Early events associated with the excited state proton transfer in 2-(2′pyridyl)benzimidazole Tarak Nath Burai, Tushar Kanti Mukherjee, Priyanka Lahiri, Debashis Panda, and Anindya Datta Citation: J. Chem. Phys. 131, 034504 (2009); doi: 10.1063/1.3177457 View online: http://dx.doi.org/10.1063/1.3177457 View Table of Contents: http://jcp.aip.org/resource/1/JCPSA6/v131/i3 Published by the American Institute of Physics. Additional information on J. Chem. Phys. Journal Homepage: http://jcp.aip.org/ Journal Information: http://jcp.aip.org/about/about_the_journal Top downloads: http://jcp.aip.org/features/most_downloaded Information for Authors: http://jcp.aip.org/authors Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions THE JOURNAL OF CHEMICAL PHYSICS 131, 034504 共2009兲 Early events associated with the excited state proton transfer in 2-„2⬘-pyridyl…benzimidazole Tarak Nath Burai, Tushar Kanti Mukherjee, Priyanka Lahiri, Debashis Panda, and Anindya Dattaa兲 Department of Chemistry, Indian Institute of Technology Bombay, Powai, Mumbai 400 076, India 共Received 28 February 2009; accepted 25 June 2009; published online 16 July 2009兲 2-共2⬘-pyridyl兲benzimidazole 共2PBI兲 undergoes excited state proton transfer 共ESPT兲 in acidic solutions, leading to a tautomer emission at 460 nm. This photoprocess has been studied using ultrafast fluorescence spectroscopic techniques in acidic neat aqueous solutions, in viscous mixtures of glycerol with water, as well as in sucrose solutions. The tautomer is found to be stabilized in the more viscous medium, leading to a greater relative quantum yield as well as lifetime. The long rise time in tautomer emission is not affected by viscosity though. Rather, it appears to have the same value as the long component of the decay of the cationic excited state 共Cⴱ兲. In addition to the subnanosecond lifetime reported earlier, Cⴱ is found to exhibit a decay time of 2 ps. This is assigned to its protonation to form the nonfluorescent dication in its excited state 共Dⴱ兲 considering the ground and excited state pKa values reported earlier. An additional rising component of 100 ps is observed in the region of Cⴱ emission. This is likely to arise from a structural change or charge redistribution in Cⴱ immediately after its creation and before the phototautomerization. © 2009 American Institute of Physics. 关DOI: 10.1063/1.3177457兴 I. INTRODUCTION Proton and hydrogen transfer in the excited state are fundamental processes that have potential applications in lasers, photostabilization of polymers, and development of fluorescence sensors. So, they have generated a considerable amount of interest.1–5 Excited state proton transfer 共ESPT兲 in hydrogen-bonded systems such as benzimidazoles, in particular, has attracted a lot of attention.6,7 The ESPT of the cationic form of 2-共2⬘-pyridyl兲benzimidazole 共2PBI兲 is manifested in a distinct dual emission in its aqueous solutions in the pH range of 3.5–0.5 共Scheme 1兲.8 The inclusion of 2PBI in cyclodextrins hinders the solvent-mediated ESPT to some extent due to shielding of the fluorophore from water.9 Systematic studies have been performed on the ESPT of 2PBI in micelles,10 reverse micelles,11 and nafion film.12 These studies reveal that the ESPT can be brought about in microheterogeneous media which contain water as a proton donor along with a nonpolar compartment, separated by a negatively charged interface. The four important species that participate in the ESPT process are the normal form of 2PBI 共N兲, the monocationic form 共C兲, the dication 共D兲, and the phototautomer 共Tⴱ, Scheme 1兲. N and C fluoresce at 380 nm, while Tⴱ fluoresces at 450 nm. D has been reported to be nonfluorescent so far. The tautomer emission is associated with the rise time of approximately 0.8 ns in the microheterogeneous media where ESPT occurs. The slow dynamics of ESPT of in aqueous micelle and reverse micelles is explained as a result of slow solvation and slowing down of the proton transfer process as such.13–15 However, a 600–800 ps growth is observed in neat aqueous solutions at pH 3 as well.8 The origin of this long rise time is not understood as proton transfer is known to occur within 150 fs in hydrogenbonded acid-base complexes.16 However, examples of long time constants in the ESPT process are available as well. Proton transfer proceeds in 250 ps, for example, in encounter pairs formed by diffusion of uncomplexed photoacid and base molecules.17,18 Proton transfer in acetonitrile-water mixture from strong photoacids 6-hydroxyquinolinium and 6-hydroxy-1-methylquinolinium is associated with a rather long time constant of approximately 1 ns, the process being diffusion controlled.19 Long time constants of approximately 100 ps have been observed in the perylene quinones, hypocrellin, and calphostin C as a result conformational change and chain dynamics associated with the proton transfer a兲 Author to whom correspondence should be addressed. Electronic mail: [email protected]. 0021-9606/2009/131共3兲/034504/8/$25.00 131, 034504-1 SCHEME 1. The species involved in the ESPT of 2PBI. © 2009 American Institute of Physics Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-2 J. Chem. Phys. 131, 034504 共2009兲 Burai et al. process.20,21 In such cases, the dynamics are expected to be strongly dependent on viscosity. With this background, one might contend that the slow dynamics of ESPT, observed in 2PBI in acidic aqueous solutions, arises from a similar molecular motion likely to be the rotation of the two ring systems with respect to each other.8 One way of studying the coupling between conformational dynamics and ESPT of 2PBI is to attempt to control the mutual rotational motion of the two rings of 2PBI by varying the viscosity of the solution.20 Glycerol has been extensively employed as a viscogenic cosolvent to vary solvent viscosity.22 The viscosity of the aqueous solution has also been increased by the addition of sucrose.23 The effect of thus varying the viscosity and polarity of the medium on the ESPT of 2PBI has been investigated. II. EXPERIMENT A. Photophysical studies 2PBI 共analytical reagent grade兲 from Aldrich and sucrose 共general reagent兲 from Merck have been used as received. Methanol and glycerol 共spectroscopic grade兲 are obtained from Spectrochem, Mumbai, India. D2O and DClO4 from Aldrich have been used as received. For viscositydependence studies, glycerol-water mixtures of varying ratios are prepared. To each, a fixed amount of 2PBI solution is added. The solutions are acidified by the addition of a fixed amount of perchloric acid 共HClO4兲 to make the proton concentration in each solution 10−3M. For viscosity dependent studies with sucrose solution, solid sucrose is added to pH = 3 solution of 2PBI. The viscosities of the mixtures are taken from Ref. 24. De-ionized water is distilled twice before being used as a solvent. The pH of the aqueous solutions is adjusted using a pH meter through the addition of HClO4 solutions. The 0.5M – 2M solutions of HClO4 are also prepared from its stock solution 共11.66M兲 by proper dilution. The absorption and fluorescence spectra are recorded on JASCO V 530 spectrophotometer and Varian Cary Eclipse fluorimeter, respectively. The excitation wavelength 共ex兲 is 290 nm for fluorescence measurements. The absorbance at this wavelength is kept below 0.1 in order to avoid inner filter effects. Fluorescence quantum yields 共 f 兲 are calculated after proper correction for changes in the absorbance using literature value of 2PBI.8 1. Time-resolved fluorescence measurements Time-resolved fluorescence is recorded in a time correlated single photon counting 共TCSPC兲 system from IBH with ex = 295 nm. The fluorescence decays are collected with the emission polarizer at a magic angle of 54.7°. The data are fitted to multiexponential functions after deconvolution of the instrument response function by an iterative reconvolution technique using the JY Horiba IBH DAS 6.2 data analysis software, where reduced 2 and weighted residuals serve as parameters for goodness of fit.12 The full width at half maximum 共FWHM兲 of the instrument response function is 750 ps and the resolution is 7 ps per channel. In the femtosecond upconversion setup 共FOG 100, CDP兲, the sample is excited at 283 nm, the third harmonic of a mode-locked Ti:sapphire laser 共Tsunami, Spectra Physics兲 pumped by a 5 W Millennia 共Spectra Physics兲 DPSS laser. The fundamental beam at 850 nm is frequency tripled in a nonlinear crystal 关1 mm -barium borate 共BBO兲, = 25°, and = 90°兴. The fluorescence emitted from the sample is upconverted in a nonlinear crystal 共0.5 mm BBO, = 38°, and = 90°兲 by mixing with the gate pulse, which consists of a portion of the fundamental beam. The upconverted light is dispersed in a monochromator and detected using photon counting electronics. A cross-correlation function obtained using the Raman scattering from ethanol has a FWHM of 300 fs. The femtosecond fluorescence decays are fitted using a Gaussian function of the same FWHM for the excitation pulse. The time-dependent anisotropy, r共t兲, is constructed from the decays at parallel and perpendicular directions to that of the excitation polarization 关I储共t兲 and I⬜共t兲, respectively兴 as25 r共t兲 = I储 − I⬜ . I储 + 2I⬜ 共1兲 The decays of r共t兲 are fitted to single and multiexponential functions as required using the formula, 冉 冊 r共t兲 = r0 兺 ai exp − i t , 共 r兲 i 共2兲 where r0 is the fundamental anisotropy in the absence of other depolarizing processes such as rotational diffusion, or energy transfer is related to displacement angle 共兲 between absorption and emission dipoles as follow: r0 = 冉 冊 2 3 cos2  − 1 . 5 2 共3兲 B. Computational details The ground state geometries of the cationic and tautomeric forms of 2PBI at different dihedral angles are optimized in gas phase using the density functional theory with the Becke3LYP functional in conjugation with 6-31Gⴱ basis set as implemented in the GAUSSIAN 03 software package.26,27 The first excited singlet state 共S1兲 geometries of cationic and tautomeric forms of 2PBI at various dihedral angles are optimized in the gas phase using configuration interaction with singlets 共CIS兲 with 6-31Gⴱ basis set.28 The default options for the self-consistent field convergence and threshold limits in the optimization are used. III. RESULTS AND DISCUSSION A. The effect of viscosity and polarity on the steady state spectra Upon increasing the viscosity of the 2PBI solution by the addition of glycerol, there is a decrease in the emission intensity of the cationic species accompanied by an increase in the emission of the tautomer 共Fig. 1兲. This trend is observed until a composition of 50% glycerol 关Fig. 1共b兲兴. This behavior can be rationalized on the basis of hindrance of the rotation of the molecule about the C1 – C1⬘ bond. The decrease in the tautomer emission at even higher concentra- Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-3 J. Chem. Phys. 131, 034504 共2009兲 Ultrafast dynamics of ESPT in 2PBI TABLE I. Ratio of the relative quantum yield of monocation and tautomer in water-methanol mixture, water-glycerol mixture, and water-sucrose mixture in pH = 3. The dielectric constants and the viscosities of the mixtures are also provided. FIG. 1. 共a兲 Absorption spectra of 2PBI in aqueous solution of pH = 3 with 共i兲 no sucrose 共solid line兲 and 共ii兲 35% sucrose solution 共dotted line兲. 共b兲 Fluorescence spectra of 2PBI in glycerol-water 共pH = 3兲 mixtures of various compositions 关% glycerol 共v/v兲兴. 共c兲 Fluorescence spectra in aqueous solution with pH = 3 with 共i兲 no sucrose and 共ii兲 35% sucrose. All fluorescence spectra are corrected for fluctuations in the absorbance at ex = 305 nm. tions of glycerol is likely to be due to the decrease in the availability of water required for the solvent-mediated ESPT to take place. Alternatively, this could be due to a change in the dielectric constant of the medium from 80 to 61. However, at lower glycerol concentrations, the more significant change is that in viscosity, which increases more than fivefold 共Table I兲.29 The interplay of these factors in affecting the proton transfer process needs to be understood as pKa values vary significantly with the polarity of the solvent. In order to address this point, the experiment has been repeated in aqueous solution at pH = 3 at different concentration of sucrose as well as methanol-water mixture.23 In the presence of 35% sucrose, the dielectric constant of the solution changes from 80 to 62, but the viscosity increases four times 共Table I兲.29 In earlier studies, this increased viscosity has been found to modify the nonradiative rates 共knr兲 and increase in the excited state lifetime of molecules such as DiA 共4-di-16-ASP; 关4-共4-dihexadecylamino兲styryl兴-N-methylpyridinium iodide兲.23,30,31 The shape of the absorption spectrum of 2PBI remains unchanged upon addition of sucrose 关Fig. 1共a兲兴, but the intensity of cation emission at 380 nm decreases and that of tautomer emission at 460 nm increases concomitantly 关Fig. 1共c兲兴. The invariance of pH of the solution is checked after the experiment. As has been proposed for the experiment with glycerol-water mixture, the increase in the relative quantum yield of tautomer is likely to have arisen from the freezing of the rotation of the two rings with respect to each other at high viscosity. Notably, the variation in the relative quantum yield 共T / C兲 共Fig. 2兲 with viscosity follows almost the same trend for the glycerol and sucrose solutions at low viscosities. However, the increase in the ratio in glycerol Percentage by weight Dielectric constant Viscosity T / C 0 10 20 30 40 Water-methanol solutions 80.37 75.84 71.02 66.01 61.24 1.00 1.33 1.60 1.79 1.84 3.61 4.12 4.59 4.77 4.95 0 10 20 30 40 50 60 Water-glycerol solutions 80.37 77.55 74.72 71.77 68.76 65.63 61.03 1.00 1.29 1.73 3.08 3.65 5.41 10.68 3.61 4.48 5.08 5.88 6.25 6.58 6.71 0 10 20 30 35 Water-sucrose solutions 80.4 78.0 75.4 72.6 71.8 1.00 1.35 1.95 2.95 4.03 3.61 4.13 4.98 5.18 5.21 FIG. 2. 共a兲 Variation in the ratio of relative fluorescence quantum yields of tautomer and cationic species in glycerol-water 共pH = 3兲 mixtures 共䊊兲, in sucrose-water solution at different percentage of sucrose 共쎲兲, and in methanol-water solution 共 ∀ 兲 as a function of logarithm of viscosity. 共b兲 Variation in the ratio of relative fluorescence quantum yields of tautomer and cationic species in glycerol-water 共pH = 3兲 mixtures 共䊊兲, in sucrosewater solution at different percentage of sucrose 共쎲兲, and in methanol-water solution 共 ∀ 兲 as a function of dielectric constant ex = 305 nm. Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-4 Burai et al. solutions of higher viscosity is much more than that for sucrose solutions of same viscosity. This may be rationalized in terms of a greater hydrogen bonding of 2PBI with sucrose, whereby “blocked structure” might be formed analogous to the case of 3-hydroxyflavone.32 Such blocked structures are known to hinder ESPT to a significant extent. Alternatively, the decrease in the dielectric constant of the solution might also play an important role here. In order to ascertain the role of polarity, the experiment has been repeated in methanolwater mixtures at pH = 3. This is a control experiment as the decrease in the dielectric constant in the presence of 40% methanol is comparable to that in the presence of 60% glycerol, but the change in viscosity is negligible 共Table I兲. In the methanol-water mixtures, the ratio of tautomer and cation fluorescence increases, indicating a role of the polarity in the process 共Fig. 2兲. The extent of change in the ratio, however, is significantly lesser in methanol than in glycerol solution with the same polarity 共Table I兲. Moreover, a plot of the ratios of quantum yield against the viscosities reveal that at the low viscosity region, the variation in the ratio with viscosity is almost the same for glycerol-water, methanol-water, and sucrose-water solutions irrespective of their differences in polarity 关Fig. 2共a兲兴. However, the plots of the ratio against dielectric constant for the two viscous systems 共glycerolwater and sucrose-water兲 are distinct from the plot for the less viscous methanol-water medium even in the water-rich solutions 关Fig. 2共b兲兴. This behavior indicates the predominance of viscosity dependence in media with polarity that is not very different from that of water. The polarity dependence is secondary in nature. B. Quantum chemical calculations in the ground and first excited singlet states Quantum chemical calculations have been performed in order to develop a better rationalization of the experimental results. As the viscosity dependence is expected to have arisen due to the involvement of the rotation of the ring systems relative to each other, conformational analysis has been performed on the ground state and on the first excited state of the cation and the first excited state of the tautomer. The energy of the molecule is computed while changing the dihedral angle 共N1–C2–C3–N4兲 of the cation, setting the energy associated with zero dihedral angle to zero. The geometries are optimized in the range of dihedral angle from ⫺180° to +180°. The conformation in which the two rings are coplanar is the most stable one in all the three cases 共Fig. 3兲. In the case of the S0 and S1 states of the cationic form 共C兲, the conformations with dihedral angle of zero and those with dihedral angles of ⫾180° are degenerate. However, this is not the case with the tautomer S1 state, where the conformation with dihedral angle of zero is more stable than those with dihedral angle of ⫾180° by 8 kcal mol−1. This is explained easily in the light of steric interaction between the N–H bonds in the latter and the absence thereof in the former configuration. In the case of C and Cⴱ, such steric interaction does not exist as both the H atoms are on the nitrogen atoms of the benzimidazole moiety, thereby imparting C2v symmetry to it. From the conformational analysis, the rotational J. Chem. Phys. 131, 034504 共2009兲 FIG. 3. Energetics for the rotation of cationic form of 2PBI; in the ground state 共S0兲 共a兲 using B3 LYP/ 6-31Gⴱ and 共b兲 in first excited singlet state 共S1兲 using CIS/ 6-31Gⴱ level of theories and 共c兲 that of tautomer form of 2PBI in first excited singlet state 共S1兲 using CIS/ 6-31Gⴱ level of theories. The geometry has been optimized at each value of dihedral angle. The energy at zero dihedral angle is set to zero. barriers have been determined in each case. These barriers are 8.5 kcal/mol in the ground state and 17.7 kcal/mol in the first excited singlet state 共Fig. 3兲, indicating that the rotation around the pivotal bond of Cⴱ in its excited state is significantly restricted than that in the ground state. This could be due to the double bond character of pivotal bond in excited state as the C2–C3 bond distances are found to be 1.46, 1.39 Å in the ground state and excited state geometries of the molecule, respectively. In highly viscous media, the rotation of two aromatic rings around the pivotal bond of 2PBI can get further restricted in the first excited single state as well as in the ground state as the movement of the rings with respect to each other is expected to be sluggish. Hence the barrier is expected to increase in media of higher viscosity. Such viscosity dependent barrier to intramolecular rotational motion have been reported for the photoisomerization of 3,3⬘-diethyloxadicarbocyanineiodide.33–35 Notably, the barrier is even greater for Tⴱ, indicating that it would be stabilized even further than Cⴱ in a viscous medium. Thus, the greater yield of the tautomer in viscous media may be rationalized in the light of the higher barrier to rotation. Moreover, the dynamics of the ESPT process may be expected to be slower in more viscous media if it is coupled with the rotation of the two rings with respect to each other. An absence of viscosity dependence in the dynamics would indicate that such rotation is not involved in the rate determining step. In order to understand the effect of viscosity, if any, on the dynamics of the process, time resolved fluorescence studies have been performed. C. The slower dynamics of the excited states The fluorescence decays are recorded over the pH range of 1–3 in neat water. They are also recorded at pH = 3 at Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-5 J. Chem. Phys. 131, 034504 共2009兲 Ultrafast dynamics of ESPT in 2PBI TABLE II. Temporal characteristics of 2PBI in pH = 3 containing different percentages of sucrose 共w/w兲 glycerol 共v/v兲 in pH = 3 solution. The data in the solution with pH = 1 are also shown in this table. These data are obtained from the TCSPC experiment. 380 nm 460 nm % glycerol 共v/v兲 1 共ns兲 2 1 共ns兲 2 共ns兲 a1 a2 2 0 20 30 50 0.83 0.74 0.64 0.58 1.07 1.13 1.17 1.12 0.80 0.56 0.58 0.55 1.50 2.21 2.44 2.96 ⫺0.44 ⫺0.42 ⫺0.38 ⫺0.40 1.44 1.42 1.38 1.40 1.05 1.16 1.02 1.08 % sucrose 共w/w兲 0 10 14 28 35 0.86 0.74 0.67 0.68 0.64 1.07 1.01 1.02 1.05 1.10 0.86 0.80 0.79 0.69 0.67 1.38 1.82 2.14 2.40 2.65 ⫺0.48 ⫺0.51 ⫺0.45 ⫺0.39 ⫺0.42 1.41 1.48 1.45 1.39 1.42 1.09 1.05 1.01 1.10 1.05 pH = 1 0.38 1.05 0.38 1.71 ⫺0.41 1.41 1.10 different glycerol and sucrose contents 共Figs. 4 and 5兲. The decays at pH = 2 – 3 are indistinguishable with a lifetime of 0.70 ns at em = 380 nm, while the decay at pH = 1 is faster 共Table II兲. This is in agreement with the previously reported trends.8 The lifetime at em = 380 nm decreases at higher concentration of glycerol and sucrose. At em = 460 nm, the trace is biexponential with a rise time of 0.80 ns and decay time of 1.60 ns at pH = 3. The initial part of the decays is superimpossible 关Fig. 4共a兲兴. This might indicate that the dynamics of the ESPT is not dependent on the viscosity of the medium unlike in hypocrellin and calphostin C.20,21 Moreover, the rise times of the tautomer emission are more or less similar to the decays of the cation emission in all the cases. On a different note, the tautomer lifetime increases by a factor of 2 at 50% glycerol 共v/v兲 and 35% sucrose 共w/w兲 solution 共Table II兲. Thus, the phototautomer is found to be stabilized to a greater extent in viscous solutions as has been predicted in the quantum chemical calculation. This is manifested further in the variation in log kNR with viscosity 共Fig. 6兲. A decrease is observed at em = 460 nm with an increase in viscosity, whereas at em = 380 nm, a slight increase in the saturation is observed. Thus the phototautomer is indeed found to be stabilized, while the excited state of the cation is found to be destabilized with increased in viscosity. This observation bolsters the contention proposed from the steady state results. The time resolved area normalized emission spectra 共TRANES兲 in 35% sucrose solution of pH = 3 pass FIG. 5. The fluorescence decays of 2PBI in aqueous solution pH = 3 containing 共i兲 0% and 共ii兲 35% sucrose solution at 共a兲 em = 380 and 共b兲 460 nm. 共c兲 The fluorescence decays of 2PBI in aqueous solutions at pH = 1 at 380 and 460 nm: ex = 294 nm in all cases. FIG. 6. Variation in the logarithm of the nonradiative rate 共log kNR兲 of 2PBI with the viscosity of aqueous sucrose solutions at pH = 3: em = 380 nm 共upper panel兲 and 460 nm 共lower panel兲. FIG. 4. 共a兲 The fluorescence decays of 2PBI in aqueous solution pH = 3 containing 共i兲 0% and 共ii兲 50% glycerol: em = 380 nm and em = 460 nm. 共b兲 The fluorescence decays of 2PBI in deuterated oxide at pH = 3 at em 380 and 460 nm: ex = 294 nm in all cases. Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-6 J. Chem. Phys. 131, 034504 共2009兲 Burai et al. TABLE III. Temporal characteristics of 2PBI in H2O and D2O at pH = 3. The acid used is HClO4 in the aqueous solution and DClO4 in the D2O solution. 380 nm H 2O D 2O 460 nm 1 共ns兲 2 1 共ns兲 2 共ns兲 a1 a2 2 0.83 2.66 1.07 1.13 0.80 2.70 1.50 4.15 ⫺0.44 ⫺0.38 1.44 1.38 1.05 1.06 through an isoemissive point. This indicates that the ESPT is a two-state process in the longer time scale.36 The slow conversion to the tautomer indicates a large energy barrier. To investigate the effects of possible tunneling and to see the kinetic isotope effect, time resolved measurements have been performed in D2O 关Fig. 4共b兲, Table III兴. At em 460 nm, the rise time is 2.7 ns against the value of 0.7 ns in H2O. In both the cases, the rise times are approximately equal to the corresponding lifetimes of monocation at em 380 nm, increasing from 0.7 ns in aqueous solution at pH = 3 − 2.7 ns in deuterated solvent at pH = 3. This indicates that the decay of the 380 nm emission state should be directly correlated with the rise of the 460 nm state in this time scale. An evidence of tunneling is not observed 共see Fig. 7兲. D. Ultrafast dynamics The ultrafast fluorescence dynamics of 2PBI have been investigated by the femtosecond fluorescence upconversion at em = 380 and 460 nm. Aqueous solutions of relatively mild acidity 共pH = 0.5– 4.5兲 as well as stronger acidity 共molar concentration of HClO4兲 have been studied in this experiment. The decays in aqueous solution at pH = 7 are superimpossible with the initial parts of the decays obtained by TCSPC and yield the same value of the time constants and amplitudes 共Table II兲. However, an ultrafast component of 2 ps emerges at lower pH values for em = 380 nm 共Fig. 8, Table IV兲. The amplitude of this ultrafast component increases with the increase in acidity until it accounts for more than 50% of the decays at 2M HClO4. This ultrafast component has not been reported in earlier studies. Since its con- FIG. 7. TRANES of 2PBI in 35% sucrose solution of pH = 3 between time 0 and 3.0 ns at intervals of 0.5 ns. The arrows indicate the direction of an increase in time. FIG. 8. Femtosecond fluorescence transients of 2PBI at em = 380 nm in aqueous solution of 共a兲 pH = 7, 共b兲 pH = 3, and 共c兲 containing 0.5M and 共d兲 containing 2M HClO4: ex = 283 nm. The solid lines denote the lines of best fit. The fitting parameters are provided in Table II. The cross-correlation function of the laser pulse is shifted for the sake of clarity. tribution increases with increase in acidity, it is likely to be a signature of the dication. However, the dication is known to be nonfluorescent and the pKa for the protonation of the monocation is ⫺1.2.7 So, it is unlikely that the 2 ps decay is due to the decay of the excited state of the dication. However, the pKⴱa for the second protonation of 2PBI is 3.19 as estimated by the Forster cycle method.7 So, Cⴱ is expected to get protonated in the pH range studied to form Dⴱ. This excited state protonation process is expected to be ultrafast and should contribute to the depletion of the Cⴱ emission. So, the 2 ps component is ascribed to the decay of Cⴱ to Dⴱ, which provides an independent depletion pathway for Cⴱ. Another additional feature is observed in the transients in the range of the Cⴱ emission in the form of a 100 ps rise time 关Fig. 9共a兲, Table IV兴, which disappears above acid concentration of 0.5M. This rise time persists over the entire range of Cⴱ emission even though at wavelengths such as 390 nm it is difficult to resolve due to the presence of contribution from the longer decay of Cⴱ as well as the long rise and decay of Tⴱ. This hitherto unreported rise time indicates the temporal evolution of the cationic excited state. The initially prepared Cⴱ state seems to undergo either a structural reorganization of a charge redistribution37 or both, before a modified state, C1ⴱ, is formed and it is this C1ⴱ state that appears to undergo the actual ESPT process and yield Tⴱ eventually. The following hypothesis may be proposed regarding the nature of Cⴱ and C1ⴱ: Due to the small activation energy of rotation of the two rings with respect to each other, Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-7 J. Chem. Phys. 131, 034504 共2009兲 Ultrafast dynamics of ESPT in 2PBI TABLE IV. Early time decay parameters of 2PBI obtained from the fluorescence upconversion experiment. The lifetimes have been classified in the columns according to their values. em 共nm兲 pH = 7.0 pH = 4.5 pH = 3.0 0.5M HClO4 2.0M HClO4 pH = 3 in D2O. The acid used is DClO4. 380 380 360 380 390 460 490 380 380 360 370 380 390 460 490 1 共ps兲 2 共ps兲 3 共ps兲 2 2 2 2 86 80 100 100 130 100 700 700 700 700 1000 1800 1650 2 2 2 2 2 2 198 48 87 220 162 127 2700 2920 2600 2600 2600 2600 4150 4000 4 共ps兲 a1 a2 a3 a4 0.06 0.31 0.17 0.28 0.90 0.45 ⫺0.19 ⫺0.26 ⫺0.36 ⫺0.04 0.10 0.49 0.88 1.09 1.08 1.44 1.50 ⫺0.40 ⫺0.50 800 1000 0.40 0.58 0.53 0.46 0.37 0.42 0.6 0.42 ⫺0.15 ⫺0.23 ⫺0.22 ⫺0.25 ⫺0.38 ⫺0.48 0.62 0.76 0.85 0.83 1.38 1.48 C is likely to exist in a distribution of conformations. Vertical excitation leads to the formation of Cⴱ in a distribution of conformations as well. However, in the excited state, the coplanar conformation is relatively more stable. So, conformational rearrangements occur to lead to a narrower distribution of conformation of Cⴱ and this is what has been referred to as the C1ⴱ state in this discussion. In this argument, the 100 ps rise in the cation emission would be the time constant of the conformational rearrangement. The argument proposed in the previous paragraph finds some support in the early time fluorescence dynamics in 35% sucrose solution 共Fig. 9兲. In this medium, the contribution of the 100 ps rise time at 380 nm is substantially less than that in the neat aqueous solution. This is what one might expect in a more viscous medium, where the barrier to rotation would increase and the process of rotation is hindered. Thus, from the fluorescence upconversion experiment, it appears that the 100 ps rise time may be assigned to the conformational relaxation of the Cⴱ state. On a different note, the transients at 460 nm in the presence and absence of sucrose are superimpossible 关Fig. 9共b兲兴. This indicates that the viscosity of the media has no role in the slow dynamics of the ESPT of 2PBI even though the extent of ESPT is greater in more viscous media due to the greater stability of Tⴱ in such media. Time-resolved fluorescence anisotropy decays r共t兲 shown in Fig. 10 of 2PBI in pH = 3 and 35% sucrose solution are measured at em = 380 nm and at em = 460 nm in order to better understand the mechanism of the rise time of tautomer. The decays fit well monoexponentially. Fundamental anisotropy at 380 and 460 nm are 0.22 and 0.10, respectively. The angle 共兲 between absorption and emission transition FIG. 9. Femtosecond fluorescence transients at pH = 3 in the absence and presence of 35% of sucrose. 共a兲 em = 380 共b兲 460 nm. ex = 283 nm. The solid lines denote the lines of best fit. FIG. 10. Fluorescence anisotropy decay at pH = 3 in the absence 共a兲 and presence 共b兲 of 35% of sucrose. em = 380 and 460 nm. ex = 283 nm. The solid lines denote the best fit to the experimental data. Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 034504-8 dipole moment is calculated using Eq. 共3兲. These angle values are 33° and 45° at 380 and 460 nm, respectively. The orientation time constant of approximately 50 ps of 2PBI in water remains the same in pH = 3 solution as well. However, in the presence of 35% sucrose solution 共 = 4 mPa s兲 in pH = 3, this increases to 135 ps at em = 380 nm and 280 ps at em = 460 nm presumably because the tautomer form is more stable in high viscosity compared to cation of 2PBI due to the hindrance in the rotation of pivotal bond. Notably, the r共0兲 value in the region of Cⴱ emission is considerably less than the expected value of 0.4, indicating a missed ultrafast component, which is likely to be associated with the early event of the conformational relaxation. IV. CONCLUSION The phototautomer of 2PBI is found to be stabilized and the excited state of the cation is destabilized at higher viscosity. Consequently, the ratio of the relative quantum yields of the tautomer and the normal species increases with an increase in viscosity, as does the ratio of lifetimes. In the acidic solutions, Cⴱ is found to have an ultrafast decay channel in the formation of the excited state of the nonfluorescent dication, in addition to ESPT. Moreover, the ESPT is found to be preceded by a faster process, which is likely to be conformational relaxation of Cⴱ. The dynamics of ESPT is found to be independent of viscosity and the formation of Tⴱ is correlated directly with the decay of conformationally relaxed Cⴱ. ACKNOWLEDGMENTS This work was supported by Grant No. SR/S1/PC-19/ 2005 of SERC, DST. T.N.B. thanks CSIR for a senior Research fellowship. The authors are grateful to Professor G. Naresh Patwari for a useful suggestion and Mr. E. Siva Subramaniam Iyer for critical reading of the manuscript. The referees are thanked for insightful comments. 1 J. Chem. Phys. 131, 034504 共2009兲 Burai et al. M. Carmen Ríos Rodríguez, M. Mosquera, and F. Rodriguez-Prieto, J. Phys. Chem. A 105, 10249 共2001兲. 2 R. S. Flom and F. P. Barbara, Chem. Phys. Lett. 94, 488 共1983兲. 3 C. Chudoba, E. Riedle, M. Pfeiffer, and T. Elsaesser, Chem. Phys. Lett. 263, 622 共1996兲. 4 M. Carmen Ríos Rodríguez, F. Rodriguez-Prieto, and M. Mosquera, Phys. Chem. Chem. Phys. 1, 253 共1999兲. 5 S. Takeuchi and T. Tahara, J. Phys. Chem. A 102, 7740 共1998兲. 6 J. Catalan, G. L. J. De Paz, J. C. Del Valle, M. R. Claramount, and T. Mas, Chem. Phys. 305, 175 共2004兲. 7 J. C. Penedo, J. L. P. Lusters, I. G. Lema, M. C. R. Rodríguez, M. Mosquera, and F. Rodríguez-Prieto, J. Phys. Chem. A 108, 6117 共2004兲. 8 F. Rodríguez-Prieto, M. Mosquera, and M. Novo, J. Phys. Chem. 94, 8536 共1990兲. 9 M. C. Rath, D. K. Palit, and T. Mukherjee, J. Chem. Soc., Faraday Trans. 94, 1189 共1998兲. 10 T. K. Mukherjee and A. Datta, J. Phys. Chem. B 109, 12567 共2005兲. 11 T. K. Mukherjee, D. Panda, and A. Datta, J. Phys. Chem. B 109, 18895 共2005兲. 12 T. K. Mukherjee and A. Datta, J. Phys. Chem. B 110, 2611 共2006兲. 13 K. Bhattacharyya and B. Bagchi, J. Phys. Chem. A 104, 10603 共2000兲. 14 S. Balasubramanian, S. Pal, and B. Bagchi, Phys. Rev. Lett. 89, 115505 共2002兲. 15 P. Dutta, P. Sen, S. Mukherjee, and K. Bhattacharyya, Chem. Phys. Lett. 382, 426 共2003兲. 16 T. Elsaesser, in Ultrafast Hydrogen Bonding and Proton Transfer Processes in the Condensed Phase, edited by T. Elsaesser and H. J. Bakker 共Kluwer, Dordrecht, 2002兲, pp. 119–153. 17 M. Rini, B. Z. Magnes, E. Pines, and E. T. J. Nibbering, Science 301, 349 共2003兲. 18 O. F. Mohammed, D. Pines, J. Dreyer, E. Pines, and E. T. J. Nibbering, Science 310, 83 共2005兲. 19 J. L. Pérez-Lustres, F. Rodriguez-Prieto, M. Mosquera, T. A. Senyushkina, N. P. Ernsting, and S. A. Kovalenko, J. Am. Chem. Soc. 129, 5408 共2007兲. 20 K. Das, D. S. English, and J. W. Petrich, J. Am. Chem. Soc. 119, 2763 共1997兲. 21 A. Datta, A. V. Smirnov, J. W. G. Chumanov, and J. W. Petrich, Photochem. Photobiol. 71, 166 共2000兲. 22 B. L. McClain, I. J. Finkelstein, and M. D. Fayer, J. Am. Chem. Soc. 126, 15702 共2004兲. 23 M. M. G. Krishna and N. Periasamy, J. Fluoresc. 8, 81 共1998兲. 24 D. R. Lide, Handbook of Chemistry and Physics, 79th ed. 共CRC, Boca Raton, 1998兲, Chap. 8, p. 65. 25 J. R. Lakowicz, Principle of Fluorescence Spectroscopy, 3rd ed. 共Springer, New York, 2006兲. 26 C. Lee, W. Yang, and R. G. Parr, Phys. Rev. B 37, 785 共1988兲. 27 M. J. Frisch, G. W. Trucks, H. B. Schlegel et al., GAUSSIAN 03, Revision C.02, Gaussian, Inc., Wallingford, CT, 2004. 28 J. B. Foresman, M. Head-Gordon, J. A. Pople, and M. J. Frisch, J. Phys. Chem. 96, 135 共1992兲. 29 G. Akerlof, J. Am. Chem. Soc. 54, 4125 共1932兲. 30 J. B. Birks, Photophysics of Aromatic Molecules 共Wiley, New York, 1970兲, p. 88. 31 P. F. Aramendia, R. M. Negri, and E. S. Roman, J. Phys. Chem. 98, 3165 共1994兲. 32 P. K. Sengupta and M. Kasha, Chem. Phys. Lett. 68, 382 共1979兲. 33 J. M. Hicks, M. Vandersall, E. V. Sitzmann, and K. B. Eisenthal, Chem. Phys. Lett. 135, 413 共1987兲. 34 S. P. Velsko and G. R. Fleming, Chem. Phys. 65, 59 共1982兲. 35 D. Waldeck and G.R. Fleming, J. Phys. Chem. 85, 2614 共1981兲. 36 A. S. R. Koti, M. M. G. Krishna, and N. Periasamy, J. Phys. Chem. A 105, 1767 共2001兲. 37 D. B. Spry and M. D. Fayer, J. Chem. Phys. 128, 084508 共2008兲. Downloaded 01 Mar 2012 to 59.162.23.76. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions
© Copyright 2025 Paperzz