Next Article in Journal
A Comparison between Enrichment Optimization Algorithm (EOA)-Based and Docking-Based Virtual Screening
Previous Article in Journal
High-Dose Benzodiazepines Positively Modulate GABAA Receptors via a Flumazenil-Insensitive Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Effect of pH, Temperature, and Surfactant Structure on Bovine Serum Albumin–Cationic/Anionic/Nonionic Surfactants Interactions in Cacodylate Buffer–Fluorescence Quenching Studies Supported by UV Spectrophotometry and CD Spectroscopy

by
Krzysztof Żamojć
*,
Dariusz Wyrzykowski
and
Lech Chmurzyński
Faculty of Chemistry, University of Gdansk, Wita Stwosza 63, 80-308 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(1), 41; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23010041
Submission received: 16 November 2021 / Revised: 12 December 2021 / Accepted: 20 December 2021 / Published: 21 December 2021
(This article belongs to the Section Physical Chemistry and Chemical Physics)

Abstract

:
Due to the fact that surfactant molecules are known to alter the structure (and consequently the function) of a protein, protein–surfactant interactions are very important in the biological, pharmaceutical, and cosmetic industries. Although there are numerous studies on the interactions of albumins with surfactants, the investigations are often performed at fixed environmental conditions and limited to separate surface-active agents and consequently do not present an appropriate comparison between their different types and structures. In the present paper, the interactions between selected cationic, anionic, and nonionic surfactants, namely hexadecylpyridinium chloride (CPC), hexadecyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (SDS), polyethylene glycol sorbitan monolaurate, monopalmitate, and monooleate (TWEEN 20, TWEEN 40, and TWEEN 80, respectively) with bovine serum albumin (BSA) were studied qualitatively and quantitatively in an aqueous solution (10 mM cacodylate buffer; pH 5.0 and 7.0) by steady-state fluorescence spectroscopy supported by UV spectrophotometry and CD spectroscopy. Since in the case of all studied systems, the fluorescence intensity of BSA decreased regularly and significantly under the action of the surfactants added, the fluorescence quenching mechanism was analyzed thoroughly with the use of the Stern–Volmer equation (and its modification) and attributed to the formation of BSA–surfactant complexes. The binding efficiency and mode of interactions were evaluated among others by the determination, comparison, and discussion of the values of binding (association) constants of the newly formed complexes and the corresponding thermodynamic parameters (ΔG, ΔH, ΔS). Furthermore, the influence of the structure of the chosen surfactants (charge of hydrophilic head and length of hydrophobic chain) as well as different environmental conditions (pH, temperature) on the binding mode and the strength of the interaction has been investigated and elucidated.

1. Introduction

Due to a huge variety of industrial, biological, medical, and technical applications of surfactants in the fields of chemicals, detergents, cosmetics, foods, and pharmaceuticals [1,2,3,4,5,6,7,8,9,10], their interactions with proteins have been in recent years extensively studied, mainly in terms of factors that determine the association extent as well as modifications of the protein physicochemical and functional properties (in the case of enzymes, transport proteins, and toxins) caused by association-elicited conformational changes [11,12,13,14,15,16,17].
The macromolecule–surfactant interactions depend on many factors, among which the type of surfactant species (monomer, micelles) presented in the system under study [18,19], the chemical structure of a surfactant (anionic, cationic, amphoteric, or nonionic) [20], its architecture [21,22], as well as experimental conditions (pH, temperature) [23,24] play crucial roles.
In this paper, we report the interactions between selected surfactants (the structures are presented in Figure 1) with bovine serum albumin studied under various experimental conditions. Contrary to many previous reports—as most of the studies concern protein denaturation at surfactants concentrations much above their critical micelle concentration [25,26,27,28,29]—we have focused on low concentrations of detergents. Furthermore, we have used simultaneously surfactants belonging to three main groups (cationic, anionic, and nonionic), although most of the comprehensive studies refer to ionic ones [30,31,32,33,34]; the interactions with nonionic surfactants, used normally in formulations due to stabilizing properties [35], are found to be weaker, and the interpretation of results is somehow more difficult [31,36]. Finally, to the best of our knowledge, this is the first report on the molecular interactions between a group of conventional surfactants and BSA elucidated in cacodylate (CACO) buffer. Its relatively wide useful buffering range enables performing the measurements in the same buffer solution and at different pH (5.0 and 7.0).
In our studies, we have used steady-state fluorescence spectroscopy (supported by UV spectrophotometry and CD spectroscopy), which is regarded as an important tool for the studies of various chemical individuals (among others metal complexes [37], inclusion complexes [38], gases [39], thiols [40], radicals [41], amino acids [42], nanoparticles [43]) as well as an effective technique to establish conformational modifications and interactions of different biomolecules in a solution [44].
The presented results have important implications to understand the influence of commonly used surfactants on the functionality of proteins in modern branches of chemistry. The investigations of structural features of the surfactants that govern their affinity toward a protein may be helpful for the development of new drug delivery systems for improved medical treatments [45,46,47]. Consequently, it was the main reason that prompted us to embark on these studies.

2. Results and Discussion

The bovine serum albumin (BSA) in cacodylate buffer (CACO) exhibits an emission peak with a maximum at 348 nm (due to tryptophan residues). On the addition of all studied surfactants, the peak becomes more or less blue shifted (up to 332 nm) with a clear decrease in the quantum yield (Figure S1 in Electronic Supplementary Information). The observed shift and alteration in the fluorescence intensity of the attained band provide essential information about events that affect the microenvironment surrounding the aromatic residues in the protein molecule (such as protein conformational transitions, ligand binding, denaturation, etc.) [48] and may be a result of conformational changes in BSA structure, which lead to the exposure of intrinsic fluorophores to more hydrophobic environment (the change in the environment of tryptophan and an increase in hydrophobicity in the vicinity of this residue due to the presence of the alkyl chains of the surfactant molecules) [49]. On the other hand, at pH 7.4, anionic D and E residues remain near W134, which lies at the surface of BSA (subdomain IB) [50]. Consequently, the possible binding of the studied surfactants by electrostatic interactions under experimental conditions (namely at pH 7.0) cannot be ruled out. Figure S2 in the Electronic Supplementary Information presents the UV absorption spectra and difference UV absorption spectra of BSA (2 µM) in the presence of increasing concentrations of the studied surfactants in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K. From the inspection of these spectra, it can be clearly observed that—in the studied range of concentrations—all examined surfactants exhibit a slight but noticeable influence on the UV spectrum of the free albumin. When the concentration of the surfactant increases, the band with a maximum at approximately 220 nm gradually appears with no or a negligible impact on the band attributed to the tryptophan residue (namely 275 nm). The findings from the analysis of both UV absorption and fluorescence emission spectra may be indicative of the formation of BSA–surfactants complexes.
To get an insight into qualitative and quantitative assessment of the interactions between the BSA and the surfactants, the obtained results were analyzed according to the Stern–Volmer equation:
F 0 F = 1 + K SV Q = 1 + k q τ 0 Q
where F0 and F denote the steady-state fluorescence intensities in the absence and presence of a quencher (surfactant), [Q] is the quencher concentration, KSV is the Stern–Volmer quenching constant, kq is the bimolecular quenching rate constant, while τ0 is the average lifetime of the fluorophore (BSA) in the absence of quencher [51,52,53,54]. The graphs of F 0 F versus [Q] plotted for the steady-state fluorescence quenching of BSA tryptophan residues by various surfactants under experimental conditions according to the Stern–Volmer equation are shown in Figure 2 and Figure 3. Consequently, Table 1 presents the newly determined values of Stern–Volmer quenching constants (KSV) along with linear correlation coefficients (R2) for Stern–Volmer plots and bimolecular quenching rate constants (kq) recovered for the studied systems. The latter ones were calculated based on the values of the average fluorescence lifetimes τ0 of free BSA equal to approximately 6 ns at pH 7.0 (phosphate buffer) and 5.7 ns at pH 5.0 (acetate buffer) [55].
In the studied concentration range, strictly straight lines were obtained with no deviations from the observed linearity. To get a deeper insight into the mechanism of fluorescence quenching, the newly determined bimolecular quenching rate constants were compared with the maximum scattering collision quenching rate constant, which for the interactions of various quenchers with biopolymer in aqueous media is equal to 2.0 × 1010 M−1 s−1 [56,57,58]. The estimated values of kq for all surfactants are of the order 1011–1013 M−1 s−1, which is approximately 23–700-fold higher than the mentioned maximum value possible for diffusion controlled quenching rate constant. Since the values of bimolecular quenching rate constants are considered to be definitive in differentiating between dynamic and static quenching mechanisms [59,60], the predominant role of ground-state complexation (static quenching) in the investigated systems was affirmed. The observed changes in fluorescence emission of BSA induced by low surfactant concentrations are probably due to the changes of W131 residue orientation and contacts with neighboring residues—some residues in close contact with the tryptophan indole group are known to effectively quench the fluorescence by a static quenching mechanism [61,62,63]. Interestingly, the results presented in Figure 2 revealed that the quenching constants and thus the stabilities of the resulting BSA–surfactant complexes increase in proportion to the increase in temperature [64,65,66]. This phenomenon suggests an endothermic nature of the investigated interactions. The above assumption has been subsequently verified by examining the thermodynamic parameter of the reactions under study.
For a static quenching interaction, when small molecules bind independently to a set of equivalent sites of a protein, the equilibrium between free and bound molecules is given by the following modified version of the Stern–Volmer equation [67,68]:
log F 0 F F = log K a + n log Q .
Thus, the apparent binding constant (Ka) and the number of binding sites (n) can be calculated from the intercepts and the slopes of the linear plots of log F 0 F F versus log[Q], respectively (Figure 4 and Figure 5). These values have been obtained (as the averages of three independent experiments) and are shown in Table 2, along with linear correlation coefficients (R2) recovered for the studied systems.
The value of Ka is particularly relevant for the understanding of the distribution of the drug in plasma, since weak binding can lead to a short lifetime or a poor distribution, while strong binding is responsible for the reduction in the plasmatic distribution of free drug [69,70]. The Ka values obtained for the studied BSA complexes suggest relatively high/medium binding affinity to the protein for ionic surfactants (mainly at higher pH) and significantly lower binding affinity for nonionic ones when compared to other known strong biomolecule–ligand complexes with binding constants ranging from 105 to 108 M−1 [71,72,73]. However, lower binding constants (102−104 M−1), which indicate a very weak interaction between the ligand and the protein, have been reported for several other protein–ligand complexes as well [74,75,76]. Since the interactions between studied ionic surfactants and BSA are quite significant (mainly at pH 7.0) and the effect of temperature is rather small, it shows that these compounds can be stored and transported by the protein in the body [77]. For most of the surfactant–serum albumin complexes, the value of n was approximately equal to one, indicating the existence of just a single binding site in BSA for the studied compounds [78,79].
Generally, there have been suggested two main types of interactions between surfactants and proteins, namely specific binding by electrostatic attraction/repulsion between oppositely/equally charged surfactant head groups and sites of the proteins or the cooperative association of alkyl chains to the protein via hydrophobic affinity [80]. It can be observed that the fluorescence quenching of BSA is clearly pH dependent; decreasing pH results in a noticeable reduction in the quenching efficiency (Figure 2, Table 1) and, consequently, binding affinity (Figure 4, Table 2). It proves that the protonation of the protein residues affects the binding of the surfactants (except using SDS, where the change in pH does not affect its affinity toward the first site of the protein). In the case of—for example—cationic surfactants, an increase in pH from 5.0 to 7.0 probably favors interactions between negatively charged groups in the protein and CPC/CTAB. At pH 5.0, the protein is packed in a more compact form (pH 5.0 is close to the isoelectric point; pI = 5.2 [81]), which reduces the accessibility of the surfactants to the hydrophobic cavities in the binding sites. Conversely, increasing the pH makes the protein more flexible and open in its structure [82]—more mobile peptide chains contribute to an increase in the accessibility of the surfactants to the hydrophobic core. On the other hand, higher values of Stern–Volmer quenching constants were obtained for SDS in comparison with the ones for CPC and CTAB (particularly at pH 7.0; above the isoelectric point). It is in a good agreement with the literature, and the possible reason may lie in the surfactant features. Anionic surfactants interact strongly with proteins and cause their denaturation, which is possible due to the unfolding of proteins induced by surfactants. Comparing with anionic surfactants, cationic ones exhibit a lower tendency to interact with proteins mainly as a consequence of a smaller relevance of electrostatic interaction at the pHs of interest [77,83,84]. The observed phenomenon can also suggest that the charges of the surfactant head group modulate the interactions only partially, and the hydrophobic interactions of surfactant methylene chains are responsible for binding as well.
Apart from pH, the impact of the structure of the surfactant on the quenching intensity was also considered. The comparison of bimolecular quenching rate constants obtained for BSA–CTAB and BSA–CPC systems led to an evaluation of the influence of the pyridinium moiety, which was found to be an efficient fluorescence quencher that could reduce the intrinsic protein fluorescence (which is in good agreement with the literature [85]). Furthermore, the results obtained for TWEEN 20, TWEEN 40, and TWEEN 80 revealed no significant changes in the strength of interactions between these surfactants and BSA, which proves that the length of the alkyl chain can be of the slightest importance. On the other hand, at pH 7.0, the following dependence can be observed: Ka TWEEN 20 < Ka TWEEN 40 < Ka TWEEN 80. This trend is expected as the stability of the resulting complexes increases with the lengthening of the hydrophobic alkyl chain of the surfactant [45,50]. Thus, the formation of the BSA complexes with nonionic surfactants seems to be an entropy-driven process.
The acting forces for binding between a small molecule and protein may include hydrophobic and electrostatic interactions, van der Waals interactions, hydrogen bonds, etc. [86,87]. The noncovalent interaction forces between various ligands and biomolecules can be described by thermodynamic parameters, which can be calculated from the van’t Hoff equation:
ln K SV = Δ H RT + Δ S R
where ΔH and ΔS are enthalpy change and entropy change, respectively, R is the gas constant, and T denotes the absolute temperature [88]. The plot of lnKSV (for CPC) measured at three different temperatures vs. 1 T is linear within the experimental error (Figure 6), which allowed one to calculate ΔH and ΔS for the quenching process.
It has been reported that the positive ΔH and ΔS values are associated with hydrophobic interactions playing a major role in the binding reaction; negative ΔH and ΔS values are correlated with van der Walls interactions and hydrogen bonding, while electrostatic forces usually make ΔH ≈ 0 or ΔH < 0 and ΔS > 0 [89]. The negative sign for free energy change of quenching ΔG ( Δ G = Δ H T Δ S = RTln K SV ) for all studied surfactants indicates the spontaneity of their binding with bovine serum albumin in the concentration range employed [49]. The positive ΔH (12.1 kJ mol−1 for pH 5.0 and 9.23 kJ mol−1 for pH 7.0) and ΔS (75.1 J mol−1 K−1 for pH 5.0 and 70.0 J mol−1 K−1 for pH 7.0) indicate that the interaction between BSA and CPC is mainly entropy-driven, the enthalpy is unfavorable for it, and the hydrophobic forces play a major role with less dominating hydrogen bonds formation. The calculated positive values of the enthalpy changes stay in line with the observed increase in the Ksv values with the increase in temperature, confirming a static quenching mechanism of the protein–surfactant complex formation [90,91].
The far-UV CD spectroscopy of BSA and the resulting BSA–surfactant complexes were employed for monitoring the changes in the secondary structure of the investigated albumin. For all systems studied, two negative bands characteristic for the α-helix structure are present in the CD spectra at 208 nm (the pp* transition) and 222 nm (the np* transition) (Figure S3 in Electronic Supplementary Information) [92]. The conformational contributions of α-helix, β-sheet, turns, and unordered structures of the albumin secondary structure under the experimental conditions are collected in Table 3. The obtained data have shown that the saturation of BSA with the nonionic surfactants does not contribute to noticeable changes in the secondary structure of the investigated protein. On the other hand, the decrease in the α-helix content is seen for the ionic surfactants. This phenomenon is enhanced for the cationic surfactants with the increase in the pH of a solution. Thus, it can be expected that the presence of the ionic surfactants in the system can affect the biological activity of BSA.

3. Materials and Methods

Bovine serum albumin (BSA, lyophilized powder, ≥96%), hexadecyltrimethylammonium bromide (CTAB, BioXtra, ≥99%), sodium dodecyl sulfate (SDS, BioUltra, for molecular biology, ≥99%), and sodium cacodylate trihydrate (CACO, ≥98%) were obtained from Merck (Warsaw, Poland) and employed as received without further purification. Hexadecylpyridinium chloride monohydrate (CPC, ≥96%), polyethylene glycol sorbitan monolaurate (TWEEN 20), polyethylene glycol sorbitan monopalmitate (TWEEN 40), and polyethylene glycol sorbitan monooleate (TWEEN 80) were kindly provided by Cerko (Gdynia, Poland). Double-distilled water with conductivity not exceeding 0.18 μS cm−1 was used for preparations of buffer solutions. The stock solutions of the protein and all surfactants were prepared in 10 mM CACO buffer of pH 5.0 and 7.0 (all subsequent dilutions were made with this buffer). The concentration of the bovine serum albumin was measured using an extinction coefficient equal to ε 280 BSA = 41,180 M−1 cm−1 calculated based on the content of tryptophan ( ε 280 W = 5690 M−1 cm−1), tyrosine ( ε 280 Y = 1280 M−1 cm−1), and cysteine ( ε 280 C = 120 M−1 cm−1) [93,94]. A maximum absorbance value of approximately 0.08 at 280 nm (corresponding to a protein concentration of 2 µM) was used to avoid the inner filter effect.
Steady-state fluorescence emission spectra were registered, and further emission measurements were carried out on a Cary Eclipse Varian (Agilent, Santa Clara, CA, USA) spectrofluorometer equipped with a temperature controller and a 1.0 cm multicell holder. The absorption spectra were recorded on a Lambda 650 (Perkin Elmer, Waltham, MA, USA) UV/Vis spectrophotometer. In the performed fluorescence and spectophotometric titration experiments, 2 mL of BSA at a fixed concentration equal to 2 µM was titrated with fifteen 2 μL aliquots of each surfactant’s solution (cSDS = 1 mM; cCPC = 4 mM; cCTAB = 4 mM; cTWEEN 20 = 10 mM; cTWEEN 40 = 10 mM; cTWEEN 80 = 10 mM). The concentration of surfactants was varied and dependent on the observed quenching efficiency. The fluorescence intensity of the band at 348 nm—corresponding to the initial maximum emission of BSA—was used to calculate the binding constants and other parameters. The excitation wavelength was always set at 280 nm. All experiments were performed at 298 K (in the case of experiments with CPC additionally at 288 K and 308 K). All fluorescence intensity values have been corrected for the inner filter effect based on the following equation:
F corr = F obs · 10 A 280 + A 348 2
where Fcorr and Fobs correspond to corrected and observed fluorescence intensity values, respectively; while A280 and A348 correspond to absorbance values measured at the excitation and emission wavelength, respectively.
Circular dichroism (CD) spectra were recorded in 10 mM CACO buffer (pH 5.0 and 7.0) on a J-715 (JASCO Inc., Easton, MD, USA) automatic recording spectropolarimeter for pure BSA (2 µM) and all BSA–surfactant mixtures (in stoichiometric ratios equal to 1:7.5 for SDS, 1:30 for CPC and CTAB, 1:75 for TWEEN 20, TWEEN 40, and TWEEN 80) at 298 K. The spectra were recorded in the 190–260 nm wavelength range in 1 mm quartz cuvettes (the volume of sample was 0.3 mL), using a sensitivity of five millidegrees and a scan speed of 50 nm min−1. The effect of surfactant binding to the protein on the BSA secondary structure content was estimated from all CD spectra using the Dichroweb online server [95] with the CDSSTR analysis algorithm [96] and reference dataset 7 [97].

4. Conclusions

In the current work, fluorescence quenching experiments supported by UV spectrophotometry have been applied to unravel and compare the nature of binding of selected cationic, anionic, and nonionic surfactants with different head groups and chains to the circulatory protein BSA (in the cacodylate buffer; at various temperatures and pHs). Furthermore, the changes in the secondary structure of BSA on account of surfactants binding were assessed based on far-UV CD spectra. The quenching mechanism of the albumin induced by SDS, CPC, CTAB, TWEEN 20, TWEEN 40, and TWEEN 80 has been attributed to a static quenching process. The Stern–Volmer quenching constants, bimolecular quenching rate constants, binding constants, and corresponding thermodynamic parameters ΔH, ΔG, and ΔS were calculated, compared, and discussed. It was found that all studied surfactants interact more or less with BSA (nonionic ones exhibit lower affinity for the protein) and hydrophobic forces play a major role during the binding interaction. This phenomenon stays in line with the CD results. The interaction of cationic surfactants with BSA is a pH-dependent. The strength of the interactions increases with the increase in the pH of a solution. This points to the fact that charge–charge electrostatic interactions established between more negatively charged functional groups of amino acids at pH 7.0 (than at pH 5.0) and positively charged CPC and CTAB surfactants play a pivotal role in the stabilization of the resulting complexes. It has been proven that the presence in the system of the low molecular ligands such as SDS, CPC, CTAB, TWEEN 20, TWEEN 40, and TWEEN 80 can affect the BSA structure and thus its biological activity. On the one hand, saturation of the binding sites of BSA with surfactants hinders the binding of other ligands that reveal the affinity to the same binding site of the protein. This finding is worth taking into account when planning the use of surfactants as solubility modifiers of biologically active compounds for analyzing their interactions with the albumins. On the other hand, surfactants can selectively block the binding sites of BSA (the important component of many cell media cultures) and thus prevent binding the tested compounds with the albumin during the cytotoxicity assays. As a consequence, this can affect the concentration of free, active species.

Supplementary Materials

Author Contributions

Conceptualization, K.Ż. and D.W.; methodology, K.Ż.; formal analysis, K.Ż. and D.W.; investigation, K.Ż.; writing—original draft preparation, K.Ż.; writing—review and editing, K.Ż. and D.W.; supervision, L.C.; project administration, K.Ż.; funding acquisition, K.Ż. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Polish National Science Centre, grant number 2016/23/D/ST4/01576.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

BSABovine serum albumin
CPCHexadecylpyridinium chloride
CTABHexadecyltrimethylammonium bromide
SDSSodium dodecyl sulfate
TWEEN 20Polyethylene glycol sorbitan monolaurate
TWEEN 40Polyethylene glycol sorbitan monopalmitate
TWEEN 80Polyethylene glycol sorbitan monooleate
CACOCacodylate buffer
CDCircular dichroism

References

  1. Menger, F.M.; Rhee, J.U.; Rhee, H.K. Applications of surfactants to synthetic organic chemistry. J Org. Chem. 1975, 40, 3803–3805. [Google Scholar] [CrossRef]
  2. Abe, M. Synthesis and applications of surfactants containing fluorine. Curr. Opin. Colloid Interface Sci. 1999, 4, 354–356. [Google Scholar] [CrossRef]
  3. Luk, Y.Y.; Abbott, N.L. Applications of functional surfactants. Curr. Opin. Colloid Interface Sci. 2002, 7, 267–275. [Google Scholar] [CrossRef]
  4. Banat, I.M.; Makkar, R.S.; Cameotra, S.S. Potential commercial applications of microbial surfactants. Appl. Microbiol. Biotechnol. 2000, 53, 495–508. [Google Scholar] [CrossRef]
  5. Singh, A.; Van Hamme, J.D.; Ward, O.P. Surfactants in microbiology and biotechnology: Part 2. Application aspects. Biotechnol. Adv. 2007, 25, 99–121. [Google Scholar] [CrossRef]
  6. Singh, P.; Cameotra, S.S. Potential applications of microbial surfactants in biomedical sciences. Trends Biotechnol. 2004, 22, 142–146. [Google Scholar] [CrossRef]
  7. Kumar, N.; Tyagi, R. Industrial applications of dimeric surfactants: A review. J. Dispers. Sci. Technol. 2014, 35, 205–214. [Google Scholar] [CrossRef]
  8. Nitschke, M.; Silva, S.S.E. Recent food applications of microbial surfactants. Crit. Rev. Food Sci. Nutr. 2018, 58, 631–638. [Google Scholar] [CrossRef]
  9. Jahan, R.; Bodratti, A.M.; Tsianou, M.; Alexandridis, P. Biosurfactants, natural alternatives to synthetic surfactants: Physicochemical properties and applications. Adv. Colloid Interface Sci. 2020, 275, 102061. [Google Scholar] [CrossRef] [PubMed]
  10. Rocha e Silva, N.M.P.; Meira, H.M.; Almeida, F.C.G.; Soares da Silva, R.D.C.F.; Almeida, D.G.; Luna, J.M.; Rufino, R.D.; Santos, V.A.; Sarubbo, L.A. Natural surfactants and their applications for heavy oil removal in industry. Sep. Purif. Rev. 2019, 48, 267–281. [Google Scholar] [CrossRef]
  11. Vasilescu, M.; Angelescu, D.; Almgren, M.; Valstar, A. Interactions of globular proteins with surfactants studied with fluorescence probe methods. Langmuir 1999, 15, 2635–2643. [Google Scholar] [CrossRef]
  12. Goddard, E.D.; Ananthapadmanabhan, K.P. Interactions of Surfactants with Polymers and Proteins; CRC Press: Boca Raton, FL, USA, 1993. [Google Scholar]
  13. Jones, M.N. Surfactant interactions with biomembranes and proteins. Chem. Soc. Rev. 1992, 21, 127–136. [Google Scholar] [CrossRef]
  14. Otzen, D.E. Biosurfactants and surfactants interacting with membranes and proteins: Same but different? Biochim. Biophys. Acta 2017, 1859, 639–649. [Google Scholar] [CrossRef] [PubMed]
  15. Grabowska, O.; Kogut, M.; Żamojć, K.; Samsonov, S.; Makowska, J.; Tesmar, A.; Chmur, K.; Wyrzykowski, D.; Chmurzyński, L. Effect of tetraphenylborate on physicochemical properties of bovine serum albumin. Molecules 2021, 26, 6565. [Google Scholar] [CrossRef]
  16. Nielsen, A.D.; Borch, K.; Westh, P. Thermochemistry of the specific binding of C12 surfactants to bovine serum albumin. Biochim. Biophys. Acta 2000, 1479, 321–331. [Google Scholar] [CrossRef]
  17. Sharma, V.; Yañez, O.; Zúñiga, C.; Kumar, A.; Singh, G.; Cantero-López, P. Protein-surfactant interactions: A multitechnique approach on the effect of Co-solvents over bovine serum albumin (BSA)-cetyl pyridinium chloride (CPC) system. Chem. Phys. Lett. 2020, 747, 137349. [Google Scholar] [CrossRef]
  18. Han, Y.; Wang, Y. Aggregation behavior of gemini surfactants and their interaction with macromolecules in aqueous solution. Phys. Chem. Chem. Phys. 2011, 13, 1939–1956. [Google Scholar] [CrossRef]
  19. Morris, S.A.; Thompson, R.T.; Glenn, R.W.; Ananthapadmanabhan, K.P.; Kasting, G.B. Mechanisms of anionic surfactant penetration into human skin: Investigating monomer, micelle and submicellar aggregate penetration theories. Int. J. Cosmet. Sci. 2019, 41, 55–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Qu, P.; Lu, H.; Yan, S.; Lu, Z. Influences of cationic, anionic, and nonionic surfactants on alkaline-induced intermediate of bovine serum albumin. Int. J. Biol. Macromol. 2010, 46, 91–99. [Google Scholar] [CrossRef] [PubMed]
  21. Zhou, T.; Ao, M.; Xu, G.; Liu, T.; Zhang, J. Interactions of bovine serum albumin with cationic imidazolium and quaternary ammonium gemini surfactants: Effects of surfactant architecture. J. Colloid Interface Sci. 2013, 389, 175–181. [Google Scholar] [CrossRef]
  22. Yakimova, L.; Padnya, P.; Tereshina, D.; Kunafina, A.; Nugmanova, A.; Osin, Y.; Evtugyn, V.; Stoikov, I. Interpolyelectrolyte mixed nanoparticles from anionic and cationic thiacalix [4] arenes for selective recognition of model biopolymers. J. Mol. Liq. 2019, 279, 9–17. [Google Scholar] [CrossRef]
  23. Faustino, C.M.; Calado, A.R.; Garcia-Rio, L. Gemini surfactant−protein interactions: Effect of pH, temperature, and surfactant stereochemistry. Biomacromolecules 2009, 10, 2508–2514. [Google Scholar] [CrossRef] [PubMed]
  24. Srivastava, R.; Alam, M.S. Effect of pH and surfactant on the protein: A perspective from theory and experiments. Int. J. Biol. Macromol. 2018, 107, 1519–1527. [Google Scholar] [CrossRef] [PubMed]
  25. Otzen, D. Protein–surfactant interactions: A tale of many states. Biochim. Biophys. Acta 2011, 1814, 562–591. [Google Scholar] [CrossRef] [PubMed]
  26. Srivastava, R.; Alam, M.S. Influence of micelles on protein’s denaturation. Int. J. Biol. Macromol. 2020, 145, 252–261. [Google Scholar] [CrossRef] [PubMed]
  27. Srivastava, R.; Alam, M.S. Role of (single/double chain surfactant) micelles on the protein aggregation. Int. J. Biol. Macromol. 2019, 122, 72–81. [Google Scholar] [CrossRef] [PubMed]
  28. Srivastava, R.; Alam, M.S. Spectroscopic studies of the aggregation behavior of Human Serum Albumin and cetyltrimethylammonium bromide. Int. J. Biol. Macromol. 2020, 158, 394–400. [Google Scholar] [CrossRef]
  29. Arsiccio, A.; McCarty, J.; Pisano, R.; Shea, J.E. Effect of surfactants on surface-induced denaturation of proteins: Evidence of an orientation-dependent mechanism. J. Phys. Chem. B 2018, 122, 11390–11399. [Google Scholar] [CrossRef]
  30. Kelley, D.J.M.D.; McClements, D.J. Interactions of bovine serum albumin with ionic surfactants in aqueous solutions. Food Hydrocoll. 2003, 17, 73–85. [Google Scholar] [CrossRef]
  31. Gelamo, E.L.; Silva, C.H.T.P.; Imasato, H.; Tabak, M. Interaction of bovine (BSA) and human (HSA) serum albumins with ionic surfactants: Spectroscopy and modelling. Biochim. Biophys. Acta 2002, 1594, 84–99. [Google Scholar] [CrossRef]
  32. Gelamo, E.L.; Tabak, M. Spectroscopic studies on the interaction of bovine (BSA) and human (HSA) serum albumins with ionic surfactants. Spectrochim. Acta A 2000, 56, 2255–2271. [Google Scholar] [CrossRef]
  33. Bordbar, A.K.; Saboury, A.A.; Housaindokht, M.R.; Moosavi-Movahedi, A.A. Statistical effects of the binding of ionic surfactant to protein. J. Colloid Interface Sci. 1997, 192, 415–419. [Google Scholar] [CrossRef]
  34. Erfani, A.; Khosharay, S.; Flynn, N.H.; Ramsey, J.D.; Aichele, C.P. Effect of zwitterionic betaine surfactant on interfacial behavior of bovine serum albumin (BSA). J. Mol. Liq. 2020, 318, 114067. [Google Scholar] [CrossRef]
  35. Ruiz-Peña, M.; Oropesa-Nuñez, R.; Pons, T.; Louro, S.R.W.; Pérez-Gramatges, A. Physico-chemical studies of molecular interactions between non-ionic surfactants and bovine serum albumin. Colloids Surf. B Biointerfaces 2010, 75, 282–289. [Google Scholar] [CrossRef]
  36. Chakraborty, T.; Chakraborty, I.; Moulik, S.P.; Ghosh, S. Physicochemical and conformational studies on BSA− surfactant interaction in aqueous medium. Langmuir 2009, 25, 3062–3074. [Google Scholar] [CrossRef] [PubMed]
  37. Soroka, K.; Vithanage, R.S.; Phillips, D.A.; Walker, B.; Dasgupta, P.K. Fluorescence properties of metal complexes of 8-hydroxyquinoline-5-sulfonic acid and chromatographic applications. Anal. Chem. 1987, 59, 629–636. [Google Scholar] [CrossRef]
  38. Yorozu, T.; Hoshino, M.; Imamura, M. Fluorescence studies of pyrene inclusion complexes with. alpha.-, beta.-, and. gamma.-cyclodextrins in aqueous solutions. Evidence for formation of pyrene dimer in. gamma.-cyclodextrin cavity. J. Phys. Chem. 1982, 86, 4426–4429. [Google Scholar] [CrossRef]
  39. Żamojć, K.; Jacewicz, D.; Zdrowowicz, M.; Chmurzyński, L. Kinetics of the reaction between 1,3-diphenylisobenzofuran and nitrogen dioxide studied by steady-state fluorescence. Res. Chem. Intermed. 2013, 39, 3023–3031. [Google Scholar] [CrossRef]
  40. Yi, L.; Li, H.; Sun, L.; Liu, L.; Zhang, C.; Xi, Z. A highly sensitive fluorescence probe for fast thiol-quantification assay of glutathione reductase. Angew. Chem. Int. Ed. 2009, 48, 4034–4037. [Google Scholar] [CrossRef]
  41. Żamojć, K.; Zdrowowicz, M.; Wiczk, W.; Jacewicz, D.; Chmurzyński, L. Dihydroxycoumarins as highly selective fluorescent probes for the fast detection of 4-hydroxy-TEMPO in aqueous solution. RSC Adv. 2015, 5, 63807–63812. [Google Scholar] [CrossRef]
  42. Cohen, B.E.; McAnaney, T.B.; Park, E.S.; Jan, Y.N.; Boxer, S.G.; Jan, L.Y. Probing protein electrostatics with a synthetic fluorescent amino acid. Science 2002, 296, 1700–1703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chen, M.M.Y.; Katz, A. Steady-state fluorescence-based investigation of the interaction between protected thiols and gold nanoparticles. Langmuir 2002, 18, 2413–2420. [Google Scholar] [CrossRef]
  44. De, S.; Girigoswami, A.; Das, S. Fluorescence probing of albumin–surfactant interaction. J. Colloid Interface Sci. 2005, 285, 562–573. [Google Scholar] [CrossRef] [PubMed]
  45. Jiao, J. Polyoxyethylated nonionic surfactants and their applications in topical ocular drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1663–1673. [Google Scholar] [CrossRef] [PubMed]
  46. Seweryn, A. Interactions between surfactants and the skin–Theory and practice. Adv. Colloid Interface Sci. 2018, 256, 242–255. [Google Scholar] [CrossRef] [PubMed]
  47. Dai, L.; Wei, Y.; Sun, C.; Mao, L.; McClements, D.J.; Gao, Y. Development of protein-polysaccharide-surfactant ternary complex particles as delivery vehicles for curcumin. Food Hydrocoll. 2018, 85, 75–85. [Google Scholar] [CrossRef]
  48. Akram, M.; Ansari, F.; Bhat, I.A. Probing interaction of bovine serum albumin (BSA) with the biodegradable version of cationic gemini surfactants. J. Mol. Liq. 2019, 276, 519–528. [Google Scholar] [CrossRef]
  49. Sharma, V.; Yañez, O.; Alegría-Arcos, M.; Kumar, A.; Thakur, R.C.; Cantero-López, P. A physicochemical and conformational study of co-solvent effect on the molecular interactions between similarly charged protein surfactant (BSA-SDBS) system. J. Chem. Thermodyn. 2020, 142, 106022. [Google Scholar] [CrossRef]
  50. Aslam, J.; Lone, I.H.; Ansari, F.; Aslam, A.; Aslam, R.; Akram, M. Molecular binding interaction of pyridinium based gemini surfactants with bovine serum albumin: Insights from physicochemical, multispectroscopic, and computational analysis. Spectrochim. Acta A 2021, 250, 119350. [Google Scholar] [CrossRef] [PubMed]
  51. Mir, M.U.H.; Maurya, J.K.; Ali, S.; Ubaid-Ullah, S.; Khan, A.B.; Patel, R. Molecular interaction of cationic gemini surfactant with bovine serum albumin: A spectroscopic and molecular docking study. Process Biochem. 2014, 49, 623–630. [Google Scholar] [CrossRef]
  52. Makowska, J.; Żamojć, K.; Wyrzykowski, D.; Żmudzińska, W.; Uber, D.; Wierzbicka, M.; Wiczk, W.; Chmurzyński, L. Probing the binding of Cu2+ ions to a fragment of the Aβ(1–42) polypeptide using fluorescence spectroscopy, isothermal titration calorimetry and molecular dynamics simulations. Biophys. Chem. 2016, 216, 44–50. [Google Scholar] [CrossRef] [PubMed]
  53. Tian, J.; Zhao, Y.; Liu, X.; Zhao, S. A steady-state and time-resolved fluorescence, circular dichroism study on the binding of myricetin to bovine serum albumin. Luminescence 2009, 24, 386–393. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, Y.Z.; Zhou, B.; Liu, Y.X.; Zhou, C.X.; Ding, X.L.; Liu, Y. Fluorescence study on the interaction of bovine serum albumin with p-aminoazobenzene. J. Fluoresc. 2008, 18, 109–118. [Google Scholar] [CrossRef] [PubMed]
  55. Gorodnichev, E.S.; Kuleshova, A.A.; Volkova, O.I.; Saletsky, A.M. The binding of bovine serum albumin with dye molecules at different pH values. Fluorescence lifetime studies. Laser Phys. 2021, 31, 065601. [Google Scholar] [CrossRef]
  56. Cui, F.L.; Fan, J.; Ma, D.L.; Liu, M.C.; Chen, X.G.; Hu, Z.D. A study of the interaction between a new reagent and serum albumin by fluorescence spectroscopy. Anal. Lett. 2003, 36, 2151–2166. [Google Scholar] [CrossRef]
  57. Makowska, J.; Żamojć, K.; Wyrzykowski, D.; Wiczk, W.; Chmurzyński, L. Copper(II) complexation by fragment of central part of FBP28 protein from Mus musculus. Biophys. Chem. 2018, 241, 55–60. [Google Scholar] [CrossRef]
  58. Liu, Z.; Guo, X.; Feng, Z.; Jia, L. Spectroscopic investigation of the interaction of pyridinium surfactant with bovine serum albumin. J. Solut. Chem. 2015, 44, 293–306. [Google Scholar] [CrossRef]
  59. Żamojć, K.; Wiczk, W.; Chmurzyński, L. The influence of the type of substituents and the solvent on the interactions between different coumarins and selected TEMPO analogues–Fluorescence quenching studies. Chem. Phys. 2018, 513, 188–194. [Google Scholar] [CrossRef]
  60. Gauthier, T.D.; Shane, E.C.; Guerin, W.F.; Seitz, W.R.; Grant, C.L. Fluorescence quenching method for determining equilibrium constants for polycyclic aromatic hydrocarbons binding to dissolved humic materials. Environ. Sci. Technol. 1986, 20, 1162–1166. [Google Scholar] [CrossRef]
  61. Lakowicz, J.R.; Weber, G. Quenching of protein fluorescence by oxygen. Detection of structural fluctuations in proteins on the nanosecond time scale. Biochemistry 1973, 12, 4171–4179. [Google Scholar] [CrossRef]
  62. Eftink, M.R.; Ghiron, C.A. Exposure of tryptophanyl residues in proteins. Quantitative determination by fluorescence quenching studies. Biochemistry 1976, 15, 672–680. [Google Scholar] [CrossRef]
  63. Eftink, M.R.; Ghiron, C.A. Fluorescence quenching of indole and model micelle systems. J. Phys. Chem. 1976, 80, 486–493. [Google Scholar] [CrossRef]
  64. Arık, M.; Çelebi, N.; Onganer, Y. Fluorescence quenching of fluorescein with molecular oxygen in solution. J. Photochem. Photobiol. A 2005, 170, 105–111. [Google Scholar] [CrossRef]
  65. Żamojć, K.; Bylińska, I.; Wiczk, W.; Chmurzyński, L. Fluorescence quenching studies on the interactions between chosen fluoroquinolones and selected stable TEMPO and PROXYL nitroxides. Int. J. Mol. Sci. 2021, 22, 885. [Google Scholar] [CrossRef] [PubMed]
  66. Hu, Y.J.; Liu, Y.; Zhang, L.X.; Zhao, R.M.; Qu, S.S. Studies of interaction between colchicine and bovine serum albumin by fluorescence quenching method. J. Mol. Struct. 2005, 750, 174–178. [Google Scholar] [CrossRef]
  67. Geng, F.; Zheng, L.; Yu, L.; Li, G.; Tung, C. Interaction of bovine serum albumin and long-chain imidazolium ionic liquid measured by fluorescence spectra and surface tension. Process Biochem. 2010, 45, 306–311. [Google Scholar] [CrossRef]
  68. Anand, U.; Jash, C.; Mukherjee, S. Spectroscopic probing of the microenvironment in a protein−surfactant assembly. J. Phys. Chem. B 2010, 114, 15839–15845. [Google Scholar] [CrossRef] [PubMed]
  69. Kandagal, P.B.; Shaikh, S.M.T.; Manjunatha, D.H.; Seetharamappa, J.; Nagaralli, B.S. Spectroscopic studies on the binding of bioactive phenothiazine compounds to human serum albumin. J. Photochem. Photobiol. A 2007, 189, 121–127. [Google Scholar] [CrossRef]
  70. Khan, A.B.; Khan, J.M.; Ali, M.S.; Khan, R.H. Interaction of amphiphilic drugs with human and bovine serum albumins. Spectrochim. Acta A 2012, 97, 119–124. [Google Scholar] [CrossRef]
  71. Butowska, K.; Żamojć, K.; Kogut, M.; Kozak, W.; Wyrzykowski, D.; Wiczk, W.; Czub, J.; Piosik, J.; Rak, J. The product of matrix metalloproteinase cleavage of doxorubicin conjugate for anticancer drug delivery: Calorimetric, spectroscopic, and molecular dynamics studies on peptide–doxorubicin binding to DNA. Int. J. Mol. Sci. 2020, 21, 6923. [Google Scholar] [CrossRef] [PubMed]
  72. Mandeville, J.S.; Froehlich, E.; Tajmir-Riahi, H.A. Study of curcumin and genistein interactions with human serum albumin. J. Pharm. Biomed. Anal. 2009, 49, 468–474. [Google Scholar] [CrossRef] [PubMed]
  73. Kumari, M.; Maurya, J.K.; Singh, U.K.; Khan, A.B.; Ali, M.; Singh, P.; Patel, R. Spectroscopic and docking studies on the interaction between pyrrolidinium based ionic liquid and bovine serum albumin. Spectrochim. Acta A 2014, 124, 349–356. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, Y.; Jiang, X.; Zhou, L.; Yang, L.; Xia, G.; Chen, Z.; Duan, M. Synthesis and binding with BSA of a new gemini surfactant. Colloids Surf. A Physicochem. Eng. Asp. 2013, 436, 1159–1169. [Google Scholar] [CrossRef]
  75. Khan, A.B.; Khan, J.M.; Ali, M.S.; Khan, R.H.; Din, K.U. Spectroscopic approach of the interaction study of amphiphilic drugs with the serum albumins. Colloids Surf. B Biointerfaces 2011, 87, 447–453. [Google Scholar] [CrossRef]
  76. Mehta, S.K.; Bhasin, K.K.; Kumar, A. An insight into the micellization of dodecyldimethylethylammonium bromide (DDAB) in the presence of bovine serum albumin (BSA). J. Colloid Interface Sci. 2008, 323, 426–434. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Y.; Wang, X.; Wang, Y. Comparative studies on interactions of bovine serum albumin with cationic gemini and single-chain surfactants. J. Phys. Chem. B 2006, 110, 8499–8505. [Google Scholar] [CrossRef] [PubMed]
  78. Lu, J.; Sun, Q.; Li, J.L.; Jiang, L.; Gu, W.; Liu, X.; Tian, J.L.; Yan, S.P. Two water-soluble copper(II) complexes: Synthesis, characterization, DNA cleavage, protein binding activities and in vitro anticancer activity studies. J. Inorg. Biochem. 2014, 137, 46–56. [Google Scholar] [PubMed]
  79. Wang, Y.Q.; Zhang, H.M.; Zhang, G.C.; Tao, W.H.; Fei, Z.H.; Liu, Z.T. Spectroscopic studies on the interaction between silicotungstic acid and bovine serum albumin. J. Pharm. Biomed. 2007, 43, 1869–1875. [Google Scholar] [CrossRef] [PubMed]
  80. Green, R.J.; Su, T.J.; Joy, H.; Lu, J.R. Interaction of lysozyme and sodium dodecyl sulfate at the air−liquid interface. Langmuir 2000, 16, 5797–5805. [Google Scholar] [CrossRef]
  81. Hazra, P.; Chakrabarty, D.; Chakraborty, A.; Sarkar, N. Probing protein-surfactant interaction by steady state and time-resolved fluorescence spectroscopy. Biochem. Biophys. Res. Commun. 2004, 314, 543–549. [Google Scholar] [CrossRef]
  82. Höök, F.; Rodahl, M.; Kasemo, B.; Brzezinski, P. Structural changes in hemoglobin during adsorption to solid surfaces: Effects of pH, ionic strength, and ligand binding. Proc. Natl. Acad. Sci. USA 1998, 95, 12271–12276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhao, X.N.; Liu, Y.; Niu, L.Y.; Zhao, C.P. Spectroscopic studies on the interaction of bovine serum albumin with surfactants and apigenin. Spectrochim. Acta A 2012, 94, 357–364. [Google Scholar] [CrossRef]
  84. Wang, Y.; Guo, R.; Xi, J. Comparative studies of interactions of hemoglobin with single-chain and with gemini surfactants. J. Colloid Interface Sci. 2009, 331, 470–475. [Google Scholar] [CrossRef]
  85. Dıaz, X.; Abuin, E.; Lissi, E. Quenching of BSA intrinsic fluorescence by alkylpyridinium cations: Its relationship to surfactant-protein association. J. Photochem. Photobiol. A 2003, 155, 157–162. [Google Scholar] [CrossRef]
  86. Tang, J.; Luan, F.; Chen, X. Binding analysis of glycyrrhetinic acid to human serum albumin: Fluorescence spectroscopy, FTIR, and molecular modeling. Bioorg. Med. Chem. 2006, 14, 3210–3217. [Google Scholar] [CrossRef]
  87. Han, X.L.; Mei, P.; Liu, Y.; Xiao, Q.; Jiang, F.L.; Li, R. Binding interaction of quinclorac with bovine serum albumin: A biophysical study. Spectrochim. Acta A 2009, 74, 781–787. [Google Scholar] [CrossRef]
  88. Mote, U.S.; Bhattar, S.L.; Patil, S.R.; Kolekar, G.B. Interaction between felodipine and bovine serum albumin: Fluorescence quenching study. Luminescence 2010, 25, 1–8. [Google Scholar] [CrossRef] [PubMed]
  89. Zhao, X.; Liu, R.; Teng, Y.; Liu, X. The interaction between Ag+ and bovine serum albumin: A spectroscopic investigation. Sci. Total Environ. 2011, 409, 892–897. [Google Scholar] [CrossRef] [PubMed]
  90. Callis, P.R. Binding phenomena and fluorescence quenching. I: Descriptive quantum principles of fluorescence quenching using a supermolecule approach. J. Mol. Struct. 2014, 1077, 14–21. [Google Scholar] [CrossRef]
  91. Aprodu, I.; Dumitras, L.; Râpeanu, G.; Bahrim, G.E.; Stănciuc, N. Spectroscopic and molecular modeling investigation on the interaction between folic acid and bovine lactoferrin from encapsulation perspectives. Foods 2020, 9, 744. [Google Scholar] [CrossRef] [PubMed]
  92. Sreerama, N. Circular dichroism of peptides and proteins. In Circular Dichroism: Principles and Applications, 2nd ed.; Berova, N., Nakanishi, K., Woody, R.W., Eds.; Wiley: New York, NY, USA, 2000; pp. 601–620. [Google Scholar]
  93. Tesmar, A.; Kogut, M.M.; Żamojć, K.; Grabowska, O.; Chmur, K.; Samsonov, S.A.; Makowska, J.; Wyrzykowski, D.; Chmurzyński, L. Physicochemical nature of sodium dodecyl sulfate interactions with bovine serum albumin revealed by interdisciplinary approaches. J. Mol. Liq. 2021, 340, 117185. [Google Scholar] [CrossRef]
  94. Pelton, J.T.; McLean, L.R. Spectroscopic methods for analysis of protein secondary structure. Anal. Biochem. 2000, 277, 167–176. [Google Scholar] [CrossRef] [PubMed]
  95. Whitmore, L.; Wallace, B.A. Protein secondary structure analyses from circular dichroism spectroscopy: Methods and reference databases. Biopolymers 2008, 89, 392–400. [Google Scholar] [CrossRef] [PubMed]
  96. Compton, L.A.; Johnson Jr, W.C. Analysis of protein circular dichroism spectra for secondary structure using a simple matrix multiplication. Anal. Biochem. 1986, 155, 155–167. [Google Scholar] [CrossRef]
  97. Sreerama, N.; Woody, R.W. Estimation of protein secondary structure from circular dichroism spectra: Comparison of CONTIN, SELCON, and CDSSTR methods with an expanded reference set. Anal. Biochem. 2000, 287, 252–260. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Molecular structures of the studied surfactants.
Figure 1. Molecular structures of the studied surfactants.
Ijms 23 00041 g001
Figure 2. Stern–Volmer plots for the steady-state fluorescence quenching of BSA by different surfactants in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at 298 K.
Figure 2. Stern–Volmer plots for the steady-state fluorescence quenching of BSA by different surfactants in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at 298 K.
Ijms 23 00041 g002
Figure 3. Stern–Volmer plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at various temperatures (288 K, 298 K, and 308 K).
Figure 3. Stern–Volmer plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at various temperatures (288 K, 298 K, and 308 K).
Ijms 23 00041 g003
Figure 4. Modified Stern–Volmer plots for the steady-state fluorescence quenching of BSA by different surfactants in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at 298 K.
Figure 4. Modified Stern–Volmer plots for the steady-state fluorescence quenching of BSA by different surfactants in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at 298 K.
Ijms 23 00041 g004
Figure 5. Modified Stern–Volmer plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at various temperatures (288 K, 298 K, and 308 K).
Figure 5. Modified Stern–Volmer plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 (A) and 7.0 (B) at various temperatures (288 K, 298 K, and 308 K).
Ijms 23 00041 g005
Figure 6. Van’t Hoff plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 and 7.0.
Figure 6. Van’t Hoff plots for the steady-state fluorescence quenching of BSA by CPC in 10 mM CACO buffer of pH 5.0 and 7.0.
Ijms 23 00041 g006
Table 1. Stern–Volmer quenching constants (KSV), linear correlation coefficients (R2), and bimolecular quenching rate constants (kq) recovered for the steady-state fluorescence quenching of BSA by various surfactants in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K (for CPC additionally at 288 and 308 K).
Table 1. Stern–Volmer quenching constants (KSV), linear correlation coefficients (R2), and bimolecular quenching rate constants (kq) recovered for the steady-state fluorescence quenching of BSA by various surfactants in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K (for CPC additionally at 288 and 308 K).
SurfactantpHKSV [M−1]R2kq [M−1 s−1]
CPC5.0288 K0.94 × 1041.0001.64 × 1012
298 K1.28 × 1040.9982.25 × 1012
308 K1.98 × 1040.9983.47 × 1012
7.0288 K3.45 × 1040.9965.75 × 1012
298 K5.23 × 1040.9928.72 × 1012
308 K6.14 × 1041.00010.2 × 1012
CTAB5.00.81 × 1040.9971.41 × 1012
7.02.95 × 1040.9964.92 × 1012
SDS5.07.97 × 1040.98914.0 × 1012
7.07.65 × 1040.99212.8 × 1012
TWEEN 205.00.31 × 1040.9780.54 × 1012
7.00.53 × 1040.9870.89 × 1012
TWEEN 405.00.27 × 1040.9380.47 × 1012
7.00.61 × 1040.9881.01 × 1012
TWEEN 805.00.34 × 1040.9870.60 × 1012
7.00.73 × 1040.9971.21 × 1012
Table 2. Apparent binding constants (Ka), linear correlation coefficients (R2) and numbers of binding sites (n) recovered for the steady-state fluorescence quenching of BSA by various surfactants in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K (for CPC additionally at 288 and 308 K).
Table 2. Apparent binding constants (Ka), linear correlation coefficients (R2) and numbers of binding sites (n) recovered for the steady-state fluorescence quenching of BSA by various surfactants in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K (for CPC additionally at 288 and 308 K).
SurfactantpHKa [M−1]R2n
CPC5.0288 K0.73 × 1041.0000.98
298 K0.37 × 1040.9990.89
308 K0.74 × 1041.0000.91
7.0288 K23.8 × 1041.0001.18
298 K98.2 × 1040.9901.27
308 K4.22 × 1040.9990.97
CTAB5.00.40 × 1040.9980.93
7.09.12 × 1040.9981.11
SDS5.077.6 × 1040.9921.20
7.085.1 × 1040.9981.20
TWEEN 205.00.02 × 1040.9930.70
7.00.07 × 1040.9980.78
TWEEN 405.00.04 × 1040.9890.55
7.00.07 × 1040.9990.78
TWEEN 805.00.03 × 1040.9920.74
7.00.19 × 1040.9940.87
Table 3. The secondary structure content (in %) of BSA in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K revealed from CD measurements.
Table 3. The secondary structure content (in %) of BSA in 10 mM CACO buffer of pH 5.0 and 7.0 at 298 K revealed from CD measurements.
SystempHα-Helix [%]Strand [%]Turns [%]Unordered [%]
BSA5.06171219
7.06315814
BSA + SDS5.056121318
7.059151017
BSA + CPC5.05891220
7.054201114
BSA + CTAB5.06091021
7.055181016
BSA + TWEEN 205.06191020
7.060151016
BSA + TWEEN 405.062101118
7.06016618
BSA + TWEEN 805.061141014
7.054201114
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Żamojć, K.; Wyrzykowski, D.; Chmurzyński, L. On the Effect of pH, Temperature, and Surfactant Structure on Bovine Serum Albumin–Cationic/Anionic/Nonionic Surfactants Interactions in Cacodylate Buffer–Fluorescence Quenching Studies Supported by UV Spectrophotometry and CD Spectroscopy. Int. J. Mol. Sci. 2022, 23, 41. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23010041

AMA Style

Żamojć K, Wyrzykowski D, Chmurzyński L. On the Effect of pH, Temperature, and Surfactant Structure on Bovine Serum Albumin–Cationic/Anionic/Nonionic Surfactants Interactions in Cacodylate Buffer–Fluorescence Quenching Studies Supported by UV Spectrophotometry and CD Spectroscopy. International Journal of Molecular Sciences. 2022; 23(1):41. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23010041

Chicago/Turabian Style

Żamojć, Krzysztof, Dariusz Wyrzykowski, and Lech Chmurzyński. 2022. "On the Effect of pH, Temperature, and Surfactant Structure on Bovine Serum Albumin–Cationic/Anionic/Nonionic Surfactants Interactions in Cacodylate Buffer–Fluorescence Quenching Studies Supported by UV Spectrophotometry and CD Spectroscopy" International Journal of Molecular Sciences 23, no. 1: 41. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23010041

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop