl-10H-phenothiazine-3-carbaldehyde (12.50 g, 37.39 mmol), 2,2-methyl-1,3-propandiol (7.79 g ,74.79 mmol) and a catalytic amount of p-toluenesulfonic acid in toluene (150 mL) was refluxed for 4 h. The reaction mixture was allowed to reach room temperature, extracted with dichloromethane and washed with brine solution. The organic layer was dried over anhydrous Na2SO4 and the solvent was removed under vacuum. The crude product was purified by basic alumina column chromatography using 40% dichromethane in hexane yielding the product as colorless crystalline solid. Yield = 11.64 g (74 %). 1H NMR (400 MHz, CDCl3, ): 7.25-7.27 (m, 2H), 7.17-7.2 (m, 2H), 6.80 (d, 1H, J = 8.2 Hz), 6.65 (d, 1H, J = 9.24 Hz), 5.28 (s, 1H), 3.84 (q, 2H, J = 6.9 Hz), 3.73 (d, 2H, J = 11.1 Hz), 3.6 (d, 2H, J = 10.7 Hz), 1.35 (t, 3H, J = 6.9 Hz), 1.26 (s, 3H), 0.77 (s, 3H); 13C{1H} NMR (100 MHz, CDCl3, ): 144.8, 144.0, 133.2, 129.8, 129.4, 126.5, 125.3, 125.2, 123.6, 116.1, 114.8, 114.2, 100.94, 77.6, 41.9, 30.1, 23.0, 21.8, 12.7. 7-(dimesitylboranyl)-10-ethyl-10H-phenothiazine-3-carbaldehyde 3-bromo-7-(5,5-dimethyl-1,3-dioxan-2-yl)-10-ethyl-10H-phenothiazine (11.64 g, 27.68 mmol) was dissolved in THF (150 mL), This solution was cooled to -78 °C, and nBuLi (19 mL, 30.45 mmol, 1.6 M in hexane) was added dropwise. The reaction mixture was allowed to stir 30 min and then Mes2BF (8.16 g, 30.45 mmol) in THF (50 mL) was added at -78 °C

dropwise. The color of the solution changed pale yellow to bright green. The reaction mixture slowly allowed to RT and then stirred overnight. The reaction mixture was reduced to one third of its volume and treated with 5% hydrochloric acid solution (THF: 5% HCl; 1:2) heated to 70 °C for 1 hour. During which time, greenish yellow precipitate formed. The reaction mixture allowed to reach room temperature and neutralized using an aqueous solution of NaHCO3 and extracted with ether. In the separating funnel, green precipitate stayed on the orange color organic layer and clear aqueous layer was discarded. Orange color solution was discarded by decantation from top and the precipitate collected and washed three times with ether to yield the product as bright green powder. Column chromatography was not required as the product formed in good purity as checked by TLC and NMR. Yield = 6.35 g (45.5 %). 1H NMR (400 MHz, CDCl3, ): 9.78 (s, 1H), 7.62 (dd, 1H, J = 1.7, 8.5 Hz), 7.52 (d, 1H, J = 1.8 Hz), 7.3 (dd, 1H, J = 1.4, 8.2 Hz), 7.21 (d, 1H, J = 1.4 Hz), 6.91 (d, 1H, J = 8.4 Hz), 6.84 (s, 1H), 6.82 (s, 4H), 3.98 (q, 2H, J = 6.9 Hz), 2.31 (s, 6H), 2.02 (s, 12H), 1.46 (t, 3H, J = 6.9 Hz); 13C{1H} NMR (100 MHz, CDCl3, ): 190.0, 149.1, 146.3, 141.2, 140.7, 138.5, 137.2, 135.6, 131.3, 130.0, 128.2, 128.1, 124.3, 121.8, 114.6, 114.6, 42.7, 23.5, 21.2, 12.7; 11B{1H} NMR (160 MHz, CDCl , ): 78.8; HRMS: calcd. m/z for 3 C33H35BNOS+ [M+H]+, 504.2532; found, 504.2531. Synthesis of -Conjugated OPV16. Typical procedure is explained for OPV1 and other compounds OPV26 were synthesized following similar procedure. 10-ethyl-phenothiazine-3carbaldehyde (0.463 g, 1.815 mmol) and 1,4-bis(diethylphosphonato methyl)-2,5-dihexyloxybenzene (0.5 g, 0.885 mmol) were dissolved in dry THF (50 mL). The mixture was stirred under an argon atmosphere about 15 min at RT. Potassium tert-butoxide (0.297 g, 2.655 mmol) dissolved in THF was added dropwise to the reaction mixture under an argon atmosphere and the stirring was continued at RT for 24 h. The resultant solution was removed under high-vacuum. Workup was done using DCM/water. Organic phase was separated, dried on Na2SO4 and the solvent was removed by rotavapor. Finally, the compound was precipitated using methanol and washed many times with diethyl ether. Yield = 0.620 g (89 %). 1H NMR (400 MHz, CDCl3, ): 7.35 – 7.24 (m, 6H), 7.17 – 7.10 (m, 4H), 7.06 (s, 2H), 6.99 (d, J = 16.4 Hz, 2H), 6.93 – 6.79 (m, 6H), 4.03 (t, J = 6.5 Hz, 4H), 3.93 (q, J = 7.0 Hz, 4H), 1.91 – 1.81 (m, 4H), 1.55 (ddt, J = 15.1, 11.9, 5.8 Hz, 4H), 1.46 – 1.36 (m, 14H), 0.97 – 0.90 (m, 6H); 13C NMR (100 MHz, CDCl3, ): 151.1, 144.7, 144.1, 132.7, 127.5, 127.4, 126.8, 125.9, 125.1, 124.5, 123.9, 122.5, 121.9, 115.1, 110.5, 69.7, 42.0, 31.8, 29.6, 26.1, 22.8, 14.3, 13.1; HRMS calcd. for C50H57N2O2S2 [M+H]+, 781.3861; found, 781.3845. Synthesis of OPV2 7-bromo-10-ethyl-10H-phenothiazine-3-carbaldehyde (0.606 g, 1.815 mmol) was reacted with 1,4-bis(diethylphosphonatomethyl)-2,5dihexyloxybenzene (0.5 g, 0.885 mmol) using potassium tert-butoxide (0.297 g, 2.655 mmol) in dry THF as described for compound 1. Yield = 0.716 g (86 %). 1H NMR (400 MHz, CDCl3, ): 7.35 – 7.25 (m, 6H), 7.25 – 7.20 (m, 4H), 7.06 (s, 2H), 6.99 (d, J = 16.4 Hz, 2H), 6.82 (d, J = 8.3 Hz, 2H), 6.72 – 6.67 (m, 2H), 4.03 (t, J = 6.5 Hz, 4H), 3.89 (q, J = 7.0 Hz, 4H), 1.87 (dq, J = 8.6, 6.5 Hz, 4H), 1.58 – 1.50 (m, 4H), 1.41 (td, J = 7.1, 4.0 Hz, 14H), 0.98 – 0.90 (m, 6H); 13C NMR (100 MHz, CDCl3, ): 151.1, 143.9, 143.7, 133.0, 130.0, 129.7, 127.3, 126.8, 126.3, 126.1, 125.1, 123.8, 122.3, 116.2, 115.2, 114.5, 110.6, 69.7, 42.2, 31.8, 29.6, 26.1, 22.8, 14.3, 13.0; HRMS calcd for C50H55Br2N2O2S2 [M+H]+, 937.2071; found, 937.2068. Synthesis of OPV3 7-(dimesitylboranyl)-10-ethyl-10H-phenothiazine-3-carbaldehyde (0.914 g, 1.815 mmol) was reacted with 1,4-bis(diethylphosphonatomethyl)-2,5dihexyloxybenzene (0.5 g, 0.885 mmol) using potassium tert-butoxide (0.297 g, 2.655 mmol) in dry THF as described for compound 1. Yield = 0.987 g (87 %). 1H NMR (400 MHz, CDCl3, ): 7.33 – 7.27 (m, 5H), 7.24 (td, J = 5.1, 4.6, 2.1 Hz, 5H), 7.06 (s, 2H), 6.98 (d, J = 16.3 Hz, 2H), 6.85 – 6.76 (m, 12H), 4.03 (t, J = 6.5 Hz, 4H), 3.96 (q, J = 6.9 Hz, 4H), 2.31 (s, 12H), 2.03 (s, 24H), 1.86 (p, J = 6.6 Hz, 4H), 1.54 – 1.49 (m, 4H), 1.44 (t, J = 6.9 Hz, 6H), 1.39 (tt, J = 6.0, 2.7 Hz, 8H), 0.95 – 0.89 (m, 6H); 13C

This article is protected by copyright. All rights reserved.

Accepted Manuscript

FULL PAPER

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER

Synthesis of OPV4 10-ethyl-phenothiazine-3-carbaldehyde (0.310 g, 1.213 mmol) was reacted with 1,4-bis(diethylphosphonatomethyl)-2,5dihexadecaloxybenzene (0.5 g, 0.592 mmol) using potassium tertbutoxide (0.199 g, 1.776 mmol) in dry THF as described for compound 1. Yield = 0.548 g (87 %). 1H NMR (400 MHz, CDCl3, ): 7.34 – 7.24 (m, 6H), 7.16 – 7.09 (m, 4H), 7.06 (s, 2H), 6.99 (d, J = 16.5 Hz, 2H), 6.86 (ddd, J = 22.6, 14.3, 7.9 Hz, 6H), 4.02 (t, J = 6.5 Hz, 4H), 3.98 – 3.88 (m, 4H), 1.86 (p, J = 6.8 Hz, 4H), 1.52 (d, J = 7.9 Hz, 4H), 1.42 (t, J = 6.9 Hz, 6H), 1.24 (br.s, 44H), 0.87 (t, J = 6.6 Hz, 6H); 13C NMR (100 MHz, CDCl3, ): 151.1, 144.1, 132.7, 127.5, 127.4, 126.9, 125.9, 125.1, 122.5, 122.0, 115.1, 110.6, 69.7, 42.0, 32.1, 29.9, 29.8, 29.7, 29.5, 26.5, 22.8, 14.3, 13.1; HRMS calcd for C70H97N2O2S2 [M+H]+, 1061.6991; found, 1061.6995. Synthesis of OPV5 7-bromo-10-ethyl-10H-phenothiazine-3-carbaldehyde (0.405 g, 1.213 mmol) was reacted with 1,4-bis(diethylphosphonatomethyl)-2,5dihexadecaloxybenzene (0.5 g, 0.592 mmol) using potassium tertbutoxide (0.199 g, 1.776 mmol) in dry THF as described for compound 1. Yield = 0.645 g (89 %). 1H NMR (400 MHz, CDCl3, ): 7.36 – 7.18 (m, 12H), 7.06 (s, 2H), 7.00 (d, J = 16.4 Hz, 2H), 6.82 (d, J = 8.4 Hz, 2H), 6.69 (d, J = 9.3 Hz, 2H), 4.03 (t, J = 6.5 Hz, 4H), 3.89 (q, J = 7.0 Hz, 4H), 1.87 (p, J = 6.7 Hz, 4H), 1.53 (td, J = 9.4, 8.6, 4.6 Hz, 4H), 1.41 (t, J = 7.0 Hz, 6H), 1.25 (br.s, 44H), 0.88 (t, J = 6.8 Hz, 6H); 13C NMR (100 MHz, CDCl3, ): 151.1, 143.9, 143.7, 133.0, 130.0, 129.7, 127.4, 126.8, 126.3, 126.1, 125.1, 123.8, 122.3, 116.2, 115.2, 114.5, 110.6, 69.7, 42.2, 32.1, 29.9, 29.8, 29.7, 29.5, 26.5, 22.8, 14.3, 13.0; MS calcd for C70H95Br2N2O2S2 [M+H]+, 1217.5; found, 1218.2. Synthesis of OPV6 7-(dimesitylboranyl)-10-ethyl-10H-phenothiazine-3-carbaldehyde (0.611 g, 1.213 mmol) was reacted with 1,4-bis(diethylphosphonatomethyl)-2,5dihexadecaloxybenzene (0.5 g, 0.592 mmol) using potassium tertbutoxide (0.199 g, 1.776 mmol) in dry THF as described for compound 1. Yield = 0.793 g (86 %). 1H NMR (400 MHz, CDCl3, ): 7.33 – 7.21 (m, 10H), 7.06 (s, 2H), 6.99 (d, J = 16.3 Hz, 2H), 6.86 – 6.77 (m, 12H), 4.03 (t, J = 6.5 Hz, 4H), 3.96 (q, J = 6.9 Hz, 4H), 2.31 (s, 12H), 2.03 (s, 24H), 1.91 – 1.81 (m, 4H), 1.55 – 1.49 (m, 4H), 1.45 (t, J = 7.0 Hz, 6H), 1.25 (br.s, 44H), 0.91 – 0.83 (m, 6H); 13C NMR (100 MHz, CDCl3, ): 151.1, 148.0, 142.8, 141.6, 140.9, 138.4, 137.4, 135.9, 128.2, 127.4, 126.8, 125.7, 125.1, 124.1, 122.3, 122.2, 115.2, 114.1, 110.6, 69.7, 32.1, 29.9, 29.8, 29.8, 29.6, 29.6, 29.5, 26.4, 23.6, 22.8, 21.3, 14.3, 13.0; MS calcd for C106H139B2N2O2S2 [M+H]+, 1558.0; found, 1558.7. Computational Details All calculations (DFT and TD-DFT) were carried out with the program package Gaussian 09 (Rev. C. 01)[35] and were performed on a parallel cluster system. The electrostatic potentials (ESP) were calculated using the Coulomb-attenuated functional CAM-B3LYP fuctional[36] in combination with the cc-pVDZ basis set.[37] The lowest-energy vertical transitions were calculated by TD-DFT, using long range dispersion corrected wB97X[38] functional in combination with the cc-pVDZ basis set. TD-DFT results were extracted using GaussSum 3.0 software.[39] Morphology To record SEM images of gels, smaller portions of the gels were mounted over a glass cover slip and were dried in air one day before the experiment. SEM images were acquired with an FE-SEM instrument (Model: HITACHI S 4800) upond the gold sputter coated samples. All the

samples before SEM were subjected to gold sputter coating for 30 seconds using a HITACHI E-1010 ion sputter. Powder X-Ray Diffraction (PXRD) Studies The gel placed on the glass cover slips and dried for more than 48 hours. Powder-XRD patterns were recorded on a Bruker D8 CuK radiation ( = 1.54178 Å). Single Crystal X-Ray Diffraction Analysis X-ray diffraction data for the crystal OPV2 collected using Bruker Kappa apexII CCD single crystal diffractometer, equipped with graphite monochromated Mo Kα (λ=0.71078Å) radiation. Data collection was carried out at 296 K using ω-φ scan modes. The collected frames were integrated followed by Lorentz and Polarization correction using the program SAINT-APEXII software. Multi-scan absorption correction has been employed for the data using SADABS program. The molecular structure was solved by direct methods procedure using SHELXT2014/7.[40] Initially isotropic refinements of non-hydrogen atoms were carried out followed by full-matrix least squares refinement with anisotropic thermal parameters for non-hydrogen atoms through SHELXL-2014/7 program. The hydrogen atoms associated with the carbon atoms were identified from the difference electron density map and were allowed to ride on the parent atom using suitable constraint, with distance 0.93Å(for aromatic CH) and 0.96Å(CH3) and thermal displacement of Ueq(H) = 1.2Uiso(C) and Ueq(H) = 1.5Uiso(C) respectively. All the interactions and molecular drawings were obtained using the program Olex2, ORTEP3 and Mercury (ver. 3.9).[41-43] Electrochemical AC Impedance Measurements Impedance spectroscopy of the films was recorded using Bio-Logic SP150 instrument at 25 ºC over the frequency range from 100 kHz to 10 mHz with an AC perturbation of 10 mV at the 0V DC level. The Nyquist plots obtained were then fitted to obtain the resistance and capacitance values using the Z-fit program (EC lab, version 11.10). The electrical conductance was calculated using the following equation. G = l/R. Where, ‘G’ is the electronic conductance and ‘R’ is the resistance, selected exclusively from the electronic contribution (no ionic contribution) only.

Acknowledgements The authors are grateful to IIT Madras for financial support through Exploratory Research Project (ERP): CHY/1617/845/RFER/SGHO. The authors thank department of chemical engineering, IIT Madras for providing FE-SEM facilities. SAIF, IIT Madras is gratefully acknowledged for single crystal X-ray diffraction analysis. P.G Senapathy Centre for computing resources, IIT Madras for providing us oppertunity to utilize their supercomputing facilities for performing all the DFT calculations throughout this work. C. A., S. S. and P. M are grateful to CSIR (Counsil of Scientific and Industrial Research), UGC (University Grants Commission) and IIT Madras, respectively for senior research fellowships. Keywords: oligo(p-phenylene vinylene)s • organogels • phenothiazine • self-assembly • supramolecular chemistry [1]

a) M. B. Avinash, T. Govindaraju, Acc. Chem. Res. 2018, 51, 414426; b) C.-F. Liu, M. Sang, W.-Y. Lai, T.-T. Lu, X. Liu, W. Huang, Macromolecules 2018, 51, 13251335; c) R. Rajamalli, D. R. Martir, E. Zysman-Colman, ACS Appl. Energy Mater. 2018, 1, 649654; d) Y. Zhao, H. Wang, W. Zeng, S. Xia, F. Zhou, H. Chen, F. He, C. Yang, J. Mater. Chem. A 2018, 6, 81018108; e) C. McDowell, K. Narayanaswamy, B. Yadagiri, T. Gayathri, M. Seifrid, R. Datt, S. M. Ryno, M. C. Heifner, V. Gupta, C. Risko, S. P. Singh, G. C. Bazan, J.

This article is protected by copyright. All rights reserved.

Accepted Manuscript

NMR (100 MHz, CDCl3, ): 151.1, 148.0, 142.8, 141.6, 140.9, 139.4, 138.4, 137.4, 135.9, 133.2, 128.2, 127.3, 126.8, 125.8, 125.0, 124.1, 122.3, 122.1, 115.2, 114.1, 110.5, 69.7, 42.4, 31.8, 29.6, 26.1, 23.6, 22.8, 21.3, 14.2, 13.0; HRMS calcd for C86H99B2N2O2S2 [M+H]+, 1277.7334; found, 1277.7324.

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER

[3]

[4]

[5]

[6]

[7]

[8] [9] [10]

[11]

[12]

[13]

[14] [15]

[16] [17]

F. Chen, Q. Chen, L. Zhu, Z. Tang, Q. Li, G. Qin, J. Yang, Y. Zhang, B. Ren, J. Zheng, Chem. Mater. 2018, 30, 17431754.

[18] [19] [20]

[21]

[22]

[23]

[24]

[25]

[26]

[27] [28] [29] [30] [31]

[32] [33] [34]

A. R. Hirst, B. Escuder, J. F. Miravet, D. K. Smith, Angew. Chem. Int. Ed. 2008, 47, 80028018; Angew. Chem. 2008, 120, 81228139. S. Zhang, C. E. Semino, Adv. Exp. Med. Biol. 2003, 534, 147163. a) K. Takimiya, M. Nakano, Bull. Chem. Soc. Jpn. 2018, 91, 121140; b) Y. Zhang, M. Li, H. Feng, W. Ni, H. Zhang, F. Liu, X. Wan, Y. Chen, Dyes Pigm. 2017, 141, 262268. a) S. Yagai, M. Gushiken, T. Karatsu, A. Kitamura, Y. Kikkawa, Chem. Commun. 2011, 47, 454456; b) S. Yagai, K. Ohta, M. Gushiken, K. Iwai, A. Asano, S. Seki, Y. Kikkawa, M. Morimoto, A. Kitamura, T. Karatsu, Chem. Eur. J. 2012, 18, 22442253. a) A. Ajayaghosh, V. K. Praveen, Acc. Chem. Res. 2007, 40, 644656; b) S. J. George, A. Ajayaghosh, Chem. Eur. J. 2005, 11, 32173227; c) V. K. Praveen, S. J. George, A. Ajayaghosh, Macromol. Symp. 2006, 241, 18; d) S. J. George, A. Ajayaghosh, P. Jonkheijm, A. P. H. J. Schenning, E. W. Meijer, Angew. Chem. Int. Ed. 2004, 43, 34223425; Angew. Chem. 2004, 116, 35043507; e) S. K. Samanta, S. Scherf, Macromol. Chem. Phys. 2017, 218, 1600374; f) C. Yao, Q. Lu, X. Wang, F. Wang, J. Phys. Chem. B 2014, 118, 46614668. a) K. Cantin, S. Rondeau-Gagné, J. R. Néabo, M. Daigle, J.-F. Morin, Org. Biomol. Chem. 2011, 9, 44404443; b) T. Ide, D. Taeuchi, K. Osakada, Chem. Commun. 2012, 48, 278280; c) S. Roy, D. Samanta, P. Kumar, T. K. Maji, Chem. Commun. 2018, 54, 275278; d) S. Kushida, E. Smarsly, L. Veith, I. Wacker, R. R. Schröder, U. H. F. Bunz, Macromolecules 2017, 50, 78807886. a) B.-K. An, D.-S. Lee, J.-S. Lee, Y.-S. Park, H.-S. Song, S. Y. Park, J. Am. Chem. Soc. 2004, 126, 1023210233; b) P. Xue, B. Yao, J. Ding, Y. Shen, P. Wang, R. Lu, X. Zhao, ChemistrySelect 2017, 2, 28412846; c) H. Kim, J. Y. Chang, Langmuir 2014, 30, 1367313679; d) D. Kim, J. E. Kwon, S. Y. Park, Adv. Funct. Mater. 2018, 1706213; e) M. Martinex-Abadía, R. Giménez, M. B. Ros, Adv. Mater. 2018, 30, 1704161; f) Y. Ren, R. Zhang, C. Yan, T. Wang, H. Cheng, X. Cheng, Tetrahedron 2017, 73, 52535259. a) V. Sidorov, S. M. Dzekunov, D. Abdallah, B. Ghebremariam, P. D. Roepe, S. Matile, Chem. Commun. 1999, 14291430; b) J.-H. Ryu, D.J. Hong, M. Lee, Chem. Commun. 2008, 10431054; c) Z. Huang, H. Lee, E. Lee, S.-K. Kang, J.-M. Nam, M. Lee, Nat. Commun. 2011, 2, 459. a) C. Vijayakumar, V. K. Praveen, A. Ajayaghosh, Adv. Mater. 2009, 21, 20592063; b) A. Ajayaghosh, C. Vijayakumar, V. K. Praveen, S. S. Babu, R. Varghese, J. Am. Chem. Soc. 2006, 128, 71747175; c) A. Ajayaghosh, S. J. George, V. K. Praveen, Angew. Chem. Int. Ed. 2003, 42, 332335; Angew. Chem. 2003, 115, 346349; d) A. Ajayaghosh, V. K. Praveen, C. Vijayakumar, S. J. George, Angew. Chem. Int. Ed. 2007, 46, 62606265; Angew. Chem. 2007, 119, 63766381; e) A. Ajayaghosh, V. K. Praveen, S. Srinivasan, R. Varghese, Adv. Mater. 2007, 19, 411415; f) S. Bhattacharjee, S. Bhattacharya, Langmuir 2016, 32, 42704277; g) J. Kumpf, J. Freudenberg, K. Fletcher, A. Dreuw, U. H. F. Bunz, J. Org. Chem. 2014, 79, 66346645. Z. Iqbal, W.-Q. Wu, H. Zhang, P.-L. Hua, X. Fang, D.-Bin Kuang, L. Wang, H. Meier, D. Cao, Dyes Pigm. 2013, 99, 299307. F. M. Lucas, J. A. Labastide, L. Wei, J. S. Tinkham, M. D. Barnes, P. M. Lahti, J. Phys. Chem. A 2015, 119, 80108020. M. R. Pinto, B. Hu, F. E. Karasz, L. Akcelrud, Polymer 2000, 41, 80958102. C. Arivazhagan, A. Maity, K. Bakthavachalam, A. Jana, S. K. Panigrahi, E. Suresh, A. Das, S. Ghosh, Chem. Eur. J. 2017, 23, 70467056. a) Y. Jeong, K. Hanabusa, H. Masunaga, I. Akiba, K. Miyoshi, S. Sakurai, K. Sakurai, Langmuir 2005, 21, 586594; b) J. Brinksma, B. L. Feringa, R. M. Kellogg, R. Vreeker, J. van Esch, Langmuir 2000, 16, 92499255. A. K. Chaudhari, I. Han, F.-C. Tan, Adv. Mat. 2015, 27, 44384446. M. Bharati, S. Dutt, R. Raturi, V. Joshi, IOP Conf. Ser.: Mater. Sci. Eng. 2017, 225, 012093. a) P. Chakraboraty, S. Das, S. Mondal, A. K. Nandi, CrystEngComm. 2015, 17, 80938104; b) P. Chakraborty, P. Bairi, S. Mondal, A. K. Nandi, J. Phys. Chem. B 2014, 118, 1396913980; c) P. Chakraborty, P. Bairi, B. Roy, A. K. Nandi, ACS Appl. Mater. Interfaces 2014, 6, 36153622.

This article is protected by copyright. All rights reserved.

Accepted Manuscript

[2]

Mater. Chem. A 2018, 6, 383394; f) Z. Hu, B. Shao, G. T. Geberth, D. A. V. Bout, Chem. Sci. 2018, 9, 11011111; g) J. Min, X. Jiao, V. Sgobba, B. Kan, T. Heumuller, S. Rechberger, E. Spiecker, D. M. Guldi, X. Wan, Y. Chen, H. Ade, C. J. Brabec, Nano Energy 2016, 28, 241249. a) Z. Hu, H. Chen, F. Qu, X. Zhong, P. Chao, M. Xie, W. Lu, A. Liu, L. Tian, Y.-A. Su, W. Chen, F. He, ACS Energy Lett. 2017, 2, 753758; b) A. Maity, A. Dey, M. Gangopadhyay, A. Das, Nanoscale 2018, 10, 14641473; c) R. Singh, J. Lee, M. Kim, P. E. Keivanidis, K. Cho, J. Mater. Chem. A 2017, 5, 210220. a) S. Bhattacharyya, B. Jana, S. Sain, M. K. Barman, S. K. Pradhan, A. Patra, Small 2015, 11, 63176324; b) M. Adinehnia, B. Borders, M. Ruf, B. Chilukuri, K. W. Hipps, U. Mazur, J. Mater. Chem. C 2016, 4, 1022310239. a) Y. Li, Y. Liu, T. Gao, B. Zhang, Y. Song, J. L. Terrell, N. Barber, W. E. Bentley, I. Takeuchi, G. F. Payne, Q. Wang, ACS Appl. Mater. Interfaces 2015, 7, 1058710598; b) X.-P. Liu, L.-W. Xue, Q. Wei, M. Liang, K. Deng, Z.-J. Zhang, P. Jiang, J. Phys. Chem. C 2016, 120, 1860718615; c) F. Ishiwari, Y. Shoji, T. Fukushima, Chem. Sci. 2018, 9, 20282041; d) Y. Peng, Y. Feng, G.-J. Deng, Y.-M. He, Q.-H. Fan, Langmuir 2016, 32, 93139320. a) S. Valtera, J. Prokesš, J. Kopecká, M. Vrňata, M. Trchova, M. Varga, J. Stejskal, D. Kopecký, RSC Adv. 2017, 7, 5146551505; b) S. P. Goskulwad, D. D. La, R. S. Bhosale, M. A. Kobaisi, L. A. Jones, S. V. Bhosale, S. V. Bhosale, ChemistrySelect 2018, 3, 14601465; c) X. Zhang, R. Lu, J. Jia, X. Liu, P. Xue, D. Xu, H. Zhou, Chem. Commun. 2010, 46, 84198421; d) C. Rest, R. Kandanelli, G. Fernandez, Chem. Soc. Rev. 2015, 44, 25432572; e) A. Maity, F. Ali, H. Agarwalla, B. Anothumakkool, A. Das, Chem. Commun. 2015, 51, 21302133. a) K. Shi, W. Zhang, D. Gao, S. Zhang, Z. Lin, Y. Zou, L. Wang, G. Yu, Adv. Mater. 2018, 30, 1705286; b) M. Pandey, S. Nagamatsu, W. Takashima, S. S. Pandey, S. Hayase, J. Phys. Chem. C 2017, 121, 1118411193. B. Yang, J. Xiao, J. I. Wong, J. Guo, Y. Wu, L. Ong, L. L. Lao, F. Boey, H. Zhang, H. Y. Yang, Q. Zhang, J. Phys. Chem. C 2011, 115, 79247927. J. Xiao, Z. Yin, B. Yang, Y. Liu, L. Ji, J. Guo, L. Huang, X. Liu, Q. Yan, H. Zhang, Q. Zhang, Nanoscale 2011, 3, 47204723. G. Zotti, B. Vercelli, A. Berlin, Acc. Chem. Res. 2008, 41, 10981109. a) P. Terech, R. G. Weiss, Chem. Rev. 1997, 97, 31333159; b) S. S. Babu, V. K. Praveen, A. Ajayaghosh, Chem. Rev. 2014, 114, 19732129. a) A. Von Lipowitz, Liebigs Ann. Chem. Pharm. 1841, 38, 348355; b) Y.-C. Lin, B. Kachar, R. G. Weiss, J. Am. Chem. Soc. 1989, 111, 55425551; c) A. Ajayaghosh, S. J. George, J. Am. Chem. Soc. 2001, 123, 51485149; d) J. H. van Esch, B. L. Feringa, Angew. Chem. Int. Ed. 2000, 39, 22632266; Angew. Chem. 2000, 112, 23512354. a) D. M. Zurcher, A. J. McNeil, J. Org. Chem. 2015, 80, 24732478; b) Y. Lan, M. G. Corradini, R. G. Weiss, S. R. Raghavan, M. A. Rogers, Chem. Soc. Rev. 2015, 44, 60356058; c) G. Yu, X. Yan, C. Han, F. Huang, Chem. Soc. Rev. 2013, 42, 66976722; d) J. W. Steed, Chem. Commun. 2011, 47, 13791383. a) N. Zanna, S. Focaroli, A. Merlettini, L. Gentilucci, G. Teti, M. Falconi, C. Tomasini, ACS Omega 2017, 2, 23742381; b) Y. Ohsedo, Gels 2016, 2, 13. J. Nanda, A. Biswas, B. Adhikari, A. Banerjee, Angew. Chem. Int. Ed. 2013, 52, 50415045; Angew. Chem. 2013, 125, 51455149. a) T. Tu, W. Fang, Z. Sun, Adv. Mater. 2013, 25, 53045313; b) S. Dong, B. Zheng, D. Xu, X. Yan, M. Zhang, F. Huang, Adv. Mater. 2012, 24, 31913195; c) H. Chen, Y. Feng, G.-J. Deng, Z.-X. Liu, Y.-M. He, Q.-H. Fan, Chem. Eur. J. 2015, 21, 1101811028; d) B. Hu, Y. Shen, J. Adamcik, P. Fischer, M. Schneider, M. J. Loessner, R. Mezzenga, ACS Nano 2018, 12, 33853396; e) Y. Feng, H. Chen, Z.-X. Liu, Y.-M. He, Q.-H. Fan, Chem. Eur. J. 2016, 22, 49804990. S. M. M. Reddy, P. Dorishetty, G. Augustine, A. P. Deshpande, N. Ayyadurai, G. Shanmugam, Langmuir 2017, 33, 1350413514.

Chemistry - A European Journal

10.1002/chem.201801810

[36] [37] [38] [39] [40] [41] [42] [43]

M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, Gaussian, Inc., Wallingford CT, 2010. T. Yanai, D. Tew, N. Handy, Chem. Phys. Lett. 2004, 393, 5157. T. H. Dunning Jr., J. Chem. Phys. 1989, 90, 100723. J.-D. Chai, M. Head-Gordon, J. Chem. Phys. 2008, 128, 084106. N. M. O’Boyle, A. L. Tenderholt, K. M. Langner, J. Comput. Chem. 2008, 29, 839845. G. M. Sheldrick, SHELXL-2014/7, Germany, 2014. O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard, H. Puschmann, J. Appl. Cryst. 2009, 42, 339341. L. J. Farrugia, J. Appl. Cryst. 2012, 45, 849854. C. F. Macrae, P. R. Edgington, P. McCabe, E. Pidcock, G. P. Shields, R. Taylor, M. Towler, J. van de Streek, J. Appl. Cryst. 2006, 39, 453457.

This article is protected by copyright. All rights reserved.

Accepted Manuscript

FULL PAPER [35]

10.1002/chem.201801810

Chemistry - A European Journal

FULL PAPER Entry for the Table of Contents (Please choose one layout)

FULL PAPER C. Arivazhagan, Sitakanta Satapathy, Arijit Jana, Partha Malakar, Edamana Prasad* and Sundargopal Ghosh*

Series of phenothiazine based oligo(phenylenevinylene) derivatives show intriguing optical properties with intramolecular charge-transfer characteristics. The Long alkyl chain OPVs without bulky substituents at periphery led gelation in a mixture of organic solvents. The gels undergo monolayer lamellar assembly revealing the formation of hierarchically ordered 1D fibers. The electrochemical AC impedance studies for the fluorescence thin films of gelator with PMMA shows an amplified charge storage ability.

Phenothiazine Based Oligo(pphenylenevinylene)s: Substituents Affected Self-Assembly, Optical Properties and Morphology Induced Transport

This article is protected by copyright. All rights reserved.

Accepted Manuscript

Page No. – Page No.