Skip to content
Publicly Available Published by De Gruyter July 1, 2015

Photophysical, electrochemical, and spectroelectrochemical investigation of electronic push–pull benzothiadiazole fluorophores

  • Monika Wałęsa-Chorab , Marie-Hélène Tremblay , Mohamed Ettaoussi and William G. Skene EMAIL logo

Abstract

Two electronic push–pull fluorophores consisting of a benzothiadiazole core and a terminal N,N-dimethylamino electron donating group were prepared. The effect of the terminal electron withdrawing group (–NO2 and –CN) on the spectroscopic, electrochemical, and spectroelectrochemical properties were examined. The fluorophores were solvatochromic with Stokes shifts upward of 9000 cm−1 being observed in aprotic solvents of varying polarity. It was found that the fluorophores fluoresced appreciably (86 % > Φfl > 18 %) in hexane, toluene, diethyl ether, dichloromethane, ethyl acetate, and THF. The fluorescence was quenched in acetonitrile, acetone, and DMSO. The fluorophores also fluoresced appreciably in thin films when embedded in poly(methyl methacrylate) and poly(dimethylsiloxane) matrices. Protonating the terminal amine with trifluoroacetic acid quenched the intramolecular charge transfer band, although the fluorophores remained fluorescent. The fluorophores could also be reversibly oxidized electrochemically. The resulting oxidized state was visually different than the neutral form. Electrochemical oxidation also led to reversible changes in both the fluorescence wavelength and intensity.

Introduction

Benzothiadiazole (BZT) has been widely used in an array of applications [1–3]. This is in part due to its intrinsic electron accepting character that gives rise to interesting opto-electronic properties. These properties are enhanced when BZT is introduced between electron rich neighbors. The resulting π-conjugated donor–acceptor–donor arrangement results in BZT derivatives having combined optical and electrochemical properties that are well-suited for their use as active functional materials in plastic electronics [4, 5], including electrochromic devices [6], organic field effect transistors [7–9], and solar cells [10–14]. BZT derivatives are also known as active components that signal pathogen attack [15] and stimulate defense reactions of plants [16].

Among the many useful properties of BZT, is its intense fluorescence [17]. Its inherent high fluorescence yield has proven beneficial for emitting applications such as organic light emitting diodes [18–20]. The intense fluorescence of BZT derivatives has also found uses in bioimaging [21] and for nucleic base molecular recognition in DNA sensing [22]. The versatility of BZT as a fluorophore is in part owing to its emission color that can be adjusted contingent on the aromatics substituted in the 4,7′-positions [23]. Emission tailoring is also possible by conjugating BZT with various electron donating and withdrawing groups. These groups perturb the excited state, resulting in intramolecular charge transfer [2] and fluorescence solvatochromism [24].

N,N-Dimethylamino and nitro groups are typically used as complementary electron donating and withdrawing functionalities, respectively, for solvatochromic push–pull fluorophores. However, the nitro group can be problematic. It can significantly quench the fluorescence by many deactivation modes, including photoinduced electron transfer, aryl–NO2 bond rotation, and solvent–solute interactions [25–27]. It would be therefore advantageous to design a fluorophore whose fluorescence yield would be consistent in a wide range of solvents, yet whose emission color would dependent on its environment. These are ideal features of a universal fluorescence probe that could potentially span a wide range of applications from biological sensors to polymer deformation probes.

Given their similar electron affinity, the nitrile group is a potential viable replacement for the nitro group. We were therefore motivated to investigate the electronic push–pull analogues 1 and 2 (Chart 1) to better understand the effect of the similar electron withdrawing groups on the photophysical properties. Although BZT derivatives have been extensively examined, to the best of our knowledge, 1 and 2 have not been previously investigated and their collective spectroscopic and electrochemical properties remain unexplored. Bearing this mind, the spectroscopic, spectroelectrochemical, and electrochemical properties of the similar BZT fluorophores are herein presented to evaluate these similar electronic push–pull derivatives as universal fluorescence probes.

Chart 1: 
          Fluorophores examined.
Chart 1:

Fluorophores examined.

Experimental section

Spectroscopy

Electronic absorption spectra were measured with a Varian Cary 500 UV-vis-NIR spectrometer. Fluorescence measurements were made with an Edinburgh Instruments FLSP-920 spectrometer. Solutions for fluorescence and quantum yields measurements were prepared having ≤0.1 at the corresponding absorption maxima and they were purged with nitrogen for at least 20 min. Absolute quantum yields were calculated using an integrating sphere that was calibrated against anthracene in ethanol (Φfl = 28 %) [28]. Quantum yields of thin films were measured with an integrating sphere using a front face excitation/emission geometry.

Transient absorption measurements

Laser flash photolysis measurements were done with an Edinburgh Instrument LP920 system exciting at the corresponding absorption maxima with a Continuum Horizon I OPO pumped with a Surelite-I Nd-YAG laser. The transient absorption signals were collected with a monochromator and a Princeton Instruments ICCD camera that were corrected for the background and sample fluorescence. The solutions were prepared having ∼0.4 at the corresponding absorption maxima and they were purged with nitrogen for 30 min. The transient absorption spectra were generated by averaging the maximum absorption over 10 shots per wavelength that were recorded at 10 ns intervals after the laser pulse.

Electrochemistry

Electrochemical measurements were performed on a multi-channel BioLogic VSP potentiostat. Compounds were dissolved in anhydrous and deaerated dichloromethane at 10−4 M with 0.1 M tetrabutylammonium hexafluorophosphate (TBAPF6) as the electrolyte. A platinum electrode was used as the working electrode with a platinum wire as the auxiliary electrode. The reference electrode was a silver wire electrode. Ferrocene was added to the solution as an internal reference and the voltammograms were calibrated against its reversible Fc/Fc+ redox couple (Eo = 0.46 V vs. SCE) in CH2Cl2 with 0.1 M TBAPF6 as electrolyte [29].

Spectroelectrochemistry

Spectroelectrochemical measurements were done using a multi-channel BioLogic VSP potentiostat connected to a Varian Cary 500 UV-vis-NIR spectrometer. The fluorescence changes with different applied potentials were measured using an Edinburgh Instruments FLSP-920 spectrometer connected to a potentiostat. The sample was rotated by 45° relative to the excitation beam. Samples were prepared by dissolving the compounds in anhydrous and deaerated dichloromethane with 0.1 M TBAPF6 as the electrolyte. A platinum mesh working electrode, platinum wire auxiliary electrode, and silver wire reference electrode were used.

Polymer thin films and encapsulation

Thin films were prepared by drop-casting dichloromethane solutions of the fluorophores onto glass microscope slides. The thicknesses of the as-cast films were 30 nm for 1 and 35 nm for 2. Encapsulation of the fluorophores in a PMMA matrix (Mn = 70 000 g/mol) was done by dissolving the fluorophores and PMMA in 1:10 and 1:100 weight ratios in dichloromethane. The solutions were drop-cast onto glass slides. The thin films formed had the following thickness: 200 nm (1:10) and 250 nm (1:100) for 1; 180 nm (1:10) and 210 nm (1:100) for 2. PDMS films were prepared by mixing 2.02 g of Sylgard® 184 silicone elastomer base, 0.20 g of Sylgard® 184 silicone elastomer curing agent with 2 mg of the fluorophore dissolved in 0.5 mL of dichloromethane. The mixture was placed in a 5 × 5 cm Teflon mold and cured at 100 °C for 20 min.

Fluorescence lifetime measurements

Lifetime measurements were done using a EPL-375 nm ps laser diode and a time correlated photon counting system (TCPCS) from Edinburgh Instruments. The kinetic decays were measured and fitted with FAST and Origin software using mono, bi-, and triexponential fits in order to obtain the singlet excited state lifetimes having the best correlations. The instrument response frequency (IRF) was measured from the excitation beam scattering on a white card. The samples were purged with nitrogen for at least 20 min before the measurements.

Theoretical calculations

Calculations were done with Gaussian 09 at the DFT level [30]. The geometries of all the compounds were optimized using the same theoretical level; B3LYP hybrid functional with 6-31G(d) basis set [31–35]. Time-dependent density functional theory (TD-DFT) was used to calculate the optical transitions using the 6-31G(d) basis set. The first six lowest transitions were calculated.

Synthesis

4-Bromo-7- ( 4-dimethylaminophenyl ) -2,1,3-benzothiadiazole ( 4 ) [36, 37]. To a solution of 3 (91 mg, 0.31 mmol) in a mixture of toluene/water (6/2 mL), were added 4-(dimethylamino)phenylboronic acid (0.103 g, 0.62 mmol), tetrabutylammonium bromide (TBAB) (20 mg, 0.062 mmol), and cesium carbonate (0.504 g, 1.55 mmol). Then the mixture was degassed for 5 min and tetrakis(triphenylphosphine)palladium(0) (0.016 g, 0.0155 mmol) was added. The mixture was stirred and heated at 90 °C overnight. The cooled reaction mixture was filtered over celite to remove the catalyst. The filtrate was concentrated under reduced pressure then purified by silica gel chromatography. The product was isolated as an orange solid (68 % yield); 1H NMR (400 MHz, CDCl3-d): δ = 3.05 (s, 6H), 6.86 (d, 2H, J = 6.9 Hz), 7.49 (d, 1H, J = 7.5 Hz), 7.85–7.87 (m, 3H).13C NMR (100 MHz, CDCl3-d): δ = 154.1, 153.5, 150.8, 134.4, 132.6, 130.2, 126.5, 124.6, 112.5, 111.1, 40.6. HR-MS = calcd 334.0008; found 334.0012 (M+H)+.

4- { 7- [ 4- ( Dimethylamino ) phenyl ] -2,1,3-benzothiadiazol-4-yl } benzonitrile ( 1 ) . To a solution of 4 (40 mg, 0.12 mmol) in a mixture of toluene/water (6/2 mL), were added 4-cyanophenylboronic acid (35 mg, 0.24 mmol), tetrabutylammonium bromide (TBAB) (10 mg, 0.028 mmol), and cesium carbonate (235 mg, 0.72 mmol). Then the mixture was degassed for 5 min and tetrakis(triphenylphosphine)palladium(0) (8 mg, 6.5 μmol) was added. The mixture was stirred and heated at 90 °C overnight. The reaction mixture was then cooled to room temperature and filtered over celite. The filtrate was then concentrated under reduced pressure and the title compound was obtained as an orange solid (26 % yield). 1H NMR (400 MHz, CDCl3-d): δ = 3.07 (s, 6H), 6.88 (d, 2H, J = 6.9 Hz), 7.76 (d, 1H, J = 7.8 Hz), 7.79–7.82 (m, 3H), 7.94 (d, 2H, J = 7.9 Hz), 8.10 (d, 2H, J = 8.1 Hz). 13C NMR (100 MHz, CDCl3-d): δ = 154.4, 153.9, 150.9, 142.3, 135.2, 132.5, 130.3, 129.8, 129.4, 129.3, 126.0, 124.9, 119.1, 112.4, 111.5, 40.5. HR-MS = calcd 377.1072; found 377.1065 (M+H)+.

N,N-Dimethyl-4- [ 7- ( 4-nitrophenyl ) -2,1,3-benzothiadiazol-4-yl]aniline ( 2 ) . To a solution of 4 (40 mg, 0.12 mmol) in a mixture of toluene/water (6/2 mL), were added 4-nitrophenylboronic acid (40 mg, 0.24 mmol), tetrabutylammonium bromide (TBAB) (10 mg, 0.028 mmol), and cesium carbonate (235 mg, 0.72 mmol). The mixture was then degassed for 5 min and tetrakis(triphenylphosphine)palladium(0) (8 mg, 6.5 μmol) was added. The mixture was stirred and heated at 90 °C overnight, afterwards it was filtered on celite. The filtrate was then concentrated under reduced pressure with the title compound being obtained as a red solid (68 % yield). 1H NMR (400 MHz, CDCl3-d): δ = 3.07 (s, 6H), 6.89 (d, 2H, J = 6.9 Hz), 7.76 (d, 1H, J = 7.8 Hz), 7.84 (d, 1H, J = 7.8 Hz), 7.95 (d, 2H, J = 7.9 Hz), 8.17 (d, 2H, J = 8.2 Hz), 8.38 (d, 2H, J = 8.4 Hz) ppm. 13C NMR (100 MHz, CDCl3-d): δ = 154.3, 153.9, 150.9, 147.3, 144.2, 135.4, 130.4, 129.9, 129.6, 128.9, 125.9, 124.7, 124.0, 112.4, 40.5. HR-MS = calcd 357.1174; found 357.1167 (M+H)+.

Results and discussion

To probe the solvent effect on the spectroscopic properties, the electronic absorption and steady-state fluorescence of 1 and 2 (Scheme 1) were measured in aprotic solvents of different polarity. As seen in Table 1, the absorption of 1 is shifted by 10 nm (475 cm−1) as a function of solvent, whereas 2 is shifted by 15 nm (685 cm−1). The larger spectral shift observed for 2 is from its ground state being more perturbed by solvent effects than 1. This implies that 2 has a greater ground state dipole than 1, courtesy of a stronger electronic push–pull effect between the –NO2 and –NMe2 termini.

Scheme 1: 
          Synthetic scheme for the preparation of 1 and 2: a) Cs2CO3, TBAB, Pd(PPh3)4, toluene/H2O (3/1), 90 °C.
Scheme 1:

Synthetic scheme for the preparation of 1 and 2: a) Cs2CO3, TBAB, Pd(PPh3)4, toluene/H2O (3/1), 90 °C.

Table 1

Spectroscopic properties of 1 and 2 in different solvents.

Solvent Δfa ET(30)a 1
2
λ abs (nm)b,c λ em (nm)b,d Stokes shift (cm−1) Φfl (%)e λ abs (nm)b,c λ em (nm)b,d Stokes shift (cm−1) Φfl (%)e
Hexane 0.002 31.0 455 (413) 605 5470 84 460 (435) 605 5190 86
Toluene 0.013 33.9 460 (422) 630 5750 69 470 (426) 635 5530 74
Ether 0.167 34.5 455 (421) 640 6280 73 460 (425) 650 6285 67
Dichloromethane 0.218 40.7 455 (430) 675 7140 44 470 (436) 680 6730 40
Ethyl acetate 0.200 38.1 455 680 7280 41 465 695 7210 24
THF 0.210 37.4 460 (428) 685 7100 42 470 (434) 700 7080 18
Acetone 0.284 42.2 455 (429) 735 8370 10 465 (435) 710 7600 <1
Acetonitrile 0.305 45.6 450 (429) 755 9030 5 460 745 8470 <1
DMSO 0.263 45.4 465 (433) 780 8720 3 475 (440) 765 8040 <1
as-castf 470 680 6570 4 500 702 5755 <1
1:10 PMMAf 465 615 5245 17 467 626 5440 29
1:100 PMMAf 465 610 5110 59 465 619 5350 65
1:1000 PDMSg 435 620 6860 40 457 592 4990 10

aValues taken from literature [38, 39]. b±3 nm. cValue in parentheses are theoretically calculated by TD-DFT methods using B3LYP/6-31G(d) basis set. dExciting at the corresponding maximum absorption. eDetermined with an integrating sphere calibrated with anthracence. fThin film thickness ca. 30 nm (as-cast) and ca. 200 nm deposited on glass substrates in a given polymer matrix. gEmbedded in a polydimethylsiloxane matrix and polymerized to form a free standing gel ca. 0.1 cm thick.

The effect of solvent polarity on the fluorescence wavelength was also examined. The fluorescence of 1 and 2 was significantly shifted in different solvents. This is evident in Fig. 1, where the emission maximum of 1 is shifted upwards of 175 nm (3710 cm−1), while 2 is shifted by 160 nm (3455 cm−1) in similar solvents. To some extent, the observed shifts are not surprising given the excited state is known to be more perturbed by solvent effects than the ground state. This is a result of the excited state being more polar than the ground state for the π to π* transition. The π to π* transition was confirmed by DFT calculations (vide infra) with the corresponding HOMO and LUMO calculated orbitals seen in Fig. 3. The conjugated electron donating and withdrawing groups can further give rise to an intramolecular charge transfer that is favored in polar solvents [40]. Although the exact nature of the excited state cannot be unequivocally identified, the excited state is significantly perturbed by different solvent polarities. This is further supported by the large Stokes shift. The effect of solvent on the emission color is qualitatively seen in the photographs of Fig. 1c and d. It is evident that the fluorescence of 1 shifts from yellow to green with increasing solvent polarity. Similarly, the emission color of 2 shifts from yellow to blue. It should be noted that the fluorescence of both 1 and 2 was poorly correlated with the Lippert–Mataga relationship (Fig. S3). This implies specific solute–solvent interactions take place in the excited state. These interactions are more pronounced with 2, whose linear regression for the Stokes shift as function of ET(30) was r2 = 0.89. To some extent, this is not surprising, owing to the nitro group that is known is sustain specific NO2–solvent interactions that are not taking into account in the general Lippert–Mataga equation. An improved linear regression was found for 1 (r2 = 0.93), confirming reduced specific solvent–solute interactions involving the CN group in the excited state compared to 2. Although ethanol would be an ideal solvent for examining the hydrogen bonding effect on the fluorescence behavior, both 1 and 2 were insufficiently soluble in alcohols for quantitative measurements.

Fig. 1: 
          Normalized emission spectra of 1 (a) and 2 (b) measured in dichloromethane (■), DMSO (□), acetonitrile (), hexane (), acetone (), ethyl acetate (), diethyl ether (), THF (), toluene () by exciting at the respective maximum absorption. Photographs showing the fluorescence of 1 (c) and 2 (d) in hexane (1), toluene (2), ether (3), dichloromethane (4), ethyl acetate (5), THF (6), acetone (7), acetonitrile (8), and DMSO (9), when excited with a UV lamp at 360 nm.
Fig. 1:

Normalized emission spectra of 1 (a) and 2 (b) measured in dichloromethane (■), DMSO (□), acetonitrile (

), hexane (
), acetone (
), ethyl acetate (
), diethyl ether (
), THF (
), toluene (
) by exciting at the respective maximum absorption. Photographs showing the fluorescence of 1 (c) and 2 (d) in hexane (1), toluene (2), ether (3), dichloromethane (4), ethyl acetate (5), THF (6), acetone (7), acetonitrile (8), and DMSO (9), when excited with a UV lamp at 360 nm.

Of particular interest is the fluorescence quantum yield (Φfl) dependence on solvent polarity. Rather than using relative actinometry to determine the Φfl, the values were calculated using an integrating sphere. The benefit of this method is that fluorescence references absorbing and emitting in the same region and having roughly the same Φfl as the compound of study are not required. This is particularly advantageous for studying the BZT fluorophores whose emission wavelength and intensity varied considerably with solvent (vide supra). The Φfl values derived from the integrating sphere method are additionally absolute values. To validate the results, the sphere was calibrated against anthracene in degassed anhydrous ethanol (Φfl = 28 %) [28]. As seen in Table 1 and Fig. 1b and c, both 1 and 2 fluoresce appreciably in apolar solvents. In fact, the Φfl of both 1 and 2 are similar within experimental error in hexane, toluene, ether, and dichloromethane. The emission of 2 decreased significantly in acetone, acetonitrile, and DMSO, such that it could not be quantified with the instrument. While the same trend was observed for 1, its emission could nonetheless be quantified using the integrating sphere method. Although 1 and 2 are electronically similar, the emission yield of the fluorophore with the nitrile group is less perturbed by solvent effects than with the nitro group.

The change in Φfl with solvent polarity (Table 1) is not surprising given the narrowing of the 1S-1T energy gap that can occur in polar solvents. This can potentially increase intersystem crossing to the triplet manifold. Transient absorption spectroscopy was therefore used to determine whether a triplet state was formed when exciting 1 and 2. The transient absorption spectra of 1 and 2 were examined in toluene, dichloromethane, and acetonitrile. These solvents were chosen because of the contrasting measured Φfl values. A transient was observed only in the case of 1 in acetonitrile having an absorption at 360 and 540 nm. The absorption in the visible region was assigned to the triplet owing to its unimolecular decay (τ = 2.6 μs) and its quenching with oxygen. The absence of detectable signal on the μs time scale suggests a triplet is not formed for 2. Its quenched fluorescence therefore arises from solute-solvent interactions. These occur to a lesser extent for 1, whose fluorescence deactivation occurs by intersystem crossing in acetonitrile.

Two short-lived transients within a 100 ns time window were observed for 1 and 2 by transient absorption spectroscopy. While these were ascribed to the singlet excited state, the interesting feature is their interdependence. For example, the transient at 530 nm was converted into the transient at 655 nm (3600 cm−1 shift) for 1 in dichloromethane. The red-shifted transient observed for 1 in acetonitrile (715 nm; 4880 cm−1), toluene (620 nm; 2740 cm−1), and dichloromethane (675 nm; 4055 cm−1) was consistent with the emission measured by steady-state fluorescence. The two states were further probed by time-resolved fluorescence. The lifetime of the red-emission of 1 in dichloromethane was 7.4 ns, whereas the blue emission was 5.4 ns. The fluorescence lifetime of 1 was consistent in the solvents examined. In contrast, the lifetime of 2 became shorter when progressing from toluene, dichloromethane, THF, acetonitrile, and DMSO (Table S1). This trend was consistent with the decrease in Φfl of 2 in the different solvents, confirming the importance of non-radiative deactivation modes involving the nitro group with increasing solvent polarity.

The emission scans of 1 and 2 confirmed the dual-emission of the two fluorophores. This is evident in Fig. 2 where an emission at 480 and 675 nm can be seen for 1 in dichloromethane. The emission intensity was contingent on the excitation wavelength. The isosbestic point at 580 nm confirms the independence of the dual-emission. 2 similarly had a dual-emission with the red-shifted emission being less intense. In contrast, the blue-emission of 1 was essentially quenched in toluene. The origin of the dual-emission cannot be from aggregates, excimers, and bimolecular processes given the ratio of the blue/red emission of 1 and 2 did not change with concentration. In fact, no change was observed when diluting the samples up to 370 fold (Figs. S13 and S16). This provides strong evidence for intramolecular effects. Given the shift was contingent on solvent polarity, the red emission is the intramolecular charge transfer band between the donating amine and accepting nitrile and nitro, respectively, whereas the blue-emission is the global emission.

Fig. 2: 
          Fluorescence spectra of 1 in dichloromethane by exciting between 380 and 450 nm.
Fig. 2:

Fluorescence spectra of 1 in dichloromethane by exciting between 380 and 450 nm.

To gain a better understanding of the spectroscopic properties, DFT calculations were done to investigate both the ground and excited states. This level of theory was chosen because it is well-known to accurately afford both the optimized geometry and corresponding properties of π-conjugated molecules [41]. Moreover, the intrinsic errors of the theoretically calculated values are consistent such that the relative values can be accurately calculated. The X-ray data of a similar compound was used as the starting geometry of 1 and 2 to accelerate the geometry optimization calculations [36]. The resulting calculated optimized geometries were in good agreement with the XRD data. Figure 3 shows the Frontier orbitals and the HOMO/LUMO energy levels calculated for 1 in different solvents (Fig. S35 for 2). Little difference in the HOMO-LUMO energy levels was calculated, suggesting these orbitals are only weakly perturbed by the solvent. The calculations provide evidence for the HOMO-1 and LUMO+1 orbitals contributing to the change in spectral properties, albeit only weakly. Time-dependent density functional theory (TD-DFT) was used to calculate the excited energies. The calculated oscillator strengths and excitation energies (Table S2) were used to calculate the major electronic transition in the theoretical absorption spectrum, according to Stephens methods [42]. As seen in Table 1, the calculated absorptions are consistent with the experimental data with the same trend in absorption shift with solvent. The Frontier orbitals provide further insight into the origin of the observed spectral changes. For example, the charge density of the HOMO is concentrated on the donor portion of the fluorophore. Similarly, the charge density of the LUMO is uniquely localized on the acceptor portion of the fluorophore. The electronic transition therefore involves intramolecular charge transfer between the HOMO to the LUMO for both fluorophores. This is consistent with the experimental results.

Fig. 3: 
          Top: calculated HOMO () and LUMO () energy levels and Frontier orbitals (right) of 1 in different solvents. Bottom: molecular Frontier orbitals and corresponding energy levels of 1 calculated in hexane.
Fig. 3:

Top: calculated HOMO (

) and LUMO (
) energy levels and Frontier orbitals (right) of 1 in different solvents. Bottom: molecular Frontier orbitals and corresponding energy levels of 1 calculated in hexane.

The change in spectral properties of 1 and 2 was further investigated in the presence of trifluoroacetic acid (TFA). This was to confirm the origin of the red-emission arising from intramolecular charge transfer. TFA was chosen because of its capacity to fully dissociate in a wide range of organic solvents. It is therefore a sufficiently strong acid to protonate the N,N-dimethylamino terminus in the solvents examined. Protonating the amine disrupts the electronic push–pull character of the fluorophore, yet the overall conjugation remains intact. As a result, only the electron withdrawing group can affect the properties. The absorption of 1 at 450 nm was blue-shifted to 370 nm (4805 cm−1) when the fluorophore was protonated with TFA. The same trend was observed for 2. As seen in Fig. 4, the red-emission of 1 is gradually quenched when titrating with TFA. The same behavior was also observed for 2 in the various solvents. The collective solvent and TFA induced spectral shifts provide evidence that the red-shifted emission is from intramolecular charge transfer, while the blue-shifted emission is from the local emission.

Fig. 4: 
          Change in fluorescence of 1 in toluene with 0 (○), 90 (), 170 (), 260 (), 350 (■), 430(), and 520 () mM TFA.
Fig. 4:

Change in fluorescence of 1 in toluene with 0 (○), 90 (

), 170 (
), 260 (
), 350 (■), 430(
), and 520 (
) mM TFA.

Given the interest of using the fluorophores in emitting applications and rheological probes in polymer composites, the emission properties of 1 and 2 were further examined in thin films. The fluorophores were drop-cast (as-cast) on to glass substrates from dichloromethane and the resulting films of 2 were non-fluorescent (Table 1). The quenched fluorescence in thin films, to some extent, was not surprising, given the known intermolecular self-quenching deactivation modes that prevail in the solid-state. The fluorophores were subsequently mixed in different ratios with the poly(methyl methacrylate) (PMMA) matrix to reduce intermolecular self-quenching (Fig. 5). The fluorescence yield increased with increasing PMMA/fluorophore ratio and the resulting Φfl values measured were comparable to the solution measurements. Consistent high emission yields were also obtained when mixing the fluorophores in a poly(dimethylsiloxane) matrix. In fact, bendable and flexible free-standing substrates ca. 5 cm × 5 cm were obtained. Consistent fluorescence was observed with repeated bending and stretching of the elastomeric film.

Fig. 5: 
          Photographs of 1 in a PMMA matrix 1:0 (left), 1:10 (middle), and 1:100 (right) on glass substrates irradiated with a UV lamp.
Fig. 5:

Photographs of 1 in a PMMA matrix 1:0 (left), 1:10 (middle), and 1:100 (right) on glass substrates irradiated with a UV lamp.

The electrochemistry of the fluorophores was investigated to determine whether they could be reversibly oxidized. This is a desired property for using the fluorophores as functional materials in plastic electronics. As seen in Fig. 6, both 1 and 2 could be reversibly oxidized. The reversible behavior is in part owing to the terminal electron groups that prevent cross-coupling of the oxidized intermediates, according to known mean [43]. Using ferrocene as an internal reference, the oxidation potentials (Eo) measured for 1 and 2 were 875 and 930 mV vs. SCE, respectively. The difference in the oxidation potentials is owing to the different electron withdrawing strengths of the CN and NO2 groups. As a result of the reduced electron withdrawing capacity of the nitrile group, the donating effect of the terminal amine of 1 is enhanced compared to 2. This results in the less positive oxidation potential observed for 1 relative to 2.

Fig. 6: 
          Anodic cyclic voltammograms of 1 () and 2 () measured in anhydrous and deaerated dichloromethane with 0.1 M TBAPF6 supporting electrolyte and ferrocene as an internal reference.
Fig. 6:

Anodic cyclic voltammograms of 1 (

) and 2 (
) measured in anhydrous and deaerated dichloromethane with 0.1 M TBAPF6 supporting electrolyte and ferrocene as an internal reference.

To further evaluate the usefulness of the fluorophores, especially as electrochromes, their spectroelectrochemistry was evaluated. This was done by gradually applying a positive potential to generate the oxidized species, while examining the change in either the absorption or fluorescence. As seen in Fig. 7a, the absorption of 1 at 455 nm decreases with increasing potential with the concurrent formation of new peaks at 365, 525, 760, and 840 nm. The isosbestic point at 405 nm confirms the presence of only two species, the neutral and oxidized states, and their independence. The resulting visible color change of 1 with electrochemical oxidation was yellow to mauve. The neutral state could be regenerated by applying a slightly negative potential. The original absorption spectrum was obtained after neutralizing the oxidized state. The same spectral trend with applied potential was also observed with 2. For this fluorophore, the color changed from yellow to pale violet with a positive potential, along with the formation of new absorptions at 370, 505, 760, and 835 nm. Similar to 1, a well-defined isosbestic point was observed at 410 nm for 2.

Fig. 7: 
          (a) Spectroelectrochemistry of 1 with applied potentials from 0 to 1 300 mV vs. Ag/Ag+ held for 30 s per potential in dichloromethane with TBAPF6 as an electrolyte. (b) Fluorescence change of 1 when excited at 390 nm with applied potentials from 0 to 1 300 mV for 30 s followed by −1000 mV.
Fig. 7:

(a) Spectroelectrochemistry of 1 with applied potentials from 0 to 1 300 mV vs. Ag/Ag+ held for 30 s per potential in dichloromethane with TBAPF6 as an electrolyte. (b) Fluorescence change of 1 when excited at 390 nm with applied potentials from 0 to 1 300 mV for 30 s followed by −1000 mV.

Given the intrinsic fluorescence of 1 and 2, their spectroelectrochemical fluorescence was also investigated. This was done by exciting the fluorophores at 390 nm and monitoring their change in fluorescence with increasing potential. The resulting fluorescence change of 1 is shown in Fig. 7b. Qualitatively, the emission color changed from purple to blue upon oxidation. Quantitatively, the emission of the intramolecular charge transfer band at 675 nm decreased when the neutral form was oxidized concomitant with an increase at 480 nm. The two neutral and oxidized states are interdependent, as per the isosbestic point at 615 nm in Fig. 7b. 2 similarly underwent fluorescence changes when oxidized. Its emission at 540 nm increased while the emission at 680 nm decreased with increasing potential, resulting in visible color changes from gray-green to green.

Conclusion

Two electronic push–pull benzothiadiazole fluorophores were prepared and the effect of the terminal electron withdrawing groups on the spectroscopic and electrochemical properties was examined. It was found that both fluorophores had similar properties, notably their solvatochromism and fluorescence quantum yields. Interestingly, the intramolecular charge transfer band could be electrochemically modulated. This resulted in a change in emission wavelength and intensity with electrochemical oxidation. The intrinsic intense fluorescence of the fluorophores was preserved when incorporated into polymer matrices. The collective solvatochromism, intense fluorescence in polymer matrices, and spectroelectrofluorescence make the electronic push–pull benzothiadiazole derivatives interesting universal fluorescence probes. Meanwhile, the similar spectroscopy and electrochemistry of the two fluorophores demonstrate that the electron withdrawing –NO2 group can be readily replaced with –CN, without affecting adversely the properties.


Article note

A collection of invited papers based on presentations at the XXVth IUPAC Symposium on Photochemistry, Bordeaux, France, July 13–18, 2014.



Corresponding author: William G. Skene, Laboratoire de caractérisation photophysique des matériaux conjugués, Département de chimie, Université de Montréal, CP 6128, Centre-ville, Montreal, QC, Canada, e-mail:

Acknowledgments

The authors acknowledge both the Natural Sciences and Engineering Research Council Canada and Canada Foundation for Innovation for operating and equipment grants, respectively. M.-H. T. acknowledges the Fonds de recherche du Québec – Nature et technologies for a graduate scholarship. WestGrid (www.westgrid.ca) and Compute Canada/Calcul Canada (www.computecanada.ca) are acknowledged for access to software and computational resources.

Funding: Natural Sciences and Engineering Research Council of Canada, (Grant/Award number: Discovery Grant). Fonds Québécois de la Recherche sur la Nature et les Technologies, (Grant/Award number: Graduate Scholarship).

References

[1] Y. Olivier, D. Niedzialek, V. Lemaur, W. Pisula, K. Müllen, U. Koldemir, J. R. Reynolds, R. Lazzaroni, J. Cornil, D. Beljonne. Adv. Mater. 26, 2119 (2014).Search in Google Scholar

[2] Y. Li, T. Liu, H. Liu, M.-Z. Tian, Y. Li. Acc. Chem. Res. 47, 1186 (2014).Search in Google Scholar

[3] C. Gao, L. Wang, X. Li, H. Wang. Polym. Chem. 5, 5200 (2014).Search in Google Scholar

[4] Z. Zhu, D. Waller, C. J. Brabec. “New construction of low-bandgap conducting polymers”, Chapter 4, in Organic Photovoltaics, pp. 129–151, Wiley-VCH Verlag GmbH & Co. KGaA (2009).10.1002/9783527623198.ch4Search in Google Scholar

[5] T. C. Parker, D. G. Patel, K. Moudgil, S. Barlow, C. Risko, J.-L. Bredas, J. R. Reynolds, S. R. Marder. Mater. Horiz. 2, 22 (2015).Search in Google Scholar

[6] G. Gunbas, L. Toppare. Macromol. Symp. 297, 79 (2010).Search in Google Scholar

[7] H. N. Tsao, H. J. Räder, W. Pisula, A. Rouhanipour, K. Müllen. “Novel organic semiconductors and processing techniques for organic field-effect transistors”, Chapter 4, in Organic Electronics, pp. 61–74, Wiley-VCH Verlag GmbH & Co. KGaA, Germany (2009).10.1002/9783527627387.ch4Search in Google Scholar

[8] Y. Yamashita. Chem. Lett. 38, 870 (2009).10.1246/cl.2009.870Search in Google Scholar

[9] X. Liu, B. B. Y. Hsu, Y. Sun, C.-K. Mai, A. J. Heeger, G. C. Bazan. J. Am. Chem. Soc. 136, 16144 (2014).Search in Google Scholar

[10] J. Chen, Y. Cao. Acc. Chem. Res. 42, 1709 (2009).Search in Google Scholar

[11] Y. Li. Acc. Chem. Res. 45, 723 (2012).10.1021/ar2002446Search in Google Scholar PubMed

[12] S. Beaupre, M. Leclerc. J. Mater. Chem. A1, 11097 (2013).10.1039/c3ta12420gSearch in Google Scholar

[13] Y. Wu, W. Zhu. Chem. Soc. Rev. 42, 2039 (2013).Search in Google Scholar

[14] Y. Kim, E. Lim. Polymers6, 382 (2014).10.3390/polym6020382Search in Google Scholar

[15] H.-S. Yi, C.-M. Ryu, M. Heil. Plant Signal. Behav. 5, 528 (2010).Search in Google Scholar

[16] A. I. Darras. “Novel elicitors induce defense responses in cut flowers”, Chapter 4, in Plant Pathology, C. J. Cumagun (Ed.), pp. 85–116, InTech Europe, Rijeka, Croatia, (2012).10.5772/30180Search in Google Scholar

[17] J. Pina, J. S. de Melo, D. Breusov, U. Scherf. Phys. Chem. Chem. Phys. 15, 15204 (2013).Search in Google Scholar

[18] S. Ayachi, A. Mabrouk, K. Alimi, M. Bouachrine. “Photophysical properties of two new donor-acceptor conjugated copolymers and their model compounds: applications in polymer light emitting diodes (PLEDs) and polymer photovoltaic cells (PPCs)”, Chapter 5, in Organic Light Emitting Devices, J. Singh (Ed.), pp. 105–142, InTech, Rijeka, Croatia (2012).Search in Google Scholar

[19] B. A. D. Neto, A. A. M. Lapis, E. N. da Silva Jr., J. Dupont. Eur. J. Org. Chem. 2013, 228 (2013).Search in Google Scholar

[20] Y. Yang, Y. Zhou, Q. He, C. He, C. Yang, F. Bai, Y. Li. J. Phys. Chem. B113, 7745 (2009).10.1021/jp900362fSearch in Google Scholar PubMed

[21] R. Zhan, B. Liu. Macromol. Chem. Phys. 216, 131 (2015).Search in Google Scholar

[22] R. Bou Zerdan, P. Cohn, E. Puodziukynaite, M. B. Baker, M. Voisin, C. Sarun, R. K. Castellano. J. Org. Chem. 80, 1828 (2015).Search in Google Scholar

[23] A. Casey, Y. Han, Z. Fei, A. J. P. White, T. D. Anthopoulos, M. Heeney. J. Mater. Chem. C3, 265 (2015).10.1039/C4TC02008ASearch in Google Scholar

[24] H. Li, Y. Guo, Y. Lei, W. Gao, M. Liu, J. Chen, Y. Hu, X. Huang, H. Wu. Dyes and Pigments112, 105 (2015).10.1016/j.dyepig.2014.06.035Search in Google Scholar

[25] J. R. Lakowicz. Principles of Fluorescence Spectroscopy, Springer, New York (2006).10.1007/978-0-387-46312-4Search in Google Scholar

[26] N. I. Rtishchev, D. V. Samoilov, V. P. Martynova, A. V. El’tsov. Russ. J. Gen. Chem. 71, 1467 (2001).Search in Google Scholar

[27] T. Ueno, Y. Urano, H. Kojima, T. Nagano. J. Am. Chem. Soc. 128, 10640 (2006).Search in Google Scholar

[28] A. M. Brouwer. Pure Appl. Chem. 83, 2213 (2011).10.1351/PAC-REP-10-09-31Search in Google Scholar

[29] N. G. Connelly, W. E. Geiger. Chem. Rev. 96, 877 (1996).Search in Google Scholar

[30] 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. J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, 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, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox. Gaussian, Inc., Wallingford CT (2009).Search in Google Scholar

[31] A. D. Becke. J. Chem. Phys. 98, 5648 (1993).10.1063/1.464913Search in Google Scholar

[32] A. Widom, T. D. Clark, G. Megaloudis. Phys. Lett. A76, 163 (1980).10.1016/0375-9601(80)90602-7Search in Google Scholar

[33] P. C. Hariharan, J. A. Pople. Mol. Phys. 27, 209 (1974).Search in Google Scholar

[34] W. J. Hehre, R. Ditchfield, J. A. Pople. J. Chem. Phys. 56, 2257 (1972).Search in Google Scholar

[35] R. Ditchfield, W. J. Hehre, J. A. Pople. J. Chem. Phys. 54, 724 (1971).Search in Google Scholar

[36] S.-I. Kato, T. Matsumoto, M. Shigeiwa, H. Gorohmaru, S. Maeda, T. Ishi-I, S. Mataka. Chem. Eur. J. 12, 2303 (2006).Search in Google Scholar

[37] S.-I. Kato, T. Matsumoto, T. Ishi-I, T. Thiemann, M. Shigeiwa, H. Gorohmaru, S. Maeda, Y. Yamashita, S. Mataka. Chem. Commun. 20, 2342 (2004).Search in Google Scholar

[38] J. A. Riddick, W. B. Bunger, T. K. Sakano. Organic Solvents: Physical Properties and Methods of Purification, 4th ed. Wiley- Interscience, New York (1986).Search in Google Scholar

[39] I. M. Smallwood. Handbook of Organic Solvent Properties, Elsevier, Burlington, MA (1996).Search in Google Scholar

[40] H. Dinçalp, G. Murat, S. İçli. Opt. Mater. 36, 1525 (2014).Search in Google Scholar

[41] D. P. Hagberg, T. Edvinsson, T. Marinado, G. Boschloo, A. Hagfeldt, L. Sun. Chem. Commun. 2006, 2245 (2006).Search in Google Scholar

[42] P. J. Stephens, N. Harada. Chirality22, 229 (2010).Search in Google Scholar

[43] J. Roncali. Chem. Rev. 92, 711 (1992).10.1021/cr00012a009Search in Google Scholar


Supplemental Material

The online version of this article (DOI: https://doi.org/10.1515/pac-2015-0404) offers supplementary material, available to authorized users.


Published Online: 2015-07-01
Published in Print: 2015-07-01

©2015 IUPAC & De Gruyter

Downloaded on 30.4.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2015-0404/html
Scroll to top button