Acessibilidade / Reportar erro

Activated Complex Approach to Describe Bovine Serum Albumin-Azure A and Bovine Serum Albumin-Azure B Intermolecular Interactions

Abstract

Azure A (AZA) and azure B (AZB) phenothiazine dyes are used for clinical and medical purposes, and their functions can be altered via interactions with proteins. However, no kinetics information on the interactions between phenothiazine dyes and bovine serum albumin (BSA) is available. Surface plasmon resonance was used to determine the energetic and dynamic of the BSA-AZA and BSA-AZB complexes formation at pH 7.4. At temperature ≤ 16 °C, the formation of activated (DHa,12°C,AZA= -310.57 kJ mol-1 and DHa,12 °C,AZB= -256.37 kJ mol-1) and thermodynamically stable (DH°12°C,AZA= -314.56 kJ mol-1 and DH°12°C,AZB= -265.73 kJ mol-1) complexes was driven by enthalpy, while at temperature ≥ 20 °C, by entropy, (TDSa,28°C,AZA= 207.49 and TDSa,28°C,AZB= 190.69; TDS°28°C,AZA= 277.50 and TDS°28°C,AZB= 257.26 kJ mol-1). Hydrophobic interactions were fundamental to the complex stability and the increase in number of -CH3 groups in the dyes do not affect kinetic and thermodynamic parameters. Our results could help optimize the medical and pharmaceutical applications of phenothiazine dyes.

Keywords:
protein-dye interaction; BSA; phenothiazine dyes; thermodynamics; kinetics; SPR


Introduction

Several organic dye molecules have drawn attention recently owing to their different applications in the pharmaceutical,11 Rajan, D.; Ilanchelian, M.; Int. J. Biol. Macromol. 2018, 113, 1052. food,22 Dotto, G. L.; Pinto, L. A. A.; Hachicha, M. A.; Knani, S.; Food Chem. 2015, 171, 1. medical,33 Blivis, D.; Falgairolle, M.; O’Donovan, M. J.; Sci. Rep. 2019, 9, 3201. and textile44 Soares, O. S. G. P.; Órfão, J. J. M.; Portela, D.; Vieira, A.; Pereira, M. F. R.; J. Hazard. Mater. 2006, 137, 1664. industries. In this context, the phenothiazine dyes, azure A and B (AZA and AZB, respectively), which feature similar molecular chemical structures (Figure 1), present broad applications in medicine owing to their high affinity for proteins.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.,66 Moreira, L. M.; Lyon, J. P.; Romani, A. P.; Severino, D.; Rodrigues, M. R.; de Oliveira, H. P. M. In Advanced Aspects of Spectroscopy; Farrukh, M. A., ed.; InTech: Rijeka, Croatia, 2012, p. 393-422.

Figure 1
Chemical structures of azure A and B dyes.

Azure A is used to diagnose amyloid accumulation, predict the extent of heart disease, as antimalarial, and as photo-chemotherapeutic agent against carcinomas.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.,77 Vennerstrom, J. L.; Makler, M. T.; Angerhofer, C. K.; Williams, J. A.; Antimicrob. Agents Chemother. 1995, 39, 2671.

8 Khan, A. Y.; Suresh Kumar, G.; J. Chem. Thermodyn. 2015, 91, 225.
-99 Diabou Gaye-Seye, M.; Aaron, J.-J.; Parkanyi, C.; Motohashi, N.; Curr. Drug Targets 2006, 7, 1083. Azure B also presents a variety of pharmacological properties, and is used to diagnose amyloid accumulation-related diseases, detect oral cancer, diagnose neurodegenerative diseases, and as inhibitor of b-amyloid protein filament.88 Khan, A. Y.; Suresh Kumar, G.; J. Chem. Thermodyn. 2015, 91, 225.,1010 Oz, M.; Lorke, D. E.; Petroianu, G. A.; Biochem. Pharmacol. 2009, 78, 927.,1111 Delport, A.; Harvey, B. H.; Petzer, A.; Petzer, J. P.; Life Sci. 2014, 117, 56. However, the functions and toxicity of AZA or AZB dyes can be altered by their interactions with proteins, such as bovine serum albumin (BSA),1212 Flarakos, J.; Morand, K. L.; Vouros, P.; Anal. Chem. 2005, 77, 1345. one of the most abundant protein in the bovine blood plasma.

BSA is responsible for maintaining the pH of the skin, helps control the colloidal osmotic pressure, and plays a dominant role in the disposal and transport of dyes, drugs, fatty acids, and endogenous and exogenous compounds in the bovine body.1313 Mathew, T. V.; Kuriakose, S.; Colloids Surf., B 2013, 101, 14. Despite the great importance of understanding BSA-AZA and BSA-AZB interactions, unfortunately only three papers55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.,1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212.,1515 Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609. have been published describing these types of bonding, in which, only one analyzed the thermodynamic interaction between these species1515 Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609. and none of them investigated the kinetics of the BSA-AZA or BSA-AZB complex formation.

The interactions of BSA with AZA or AZB were studied using spectroscopic techniques (ultraviolet-visible, Fourier-transform infrared, circular dichroism, and time-resolved fluorescence),55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.,1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212.,1515 Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609. and calorimetry.1515 Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609. The results indicated that the AZA or AZB dyes were strongly bound to the hydrophobic Sudlow’s site I of BSA and this process was enthalpy driven (DH°AZA= −4.12 ± 0.081 kcal mol-1 and DH°AZB= −2.38 ± 0.041 kcal mol-1).1515 Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609. However, to the best of our knowledge, no kinetic studies on the interactions between BSA and AZA or AZB have been performed so far. Therefore, it is of great importance to describe the molecular dynamics of the BSA-AZA and BSA-AZB interactions, and to measure the thermodynamic parameters associated with these interactions using surface plasmon resonance (SPR).1616 Amaral, M.; Kokh, D. B.; Bomke, J.; Wegener, A.; Buchstaller, H. P.; Eggenweiler, H. M.; Matias, P.; Sirrenberg, C.; Wade, R. C.; Frech, M.; Nat. Commun. 2017, 8, 2276.

SPR is a free-labeled optical method for the real-time detection of interactions between proteins and different solutes,1717 Hudson, E. A.; de Paula, H. M. C.; da Silva, R. M.; Pires, A. C. S.; da Silva, L. H. M.; Int. J. Biol. Macromol. 2019, 133, 860.

18 Fathi, F.; Mohammadzadeh-Aghdash, H.; Sohrabi, Y.; Dehghan, P.; Dolatabadi, J. E. N.; Food Chem. 2018, 246, 228.
-1919 Taghipour, P.; Zakariazadeh, M.; Sharifi, M.; Dolatabadi, J. E. N.; Barzegar, A.; J. Photochem. Photobiol., B 2018, 183, 11. and its generated data would allow us to calculate the kinetic binding constants, such as the association rate constant (ka), dissociation rate constant (kd), as well as the thermodynamic binding constant (Kb). From the temperature dependence of these kinetic and thermodynamic constants it would be possible to determine the following parameters: activation energy, changes in the Gibbs free energy, enthalpy, and entropy for the transition process, and changes in the standard Gibbs free energy, enthalpy, and entropy (Eact, ΔG, ΔH, ΔS, ΔG°, ΔH°, and ΔS°, respectively). Recently, many researchers have reported the use of the SPR technology to investigate drug-protein,2020 Juhász, Á.; Luty-Błocho, M.; Wojnicki, M.; Tóth, G. K.; Csapó, E.; Microchem. J. 2019, 147, 311.

21 Lima, C. H. M.; de Paula, H. M. C.; da Silva, L. H. M.; Rocha, M. S.; Biopolymers 2017, 107, e23071.

22 Mowla, R.; Wang, Y.; Ma, S.; Venter, H.; Biochim. Biophys. Acta, Biomembr. 2018, 1860, 878.
-2323 Patching, S. G.; Biochim. Biophys. Acta, Biomembr. 2014, 1838, 43. dye-protein,2424 de Paula, H. M. C.; Coelho, Y. L.; Agudelo, A. J. P.; Rezende, J. P.; Ferreira, G. M. D.; Ferreira, G. M. D.; Pires, A. C. S.; da Silva, L. H. M.; Colloids Surf., B 2017, 159, 737.,2525 Coelho, Y. L.; de Paula, H. M. C.; Agudelo, A. J. P.; de Castro, A. S. B.; Hudson, E. A.; Pires, A. C. S.; da Silva, L. H. M.; Int. J. Biol. Macromol. 2019, 136, 559. protein-food additive,1818 Fathi, F.; Mohammadzadeh-Aghdash, H.; Sohrabi, Y.; Dehghan, P.; Dolatabadi, J. E. N.; Food Chem. 2018, 246, 228. protein-nanoparticle,2626 Atale, S. S.; Dyawanapelly, S.; Jagtap, D. D.; Jain, R.; Dandekar, P.; Int. J. Biol. Macromol. 2019, 123, 97. low molecular substance-protein,2727 Patil, S.; Sistla, S.; Jadhav, J.; Int. J. Biol. Macromol. 2016, 92, 1123. polyphenol-protein2828 Rezende, J. P.; Hudson, E. A.; de Paula, H. M. C.; Coelho, Y. L.; da Silva, L. H. M.; Pires, A. C. S.; Food Hydrocolloids 2019, 95, 526.,2929 Rezende, J. P.; Hudson, E. A.; de Paula, H. M. C.; Meinel, R. S.; da Silva, A. D.; Pires, A. C. S.; da Silva, L. H. M.; Food Chem. 2020, 307, 125514. interactions.

In this study, we investigated the interactions between BSA and AZA and AZB using the SPR technique and determined all kinetic and thermodynamic parameters for a better understanding of the molecular dynamics and energetic changes associated with these interactions.

Experimental

Chemicals and apparatus

Azure A (purity ≥ 70 wt.%), azure B (purity ≥ 89 wt.%), and BSA (purity ≥ 99 wt.%) were purchased from Sigma-Aldrich (Saint Louis, USA). All chemicals used in this study were of analytical grade and were used without further purification. Deionized water was used for all experiments.

Research-grade CM5 sensor chips and coupling reagents (N-ethyl-N’,N’dimethylaminopropyl carbodiimide (EDC), N-hydroxysuccinimide (NHS), and 1 M ethanolamine hydrochloride, pH 8.5) were purchased from GE Healthcare (Pittsburgh, USA). All SPR experiments to investigate the kinetic and thermodynamic parameters of the BSA-AZA and BSA-AZB interactions were performed using a Biacore X100 instrument (GE Healthcare, Pittsburgh, USA).

Solution preparation and method

The CM5 chips were activated for 7 min using EDC/NHS. Afterward, the excess activated carboxyl groups were blocked using ethanolamine for 7 min. Subsequently, BSA was immobilized onto the chips (15 µg mL-1) in 10 mM sodium acetate, pH 4.0. The immobilization of BSA was performed at low density (3728 resonance units (RUs)) to reduce any potential mass transport and crowding artifacts. During experiments using immobilized BSA, one flow cell was used as reference surface; this surface was prepared as describe above, but without BSA immobilization.3030 Drake, A. W.; Tang, M. L.; Papalia, G. A.; Landes, G.; Haak-Frendscho, M.; Klakamp, S. L.; Anal. Biochem. 2012, 429, 58.

Dye solutions were prepared at working concentrations (1-8 µM) in flow buffer. Each dye concentration was injected under the surface of the chip with immobilized BSA (channel 1) for 15 s at the flow of 30 µL min-1 to allow binding. Then, the flow buffer was allowed to run through the surface to dissociate for 35 s at the flow of 30 µL min-1, to regenerate the surface of the chip. The procedure was performed for each AZA and AZB concentration. The channel without immobilized BSA (channel 2) underwent the same procedure. The RU response was obtained by subtracting the channel 2 response from the channel 1. The BSA-dye binding experiments were performed at pH 7.4 at temperatures ranging from 12 to 28 °C.

Results and Discussion

Analysis of BSA-AZA and BSA-AZB binding kinetics using SPR

The analysis of the SPR signal, i.e., RU dependence of time (RU vs t), allowed us to obtain the ka and kd kinetic constants values of the intermolecular interactions between different compounds. In addition, the temperature dependence of ka and kd allowed us to analyze the energetic and molecular dynamics of the activated (transition) complex formation process.3131 Schubert, F.; Zettl, H.; Häfner, W.; Krauss, G.; Krausch, G.; Biochemistry 2003, 42, 10288. Figures 2a and 2b illustrate the sensorgrams (RU vs. t) of the BSA-AZA and BSA-AZB interactions, respectively, in the concentration range of 1-8 µM, using immobilized BSA at low density (3728 RU) on the surface of Biacore CM5 chips at 25 °C. Similar results were obtained at other temperatures (Figures S1 and S2, Supplementary Information (SI) section Supplementary Information Supplementary information is available free of charge at http://jbcs.sbq.org.br as PDF file. ).

Figure 2
Sensorgrams of immobilized bovine serum albumin interactions with (a) azure A and (b) azure B at low density (3728 resonance units (RUs)) on surface of Biacore CM5 chips, at 25 °C; 1 ≤ dye concentration ≤ 8 µM.

The sensorgrams were obtained by subtracting the signal of the reference cell from the signals of the sample cells, and could be described using three regions (I, II, and III). In region I, called the baseline region, the buffer entered both the sample and reference cells, and therefore the difference between signals was close to zero. For region II (0 ≤ t ≤ 25 s), the buffer-dye solutions entered both cells causing the RU signal to grow continuously over time. During these 25 s, associations between the free BSA and dye molecules and dissociations of the BSA-dye thermodynamic stable complexes adsorbed on the surfaces of the Biacore CM5 chips occurred simultaneously. After 25 s, region III (25 < t ≤ 50 s) started, and only pure buffer entered the reference and sample cells, again, which led to the decrease in the RU signals over time until the RU values became equal to the baseline signal. The decrease in the RU signals occurred because in region III the dissociation of the BSA-AZA and BSA-AZB complexes occurred predominantly.

The complex formation process attributed to the BSA-AZA and BSA-AZB interactions could be described using a 1:1 reversible interaction model. The kinetics of the BSA-dye association and BSA-dye complex dissociation processes could be described using equation 1.

(1) BSA + dye k d k a BSA dye

where ka is the kinetic association constant of the free BSA and dye molecules for complex formation and kd is the kinetic dissociation constant of the BSA-dye complex. To elucidate the kinetics features of the process, calculating the ka and kd rate constants was required.3232 Fathi, F.; Dolatanbadi, J. E. N.; Rashidi, M.; Omidi, Y.; Int. J. Biol. Macromol. 2016, 91, 1045.,3333 Ehteshami, M.; Rasoulzadeh, F.; Mahboob, S.; Rashidi, M.; J. Lumin. 2013, 135, 164. Using global fitting analyses, all sensorgrams were fitted to equations 2 and 3, which were, then, used to calculate the observed rate constant, kobs, and kd, respectively.

(2) RU t = RU max 1 e k obs t
(3) RU t = RU t f e k d t t f

where RU(t) is the resonance response at time t, RUmax is the maximum resonance response at t = ∞, and RU(tf) is the amplitude of the response at the end time of the buffer-dye flow. The kobs values were linearly dependent on the AZA and AZB concentrations (Figures S3a and S3b) for BSA-AZA and BSA-AZB, respectively), i.e., kobs= ka[dye] + kd, where [dye] is the concentration of dye. This allowed us to calculate ka from the slope of the kobs vs.[dye] plot. Table 1 summarizes the kinetic parameters (ka and kd) values for the BSA-AZA and BSA-AZB complex formation processes at six different temperatures.

Table 1
Association (ka) and dissociation (kd) kinetic rate constants for bovine serum albumin-azure A (BSA-AZA) and bovine serum albumin-azure B (BSA-AZB) interactions at six different temperatures

To the best of our knowledge, the kinetic constants of the BSA-AZA and BSA-AZB interactions have not been reported yet. The values of ka for the formation of the thermodynamically stable BSA-AZB complex and kd for the dissociation of that complex were higher than those for the BSA-AZA complex, which indicated that the increase in the number of -CH3 groups (AZA and AZB comprise 2 and 3 -CH3 groups, respectively) in the chemical structure of the dye sped up the formation and dissociation of complexes. The dissociation processes of both BSA-dye complexes were faster at higher temperature than at lower temperature. However, the temperature effect on the association process was different from that observed for the dissociation process. In the 12-20 °C temperature range, ka decreased as the temperature increased, while in the 20-28 °C temperature interval the opposite behavior was observed.

Sharma et al.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267. and Sharma et al.1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212. reported that AZB and AZA bound to the same hydrophobic BSA site located in subdomain IIA of site I, and therefore, the differences in the kinetic parameters values for the formation and dissociation of the BSA-dye complexes were due to the dynamic structural change of the dyes, as well as, of the BSA interacting site.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.,1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212.

At pH 7.4, the dyes molecules were positively charged while BSA carried negative charge, which promoted favorable electrostatic interactions between BSA and the dyes. Despite the electrostatic repulsion between Congo red (CR) and BSA, as well as the large size of CR compared with those of AZA or AZB, the interaction between the free CR and BSA molecules occurred two to ten times faster than those between BSA-AZA and BSA-AZB.2424 de Paula, H. M. C.; Coelho, Y. L.; Agudelo, A. J. P.; Rezende, J. P.; Ferreira, G. M. D.; Ferreira, G. M. D.; Pires, A. C. S.; da Silva, L. H. M.; Colloids Surf., B 2017, 159, 737. The higher rate for the CR-BSA complex formation was caused by the six free-rotating benzene rings present in the structure of CR, while AZA and AZB only comprise three condensed rings that were prevented from rotating to gain favorable conformations. Then, we could conclude that the CR molecules could self-orient faster than the AZA or AZB molecules, to form p-p stacking interactions between the benzene rings and aromatic amino acids present at the BSA binding sites. However, these p-p stacking interactions between CR and BSA caused the kd value for the BSA-CR complex formation to decrease six times compared with the values for the formation of the BSA-AZA and BSA-AZB complexes. To determine the energetic parameters associated with the BSA binding site conformation fit during the interaction with AZA and AZB, we investigated the temperature dependences of ka and kd. The Arrhenius plots (Figures 3a and 3b for BSA-AZA and BSA-AZB, respectively) are usually utilized to describe the temperature dependences of ka and kd.

Figure 3
Arrhenius plots of ln ka and ln kd for (a) bovine serum albumin-azure A and (b) bovine serum albumin-azure B interactions as functions of reciprocal temperature.

The temperature effect on the BSA-dye association processes was not linear, which suggested that the BSA-AZA and BSA-AZB interactions occurred via multi-step processes that were probably controlled by the site fitting on the BSA site I structure. However, the kd values for both thermodynamically stable complexes was linearly dependent on temperature, which indicated that the dissociation processes occurred via single-step processes.

The energetic parameters related to the formation of the activated complexes from the associations of the free BSA and AZA and AZB molecules or from the dissociation of the BSA-AZA and BSA-AZB thermodynamically stable complexes, were obtained using the Arrhenius relationship (ln kx vs. 1/T).

(4) E act = R × d ln k x d 1 / T

where Eact is the activation energy (J mol-1), R the universal gas constant (8.314 J mol-1 K-1), T is the temperature (K), and kx= ka or kd.

To determine the activation complex energy parameters (ΔGx, ΔHx, and ΔSx) involved in the formation of the transition complexes, we used equations 5-7.

(5) Δ G x = RT ln k x h K B T
(6) E act = Δ H x + RT
(7) T Δ S x = Δ H x Δ G x

where kx is the kinetic constant (ka or kd), h is Planck’s constant (6.6262 × 10-34 J s), and KB is Boltzmann’s constant (1.38 × 10-23 J K-1). Table 2 summarizes the values of all parameters at six different temperatures.

Table 2
Thermodynamic parameters for (a) formation of transition BSA-AZA and BSA-AZB complexes and (d) dissociation of complexes in 12-28 °C temperature range

The Eact values for the formation of the activated complexes ([BSA-AZA] and [BSA-AZB]) via the interactions of the free BSA and dyes molecules, were negative at lower temperatures and increased as the temperature increased, while Eact associated with the dissociation of the BSA-dye thermodynamic stable complexes was temperature independent.

Typically, Eact can be expressed as the sum of three energetic contributions, as follows: (i) desolvation energy, EDes, which is the energy required to remove the solvating layer of the free reagents; (ii) conformational change energy, Econf, which is associated with the biopolymer site fitting and dye chemical structure adjustment; and (iii) interaction energy, Eint, attributed to the interactions of the reactants (equation 8). Each one of these contributions could be temperature dependent.

(8) E act = E Des + E inf + E conf

The EDes term is positive, because to release the solvent from the solute solvation shell, energy should be absorbed from the surroundings to break up the solvent-solute interactions. Moreover, Econf is positive, because it is associated with the breaking down of the amino acid-amino acid interactions that occur at BSA sites. However, Eint is negative, owing to the new intermolecular bonds formed by the BSA-dye interactions.

When the free BSA and dye molecules associated at T < 20 °C, Eact,a was negative owing to two causes. (i) The magnitude of Econf was low because at low temperature the molecular kinetic average energy was not high enough, and therefore, the energy transfer from particle collisions was not sufficient to overcome the potential energy barrier associated with the BSA site conformational change. This conferred a less flexible character to this interacting site. (ii) Consequently, the dye molecules could not penetrate much inside the BSA molecules sites, which promoted the release of small amounts of water molecules from the solvation shells of BSA and dyes, and therefore, caused the small EDes values. Subsequently, |EDes + Econf| < |Eint|.

As the temperature increased (and consequently the molecular average kinetic energy increased), the energy transferred via molecular collisions increased, which caused intense conformational changes to the BSA interacting sites, and therefore induced higher degrees of dye penetration, as well as, the intense desolvation of interacting molecules. This intensification of the conformational site changes and the desolvation process promoted by the increase in temperature caused Eact,a to became positive, and therefore, |EDes + Econf| > |Eint|. The same analysis could be applied to ΔHa and TΔSa.

During the association process of the free BSA and dyes molecules in the 12-20 °C temperature range, increased as the temperature increased, while the opposite behavior was observed in the 20-28 °C temperature range. In addition, ΔGa remained fairly constant (approximately 50 kJ mol-1), and ΔHa and TΔSa changed much more than ΔGa, which suggested the possible activation enthalpy-entropy compensation. To verify the occurrence of this compensation process we have plotted ΔHa as function of TΔSa for the formation of the [BSA-AZA] and [BSA-AZB] transition complexes (Figure 4).

Figure 4
Enthalpy-entropy (ΔHa-TΔSa) plots for [BSA-AZA] and [BSA-AZB] transition complexes (here, BSA, AZA, and AZB are bovine serum albumin, azure A, and azure B, respectively).

The relationships between ΔHa and TΔSa were linear, and the correlation coefficients, r2, were equal to 0.999 for the BSA-dye association processes, which demonstrated that the BSA-AZA and BSA-AZB interactions occurred via an iso-kinetic enthalpy-entropy compensation process. The linear relationship between ΔHa and TΔSa could be expressed using equation 9.

(9) Δ H a = α β T Δ S a

where α and β are adjustable constants. Here, a ca. 54 kJ mol-1 and β ca. 291 K are ΔGa and the compensation temperature, respectively. This iso-kinetic compensation revealed the typical activated energetic signature for the hydrophobic interactions, which was promoted by the changes in the structure of the solvation layers of the interacting molecules, i.e., during the association process of the free BSA and dyes molecules, two sub-process could occur: (i) BSA and dye desolvation and (ii) direct interactions between the desolvated BSA and dyes molecules. The ΔHa and TΔSa values associated with the first and second sub-processes were positive and negative, respectively. During the BSA-dye association process at T < 20 °C, the desolvated BSA-dye direct interaction dominated the magnitudes of ΔHa and TΔSa, mainly because in this temperature range, the desolvation process occurred at low extent. However, as the temperature increased, the release of water molecule solvating the free BSA and dyes molecules determined the ΔHa and TΔSa values, and caused both activated kinetic parameters to be positive at higher temperatures.3434 Paul, B. K.; Ghosh, N.; Mukherjee, S.; RSC Adv. 2016, 6, 9984.,3535 Liu, L.; Guo, Q.; Chem. Rev. 2001, 101, 673.

The transformation processes of the [BSA-AZA] and [BSA-AZB] transition complexes into the final BSA-AZA and BSA-AZB thermodynamically stable complexes, respectively, were accompanied by corresponding enthalpy releases of 7.07 and 9.24 kJ mol-1. However, these processes caused large entropy increases for BSA-AZA and BSA-AZB (69 and 67 kJ mol-1, respectively), and therefore the transformation of the [BSA-dyes] transition complexes into thermodynamically stable BSA-dye complexes was driven by biopolymer and dyes conformational changes.

Analysis of thermodynamic parameters of BSA-AZA and BSA-AZB binding

To determine the stability and driving force for the formation of the thermodynamically stable BSA-dye complexes, it was necessary to measure the following standard thermodynamic parameters for the complex formation: DG°, DH°, and (DS°). All these thermodynamic parameters represented the changes in the thermodynamic properties of the system when 1 mol complex was formed after 1 mol BSA and 1 mol dye interacted. By considering that the BSA-dye complexes were formed via unimolecular mechanisms, Kb could be calculated using the classic relationship, Kb= ka/kd.2020 Juhász, Á.; Luty-Błocho, M.; Wojnicki, M.; Tóth, G. K.; Csapó, E.; Microchem. J. 2019, 147, 311. From the Kb values (equation 10) and its temperature dependence (equation 11) (Figure S4, SI section Supplementary Information Supplementary information is available free of charge at http://jbcs.sbq.org.br as PDF file. ), we could calculate DG° and DH°, while TDS° could be determined using the fundamental thermodynamic equation (equation 12).

(10) Δ G = RT ln K b
(11) ln K b 2 K b 1 = Δ H R 1 T 2 1 T 1
(12) T Δ S = Δ H Δ G

Table 3 lists the values of the thermodynamic parameters for the formation of the BSA-AZA and BSA-AZB complexes at different temperatures.

Table 3
Standard Gibbs free energy, enthalpy, and entropic parameter changes (ΔG°, ΔH°, and TΔS°) for formation of BSA-AZA and BSA-AZB complexes at temperatures in the 12-28 °C range and pH 7.4

All thermodynamic parameters values were temperature dependent and followed second order polynomial trends as the temperature increased. As the temperature increased, DG° increased until it reached the maximum value at T = 20 °C, and then decreased as the temperature continued to increase. Both DH° and TDS° increased monotonically as the temperature increased. The DG° values were negative at all temperatures, which demonstrated that for the mixtures of BSA and AZA or AZB aqueous solutions, the thermodynamic equilibrium conditions promoted higher concentrations of the BSA-AZA and BSA-AZB complexes. The chemical structure differences between dyes did not affect the magnitude of DG°. Sharma et al.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267. and Sharmaet al.1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212. used fluorescence spectroscopy to determine the DG° values for the BSA-AZA and BSA-AZB complexes (-30.35 and -33.61 kJ mol-1, respectively) which differed by approximately 20% from our results. Unfortunately, the authors did not calculate the DH° or TDS° parameters for these complexes. The small differences between the results reported by Sharma et al.55 Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267. and Sharma et al.1414 Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212. and DG° values obtained in this study could be attributed to the smaller configuration entropy contribution during the SPR experiment, which was mainly due to the protein molecules attached to the surfaces of these chips. To investigate the driving force of the BSA-dye complexes formation, we needed to analyze the thermodynamic potential components DH° and TDS°. Both thermodynamic parameters were negative at low temperatures and became positive at T = 20 °C. Both the BSA-AZA and BSA-AZB interactions occurred via enthalpy-entropy compensation (Figure S5, SI section Supplementary Information Supplementary information is available free of charge at http://jbcs.sbq.org.br as PDF file. ), which demonstrated that, as observed for the activated complex formation, during the synthesis of the thermodynamically stable complexes it was the hydrophobic interaction (release of water from the solvation shells of the BSA and dyes molecules and direct desolvated BSA and dye interactions) that determined the molecular mechanism for the formation of the BSA-dye complexes. This temperature increase effect on the BSA-AZA and BSA-AZB hydrophobic interaction could be associated with the partial destruction of the folded protein structure, exposing the inner hydrophobic core.

Conclusions

The SPR technique confirmed that BSA interacted with phenothiazine dyes (AZA and AZB) to forming thermodynamically stable complexes (Kb of 2.1 × 104 and 2.39 × 104 M-1, respectively, at 25 °C) at pH 7.4. However, the driving force for the formation of the BSA-phenothiazine dye complexes was temperature dependent. At temperature ≤ 16 °C, enthalpic forces, mainly owing to hydrogen bonding and van der Waals interactions, governed the formation of complexes (DH°12°C,AZA = -314.56 kJ mol-1 and DH°12°C,AZB = -265.73 kJ mol-1), while at temperature ≥ 20 °C, owing to the predominance of the hydrophobic interactions, the driving force was the entropy (TDS°28°C,AZA = 277.50 kJ mol-1 and TDS°28°C,AZB = 257.26 kJ mol-1). The results of this study suggested that BSA interacted with AZA and AZB via transition complexes. Protein-dye associations occurred via multi-step processes and the activation energy values increases as the temperature increased (Eact,a,12°C,AZA = -308.20 kJ mol−1 increased to Eact,a,28°C,AZA= 259.64 kJ mol−1, and Eact,a,12°C,AZB = -254.00 kJ mol−1 increased to Eact,a,28°C,AZB = 242.67 kJ mol−1). On the other hand, during the dissociation of the thermodynamically stable BSA-AZA and BSA-AZB complexes into activated complexes, single-step processes occurred (Eact,d,AZA = 9.54 ± kJ mol−1 and Eact,d,AZB= 11.70 kJ mol−1). The thermodynamically stable BSA-AZA and BSA-AZB complexes formed via enthalpy-entropy compensation processes, and the [BSA-AZA] and [BSA-AZB] activated complexes were produced via iso-kinetic compensation. The kinetic and thermodynamic analyses of the interactions between the AZA and AZB dyes and BSA could help elucidate the molecular mechanism of complex formation between organic dyes and different proteins, which could prove useful for designing protein-binding therapeutic molecules.

Acknowledgments

This work was supported by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and Fundação de Apoio à Pesquisa de Minas Gerais (FAPEMIG).

Supplementary Information

Supplementary information is available free of charge at http://jbcs.sbq.org.br as PDF file.

References

  • 1
    Rajan, D.; Ilanchelian, M.; Int. J. Biol. Macromol. 2018, 113, 1052.
  • 2
    Dotto, G. L.; Pinto, L. A. A.; Hachicha, M. A.; Knani, S.; Food Chem. 2015, 171, 1.
  • 3
    Blivis, D.; Falgairolle, M.; O’Donovan, M. J.; Sci. Rep. 2019, 9, 3201.
  • 4
    Soares, O. S. G. P.; Órfão, J. J. M.; Portela, D.; Vieira, A.; Pereira, M. F. R.; J. Hazard. Mater. 2006, 137, 1664.
  • 5
    Sharma, A. S.; Anandakumar, S.; Ilanchelian, M.; RSC Adv. 2014, 4, 36267.
  • 6
    Moreira, L. M.; Lyon, J. P.; Romani, A. P.; Severino, D.; Rodrigues, M. R.; de Oliveira, H. P. M. In Advanced Aspects of Spectroscopy; Farrukh, M. A., ed.; InTech: Rijeka, Croatia, 2012, p. 393-422.
  • 7
    Vennerstrom, J. L.; Makler, M. T.; Angerhofer, C. K.; Williams, J. A.; Antimicrob. Agents Chemother. 1995, 39, 2671.
  • 8
    Khan, A. Y.; Suresh Kumar, G.; J. Chem. Thermodyn. 2015, 91, 225.
  • 9
    Diabou Gaye-Seye, M.; Aaron, J.-J.; Parkanyi, C.; Motohashi, N.; Curr. Drug Targets 2006, 7, 1083.
  • 10
    Oz, M.; Lorke, D. E.; Petroianu, G. A.; Biochem. Pharmacol. 2009, 78, 927.
  • 11
    Delport, A.; Harvey, B. H.; Petzer, A.; Petzer, J. P.; Life Sci. 2014, 117, 56.
  • 12
    Flarakos, J.; Morand, K. L.; Vouros, P.; Anal. Chem. 2005, 77, 1345.
  • 13
    Mathew, T. V.; Kuriakose, S.; Colloids Surf., B 2013, 101, 14.
  • 14
    Sharma, A. S.; Subramanian, N.; Malaichamy, I.; J. Photochem. Photobiol., B 2016, 164, 212.
  • 15
    Das, S.; Islam, M. M.; Jana, G. C.; Patra, A.; Jha, P. K.; Hossain, M.; J. Mol. Recognit. 2017, 30, e2609.
  • 16
    Amaral, M.; Kokh, D. B.; Bomke, J.; Wegener, A.; Buchstaller, H. P.; Eggenweiler, H. M.; Matias, P.; Sirrenberg, C.; Wade, R. C.; Frech, M.; Nat. Commun. 2017, 8, 2276.
  • 17
    Hudson, E. A.; de Paula, H. M. C.; da Silva, R. M.; Pires, A. C. S.; da Silva, L. H. M.; Int. J. Biol. Macromol. 2019, 133, 860.
  • 18
    Fathi, F.; Mohammadzadeh-Aghdash, H.; Sohrabi, Y.; Dehghan, P.; Dolatabadi, J. E. N.; Food Chem. 2018, 246, 228.
  • 19
    Taghipour, P.; Zakariazadeh, M.; Sharifi, M.; Dolatabadi, J. E. N.; Barzegar, A.; J. Photochem. Photobiol., B 2018, 183, 11.
  • 20
    Juhász, Á.; Luty-Błocho, M.; Wojnicki, M.; Tóth, G. K.; Csapó, E.; Microchem. J. 2019, 147, 311.
  • 21
    Lima, C. H. M.; de Paula, H. M. C.; da Silva, L. H. M.; Rocha, M. S.; Biopolymers 2017, 107, e23071.
  • 22
    Mowla, R.; Wang, Y.; Ma, S.; Venter, H.; Biochim. Biophys. Acta, Biomembr. 2018, 1860, 878.
  • 23
    Patching, S. G.; Biochim. Biophys. Acta, Biomembr. 2014, 1838, 43.
  • 24
    de Paula, H. M. C.; Coelho, Y. L.; Agudelo, A. J. P.; Rezende, J. P.; Ferreira, G. M. D.; Ferreira, G. M. D.; Pires, A. C. S.; da Silva, L. H. M.; Colloids Surf., B 2017, 159, 737.
  • 25
    Coelho, Y. L.; de Paula, H. M. C.; Agudelo, A. J. P.; de Castro, A. S. B.; Hudson, E. A.; Pires, A. C. S.; da Silva, L. H. M.; Int. J. Biol. Macromol. 2019, 136, 559.
  • 26
    Atale, S. S.; Dyawanapelly, S.; Jagtap, D. D.; Jain, R.; Dandekar, P.; Int. J. Biol. Macromol. 2019, 123, 97.
  • 27
    Patil, S.; Sistla, S.; Jadhav, J.; Int. J. Biol. Macromol. 2016, 92, 1123.
  • 28
    Rezende, J. P.; Hudson, E. A.; de Paula, H. M. C.; Coelho, Y. L.; da Silva, L. H. M.; Pires, A. C. S.; Food Hydrocolloids 2019, 95, 526.
  • 29
    Rezende, J. P.; Hudson, E. A.; de Paula, H. M. C.; Meinel, R. S.; da Silva, A. D.; Pires, A. C. S.; da Silva, L. H. M.; Food Chem. 2020, 307, 125514.
  • 30
    Drake, A. W.; Tang, M. L.; Papalia, G. A.; Landes, G.; Haak-Frendscho, M.; Klakamp, S. L.; Anal. Biochem. 2012, 429, 58.
  • 31
    Schubert, F.; Zettl, H.; Häfner, W.; Krauss, G.; Krausch, G.; Biochemistry 2003, 42, 10288.
  • 32
    Fathi, F.; Dolatanbadi, J. E. N.; Rashidi, M.; Omidi, Y.; Int. J. Biol. Macromol. 2016, 91, 1045.
  • 33
    Ehteshami, M.; Rasoulzadeh, F.; Mahboob, S.; Rashidi, M.; J. Lumin. 2013, 135, 164.
  • 34
    Paul, B. K.; Ghosh, N.; Mukherjee, S.; RSC Adv. 2016, 6, 9984.
  • 35
    Liu, L.; Guo, Q.; Chem. Rev. 2001, 101, 673.

Publication Dates

  • Publication in this collection
    14 Dec 2020
  • Date of issue
    Dec 2020

History

  • Received
    26 Mar 2020
  • Accepted
    14 July 2020
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br