Excitation Energy Transfer from Sulforhodamine ... - ACS Publications

Jul 10, 1999 - curve should be analyzed by a single-exponential function. On the ... lation trace (Cr) for each single-exponential fit. In the case of...
0 downloads 0 Views 109KB Size
Anal. Chem. 1999, 71, 3382-3389

Excitation Energy Transfer from Sulforhodamine 101 to Acid Blue 1 at a Liquid/Liquid Interface: Experimental Approach To Estimate Interfacial Roughness Shoji Ishizaka, Satoshi Habuchi, Haeng-Boo Kim, and Noboru Kitamura*

Division of Chemistry, Graduate School of Science, Hokkaido University, Kita-ku, Sapporo 060-0810, Japan

Dynamic fluorescence anisotropy of Sulforhodamine 101 (SR101) and excitation energy transfer from SR101 to Acid Blue 1 (AB1) at water/carbon tetrachloride (CCl4) and water/1,2-dichloroethane (DCE) interfaces were studied by using time-resolved total-internal-reflection (TIR) fluorometry. A magic-angle dependence of the TIR fluorescence dynamic anisotropy revealed that rotational reorientation of SR101 at the water/CCl4 interface was restricted in the two-dimensional plane of the interface, while that at the water/DCE interface took place rather freely, similar to reorientation in an isotropic medium. Furthermore, the structural dimension derived from the excitation energy transfer dynamics measurements at the water/CCl4 or water/DCE interface was 1.93 or 2.48, respectively. These results indicated that the water/CCl4 interface was sharp with respect to the molecular size of SR101 (∼10 Å), while the water/DCE interface was relatively rough compared to the water/CCl4 interface. Structural dimension analysis of excitation energy transfer dynamics between dye molecules adsorbed on the water/ oil interface measured under the TIR conditions was shown to be a potential means to estimate roughness/ thickness of the interface at a molecular level. The present results were also compared with those predicted from molecular dynamics simulations. The chemistry at liquid/liquid interfaces is the fundamental basis for analytical and separation sciences,1-3 and so far, various techniques have been employed to elucidate characteristics of these interfaces.4-9 In particular, recent advances in surfaceselective spectroscopies such as nonlinear optical spectroscopy, quasi-elastic laser scattering, and so forth have provided molecular(1) Shi, C.; Anson, F. C. Anal. Chem. 1998, 70, 3114-3118. (2) Shao, Y.; Mirkin, M. V. Anal. Chem. 1998, 70, 3155-3161. (3) Ma, M.; Cantwell, F. F. Anal. Chem. 1999, 71, 388-393. (4) Volkov, A. G., Deamer, D. W., Eds. Liquid-Liquid Interfaces; CRC Press: Boca Raton, FL, 1996. (5) Girault, H. H.; Schiffrin, D. J. In Electroanalytical Chemistry; Bard, A. J., Eds.; Marcel Dekker: New York, 1989; Vol. 15, Chapter 1. (6) Nagatani, H.; Watarai, H. Anal. Chem. 1996, 68, 1250-1253. (7) Watarai, H.; Gotoh, M.; Gotoh, N. Bull. Chem. Soc. Jpn. 1997, 70, 957964. (8) Webster, R. D.; Dryfe, R. A. W.; Coles, B. A.; Compton, R. G. Anal. Chem. 1998, 70, 792-800. (9) Nagatani, H.; Watarai, H. Anal. Chem. 1998, 70, 2860-2865.

3382 Analytical Chemistry, Vol. 71, No. 16, August 15, 1999

level insights about interfaces.10-12 In a recent publication, we also demonstrated that total-internal-reflection (TIR) fluorescence dynamic anisotropy was a powerful means to study characteristics of dynamic motions of a molecule at water/oil interfaces.13 It is well-known that dynamic fluorescence anisotropy is based essentially on rotational reorientation of the excited dipole of a probe molecule, and its correlation time depends on local environments around the molecule. For a dye molecule in an isotropic medium, rotational reorientation of the excited dipole takes place freely. At a liquid/liquid interface, on the other hand, the out-of-plane motion of a probe molecule should be frozen when the dye is adsorbed on a sharp liquid/liquid interface (i.e., two-dimensional with respect to the molecular size of a probe), while such a motion will be allowed for a relatively thick liquid/liquid interface (i.e., three-dimensional-like). Thus, by observing rotational freedom of a dye molecule (i.e., excited dipole), one can discuss the thickness of a liquid/liquid interface. Indeed, such an approach was very successful to estimate the structure of water/phthalate ester interfaces.13 It is worth noting, however, that a liquid/liquid interfacial structure will be governed by several factors, so that a complementary study other than fluorescence dynamic anisotropy is required to obtain further detailed information about the characteristics at a liquid/liquid interface. As a new and novel approach, therefore, we introduced excitation energy transfer dynamics and the relevant structural (fractal) dimension analysis to elucidate the structure of a liquid/liquid interface. We consider here dipole-dipole (Fo¨rster-type) excitation energy transfer between an energy donor (D) and an acceptor (A) both adsorbed on a liquid/liquid interface. When diffusional motions of D and A are inhibited, as in the case for strong binding of the molecules to the surface by adsorption, excitation energy (10) (a) Grubb, S. G.; Kim, M. W.; Rasing, T.; Shen, Y. R. Langmuir 1988, 4, 452-454. (b) Conboy, J. C.; Daschbach, J. L.; Richmond, G. L. J. Phys. Chem. 1994, 98, 9688-9692. (c) Corn, R. M.; Higgins, D. A. Chem. Rev. 1994, 94, 107-125. (11) (a) Du, Q.; Freysz, E.; Shen, Y. R. Science 1994, 264, 826-828. (b) Messmer M. C.; Conboy, J. C.; Richmond, G. L. J. Am. Chem. Soc. 1995, 117, 80398040. (c) Gragson, D. E.; Richmond, G. L. J. Am. Chem. Soc. 1998, 120, 366-375. (d) Gragson, D. E.; Richmond, G. L. J. Phys. Chem. B 1998, 102, 569-576. (12) (a) Zhang, Z. H.; Tsuyumoto, I.; Takahashi, S.; Kitamori, T.; Sawada, T. J. Phys. Chem. A 1997, 101, 4163-4166. (b) Tsuyumoto, I.; Noguchi, N.; Kitamori, T.; Sawada, T. J. Phys. Chem. B 1998, 102, 2684-2687. (13) Ishizaka, S.; Nakatani, K.; Habuchi, S.; Kitamura, N. Anal. Chem. 1999, 71, 419-426. 10.1021/ac990213a CCC: $18.00

© 1999 American Chemical Society Published on Web 07/10/1999

transfer quenching dynamics of D by A reflect the structural dimension around D and A through spatial distributions of the components.14 Actually, the method has been applied to study nanometer-scale morphologies in Langmuir-Blodgett films,15,16 vesicles,16,17 polymers,18-21 and silica gels.22-24 In such a case, fluorescence dynamics of D (ID(t)) should obey the following equation as reported by Klafter and Blumen:14

ID(t) ) A exp(-(t/τD) - P(t/τD)dh /6)

(1)

where A is a preexponential factor and τD is the excited-state lifetime of D without A. P is a parameter proportional to the probability that A resides within the critical energy transfer distance (R0) of the excited donor. dh is called the fractal dimension, and reflects a spatial distribution of the acceptor around the donor. If the donor and acceptor molecules adsorb uniformly on a sharp liquid/liquid interface (two-dimensional), dh should be 2.0, since excitation energy transfer takes place exclusively along the lateral direction at the interface. On the other hand, if a liquid/liquid interface is thick and rough with respect to the molecular size of a probe molecule, dh should be 2.0-3.0 since a possibility of energy transfer along the direction other than the lateral direction cannot be neglected. Therefore, we expected that a study on excitation energy transfer dynamics would provide invaluable information about the characteristics at a liquid/liquid interface, along with a complementary study on the same system by fluorescence dynamic anisotropy. In the present study, we focus our attention on revealing the structures of water/carbon tetrachloride (CCl4) and water/1,2dichloroethane (DCE) interfaces. So far, water/CCl4 and water/ DCE interfaces have been studied by various techniques, and it has been reported that the water/CCl4 and water/DCE interfaces are very thin: thickness ∼1 nm.25-31 Furthermore, molecular dynamics simulations reported by Michael and Benjamin indicate that the water/DCE interface is thick and rough compared to the water/CCl4 interface.30 Molecular dynamics simulations give molecular-level insight about a liquid/liquid interface and are the indispensable basis for understanding the characteristics at the interface. Nonetheless, complementary studies by experiments are absolutely necessary. Therefore, we tested the present idea (14) (a) Klafter, J.; Blumen, A. J. Chem. Phys. 1984, 80, 875-877. (b) Klafter, J.; Blumen, A. J. Lumin. 1985, 34, 77-82. (15) Tamai, N.; Yamazaki, T.; Yamazaki, I. Chem. Phys. Lett. 1988, 147, 25-29. (16) Yamazaki, I.; Tamai, N.; Yamazaki, T. J. Phys. Chem. 1990, 94, 516-525. (17) Tamai, N.; Yamazaki, T.; Yamazaki, I.; Mizuma, A.; Mataga, N. J. Phys. Chem. 1987, 91, 3503-3508. (18) Lin, Y.; Nelson, M. C.; Hanson, D. M. J. Chem. Phys. 1987, 86, 1586-1592. (19) Tcherkasskaya, O.; Spiro, J. G.; Ni, S.; Winnik, M. A. J. Phys. Chem. 1996, 100, 7114-7121. (20) Yekta, A.; Spiro, J. G.; Winnik, M. A. J. Phys. Chem. B 1998, 102, 79607970. (21) Kim, H.-B.; Habuchi, S.; Kitamura, N. Anal. Chem. 1999, 71, 842-848. (22) Rojanski, D.; Huppert, D.; Bale, H. D.; Dacai, X.; Schmidt, P. W.; Farin, D.; Seri-Levy, A.; Avnir, D. Phys. Rev. Lett. 1986, 56, 2505-2508. (23) Levitz, P.; Drake, J. M. Phys. Rev. Lett. 1987, 58, 686-689. (24) Rojanski, D.; Huppert, D.; Avnir, D. Chem. Phys. Lett. 1987, 139, 109-115. (25) Benjamin, I. J. Chem. Phys. 1992, 97, 1432-1445. (26) Benjamin, I. Science 1993, 261, 1558-1560. (27) Michael, D.; Benjamin, I. J. Phys. Chem. 1995, 99, 1530-1536. (28) Schweighofer, K. J.; Benjamin, I. J. Phys. Chem. 1995, 99, 9974-9985. (29) Michael, D.; Benjamin, I. J. Phys. Chem. 1995, 99, 16810-16813. (30) Michael, D.; Benjamin, I. J. Chem. Phys. 1997, 107, 5684-5693. (31) Chang, T.-M.; Dang, L. X. J. Chem. Phys. 1998, 108, 818-819.

to elucidate the structures of water/CCl4 and water/DCE interfaces and compared the results with those predicted by molecular dynamics simulations. In the following, we report fluorescence dynamic anisotropy and excitation energy transfer quenching of Sulforhodamine 101 (SR101) at water/CCl4 and water/DCE interfaces, with special attention to elucidating thickness and/or roughness of the interface. EXPERIMENTAL SECTION Chemicals and Sample Preparation. Water was purified by distillation and deionization prior to use (GSR-200, Advantec Toyo Co., Ltd.). Carbon tetrachloride (Kanto Chemical Co., Inc.) was purified by distillation after washing successively with an aqueous potassium hydroxide solution and water. 1,2-Dichloroethane (DCE; Merck, Uvasol) and Sulforhodamine 101 (SR101; ACROS Organics, laser grade) were used without further purification. Acid Blue 1 (AB1; Tokyo Kasei Kogyo Co., Inc) was purified by recrystallizations from an acetone-water mixture. The chemical structures of SR101 and AB1 are inserted in Figure 2. For dynamic anisotropy experiments, an aqueous SR101 solution ([SR101] ) 1.0 × 10-9 M) saturated with CCl4 or DCE was poured carefully onto water-saturated carbon tetrachloride or DCE in a Pyrex cell (inner diameter ) 40 mm). As a sample for excited energy transfer measurements, a mixture of SR101 (1.0 × 10-9 M) and AB1 ((1.9-5.7) × 10-9 M) was used as the water phase and a water/oil interface was constructed by procedures analogous to those mentioned above. The lower inside of the cell was treated with dichlorodimethylsilane to construct a flat water/oil interface. The cell was washed thoroughly with CCl4 or DCE and then with water prior to a sample preparation. Spectroscopic measurements were carried out after the sample solution in the cell had been allowed to stand for over 100 min. Measurements. TIR fluorescence decay profiles of SR101 were measured as described below. Fundamental laser pulses from a mode-locked Ti-sapphire laser (Coherent, Mira model 900F), pumped by a diode laser (Verdi), were amplified by a regenerative amplifier (RegA model 9000) pumped by an Ar+ ion laser (Innova 300). Optical parametric amplification (Coherent, model 9400) of the output gave 580 nm pulses as an excitation light source (repetition rate 100 kHz; fwhm 200 fs autocorrelation trace). The excitation laser beam, polarized perpendicular to the plane of incidence (s-polarized) by using a Glan-laser prism, was used to irradiate a water/oil interface through the oil phase, as reported previously.13 In the present experiments, the incident angle of the laser beam (θi) was set (80°) larger than the critical total-reflection angle at the water/CCl4 (θc ) 66°) or water/DCE (θc ) 67°) interface. The fluorescence from the sample was collected along the surface normal, and its polarization was selected with a polarizer (Polaroid, HNP′B). The polarized fluorescence was detected by a microchannel-plate photomultiplier (Hamamatsu, R3809U-50) equipped with a monochromator (Jobin Ybon, H-20) and analyzed by a single-photon-counting module (Edinburgh Instruments, SPC-300). The monitoring wavelength of the fluorescence from SR101 was set at 615 nm throughout the study. The steady-state absorption and fluorescence spectra were measured by using Hitachi U-3300 and Hitachi F-4500 spectrometers, respectively. All experiments were conducted at room temperature (∼23 °C). Analytical Chemistry, Vol. 71, No. 16, August 15, 1999

3383

Figure 1. TIR fluorescence decay curves of SR101 at water/CCl4 (a, b) and water/DCE (c, d) interfaces. The angle of the emission polarizer was set at 45° (a, c) or 54.7° (b, d) with respect to the direction of excitation polarization. The upper trace and the inner panel of each figure represent the plots of the weighted residuals (Re) and the autocorelation trace (Cr) for a single-exponential fitting, respectively.

The interfacial tension (γ) in the water/oil system was measured by using a pendant-drop method. The shape of a pendant droplet was monitored by a CCD camera, and the digitized images were transferred to a microcomputer to calculate the γ value. The γ value at each AB1 concentration ([AB1]) was determined as reported previously.32 RESULTS AND DISCUSSION A Magic-Angle Dependence of the TIR Fluorescence Decay Profiles of SR101 at the Water/Oil Interfaces. Before describing the results on excitation energy transfer quenching of SR101 by AB1 at water/CCl4 and water/DCE interfaces, we explain a magic-angle dependence of the total-internal-reflection (TIR) fluorescence decay profile of SR101 at the interface. The studies are quite important to compare the results with those by excitation energy transfer quenching, as discussed later in detail. As reported previously, a fluorescence decay profile of SR101 observed at a water/oil interface should be analyzed by two limiting cases, depending on the structure of the interface.13 If the thickness or roughness of a water/oil interface is comparable to the molecular size of SR101 and the dye molecules adsorb strongly on the interface, the rotational motions of SR101 are strongly restricted in the interfacial layer, and the emission dipole (32) Nakatani, K.; Ishizaka, S.; Kitamura, N. Anal. Sci. 1996, 12, 701-705.

3384 Analytical Chemistry, Vol. 71, No. 16, August 15, 1999

moment of SR101 is directed within the X-Y plane of the interface. In such a case, the time profile of the total fluorescence intensity of SR101 observed from the interface should be proportional to I||(t) + I⊥(t), where I||(t) and I⊥(t) represent the fluorescence decay profiles observed with emission polarization parallel and perpendicular to the direction of excitation polarization, respectively. When the angle of an emission polarizer is set at 45° with respect to the direction of excitation polarization (magic angle), fluorescence anisotropy is canceled, so that the TIR fluorescence decay curve should be analyzed by a single-exponential function. On the other hand, if the interfacial layer is thick and rough enough compared to the molecular size of SR101 and the molecules adsorbed on the interface are weakly oriented, the rotational motions of SR101 take place rather freely, similar to those in a bulk phase. If this is the case, the time profile of the total fluorescence intensity must be proportional to I||(t) + 2I⊥(t); magic angle ) 54.7°. Figure 1 shows fluorescence decay profiles of SR101 observed from water/CCl4 and water/DCE interfaces at an emission polarization angle of 45° ((a) and (c)) or 54.7° ((b) and (d)), together with the relevant weighted residuals (Re) and autocorelation trace (Cr) for each single-exponential fit. In the case of the water/CCl4 interface ((a) and (b)), Re and Cr of the data observed at 54.7° exhibited nonrandom distributions compared to those

Table 1. Fluorescence Decay Parameters of SR101 Adsorbed on Water/Oil Interfaces Observed under TIR Conditions and in an Aqueous Solution

water/CCl4 water/DCE aqc

anglea/deg

τ/ns

χ2 a

DWb

45 54.7 45 54.7

4.06 ( 0.02 4.21 ( 0.03 3.62 ( 0.02 4.12 ( 0.03 4.12 ( 0.01

1.11 1.61 2.28 1.09 1.16

1.88 1.37 0.90 1.73 1.94

a Angle of the emission polarizer with respect to that of the excitation laser beam. b χ2 and DW represent the χ2 and Darbin-Watson parameters for the fitting, respectively. c Determined in an aqueous SR101 solution ([SR101] ) 1.7 × 10-7 M).

predicted by the best fit (b), particularly, those in the initial stage of excitation (