Joint spectroscopic and theoretical investigation of

0 downloads 0 Views 3MB Size Report
the merocyanines are very famous for their photochromic behavior,11–14 some ... the dying of textiles,20,21 or in optical filters,22 and it bears a high absorption ...
PCCP View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

PAPER

Cite this: DOI: 10.1039/c4cp05103c

View Journal

Joint spectroscopic and theoretical investigation of cationic cyanine dye Astrazon Orange-R: solvent viscosity controlled relaxation of excited states† Christian Ley,*a Patrice Bordat,b Luciano H. di Stefano,a Lola Remongin,a Ahmad Ibrahim,a Patrice Jacquesa and Xavier Allonasa

Received 4th November 2014, Accepted 12th January 2015

In this paper, the first study of cationic cyanine dye Astrazon Orange-R by combined spectroscopic and

DOI: 10.1039/c4cp05103c

theoretical investigation is presented. It is shown that molecular modeling of Astrazon Orange-R is in very good agreement with experiment, allowing us to gain insight into its complicated photophysics. A solvent

www.rsc.org/pccp

viscosity controlled relaxation of excited states, involving cyanine isomerization, is also outlined.

Introduction Nowadays, photochemistry is a popular topic, with good and clean control of the chemical processes without the need for thermal heating. The development and optimization of radical generator systems, which use light is important and has applications in many different fields such as cosmetics, medicals, microelectronics, optics, and photoresists.1–6 The first step in photochemistry is the absorption of photons by molecules. Thus, in order to enhance sensitivity, especially if low intensity sources are used, molecules bearing high molar absorption coefficients are of prime interest.7,8 Moreover, the ability to work with low dye concentrations reduces the cost and environmental impact. Among commonly available dyes with low cost, the cyanine family presents interesting molecules bearing high absorption coefficients in the visible light region.9,10 While the merocyanines are very famous for their photochromic behavior,11–14 some other cyanines present interesting reactivity in the field of radical photogeneration.15–18 After light absorption, the dyes go into their excited state from where they can deactivate towards triplets states, begin a fragmentation reaction, react with other molecules either by energy transfer or electron transfer. These mechanisms are very important because they determine the reactivity and efficiency of the photochemistry of the dye.7,8,19 The molecule studied herein, Astrazon Orange-R (AO-R), is from the family of cationic cyanine dyes. AO-R is used for example in a

LPIM, University of Haute-Alsace, 3 rue A. Werner, 68200 Mulhouse, France. E-mail: [email protected] b Institut des Sciences Analytiques et de Physico-Chimie sur l’Environnement et les Mate´riaux, IPREM, UMR 5254 du CNRS et de l’Universite´ de Pau et des Pays de l’Adour, Avenue du Pre´sident Angot, 64053 Pau cedex, France † Electronic supplementary information (ESI) available: Fig. S1–S3 and Table S1. See DOI: 10.1039/c4cp05103c

This journal is © the Owner Societies 2015

the dying of textiles,20,21 or in optical filters,22 and it bears a high absorption coefficient and a solvent dependent fluorescence emission. However, even though the photophysics of the spiropyran ring opening to merocyanine is well documented,23–31 it appears that the challenging photophysics of the excited states of the corresponding open form, merocyanine, gained interest only recently.32–34 Moreover, while some cyanine dyes have been studied for decades,35–37 to our knowledge, the photophysics of AO-R, which has the advantage that only one isomer is populated in the ground state, has not yet been reported. Therefore, we present the first experimental and computational study of this molecule with the focus on its photophysics. It is outlined that part of the photophysics is governed by the viscosity of the medium and that the molecule does not exhibit a long lived excited triplet state. Thus, its photochemical reactivity arises from its singlet excited states.

Methods Experimental section Steady state UV-visible spectra were obtained on a Cary 4000 spectrophotometer. Fluorescence spectra were performed with a Horiba-Jobin-Yvon fluoromax 4. The emission, excitation and absorption spectra of AO-R were recorded in all the solvents mentioned in this paper. Before comparison of the measured spectra, they were all corrected in order to take into account the non reciprocity of absorption and emission phenomena according to38,39 a transition dipole moment representation (TDM). The femtosecond laser excitation wavelength was adjusted to 500 nm using a collinear 800 nm pumped CDP2017 (CDP Corp.) optical parametric amplifier. 100 fs laser pulses (800 nm) were provided by a Spectra-Physics Tsunami Ti:Sa oscillator coupled to a Spitfire pro Spectra-physics regenerative amplifier.

Phys. Chem. Chem. Phys.

View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

Paper

Pump–probe measurements were performed on a CDP Excipro system. The resulting pump–probe cross-correlation of the setup was found to be about 200 fs. Astrazon Orange-R was purchased from TCI, its structure was checked by NMR, and used as received. Solvents of spectrophotometric or HPLC grade were purchased from Sigma-Aldrich and used without further purification. Laser flash photolysis (LFP) experiments were realized with a continuum Surelite YAG laser coupled to a continuum optical parametric oscillator SLOPO. The excitation pulses were adjusted to 500 nm. LFP kinetics were recorded on an Applied Photophysics LKS80 system.

PCCP

Fig. 1

Molecular structure of Astrazon Orange-R.

Computational methods Quantum mechanical (QM) calculations have been carried out using Gaussian 09.40 Standard calculations for structural optimization were performed at the B3LYP/6-311G** level and electronic transitions (both absorption and emission) have been calculated by TDDFT (time-dependent density functional theory) at the same level of consistency. Fluorescence emissions were carried out by calculating the relaxation of the excited state geometry followed by the emission state-specific solvation and by the emission to the final ground state41,42 as implemented in Gaussian 09. The solvent was taken into account with the PCM model implemented in Gaussian 09.43 Vibrational frequencies have been checked to ensure that optimized conformations correspond to energy minima. Moreover, extra-calculations were performed with CAM-B3LYP, wB97XD and M062X DFT functionals and extended bases as 6-311++G(3df,3dp) and cc-pVTZ. These extra-calculations confirm the results found with the previous modest strategy (B3LYP/6-311G**). Thus, in the following discussion, interpretation is based on the results determined at the B3LYP/6-311G** level. Charge distribu¨lliken population analysis and the tion was also investigated by Mu electrostatic potential fit methods.

Results Spectroscopic properties and solvent effect The molecular structure of Astrazon Orange-R (AO-R) is given in the following Fig. 1. It is a positively charged dye and is purchased as a chloride salt. The evolution of the UV-visible spectra of the dye in water, dimethyl sulfoxide (DMSO), acetonitrile (MeCN), acetone and ethanol (EtOH) as a function of AO-R concentration was checked. All the spectra present the same shape and evolve homothetically. This indicates that no aggregation occurs in the concentration range that was used (i.e. up to 2  104 M). Regardless of the solvent used, two bands could be observed: a very intense one with a visible maximum around 500 nm (see Fig. 2) and a small one in the UV region below 300 nm. Furthermore, this allows the determination of molar absorption coefficients in the investigated solvents (see Table 1).44,45 It should be mentioned here that the dye was not soluble in low polar and non polar solvents such as THF, toluene and cyclohexane.

Phys. Chem. Chem. Phys.

Fig. 2 TDM representation of absorption (plain line), emission (dotted line) and excitation (circles) spectra of AO-R in EtOH.

Table 1 Dielectric constants D, p* Kamlet and Taft polarity parameter of solvents,46,47 maximum absorption wavelength lmax in nm and measured molar absorption coefficients of AO-R

H2O DMSO MeCN Acetone EtOH

D

p*

lmax (nm)

e (L mol1 cm1)

78.36 46.45 35.94 20.56 24.55

1.09 1.00 0.75 0.71 0.54

493 499 496 496 499

46 424 45 912 51 517 55 561 58 192

All the molar absorption coefficients are quite high and close to 50 000 L mol1 cm1 (and up to 60 000 in EtOH), in good agreement with molecular calculations (vide infra). Moreover, a small hypochromic effect is observed in the higher p* solvents. Singlet state photophysics Some low intensity emission was detected in each solvent in which the dye is soluble. The obtained transition dipole moment (TDM) representations of AO-R absorption, emission and excitation spectra in EtOH are displayed in Fig. 2. Good mirror image symmetry can be seen between the absorption and emission spectra. Moreover, excitation and absorption spectra overlap almost perfectly, indicating that the excited state, reached after light absorption, should be very similar to the emissive state.48,49 The same behavior was obtained in the other

This journal is © the Owner Societies 2015

View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

PCCP

Paper

solvents and the maximum absorption and emission values, together with other parameters can be found in Table 2. The position of the emission and absorption maxima are not correlated to the solvent polarity as expected for a cationic molecule. Thus, it seems that neither polarity, nor hydrogen bonds play an important role in the steady state spectroscopic properties. During fluorescence experiments, it appeared that the intensity of fluorescence increased with the solvent viscosity. Therefore, quantum yields of fluorescence were measured in all solvents (with Rhodamine 6G in EtOH as a reference48). All quantum yields are very low (see Table 2), in agreement with the low emission intensity observed. However, it can be seen that the quantum yield increases when going from a low viscosity solvent (e.g. acetone) to a more viscous ones (e.g. DMSO) and even more in triethylene glycol (TEG). Knowing both the quantum yield of fluorescence and the singlet state lifetime, the determination of the natural radiative lifetime trad was possible according to: Ff = t(S1)/trad.49 The calculated values are listed in Table 2. The radiative lifetimes in the different solvents are very similar and in the range of 3 to 5 ns (the deviations being easily explained by the possible experimental errors on the quantum yields). The relatively long radiative lifetime in all the solvents associated to a very short lifetime of the first singlet state provides evidence for an excited state deactivation controlled by internal conversion (IC), intersystem crossing (ISC), charge transfer (CT) and isomerization.44,48–50 However, the viscosity of the solvents seems to influence the photophysics of AO-R. Indeed, the higher emission quantum yields and longer lifetimes are obtained in the most viscous DMSO and TEG solvents. This last remark indicates the involvement of structural changes in the present dye rather than internal charge transfer/ recombination mechanisms due to the absence of dependence on the solvent polarity.51,52 Primary events in AO-R As the photophysics of the AO-R singlet state is fast, femtosecond pump–probe spectroscopy experiments were used. Typical spectro-temporal data are displayed in Fig. 3. The excitation was realized at 500 nm in order to excite the molecule at not too high vibrational levels (to prevent internal cooling of the molecule) and to gain visibility in the stimulated emission (SE) region.

In all the solvents the spectra show similar trends. In the blue region (i.e. up to 500 nm) of the time resolved spectra it is possible to observe: (i) a positive band centered around 400 nm whose amplitude monotonically decreases down to zero, (ii) the negative bleaching band around 490 nm matching the steady state absorption of the dye (recalled as a black line in the bottom of the figure), and (iii) that the bleaching never goes back to zero. The red region of the spectra exhibits a more complicated evolution: (i) at a short time length, we can observe the stimulated emission band, which is immediately overlapped by a growing positive peak around 530 nm, (ii) at a longer time scale, the positive 530 nm band reaches a final positive value while the SE band continues to increase up to zero with apparently the same kinetics as the positive 400 nm absorption band (see inset Fig. 4), and (iii) together with the permanent negative bleaching, this is a strong indication of the fast formation of a photoproduct, which does not return back to the ground state within the time window of our experiment (o4 ns). In order to get more insight into the kinetics, a global multiexponential analysis of the data was performed by singular value decomposition (SVD)53–55 and by fitting the first three extracted orthogonal kinetics (KOP1 to 3), which are displayed in Fig. 4. Single wavelength kinetics analysis was also performed and single wavelength kinetics at 400, 525 and 575 nm can be seen in the inset of Fig. 4. The best fits of the KOP required a sum of two exponentials plus a step function. It must be noted that a delayed growth of the KOP is observed in all the solvents. Moreover, the same delayed growth is also observed in the single wavelength kinetics. These last remarks indicate a sub 100 fs phenomenon not fully resolved by the experimental setup. The time constants obtained by the global analysis, t(SVD1) and t(SVD2) are given in Table 3 (corresponding decay associated differential spectra (DADS) are displayed in Fig. S1, ESI†). The kinetics at 400 and 575 nm were simultaneously fitted with a monoexponential function by sharing the same time constant, t(400/575), while the kinetics at 525 nm were separately fitted and needed up to two time constants t(5251) and t(5252) in DMSO and water due to non-exponential behavior. All time constants are listed in Table 3. From Table 3 it is obvious that the long time constant, t(SVD2), depends on the viscosity of the solvent going from around

Table 2 Photophysical experimental parameters of AO-R: maximum absorption, emission and Stokes shift (Dn) after TDM correction (cm1), singlet state energy E(S1) (eV), fluorescence quantum yield Ffluo, singlet state and natural radiative lifetimes t(S1) and trad (in ps and ns, respectively), dielectric constant D and solvent viscosity Z (cP)

Solvent

nabs (cm1)

nfluo (cm1)

Dn (cm1)

E(S1) (eV)

Ffluo

Acetone MeCN H2O EtOH DMSO TEG

20 141 20 165 20 285 20 000 20 022 19 932

18 653 18 620 18 553 18 556 18 465 18 581

1488 1545 1732 1444 1557 1351

2.38 2.41 2.38 2.39 2.38 2.38

2.6 2.4 3.3 4.5 11.8 31.1

     

103 103 103 103 103 103

t(S1)a (ps)

trad (ns)

D

Z(cP)

9.15 7.6 9.8 13.2 35.8 154

3.6 3.2 2.95 3.5 3.0 4.9

20.56 35.94 78.36 24.55 46.45 23.69

0.31 0.37 0.89 1.07 1.99 49

a

Time Correlated Single Photon Counting (TCSPC) experiments were performed in order to determine the fluorescence lifetime. Unfortunately, the lifetimes were shorter than the time response of the TCSPC module. Therefore, the reported lifetimes in Table 2 were obtained by femtosecond pump–probe experiments (vide infra).

This journal is © the Owner Societies 2015

Phys. Chem. Chem. Phys.

View Article Online

Paper

PCCP

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

Table 3 Kinetic parameters (time constants in ps) obtained by femtosecond pump–probe spectroscopy

Z(cP) t(SVD1) t(SVD2) t(400/575) t(5251) t(5252)

Fig. 3 Pump–probe time resolved femtosecond spectroscopy of AO-R in MeCN. (Up) short time scale from 0.2 to 10 ps. (Down) longer time scale from 10 ps to 1.5 ns. The steady state absorption and emission spectra are recalled and are shown by the black lines.

Acetone

MeCN

H2O

EtOH

DMSO

TEG

0.31 5.42 9.55 9.15 5.32

0.37 5.69 8.56 7.56 4.00

0.89 2.54 11.3 9.8 3.4 12.4

1.07 8.14 16.4 13.2 8.12

1.43 1.43 37.4 35.8 2.01 27.8

49 11.8 156 153 110

The single wavelength kinetics at 400 and 575 nm exhibit the same behavior and decay homothetically. This is confirmed by fitting the decays: a single exponential function with the same time constant, t(400/575), was enough to fit simultaneously both wavelength kinetics. This strongly indicates that the transient absorption around 400 nm originates from the singlet excited state of AO-R. Moreover, this monoexponential behavior exhibits a time constant, t(400/575), which is very close to t(SVD2) and bears the same viscosity dependence with solvent. Thus, in line with fluorescence observations and quantum yields measurements, t(SVD2) reflects the lifetime of an emissive singlet excited state of AO-R. In addition, it can be seen that this lifetime increases with the solvent viscosity (see Table 3). The rising band around 530 nm is more difficult to interpret and exhibits strong non exponential behavior in DMSO and water. We can probably argue that the overlap of absorption and emission bands of different species induces complex kinetics. However, it is clear that the rise time of this absorption is not correlated to the singlet state lifetime, as it grows much faster than the S1 state decays (see inset in Fig. 4). Although, the fitting time constants given in Table 4 are indicative, they underline a less viscositydependent kinetics contrarily to t(SVD2). Moreover, the signal becomes positive and permanent in the 3–4 ns timescale of the femtosecond setup (i.e. a step function) which together with the permanent bleaching indicates the formation of a stable photoproduct. The time constant obtained from the fits are very close to t(SVD1) and one could postulate that t(SVD1) mainly corresponds to the time formation of the photoproduct, while t(SVD2) is related to the lifetime of AO-R excited singlet state S1. Laser flash photolysis

Fig. 4 Orthogonal kinetics (KOP) of AO-R in MeCN obtained by SVD of the spectro-temporal data. Inset: single wavelength at 400, 525 and 575 nm.

9 ps in lower viscosity acetone and MeCN to a 153 ps in the highly viscous TEG. This behavior is in line with the increasing intensity of emission observed in fluorescence experiments. It can also be noticed that the short t(SVD1) is very close to t(5251).

Phys. Chem. Chem. Phys.

In order to indentify the photoproduct at the ns–ms timescale, laser flash photolysis (LFP) experiments were performed in Ar saturated solution. In Fig. 5, the ms LFP transient spectra of AO-R in Ar saturated MeCN can be seen. They are very similar in shape to the 1.54 ns spectra obtained in femtosecond pump–probe experiments: on the red side of the permanent bleaching, a positive absorption band can be observed. The kinetics of the bleaching (at 450 nm) and the transient (at 550 nm) are given in the inset of the Fig. 5. A long lived transient was detected with lifetime longer than 180 ms. In order to check the possible triplet state nature of this transient the same measurements were performed under air: neither the kinetics nor the spectra were modified by the presence of oxygen in the solution. At this stage, we can claim that the formation of a triplet state from AO-R is excluded.

This journal is © the Owner Societies 2015

View Article Online

PCCP

Paper

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

Table 4 First electronic transition (nm) and oscillator strength (in brackets) for all conformers in vacuum, acetone, acetonitrile, methanol (MeOH), ethanol, DMSO and water

Vacuum Acetone EtOH MeOH MeCN DMSO H2O

E1

E2

E1anti

E2anti

Z1

Z2

Z3

Z4

441.6(0.886) 447.8(1.090) 447.7(1.091) 446.3(1.082) 446.8(1.087) 449.3(1.110) 446.1(1.085)

439.2(0.889) 445.7(1.069) 445.5(1.061) 444.1(1.061) 444.6(1.066) 446.9(1.087) 443.9(1.064)

454.8(0.784) 456.1(0.991) 455.9(0.992) 454.5(0.983) 455.0(0.988) 457.3(1.010) 454.1(0.987)

453.5(0.816) 456.1(1.035) 455.8(1.038) 454.4(1.028) 454.9(1.033) 457.3(1.056) 454.1(1.032)

463.23(0.515) 469.81(0.611) 469.78(0.612) 468.77(0.605) 469.17(0.608) 471.02(0.622) 468.69(0.606)

482.7(0.366) 479.2(0.464) 479.1(0.464) 478.2(0.457) 478.5(0.460) 480.0(0.473) 477.9(0.457)

515.3(0.277) 494.8(0.281) 494.6(0.283) 493.4(0.277) 493.4(0.279) 494.0(0.287) 492.0(0.276)

489.2(0.451) 486.3(0.594) 486.1(0.595) 484.8(0.589) 485.2(0.592) 487.0(0.608) 484.4(0.591)

Fig. 5 Transient absorption spectra of AO-R obtained by laser flash photolysis (Ar saturated solution). Inset: kinetic at 550 and 450 nm.

As a first conclusion, AO-R exhibits quite challenging photophysics: no triplet state was detected in LFP, the radiative lifetime did not depend on the medium polarity, no steady state nor dynamic solvatochromism was observed and picosecond experiments revealed viscosity dependent kinetics. In order to find out the subtle nature of the formed photoproduct quantum chemical calculations of AO-R in vacuum and in solvents have been jointly performed.

Molecular modeling and discussion Entgegen and zusammen conformers Given the molecular structure of AO-R (Fig. 1), we can expect 8 conformers: 4 entgegen (E) conformers and 4 zusammen (Z) conformers. The 4 (E) conformers come from the flip up/down or down/up of the two central hydrogens of the C–CHQCH–C moiety combined with a 1801 flip of the phenyl ring moiety. The 4 (Z) conformers result from the C–CHQCH–C moiety making an elbow on the side of the dimethyl groups or on the side of the methyl-amino group, combined to a 1801 flip of the phenyl ring moiety (conformers are displayed on Fig. S2, ESI†). The most stable conformer is displayed in Fig. 6 and corresponds to an entgegen conformer with the phenyl ring moiety on the same side as the dimethyl group. This (E) conformer will be called E1 in the following discussion. In vacuum, other entgegen conformers are 1.78, 2.12 and 3.68 kcal mol1 above E1 (for E2, E2anti and E1anti, respectively).

This journal is © the Owner Societies 2015

Fig. 6 Most stable entgegen conformer (called E1) found from QM calculations.

Zusammen conformers have considerably higher energies above E1: 8.41, 12.93, 14.37 and 17.57 kcal mol1 (for Z1, Z2, Z4, Z3, respectively). Moreover, QM calculations on all conformers in different solvents (acetone, MeCN, methanol (MeOH), EtOH, DMSO, water) have been carried out. Although, the solvents lead to stabilization by about 17 to 35 kcal mol1 depending on the nature of the solvent and on the nature of the conformers, the sequence of the conformers found in vacuum remains unchanged. Thus, the polarity of the solvent has a small influence on the conformers, as it has been unambiguously demonstrated experimentally in the previous sections. A simple calculation using a Boltzmann distribution show that E1 represents 92.5% of the ground state population at 20 1C. Fig. 7 shows the highest occupied (HOMO) and lowest unoccupied (LUMO) molecular orbitals of E1. The delocalization of both molecular orbitals combined to make a strong overlap, which is in full agreement with the strong molar absorption coefficient and is expected to provide a visible absorption. Indeed, QM calculations give an absorption for E1 at 441.65 nm in vacuum with an oscillator strength of 0.886. A second consequence is the delocalization of the net charge of +1e on the ¨lliken population analysis skeleton of AO-R. Indeed, from the Mu (as well as the ESP fit method), nitrogen atoms have a partial charge of about 0.48e, the charges on carbon are in the [0.3 : +0.3]e range and hydrogen atoms exhibit a partial charge of about +0.1e. A third consequence of the delocalization of the electronic density

Phys. Chem. Chem. Phys.

View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

Paper

Fig. 7 (a) HOMO and (b) LUMO images of E1.

consists in the majority of the C–C bond lengths around 1.42 Å revealing a mixture of single and double bond characters. However, the central C–C bond length of the C–CHQCH–C moiety is 1.38 Å (even slightly lower with DFT functionals including electrostatic and/or dispersion corrections), highlighting the predominant double bond character of the central part of AO-R. Electronic properties: absorption and emission We have also investigated the first absorption band of all 8 conformers in the 6 solvents mentioned previously. Results are collected in Table 4. From this table, we see that E1 and E2 have a similar absorption transition both in terms of maximum wavelength absorption and oscillator strengths. Next, E2anti and E1anti exhibit similar electronic transitions, which are slightly red shifted by about 12 nm (600 cm1) compared to E1. All the (E) conformers have strong oscillator strengths. For (Z) conformers, the red shift is even more important from 22 to 70 nm (i.e. from 1000 to 3000 cm1). Nevertheless, absorption of (Z) conformers is weaker with oscillator strengths typically between 0.27 and 0.62. Another feature is that the solvents induce a small red shift of about 1–8 nm (63–400 cm1) for all the (E) conformers. For Z1, the absorption wavelength is red shifted by 22 nm (1050 cm1) compared to E1 and the solvents further induce a red shift of 5 to 8 nm (300 to 360 cm1) compared to vacuum values. The behavior of other (Z) conformers is slightly more complicated: the interaction of Z2, Z3 and Z4 with the solvents leads to an unexpected blue shift compared to the vacuum values by 3 to 5 nm (100 to 150 cm1) for Z2 and Z4 and 20 to 23 nm (800 to 920 cm1) for Z3, which is the most polar of the Z conformers. However, the absorption wavelength for Z2, Z3 and Z4 is still red shifted by 30 to 45 nm (1420 to 2130 cm1) compared to E1 in the investigated solvents. From Tables 2 and 4, it can be seen that QM calculations underestimate the experimental absorption wavelength by about 50 nm (2200 cm1). Nevertheless, QM calculations give an electronic transition for AO-R, which is rather independent of the nature of the solvent as it has already been pointed out experimentally (cf. Table 2). The trend discussed above is in satisfactory agreement with the experimental data. We have also investigated the first excited states of all 8 conformers in the solvents mentioned previously. The relaxed excited state of E1, E1* is the most stable irrespective of the solvent used. It is followed by E2*, E2anti* and E1anti*, which are around 1.94, 4.42 and 6.07 kcal mol1 higher, respectively. Thus, this sequence is similar to the sequence of (E) conformers

Phys. Chem. Chem. Phys.

PCCP

in the ground state. This is also true for the sequence of excited Z conformers, i.e. Z* conformers, which present the same order as the sequence found for the ground state: Z1*, Z2*, Z4* and Z3*. The energy difference between Z1* and Z3* is 5.14 kcal mol1 and Z1* is 10.82 kcal mol1 above E1*. Fluorescence emission of AO-R in the different solvents mentioned previously has been calculated and the results are summarized in Table 5. Maximum emission wavelengths of (Z) conformers is not listed in Table 5 because they exhibit either no or a very weak fluorescence emission out of the visible range. A possible explanation of the non-fluorescence of (Z) conformers lies in the fact that (Z) conformers undergo a strong structural change between the ground state and the excited state as shown in Fig. 8 for Z3. On the contrary to (Z) conformers, the structures of (E) conformers in the ground states and in the excited states are very similar (for example E1 and E1* displayed in Fig. 8). This result is in good agreement with the steady-state fluorescence and absorption results. Therefore, they present a strong fluorescence emission (with wavelengths given in Table 5). Once again, the sensitivity of the emission to the solvent is very weak as observed experimentally. Moreover, a close view to Table 5 allows us to say that the calculated emission of E1* is in good agreement with the measured emission with only a deviation of 10 nm. Relaxation pathways of AO-R From experimental and computational results, a possible mechanism can be drawn explaining the difficult photophysics of AO-R. In Fig. 9, the position (relative to E1 see Table S1, ESI†) of all optimized ground and excited conformers in MeCN is given. The blue arrow indicates the laser excitation energy used. It can be seen that E1* and E2* are very close. Thus, we can postulate, as AO-R is in the E1 form in the ground state, that upon light absorption the Franck–Condon (FC) initial excited state reached has two relaxation pathways with unresolved sub-100 fs kinetics: either to the stabilized excited state E1* or towards the excited state of E2, E2*. Indeed, taking into account the very weak fluorescence quantum yield determined experimentally, it is possible that a small fraction of the (FC) initial population decays towards the relaxed excited state E1* (giving then a fluorescence emission), while the other fraction of (FC) population is implied into the population of E2*. Then, the two populations evolve independently according to the decoupled experimental kinetics at 400/575 nm and 530 nm. Furthermore, QM calculations show that the central

Table 5 Maximum emission wavelengths (nm) for entgegen conformers in acetone, acetonitrile, ethanol, methanol, DMSO and water

Acetone EtOH MeOH MeCN DMSO H2O

E1

E2

E1anti

E2anti

545.61 546.21 546.92 547.12 547.62 548.26

541.67 542.28 542.98 543.18 543.72 544.41

571.81 572.37 573.02 573.19 573.67 574.30

570.57 571.14 571.82 572.01 572.52 573.17

This journal is © the Owner Societies 2015

View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

PCCP

Fig. 8 Structures of Z3 (top) and E1 (bottom) in the optimized ground state (blue) and relaxed excited state (red).

Paper

do not permit any observation. This postulated scheme is also supported by QM calculations as they indicate that Z isomers do not present fluorescence emission, which is compatible with the experimental results. The postulated mechanism is displayed on Fig. 10. The possibility of cyclization such as in the well-known spiropyran–merocyanine photophysics must also be considered.23–31 QM calculations predict a cyclized conformation that is 79.66 kcal mol1 above E1 (Fig. S3, ESI†), rendering the cyclization process quite unlikely. Moreover, this product has only a weak absorption near 334 nm with an oscillator strength of 0.24, which is not compatible with the experimental results. Therefore, it seems more reasonable to consider a photoisomerization between (E) and (Z) conformers to explain the complex relaxation of AO-R. Keeping in mind that the radiative relaxation (krad) is almost constant irrespective of the solvent, the viscosity dependence of the lifetime of S1 (assigned to the relaxed excited state of E1, E1*) indicates that E1* deactivation arises from competition between fluorescence and another route, which should be viscosity dependent.52–55 It is likely that E1* undergoes an isomerization involving more important rearrangement: the most simple route of isomerization is towards Z3 and this induces greater structural changes than the E2* - Z1 isomerization (see Fig. S2, ESI†). This reaction induces a higher molecular reorganization, which becomes viscosity dependent.56,57 If we postulate that the Z3 isomer is formed with a rate constant k(E1Z), which is solvent viscosity dependent (see proposed scheme on Fig. 10), k(E1Z) will govern the emissive state lifetime according to the following equation:58–61 tS1 ¼

1 krad þ kðE1Z Þ

(1)

Therefore, k(E1Z) can be extracted from experiments. Taking into account that krad is in the ns time scale while singlet state

Fig. 9 Relative energy levels of the 8 optimized AO-R isomers in ground and relaxed excited state computed in acetonitrile.

CQC bond in the C–CHQCH–C moiety is lengthened by 0.04 Å in the E2* excited states indicating a decrease in the double bond character and favoring of isomerization from E2* to one Z isomer. The most simple isomerization in term of atoms movements occurs towards Z1 (see Fig. S2, ESI†) and is performed in a quasi constant molecular volume with small molecular rearrangement. This fact leads to a fast Hula-twist isomerization reaction as observed elsewhere56,57 and occurs in the 5–10 ps time scale giving rise to the positive peak observed around 535 nm and to low fluorescence intensity. Indeed, if E2* is fluorescent, its very short lifetime (o10 ps) and spectral recovery with E1* emission

This journal is © the Owner Societies 2015

Fig. 10 Postulated mechanism explaining AO-R photophysics.

Phys. Chem. Chem. Phys.

View Article Online

Paper

PCCP

lifetime t(S1) is in the ps time scale, k(E1Z) is directly related to t(S1):

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

kðE1Z Þ ¼

1 1  krad  tS1 tS1

(2)

It is also important to mention that the possibility of the relaxed excited state E1* to give other isomer is in favor of a weak fluorescence quantum yield. Once AO-R is in one of the Z states, a series of structural rearrangements possibly involving other (Z) and (E) conformers in order to go back to E1 can be proposed. First, the ground state of Z3 could thermally relax to the ground state of Z4, then to Z1. Finally, the last step could be the thermal relaxation of E2 towards the ground state of E1. As no experimental proofs are available, one could also propose a direct relaxation to the ground state of E1.

Conclusion In the present work, AO-R shows interesting photophysics that are partly explained through a combination of many complementary experimental and computational techniques. Molecular modeling of Astrazon Orange-R is in a very good agreement with the experimental spectroscopic results allowing us to gain insight into its challenging photophysics. It is demonstrated that 8 isomers are available for this cationic cyanine dye, and that the most stable entgegen conformer E1 represents more than 92% of the ground state population at 20 1C. Molecular modeling also rules out any cyclization mechanism in AO-R, contrary to well-known merocyanine–spyropirane dyes,17–25 due to steric hindrance. Associating computational results and experimental data permit to identify the formed photoproduct as one of the Z isomers and allow us to postulate a possible mechanism in which two isomerizations could be involved: a very fast one, and a slower, viscosity-dependent one, which governs AO-R fluorescence lifetime.

Notes and references 1 X. Wang, S. Werner, T. Weiss, K. Liefeith and C. Hoffmann, RSC Adv., 2012, 2, 156. 2 T. Muraoka, C. Y. Koh, H. Cui and S. I. Stupp, Angew. Chem., 2009, 121, 6060–6063. 3 A. Herrmann, Photochem. Photobiol. Sci., 2012, 11, 446–459. 4 K. Dietliker, A Compilation of Photoinitiators Commercially Available for UV Today, SITA Technology Limited, 2002. 5 F. Ercole, T. P. Davis and R. A. Evans, Polym. Chem., 2010, 1, 37. 6 J. P. Fouassier, X. Allonas, J. Laleve and C. Dietlin, in Photochemistry and Photophysics of Polymer Materials, ed. N. S. Allen, Wiley, Hoboken, 2010, p. 351. 7 K. Kawamura, J. Schmitt, M. Barnet, H. Salmi, C. Ley and X. Allonas, Chem. – Eur. J., 2013, 19, 12853–12858. 8 K. Kawamura, C. Ley, J. Schmitt, M. Barnet and X. Allonas, J. Polym. Sci., Part A: Polym. Chem., 2013, 51, 4325–4330.

Phys. Chem. Chem. Phys.

9 A. V. Metelisa, V. Lokshin, J. C. Micheau, A. Samt, R. Guglielmetti and V. I. Minkin, Phys. Chem. Chem. Phys., 2002, 4, 4340. 10 J. Hobley, V. Malatesta, R. Millini, L. Montanari and W. O Neil Parker, Phys. Chem. Chem. Phys., 1999, 1, 329. 11 F. M. Raymo and S. Giordani, Proc. Natl. Acad. Sci. U. S. A., 2002, 99, 4941. 12 F. M. Rayno, Adv. Mater., 2002, 14, 401. 13 P. Remon, M. Hammarson, S. Li, A. Kahnt, U. Pischel and J. Andreasson, Chem. – Eur. J., 2011, 17, 6492. 14 A.-K. Holm, O. F. Moahmmed, M. Roini, E. Mukhtar, E. T. J. Nibbering and H. Fidder, J. Phys. Chem. A, 2005, 109, 8962. 15 J. Kabatc, B. Jedrzejewska and J. Paczkowski, J. Appl. Polym. Sci., 2006, 99, 207. 16 J. Kabatc, M. Pietrzak and J. Paczkowski, J. Chem. Soc., Perkin Trans. 2, 2002, 287. 17 J. Kabatc, B. Jedrzejewska and J. Packowski, J. Polym. Sci., Part A: Polym. Chem., 2000, 38, 2365. 18 J. Kabatc, Polymer, 2010, 51, 5028. 19 H. Tar, D. Sevinc Esen, M. Aydin, C. Ley, N. Arsu and X. Allonas, Macromolecules, 2013, 46, 3266. ¨ zkan, J. Photochem. Photobiol., A, 2002, ¨kmen and A. O 20 M. So 147, 77. 21 H. Kellet, J. Soc. Dyers Colour., 1968, 84, 257. 22 Y. Dattner and O. Yadid-Pecht, Sensors, 2010, 10, 5014. 23 M. Rini, A.-K. Holm, E. T. J. Nibbering and H. Fidder, J. Am. Chem. Soc., 2003, 125, 3028–3034. 24 L. Poisson, K. D. Raffael, B. Soep, J.-M. Mestdagh and G. Buntinx, J. Am. Chem. Soc., 2006, 128, 3169. 25 A.-K. Holm, M. Rini, E. T. J. Nibbering and H. Fidder, Chem. Phys. Lett., 2003, 376, 214. 26 G. Buntinx, S. Foley, C. Lefumeux, V. Lokshin, O. Poizat and A. Samat, Chem. Phys. Lett., 2004, 391, 33. 27 A. Eilmes, J. Phys. Chem. A, 2013, 117, 2629. 28 Y. Sheng, J. Leszcsynski, A. A. Garcia, R. Rosario, D. Gust and J. Springer, J. Phys. Chem. B, 2004, 108, 16233. 29 H. Takahashi, H. Murakawa, Y. Dakaino, T. Ohzeki, J. Abe and O. Yamada, J. Photochem. Photobiol., A, 1988, 45, 233. 30 C. J. Wohl and D. Kuciauskas, J. Phys. Chem. B, 2005, 109, 21893. 31 S. Basu, S. De and B. B. Bhowmik, Spectrochim. Acta, Part A, 2007, 66, 1255. 32 S. Ruetzel, M. Diekmann, P. Nuemberger, B. Engels and T. Brixner, J. Chem. Phys., 2014, 140, 224310. 33 C. Walter, S. Ruetzel, M. Diekmann, P. Nuemberger, T. Brixner and B. Engels, J. Chem. Phys., 2014, 140, 224311. 34 A. V. Kulinich, A. A. Ishchenko, A. K. Chibisov and G. V. Zakharova, J. Photochem. Photobiol., A, 2014, 274, 91. ¨m and T. Gillbro, Chem. Phys., 1981, 61, 257. 35 V. Sundstro 36 J. Cao, C. Hu, W. Sun, Q. Xu, J. Fan, F. Song, S. Suna and X. Peng, RSC Adv., 2014, 4, 13385. 37 M. Henary and A. Lewitz, Dyes Pigm., 2013, 99, 1107. 38 G. Angulo, G. Grammp and A. Rosspeinter, Spectrochim. Acta, Part A, 2006, 65, 727. 39 G. Angulo, EPA Newsl. (Online), 2007, 26.

This journal is © the Owner Societies 2015

View Article Online

Published on 12 January 2015. Downloaded by Université du Québec à Rimouski on 03/02/2015 09:38:00.

PCCP

40 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, R. Koba-yashi, 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. Jaramil-lo, 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. ¨ . Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski Daniels, O and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009. 41 G. Scalmani, M. J. Frisch, B. Mennucci, J. Tomasi, R. Cammi and V. J. Barone, J. Chem. Phys., 2006, 124, 09107. 42 R. Improta, V. Barone, G. Scalmani and M. J. Frisch, J. Chem. Phys., 2006, 125, 054103. 43 S. Miertusˇ, E. Scrocco and J. Tomasi, Chem. Phys., 1981, 55, 117. 44 P. Suppan, Chemistry and light, RSC, Cambridge, UK, 1994. 45 A. M. Braun, M.-T. Maurette and E. Oliveros, Technologie Photochimique, Presses Polytechniques Romandes, Lausanne, Switzerland, 1986. 46 Y. Marcus, The Properties of Solvents, Wiley Series in Solution Chemistry, John Wiley & Sons Ltd, Chichester, England, vol. 4, 1998.

This journal is © the Owner Societies 2015

Paper

47 M. J. Kamlet, J.-L. M. Abboud, M. H. Abraham and R. W. Taft, J. Org. Chem., 1983, 48, 2877. 48 J. R. Lakowicz, Principles of fluorescence spectroscopy, Kluwer Academic/Plenum publishers, New York, USA, 1999. 49 B. Valeur, Molecular Fluorescence Principles and Applications, Wiley-VCH Verlag GmbH, Weinheim, 2002. 50 N. J. Turro, V. Ramamurthy and J. C. Scaiano, Modern molecular Photochemistry of Organic Molecules, University science book, Sausalito, California, 2010. 51 C. Cornelissen-Gude, W. Rettig and R. Lapouyade, J. Phys. Chem. A, 1997, 101, 9673. 52 C. Gude and W. Rettig, J. Phys. Chem. A, 2000, 104, 8050. 53 N. P. Ernsting, S. A. Kovalenko, T. Senyushkina, J. Saam and V. Farztdinov, J. Phys. Chem. A, 2001, 105, 3443. 54 I. H. M. van Stokkum, D. S. Larsen and R. van Grondelle, Biochim. Biophys. Acta, 2004, 1657, 82. 55 J. Brazard, C. Ley, F. Lacombat, P. Plaza, M. M. Martin, G. Checcucci and F. Lenci, J. Phys. Chem. B, 2008, 112, 15182. 56 R. S. H. Liu and G. Hammond, Chem. – Eur. J., 2001, 7, 4536. 57 R. S. H. Liu, Acc. Chem. Res., 2001, 34, 555. 58 L.-Y. Yang, M. Harigai, Y. Imamoto, M. Kataoka, T.-I. Ho, E. Andrioukhina, O. Federova, S. Shevyakov and R. S. H. Liu, Photochem. Photobiol. Sci., 2006, 5, 874. 59 B. Dellinger and M. Kasha, Chem. Phys. Lett., 1976, 38, 9. 60 Y.-P. Sun, J. Saltiel, N. S. Park, E. A. Horburg and D. H. Waldeck, J. Phys. Chem., 1991, 95, 10336. 61 S. Malkin and E. Fischer, J. Phys. Chem., 1964, 68, 1153.

Phys. Chem. Chem. Phys.