10442 J. Phys. Chem. B 2010, 114, 10442–10450 Reverse Micelle Induced Flipping of Binding Site and Efficiency of Albumin Protein with an Ionic Styryl Dye Dibakar Sahoo, Prosenjit Bhattacharya, and Sankar Chakravorti* Department of Spectroscopy, Indian Association for the CultiVation of Science, JadaVpur, Kolkata 700032, India ReceiVed: April 1, 2010; ReVised Manuscript ReceiVed: July 6, 2010 The effect of reverse micelle environment on the binding mechanism of 2-(4-(dimethylamino)styryl)-1methylpyridinium iodide (DASPMI) with Bovine Serum Albumin (BSA) compared with that in buffer solution has been investigated in this paper with the help of steady state and time-resolved emission spectroscopy along with molecular docking to have a correct picture about binding. The binding of DASPMI with attachment efficiency of 30% and 70% at site I (subdomain IIA) and site II (subdomain IIIA) of BSA, respectively, in buffer solution gets reversed inside a reverse micelle. The bigger cavity size of site II in buffer solution ushers the dye with increased attachment efficiency and in reverse micelle change in π-stacking and hydrophobic interaction control the attachment efficiency. The calculated Förster distance gets curtailed as the environment changes from buffer to reverse micelle. The binding becomes stronger with a smaller gap between the probe and Trp-214 inside the reverse micelle than that in buffer solution. Introduction Reverse micelles are self-organized aggregates formed by surfactants in organic solvent, and nanometer-sized water pools are formed by the solubilization of water in their polar cores. The reverse micelles that are mostly with spherical shape are usually formed in ternary surfactant-water-organic solvent mixtures including surfactants (<10%), water (0-10%), and organic solvent (80%-90%), so the reverse micelles are also called water-in-oil emulsions or Winsor II emulsions.1 Reverse micelles are generally smaller than their hydrophilic counterparts (micelles), and their aggregation number is commonly lower than 50.2 Moreover, reverse micellar systems are colloidal solutions, so characteristic properties of these systems are thermodynamic stability (no phase separation with time), spontaneous formation, low interfacial tension (<10-2 mN · m-1), transparent nature (nanometer size < 100 nm), large surface area (102-103 m2 · cm-3), viscosity comparable with pure organic solvents, and highly dynamic (constantly a collision and a fusion with each other, and occasionally the fusion surfactant molecules and the contents inside reverse micelles exchange).3 The microwater pools inside reverse micelles are stabilized by a surfactant monolayer within an organic continuum, which can solubilize hydrophilic biomolecules such as proteins, enzymes, DNA, and amino acids. In reverse micellar systems, the biomolecules inside the polar core of the surfactant monolayer are protected from denaturation by organic solvent. Therefore, protein solubilization in reverse micelles plays a key role in a number of topics of biotechnology research. Currently, reverse micelles are used as reaction systems for enzymatic catalysis,4 as models of membrane system separation of proteins,5 for solvent-based extraction of proteins,6 as a microsurrounding for protein structure discovery,7 and for protein refolding.8,9 In the past three decades, the application of reverse micelles for bioseparation has attracted considerable attention because the technique is considered to be potentially useful in downstream processing for a large-scale separation of biomolecules from * Corresponding author. E-mail: [email protected]. fermentation mixtures. The encapsulation of molecules and particles by reverse micelles can provide separation media, and study of these systems is useful for information in drug designing and delivery.10 Proteins and enzymes are solubilized into reverse micelles, maintaining their activities and native structures, and can be back-extracted.11,12 It has been found that the factors such as water content (the molar ratio of water to surfactant, namely, W0) and micelle size,13 aqueous phase pH and ionic strength,14 surfactant type and concentration,15 and cosurfactant16 affect the protein solubilization and characteristics based on the interactions between reverse micelles and proteins. Moreover, adding proteins to the reverse micelles can alter surfactant self-assembly and phase behavior which can affect the protein efficiency. Therefore, it is intriguing to explore the nature of the protein kinetics in the well-known reverse micellar systems. The qualitative and quantitative detection of binding characteristic in reverse micelles may help in designing efficient drug sensitizers for photodynamic therapy (PDT).17 The extended photodynamic action depends on the biodistribution of the probe molecule in the cytoplasmic and mitochondrial membranes,18,19 the retention, and the nature of the binding inside the cell. Albumins being the most abundant proteins in plasma have the ability to carry drugs20 as well as to be used as endogenous and exogenous substances. Bovine Serum albumin (BSA) has three major domains, each with two subdomains (Figure 1). Major binding sites, namely, site I and site II, are located at subdomains IIA and IIIA.21 BSA has three intrinsic fluorophores, tryptophan (Trp), tyrosine (Tyr), and phenylalanine, of which Trp 214 is located in site I (subdomain IIA) and Tyr 411 in site II (subdomain IIIA). Of the three chromophores of BSA, Trp is the key player; phenylalanine has feeble fluorescence; and the fluorescence of Tyr is almost totally quenched if it is ionized or near an amino group, a carboxyl group, or a tryptophan residue.22,23 Ionic styryl dyes are known to respond to changes in transmembrane potential by a fast electrostatic mechanism.24 The shift of charge of the ionic dyes from ground state to excited 10.1021/jp102937y 2010 American Chemical Society Published on Web 07/22/2010 Albumin Protein with an Ionic Styryl Dye J. Phys. Chem. B, Vol. 114, No. 32, 2010 10443 CHART 1: Molecular Structure of DASPMI Figure 1. Crystal structure of HSA (which is like BSA) and the location of different domain binding sites. The location of hydrophobic binding sites (site I and site II) is indicated. The position of the Tryptophan residue (Trp-214) is shown. state coupled with electric field within a cell membrane results in electrochromism.25,26 Among these dyes, the fluorescence intensity of 2-(4-(dimethylamino)styryl)-1-methylpyridinium iodide (DASPMI) is a dynamic measure for the membrane potential of mitochondria27,28 in living cells. DASPMI has an interesting multibond rotation involved intramolecular charge transfer photophysical property,29,30 which is dependent on both polarity and microviscosity of the medium.31,32 This dye has also been used in polymer science and in cell biology33-35 because of the strong dependence of its photophysics on viscosity and polarity. DASPMI can interact with DNA as a groove binder.36 One would think here of investigating the nature of binding of DASPMI with protein environment. Considering interesting photophysical properties of DASPMI and its possible interaction with protein, we intend to explore in this article the binding sites and attachment efficiencies. It is also envisaged to monitor the effect of reverse micelle, a biomimetic for cells, on the DASPMI-protein binding mechanism and efficiency of attachment once it is inserted inside the reverse micelle. Experimental Section Materials and Methods. 2-(4-(Dimethylamino)styryl)-1methylpyridinium iodide (DASPMI) (Chart 1) was received from Aldrich Chemical, USA, and purified by column chromatography (on silica gel 60-120; 5% ethyl acetate in petroleum ether). The purity of the compound was checked by thin layer chromatography (TLC). The compound was then subjected to vacuum sublimation before use. n-Heptane (spectroscopy grade), from Aldrich, and AOT (ultra grade) from Sigma were used as received. The molar ratio of residual water/ AOT, as determined by Karl Fischer titration, was found to be 0.1. Millipore water was used in the preparation of water-in-oil microemulsion. The absorption spectra were taken with a Shimadzu UV-vis absorption spectrophotometer model UV-2401PC. The fluorescence spectra were obtained with a Hitachi F-4500 fluorescence spectrophotometer. Quantum yields were determined by using the secondary standard method (φf ) 0.23) with recrystallized β-naphthol in MCH (methylcyclohexane), and details of the process are described elsewhere.37,38 For lifetime measurement, the sample was excited (at 440 nm: optical density ∼0.15) with a picosecond diode (IBH Nanoled-07). The time-correlated single photon counting (TCSPC) setup consists of Ortec 9327, TBX-04 detector, DataStation measurement software, and DSA6 Foundation Package. The data were collected with a DAQ card as a multichannel analyzer. The typical fwhm of the system response is about 80 ps. Typical slit width ∼30 nm, monochromator type Jobin-Yvon, number of channels 4000 (6 ps per channel), window width ∼24.5 ns, and number of counts ∼10 000 were used in taking decay profiles. Data analysis was carried out using the curve-fitting program supplied by the manufacturer. The quality of the fit was determined by the reduced χ2 and a high Durbin-Watson parameter (>1.7).39 For measurement of fluorescence, depolarization decays parallel (IVV) and perpendicular (IVH) were collected in an alternating manner for equal amounts of time until at least 10 000 fluorescence counts were collected in the peak channel of IVV. The time -dependent fluorescence anisotropy, r(t) was then calculated from the above data using the following relation r(t) ) [IVV - GIVH]/[IVV + 2GIVH] (1) where G is the ratio between the fluorescence intensity at parallel and perpendicular polarizations of the emission with respect to the excitation beam. The value of G has been used as 0.56 as identified for the instrument. Sample Preparation. Microemulsion solutions of desired w0 were obtained by adding concentrated protein solutions, drug solutions, or plain buffer (pH 7.0) to a 0.1 M AOT solution in n-heptane. Volume of additivity was assumed in calculating AOT concentration and w0 values. The samples were gently shaken until complete clarification. The final sample concentration was calculated according to the total volume of the microemulsion. A solution of DASPMI (5 × 10-5 M) was prepared in n-heptane. Binary mixtures of n-heptane/AOT were prepared by adding the required amount of AOT into 5 mL each of the second stock solution. We used the stock solution containing 5 × 10-5 M dye in a binary solution of 0.1 M AOT in n-heptane for the preparation of n-heptane/AOT/water mixtures. For a different value of W, the solution was prepared by adding an appropriate amount of Millipore water, using calibrated micropipets, to the binary stock solution. To get a homogeneous mixture, we sonicated the solutions for 3-5 min, and the samples were kept for 15-20 h at room temperature before carrying out all the measurements. Calculation of Binding Constant. The binding constant values (total) due to attachment with Trp (site I) and Tyr (site 10444 J. Phys. Chem. B, Vol. 114, No. 32, 2010 Sahoo et al. [ knr ) kr Figure 2. Emission spectra of DASPMI as a function of BSA concentration (1) 0 µM, (2) 2 µM, (3) 7 µM, (4) 10 µM, (5) 14 µM, (6) 15 µM, and (7) 17 µM. Inset shows the plot of (FR - F0)/(Fx - F0) against [L]-1 for BSA. II) have been determined from the fluorescence intensity data of DASPMI considering the following rearranged equation of the original one developed by Benesi and Hildebrand based on 1:1 probe protein complexation40 [(F∞ - F0)/(Fx - F0)] - 1 ) (Kb[L])-1 (2) where F0, Fx, and F∞ are the fluorescence intensities of DASPMI in the absence of protein, at an intermediate protein concentration, and at a protein concentration when the interaction is complete, respectively. Kb is the binding constant, and [L] is the free concentration of protein. A plot of [(F∞ - F0)/(Fx - F0)] against [L]-1 for BSA shows linear variations (Figure 2), justifying the validity of the above equation and hence confirming one-to-one interaction between the probe and proteins. The binding constant is determined from the slope of the plot. To get the free protein concentration from total protein concentration, we adopted a self-consistent approach similar to the one used during the orbital calculations. Calculation of Thermodynamic Parameters. The thermodynamic parameter associated with the complexation between SQ and BSA was determined using the following equations41 lnK ) -∆H/(RT) + ∆S/R (3) ∆G ) ∆H - T∆S ) -RTln K (4) ] 1 -1 φf (6) Cyclic voltammograms (CVs) were recorded on a BAS-CV50W cyclic voltameter. Circular dichroism (CD) spectra were recorded on a Jasco Corporation, J-815, spectrophotometer. The fluorescence picosecond lifetime measurement was done with a Horiba Jobin Yvon Fluoro Cube 01-NL time-resolved Fluorescence Lifetime Spectrometer with TBX-04 detector, DataStation measurement software, and DSA6 Foundation Package, and the excitation was done at 440 nm (diode laser). Millipore water was used in all the studies. All experiments were carried out at room temperature (26 ( 1 °C). Nuclear magnetic resonance data were taken with Bruker Avance DPX 300 in deuterium oxide as solvent. Scanning Electron Microscopy (SEM). BSA solution in the presence and absence of DASPMI was placed on sample studs and coated with platinum by ion sputtering. FESEM images were obtained on a JSM-6700F, from JEOL, scanning electron microscope with an accelerating voltage of 15 KV. The sample (DASPMI interacted with BSA) was dried in a vacuum and kept overnight before taking the SEM. The crystal structure of HSA is taken from the Brookhaven Protein Data Bank (entry code 1H9Z)43 as it gives a nearly similar structural nature to BSA (Figure 1). The potential of the 3D structure of HSA was assigned according to the Amber 4.0 force field with Kollman-all-atom charges. We took help from molecular modeling software Sybyl 6.944 for generating the initial structure of all molecules. The geometry of the molecule was subsequently optimized to minimal energy using the Tripos force field with Gasteiger-Marsili charges. The FlexX program was used to build the interaction modes between DASPMI and HSA. Results and Discussion The effect of BSA on the fluorescence of DASPMI in 0.01 M phosphate buffer at 26 °C is shown in Figure 2. The increase in DASPMI fluorescence as a function of protein concentration reaches a plateau, which indicates the increment of rigidity of the surrounding microenvironments. The BSA binding sites are very effective in preventing “free rotator motions” in the DASPMI moiety. The excited state of DASPMI when complexed with BSA is less stabilized than the corresponding ground state compared to that in uncomplexed form in water, resulting in an increase of energy gap between the excited state and where ∆H is enthalpy change; ∆S is entropy change; ∆G is change in free energy; and K is the binding constant. Calculation of Radiative and Nonradiative Rate Constant. To calculate the radiative decay rate constant (kr), eq 5 was used42 kr ) φf τf (5) where kr is radiative decay rate constant; φf ) fluorescence quantum yield of DASPMI; and τf is lifetime of DASPMI. Nonradiative decay rate constant knr can be determined using eq 6 Figure 3. Time-resolved fluorescence decays of DASPMI, in aqueous buffer, with BSA in AOT solution, in BSA of buffer solution [BSA] ) 15 µM. Albumin Protein with an Ionic Styryl Dye J. Phys. Chem. B, Vol. 114, No. 32, 2010 10445 TABLE 1: Photophysical Properties and Binding Constant of DASPMI in Buffer and AOT Solution in the Absence and Presence of BSA ([BSA] ) 15 µM) DASPMI DASPMI + BSA in buffer DASPMI + BSA in AOT τ1 τ2 Kr × 10-9 Knr × 10-9 Kb Φf (ns) (ns) (S-1) (S-1) (M-1) 0.018 0.04 1.4 ( 0.03 (20%) 4.8 ( 0.05 (30%) 0.06 ( 0.05 (80%) 2.4 ( 0.1 (70%) 0.056 0.013 3.06 0.3 (2.15 ( 0.2) × 104 0.06 4.2 ( 0.07 (60%) 2.2 ( 0.09 (40%) 0.017 0.27 (5.1 ( 0.3) × 105 TABLE 2: Thermodynamic Parameters of the System BSA-DASPMI in Buffer Solution at Different Temperatures T binding constant (K) (Kb × 10 ) (kJ mol ) (kJ mol ) (J mol-1 K-1) 298 303 310 315 2.15 ( 0.2 1.56 ( 0.1 1.31 ( 0.3 1.21 ( 0.1 -14.54 -24.81 -24.78 -24.74 -24.72 34.34 4 ∆H ∆G -1 ∆S -1 ground state and a consequent blue shift. This increase of energy gap also decreases the efficiency of radiationless deactivation of DASPMI, a process involving a low-lying twisted intramolecular charge transfer state (TICT),45 and leads to a remarkable enhancement in its fluorescence quantum yield. Picosecond timeresolved fluorescence analysis indicates that in buffer, at pH 7, DASPMI exhibits a biexponential decay with a lifetime of 1.4 ns (20%) and 0.06 ns (80%), but addition of BSA causes a change in decay time at 4.8 ns (30%) and 2.4 ns (70%) (Figure 3). The computed nonradiative and radiative decay constants of DASPMI from the quantum yield and lifetimes are found to be 3.06 × 109 and 0.056 × 109 in the absence of BSA, and in the presence of BSA, they are 0.3 × 109 and 0.012 × 109, respectively (Table 1). Considering 1:1 stoichiometry for the complex formation, the binding constant between DASPMI and protein (BSA) is calculated to be 2.15 × 104 M-1, and the corresponding free energy change is -24.81 kJ mol-1. The ∆H0 and ∆S0 values for the binding reaction between DASPMI and BSA were found to be -14.54 kJ mol-1 and 34.34 J mol-1 K-1 (Table 2). The negative ∆G means that the binding process was spontaneous, and formation of the DASPMI-BSA coordination compound was an exothermic reaction accompanied by a positive ∆S0 value. However, ∆H0 might play a role in the electrostatic reaction. So, the binding process of DASPMI to BSA involves hydrophobic interaction strongly as evidenced by positive values of ∆S0, but electrostatic interaction could also not be excluded. The binding of DASPMI with BSA is further confirmed by Field Emission Scanning Microscope (FESEM), Cyclic Voltametry (CV), Circular Dichroic Spectra (CD), and 1H NMR techniques. The FESEM images show that BSA alone has a regular structure, and this drastically changes upon addition of DASPMI, which is reflected in a drastic increase of width of the FESEM image of BSA alone of 2 ( 0.1 nm to a mean width of 70 ( 10 nm upon complexation of BSA with DASPMI (Figure 4). In the present case, the repulsion between DASPMI molecules is less as the cationic dye DASPMI loses its positive charges after interaction with negative residue Glu 292 and Glu 450 in site I and site II of BSA. The spots observed in the FESEM image point to the binding of protein and dye which is very similar to the binding of protein and DP/DNSA (Figure 4). In CV there is a noticeable decrease in current intensity (314 µA) upon addition of BSA with DASPMI compared to that of DASPMI alone (Figure 5). In CD spectra the significant decrease of negative ellipticity after addition of DASPMI into BSA indicates that the binding of DASPMI induces the R-helical structure of protein. The decrement of R-helices by 20% from free BSA in buffer to bound BSA with DASPMI suggests that binding of DASPMI to BSA brings forth an alteration of the secondary structure of the protein significantly (Figure 6). An upfield shift of about ∆δ 0.03 ppm due to the H atom of CdC in the 1H NMR spectrum was observed in the presence of BSA (Supporting Information). All of these above-mentioned phenomena confirm the formation of a stable and noncovalent complex between DASPMI and BSA. In all the measurements, we have done the control experiments with site-selective binding ligands, viz., DP and DNSA. In BSA, when we add DP or DNSA, we get the same data/images as that with DASPMI which confirms binding of the probe with BSA (Supporting Information). The anisotropy measurement of DASPMI in varying concentration of BSA also gives an indication of binding with BSA. An increment in anisotropy value of the emission of DASPMI with BSA concentration (Figure 7) implies an imposed motional restriction on the fluorophore in the protein environment, leading to the reduction of tumbling motion and greater binding interaction between the probe and BSA (Table 3). Exploring the binding site of any biologically active probe in proteins is the crucial factor for understanding the efficiency of the probe as a therapeutic agent. To know the sites in which the dye gets attached, we determined the micropolarity around the probe in different states of the proteins, that is, at native (N), intermediate (Int), and unfolded (U) states. Figure 8 and Table 4 show that the difference in micropolarity values for the N-Int transition (involving domains I and II) is nearly same as for the Int-U transition (involving domain II). The micropolarity of domain II and III is about 51.7 in terms of ET(30),46 which is very close to our measured values. So it may be inferred that the probe is located in domains II and III, as they are more hydrophobic than domain I.47 As Trp is the most active moiety in BSA, the Förster resonance energy transfer (FRET) from the Tryptophan (Trp) moiety in BSA to DASPMI has been computed as 27%, which suggests that the probe is located near the Trp moiety of domain II. Quenching of BSA fluorescence in buffer as well as in AOT reverse micelles (to be discussed in a later section) indicates that the probe molecule binds in site I. The distance between the donor (BSA) and acceptor (probe) can be calculated according to Förster’s theory for resonance energy transfer (FRET).48 The efficiency of energy transfer, E, is related to the distance (rAD) between the donor and acceptor probe by E) R60 6 R60 + rAD )1- ( ) F F0 (7) where R0 is the Förster distance (critical distance) when the efficiency of energy transfer is 50%. F and F0 are the 10446 J. Phys. Chem. B, Vol. 114, No. 32, 2010 Sahoo et al. Figure 4. SEM images of (a) BSA (15 µM) alone and (b) BSA (15 µM) with DASPMI and (c) BSA (15 µM) with DNSA. Figure 5. Cyclic voltammogram (CV) of (1) DASPMI (0.05 mM) alone and (2) DASPMI in the presence of BSA (15 µM). Figure 7. Variation of fluorescence anisotropy (r) of DASPMI with increasing concentration of BSA. TABLE 3: Fluorescence Anisotropy (r) of DASPMI in Different Concentration of BSA in Buffer concn of BSA (µM) anisotropy value (r) 2 0.12 4 0.17 7 0.22 10 0.26 15 0.3 where F(λ) is the fluorescence intensity of the donor at the wavelength λ, and ε(λ) is the molar absorption coefficient of the acceptor at wavelength λ. J can be evaluated by integrating the overlapped portion of the spectra in Figure 9, and the value of κ2 equals 2/349 if both the donor and acceptor tumble rapidly and free to assume to any orientation. Here n ) 1.4 for proteins generally, and φD ) Figure 6. CD spectra of BSA in buffer solution as a function of DASPMI concentration at (a) 0 mM, (b) 0.02 mM, (c) 0.07 mM, and (d) DASPMI alone. fluorescence intensities of BSA in the presence and absence of quencher, respectively. The value of R0 can be calculated from R0 ) 0.211(κ2n-4φDJ)1/6 (8) where κ2 is the special orientation factor between the emission dipole of the donor and the absorption dipole of the acceptor; n is the refractive index of the medium; φD is the fluorescence quantum yield of the donor; and J is the overlap integral of the fluorescence emission spectrum of the donor and the absorption spectrum of the acceptor and is given by J) ∑ F(λ)ε(λ)λ4 ∆λ ∑ F(λ)∆λ (9) Figure 8. Fluorescence spectra of BSA-bound DASPMI as a function of added urea. Curves 1f6 correspond to 0.0, 3.0, 5.0, 7.0, 8.0, and 9.0 M urea. Inset shows the variation of emission maximum (λemmax) of DASPMI in different solvents against ET(30). (1), (2), and (3) give the interpolated λemmax values of native (N), intermediate (int), and unfolded (U) states of BSA. Albumin Protein with an Ionic Styryl Dye J. Phys. Chem. B, Vol. 114, No. 32, 2010 10447 TABLE 4: Micropolarity Values in Terms of ET(30) at Different States different states of protein BSA native (N) intermediate (int) unfolded (U) 51.8 54.8 59.1 0.11 using the above-mentioned values R0 ) 1.1 nm, E ) 27%, and rAD) 1.2 nm in buffer solution (Table 5). The donor to acceptor distance in both environments is less than 7 nm, indicating a static quenching interaction between the donor and acceptor according to Förster’s nonradiative energy transfer theory.48 To understand and confirm the site-selective binding of DASPMI with BSA, known site selective binding ligands, dansylamide (DNSA) for site I (subdomain IIA) and dansylproline (DP) for site II (subdomain IIIA)50 were used. Initial addition of DNSA in the BSA-DASPMI complex showed a gradual decrease in fluorescence intensity and reached saturation at 0.7 mM (Figure 10a). In this process, DNSA effectively displaced 30% of DASPMI from the BSA-DASPMI complex. A similar experiment was done on the BSA-DASPMI complex with DP which showed a displacement of 70% of DASPMI from the BSA-DASPMI complex by DP (Figure 10b). No noticeable change in intensity due to the titration of DNSA and DP with DASPMI (Figures 11a and b) indicates all the changes that were observed earlier were due to displacement of DASPMI from the BSA-DASPMI complex by binding ligands. The above observation unambiguously helps us to conclude that the dye can bind with site I as well as site II of BSA. This is further corroborated by the observation of the biexponential lifetime having values 4.8 ns (30%) and 2.4 ns (70%). The increase in fluorescence quantum yield along with a hypsochromic shift of emission spectra with the addition of BSA reflects that the microenvironments around the fluorophore in the protein solutions are quite different from those in pure aqueous solution. Possibly the small cavity size (2.53 Å) of the Figure 10. (a) Fluorescence spectra of DASPMI in buffer solution with the addition of BSA (1) 0 µM, (2) 2 µM, (3) 7 µM, (4) 10 µM, and (5) 15 µM followed by the addition of DNSA (6) 0.01 mM, (7) 0.03 mM, (8) 0.05 mM, (9) 0.07 mM, (10) 0.08 mM, and (11) 0.09 mM. (b) Fluorescence spectra of DASPMI in buffer solution with the addition of BSA (1) 0 µM, (2) 2 µM, (3) 7 µM, (4) 10 µM, and (5) 15 µM followed by the addition of DP (6) 0.01 mM, (7) 0.03 mM, (8) 0.05 mM, (9) 0.07 mM, (10) 0.08 mM, and (11) 0.09 mM. (c) Percentage of DASPMI attached in site I and site II of BSA in buffer solution. Figure 9. Overlap of the fluorescence spectrum of BSA (dotted line) with absorption spectra (solid line) of DASPMI. The concentration of BSA and DASPMI are 15 µM and 5 × 10-5 M, respectively. TABLE 5: Calculated Parameters for the BSA/DASPMI Complex in Buffer Solution and in the Reverse Micelle parameters in buffer in reverse micelle E R0 (nm) rAD (nm) 0.27 1.1 1.2 0.36 0.94 0.99 site (I) causes the formation of a tight complex with DASPMI having π stacking and hydrophobic interactions. The aromatic rings form hydrophobic interaction with residues Leu 219, Phe 223, Leu 234, Leu 238, Leu 260, Ala 261, Ile 264, Ile 290, Ala 291, and the hydrocarbon chain of Glu 292.51,52 Also from lifetime data the 30% longer lifetime (4.8 ns) indicates that DASPMI binds with BSA in site I by 30%. This is further confirmed by binding of the ligand DNSA by displacing 30% of DASPMI from the BSA-DASPMI complex. The relatively larger cavity size of 2.6 Å of site II involves binding with DASPMI by hydrogen bonding and hydrophobic and electrostatic interaction. Site II binds DASPMI twice as much as site 10448 J. Phys. Chem. B, Vol. 114, No. 32, 2010 Figure 11. (a) Emission spectra of DASPMI as a function of dansylamide (DNSA) concentration (1) 0 mM, (2) 0.4 mM, and (3) 0.7 mM. (b) Emission spectra of DASPMI as a function of dansylproline (DP) concentration (1) 0 mM, (2) 0.4 mM, and (3) 0.7 mM. I does because site II contains tyrosine residue. The aromatic ring of tyrosine contains a hydroxyl group. This hydroxyl group is very much reactive; however, in site I the tryptophan has no hydroxyl group, so the side chain is inert. In site II, the binding Figure 12. Molecular docking of the DASPMI-BSA complex. Sahoo et al. pocket of the hydrophobic portion of the aromatic ring of DASPMI is packed against Pro 384, Leu 387, Ile 388, Phe 395, Leu 407, Leu 430, Val 433, Ala 449, Leu 453 and also with hydrocarbon chains of Arg 485 and Glu 450. As there is no tryptophan residue in the site II, a relatively loose complex is formed with DASPMI at this site. The strong binding in site I is due to the larger π-system of the tryptophan residue allowing a more pronounced π-stacking effect. This loose binding led us to observe the major component of 70% with a shorter lifetime of 2.4 ns, which is also further corroborated by the 70% displacement of DASPMI from the BSA-DASPMI complex by DP. As binding sites share a common interface, the ligand bound to site II affects the conformational changes, as well as binding affinities in site I. Trp 214, conserved in mammalian albumins, plays an important structural role in the formation of site I by solvent accessibility, and it participates in an additional hydrophobic packing interaction between the site I and site II interface. Molecular Docking of the DASPMI-BSA Complex. To establish which binding site of BSA DASPMI is located in, the complementary applications of molecular docking computed by Auto Dock 3.05 have been employed to improve the understanding of the interaction of DASPMI and BSA. For more accurate docking, we modified the parameters which were 150 for ga_pop_size and 10 000 000 for ga_num_evals. The distance between the Trp and DASPMI is 6.7 Å (0.67 nm) from molecular docking, which is close to that from FRET calculation of experimental data (0.1.2 nm). The DASPMI molecule is situated in a cavity formed by Trp 214, Lys 195, Tyr 150, Lys 190, Glu 153, Gly 188, Leu 198, His 288, Ser 454, and Arg 197 (Figure 12). The residues close to the probe are mainly hydrophobic which indicates the existence of hydrophobic interaction between the residue and probe. The interaction between the BSA and the ligand is not exclusively hydrophobic in nature since several ionic and polar residues in the proximity of the ligand play an important role in stabilizing the probe molecule via hydrogen bond and electrostatic interaction. The formation of the hydrogen bond causes a decrease in the hydrophilicity and an increase in the hydrophobicity which Albumin Protein with an Ionic Styryl Dye J. Phys. Chem. B, Vol. 114, No. 32, 2010 10449 TABLE 6: Thermodynamic Parameters of the System BSA-DASPMI in Reverse Micellar Environment at Different Temperatures T binding constant (K) (Kb × 10 ) (kJ mol ) (kJ mol ) (J mol-1 K-1) 298 303 310 315 5.1 ( 0.3 4.7 ( 0.1 4.1 ( 0.3 3.8 ( 0.2 -15.09 -32.67 -33.0 -33.4 -3.8 58 5 ∆H ∆G -1 ∆S -1 stabilize in the DASPMI-BSA system. The calculated binding free energy (G) for the Trp moiety is (-26.6 kJ mol-1), which is near the experimental data (-24.81 kJ mol-1). The difference in Gibbs free energy comes from limitation of the docking software, which can only handle docking in the single position, whereas in reality DASPMI binds BSA in two different sites which has already been proved by site-detecting probes (vide supra). Interaction in Reverse Micelle Environment. As the objective of our study is for a possible application in biological systems, we investigated all the binding characteristics of DASPMI with BSA within a model biological system, AOT reverse micelle. The suitable value of water pool (W0) of microemulsion for BSA’s stability was found to be by trial method better for W0 ) 25. Within the water pool, the quenching rate of BSA with addition of DASPMI is higher than that in buffer solution (outside the RM). Interestingly, a large change in spectroscopic parameters like quantum yield, lifetimes, and radiative and nonradiative decay constant could be observed inside the reverse micelle compared to that in buffer solution (Table 1). The binding constant (5.1 × 105) of the BSA-DASPMI complex is higher in microemulsion than that in buffer solution, and the change in free energy is -32.67 kJ mol-1 in microemulsion environment. The ∆H0 and ∆S0 values for the binding reaction between DASPMI and BSA in the microemulsion environment were found to be -15.09 kJ mol-1 and 58 J mol-1 K-1 (Table 6). In buffer solution, we used site-selective binding ligand DNSA and DP in the BSA-DASPMI complex to know the attachment efficiency; similarly, DNSA was added to the BSA-DASPMI complex in microemulsion (RM), and we have found an effective 60% displacement of DASPMI from the BSA-DASPMI complex (Figure 13a). A similar experiment with dansylproline (DP) showed that about 40% of DASPMI has been displaced by DP (Figure 13b). The greater binding constant in microemulsion than that in buffer solution indicates greater accessibility of DASPMI toward BSA when BSA is encapsulated in microemulsion, which possibly is due to the different nature of water in microemulsion. In AOT microemulsion in the presence of BSA, DASPMI emission shows a biexponential decay with lifetime values 4.2 ns (60%) and 2.2 ns (40%) compared to the values 4.8 ns (30%) and 2.4 ns (70%) in buffer. The different attachment efficiency in two sites as the protein and dye are placed inside RM compared to that in buffer may be due to one or the combination of the following reasons: (i) the water activity may change due to the specific RM milieu, which may have a profound effect on different sites of protein, as water is considered as an integral part of proteins; (ii) the dielectric constant of water in RM is different than that of bulk water (for W0 ) 25, dielectric constant is ∼23); (iii) the small scale of inner volume of the reverse micelle provides a “confined space” effect on the cavity of different sites of protein; (iv) the concentration of ions inside the water core is higher than that of bulk water which differs the π stacking, Figure 13. (a) Fluorescence spectra of DASPMI in AOT solution with the addition of BSA (1) 0 µM, (2) 2 µM, (3) 7 µM, (4) 10 µM, and (5) 15 µM followed by the addition of DNSA (6) 0.01 mM, (7) 0.03 mM, (8) 0.05 mM, (9) 0.07 mM, (10) 0.08 mM, and (11) 0.09 mM. (b) Fluorescence spectra of DASPMI in AOT solution with the addition of BSA (1) 0 µM, (2) 2 µM, (3) 7 µM, (4) 10 µM, and (5) 15 µM followed by the addition of DP (6) 0.01 mM, (7) 0.03 mM, (8) 0.05 mM, (9) 0.07 mM, (10) 0.08 mM, and (11) 0.09 mM. (c) Percentage of DASPMI attached in site I and site II of BSA in AOT solution. hydrophobic interaction, hydrogen bonding, and electrostatic interaction of protein. According to Förster’s theory for resonance energy transfer (FRET) from BSA to DASPMI in microemulsion, the value of R0, the Förster distance, and rAD, the distance between the donor and acceptor probe, are found to be 0.94 and 0.99 nm, respectively. The calculated FRET efficiency in microemulsion is also seen to be increased to 36% from that of 27% in buffer solution (Table 4). The greater FRET efficiency and smaller donor-acceptor distance in RM also indicate that there is a stronger binding between the BSA-DASPMI complex in RM than in buffer. 10450 J. Phys. Chem. B, Vol. 114, No. 32, 2010 Conclusions The most important upshot of the investigation is that we have demonstrated the site-selective binding of the probe in both water and microemulsion and reversal of efficiency of interaction of BSA with DASPMI as BSA moves inside microemulsion, a realistic biological environment; i.e., subdomain IIA is doubly active inside RM compared to that in buffer solution, and the activity of subdomain IIIA in RM is halved compared to that in buffer solution. In general, binding efficiency also increases along with a decrease in donor-acceptor distance inside RM. This in situ knowledge is very important for drug designing and its delivery, particularly in PDT, and this investigation helps us to design an efficient drug sensitizer. Acknowledgment. The authors express thanks to Mr. Subrata Das, Department of Spectroscopy, I.A.C.S, for taking picosecond time-resolved data. Supporting Information Available: Additional experimental details. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes (1) Mehta, C. S.; Somasundaran, P.; Kulkarni, R. J. Colloid Interface Sci. 2009, 333, 635–640. (2) Street, G. Highly SelectiVe Separations in Biotechnology; Blackie Academic and Professional: Glasgow, 1994. (3) Bertolini, D.; Cassettari, M.; Salvetti, G.; Tombari, E.; Veronesi, S.; Squadrito, G. Prog. Colloid Polym. Sci. 1992, 89, 278–280. (4) Zhang, D. H.; Guo, Z.; Dong, X. Y.; Sun, Y. Biotechnol. Prog. 2007, 23, 108–115. (5) Nishii, Y.; Kinugasa, T.; Nii, S.; Takahashi, K. J. Membr. Sci. 2002, 195, 11–21. (6) Liu, Y.; Dong, X. Y.; Sun, Y. Sep. Purif. Technol. 2007, 53, 289– 295. (7) Naoe, K.; Noda, K.; Kawagoe, M.; Imai, M. Colloids Surf. B 2004, 38, 179–185. (8) Ono, T.; Nagatomo, M.; Nagao, T.; Ijima, H.; Kawakami, K. Biotechnol. Bioeng. 2005, 89, 290–295. (9) Wu, X. Y.; Liu, Y.; Dong, X. Y.; Sun, Y. Biotechnol. Prog. 2006, 22, 499–504. (10) Varshney, M.; Khanna, T.; Changez, M. Colloid Surf. B, Biointerfaces 1999, 13, 1–11. (11) Hebbar, H. U.; Raghavarao, K. S. M. S. Process Biochem. 2007, 42, 1602–1608. (12) Jarudilokkul, S.; Poppenborg, L. H.; Stuckey, D. C. Biotechnol. Bioeng. 1999, 62, 593–601. (13) Hai, M.; Kong, F. J. J. Chem. Eng. Data 2008, 53, 765–769. (14) Rho, S. G.; Kang, C. H. J. Ind. Eng. Chem. 2004, 10, 247–251. (15) Shin, Y. O.; Vera, J. H. Biotechnol. Bioeng. 2002, 80, 537–543. (16) Lee, B. K.; Hong, D. P.; Lee, S. S.; Kuboi, R. Biochem. Eng. J. 2004, 22, 71–79. (17) Szacilowski, K.; Macyk, W.; Drzewiecka-Matuszek, A.; Brindell, M.; Stochel, G. Chem. ReV. 2005, 105, 2647–2694. (18) Bonnett, R. Chemical Aspects of Photodynamic Therapy; Gordon and Breach Science Publishers: The Netherlands, 2000. (19) Henderson, B.; Dougherty, T., Eds. Photodynamic Therapy: Basic Principles and clinical applications; Marcle Dekker Inc.: New York, 1992. Sahoo et al. (20) Cheema, M. A.; Taboada, P.; Barbosa, S.; Juarez, J.; GutierrezPichel, M.; Siddiq, M.; Mosquera, V. J. Chem. Thermodyn. 2009, 41, 439– 447. (21) He, X. M.; Carter, D. C. Nature 1992, 358, 209–215. (22) Dockal, M.; Carter, D. C.; Ruker, F. J. Biol. Chem. 1999, 274, 29303–29310. (23) Sulkowska, A. J. Mol. Struct. 2002, 614, 227–232. (24) Loew, L. M.; Bonneville, G. W.; Surow, J. Biochemistry 1978, 17, 4065–4071. (25) Loew, L. M.; Scully, S.; Simpson, L.; Waggoner, A. S. Nature 1979, 281, 497–499. (26) Loew, L. M.; Simpson, L. L. Biophys. J. 1981, 34, 353–365. (27) Bereiter-Hahn, J. Biochim. Biophys. Acta 1976, 423, 1–14. (28) Bereiter-Hahn, J.; Seipel, K. H.; Vöth, M.; Ploem, J. S. Cell Biochem. Funct. 1983, 1, 147–155. (29) Strehmel, B.; Seifert, H.; Rettig, W. J. Phys. Chem. B 1997, 101, 2232–2243. (30) Strehmel, B.; Rettig, W. J. Biomed. Opt. 1996, 1, 98–109. (31) Sahoo, D.; Chakravorti, S. Photochem. Photobiol. 2009, 85, 1103– 1109. (32) Sahoo, D.; Chakravorti, S. J. Photochem. Photobiol. A: Chem. 2009, 205, 129–138. (33) Spooner, S. P.; Whitten, D. G. In Photochemistry in Organized & Constrained Media; Ramamurthy, V., Ed.; Wiley-VCH: Weinheim, Germany, 1991; pp 691-739. (34) Ulmann, A. An Introduction to Ultrathin Organic Films: FromLangmuir-Blodgett to Self-Assembly; Academic Press: San Diego, CA, 1991; Chapters 3 and 5. (35) Sahoo, D.; Bhattacharya, P.; Chakravorti, S. J. Phys Chem. B 2009, 113, 13560–13565. (36) Sahoo, D.; Bhattacharya, P.; Chakravorti, S. J. Phys. Chem. B 2010, 114, 2044–2050. (37) Chowdhury, P.; Panja, S.; Chakravorti, S. J. Phys. Chem. A 2003, 107, 83–90. (38) Mitra, S.; Tamai, N. Chem. Phys. 1999, 246, 463–475. (39) O’Connor, D. V.; Phillips, D. Time-correlated single photon counting; Academic Press: New York, 1984; Chapter 6. (40) Benesi, M. L.; Hildebrand, J. H. J. Am. Chem. Soc. 1949, 71, 2703– 2707. (41) Tian, J.; Liu, J.; He, W.; Hu, Z.; Yao, X.; Chen, X. Biomacromolecules 2004, 5, 1956–1961. (42) Kamat, P. V.; Das, S.; Thomas, K. G.; George, M. V. J. Phys. Chem. 1992, 96, 195. (43) Petitpas, I.; Bhattacharya, A. A.; Twine, S.; East, M.; Curry, S. J. Biol. Chem. 2001, 30, 243–249. (44) Morris, G. SYBYL Software, Version 6.9; Tripos Associates: St. Louis, 2002. (45) Ramadass, R.; B-Hahn, J. J. Phys. Chem. B 2007, 111, 7681–7690. (46) Mallick, A.; Haldar, B.; Chattopadhya, N. J. Phys. Chem. B 2005, 109, 14683–14687. (47) Peters, T. Serum albumin. AdVances in protein Chemistry; Academic Press: New York, 1985; Vol. 37, pp 161-245. (48) Förster, T., Sinanoglu, O., Eds. Modern Quantum Chemistry; Academic Press: New York, 1965; Vol. 3. (49) Förster, T. Discuss. Faraday Soc. 1959, 27, 7–17. (50) Pandey, R. K.; Constantine, S.; Tsuchida, T.; Zheng, G.; Medforth, C. J.; Aoudia, M.; Kozyrev, A. N.; Rodgers, M. A. J.; Kato, H.; Smith, K. M.; Dougherty, T. J. J. Med. Chem. 1997, 40, 2770–2779. (51) Sudlow, G.; Birkett, D. J.; Wade, D. N. Mol. Pharmacol. 1975, 11, 824–832. (52) Sudlow, G.; Birkett, D. J.; Wade, D. N. Mol. Pharmacol. 1977, 12, 1052–1061. JP102937Y
© Copyright 2025 Paperzz