molecules

Article Polymer– Complexes Based on Cationic Imidazolium Amphiphile, Polyacrylic Acid and DNA Decamer

Darya A. Kuznetsova 1 , Dinar R. Gabdrakhmanov 1, Denis M. Kuznetsov 1, Svetlana S. Lukashenko 1, Valery M. Zakharov 2, Anastasiia S. Sapunova 1, Syumbelya K. Amerhanova 1, Anna P. Lyubina 1, Alexandra D. Voloshina 1 , Diana V. Salakhieva 3 and Lucia Ya. Zakharova 1,*

1 FRC Kazan Scientific Center, Russian Academy of Sciences, Arbuzov Institute of Organic and Physical Chemistry, Arbuzov str. 8, 420088 Kazan, Russia; [email protected] (D.A.K.); [email protected] (D.R.G.); [email protected] (D.M.K.); [email protected] (S.S.L.); [email protected] (A.S.S.); [email protected] (S.K.A.); [email protected] (A.P.L.); [email protected] (A.D.V.) 2 Kazan National Research Technological University, Karl Marx str., 68, 420015 Kazan, Russia; [email protected] 3 Institute of Fundamental Medicine and Biology, Kazan (Volga Region) Federal University, Kremlyovskaya St. 18, 420008 Kazan, Russia; [email protected] * Correspondence: [email protected]; Tel.: +7-(843)-273-22-93

  Abstract: The behavior and physicochemical characteristics of polymer–colloid complexes based on cationic imidazolium amphiphile with a dodecyl tail (IA-12) and polyacrylic acid (PAA) or Citation: Kuznetsova, D.A.; Gabdrakhmanov, D.R.; Kuznetsov, DNA decamer (oligonucleotide) were evaluated using tensiometry, conductometry, dynamic and D.M.; Lukashenko, S.S.; Zakharov, electrophoretic light scattering and fluorescent spectroscopy and microscopy. It has been established V.M.; Sapunova, A.S.; Amerhanova, that PAA addition to the system resulted in a ca. 200-fold decrease in the aggregation S.K.; Lyubina, A.P.; Voloshina, A.D.; threshold of IA-12, with the hydrodynamic diameter of complexes ranging within 100–150 nm. Salakhieva, D.V.; et al. Electrostatic forces are assumed to be the main driving force in the formation of IA-12/PAA complexes. Polymer–Colloid Complexes Based Factors influencing the efficacy of the complexation of IA-12 with oligonucleotide were determined. on Cationic Imidazolium Amphiphile, The nonconventional mode of binding with the involvement of hydrophobic interactions and the Polyacrylic Acid and DNA Decamer. intercalation mechanism is probably responsible for the IA-12/oligonucleotide complexation, and Molecules 2021, 26, 2363. https:// a minor contribution of electrostatic forces occurred. The latter was supported by zeta potential doi.org/10.3390/molecules26082363 measurements and the gel electrophoresis technique, which demonstrated the low degree of charge neutralization of the complexes. Importantly, cellular uptake of the IA-12/oligonucleotide complex Academic Editors: Ramón G. Rubio and Tamar Greaves was confirmed by fluorescence microscopy and flow cytometry data on the example of M-HeLa cells. While single IA-12 samples exhibit roughly similar cytotoxicity, IA-12–oligonucleotide complexes

Received: 2 February 2021 show a selective effect toward M-HeLa cells (IC50 1.1 µM) compared to Chang liver cells (IC50 Accepted: 15 April 2021 23.1 µM). Published: 19 April 2021 Keywords: imidazolium; amphiphile; polyacrylic acid; oligonucleotide; complexation Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affil- iations. 1. Introduction Interactions between oppositely charged and synthetic polyelectrolytes are currently the focus of modern research [1–6]. This is mainly due to the possibility of the control of the properties of these systems by the variation of the chemical struc- Copyright: © 2021 by the authors. ture, ratio and of components, molecular weight of the polymer, pH of the Licensee MDPI, Basel, Switzerland. medium, temperature and ionic strength [7–9]. In addition, the fabrication of complexes This article is an open access article can be mediated by different mechanisms, including the , van der Waals distributed under the terms and interactions and hydrogen bonding, with the contribution of dominating electrostatic in- conditions of the Creative Commons teractions [10–13]. For these reasons, of polyelectrolytes and oppositely charged Attribution (CC BY) license (https:// surfactants are characterized by the complicated behavior of [14–16]. The addi- creativecommons.org/licenses/by/ tion of a high-molecular-weight component can change the aggregation thresholds of the 4.0/).

Molecules 2021, 26, 2363. https://doi.org/10.3390/molecules26082363 https://www.mdpi.com/journal/molecules Molecules 2021, 26, x 2 of 17

surfactants are characterized by the complicated behavior of solutions [14–16]. The addi- tion of a high-molecular-weight component can change the aggregation thresholds of the system [17–20]. Moreover, the investigation of polymer–colloid systems allows simulating the interaction of cationic amphiphiles with natural polymers such as DNA, and so on [7,21–23]. This is of importance in gene therapy [24,25] because the application of naked DNA molecules is complicated due to their huge size, the electrostatic repulsion between nucleotides and negatively charged membranes and the immune response Molecules 2021, 26, 2363 2 of 17 observed [26]. Therefore, efficient transfection can be achieved using optimized carriers of polynucleotides, which allow compacting large DNA molecules and recharging them. systemThis allows [17–20]. Moreover,genetic material the investigation to be ofrecognized polymer–colloid by the systems cell allows [27,28], simulating pass through the cell themembrane interaction by of cationicthe endocytosis amphiphiles mechanism with natural polymers[29] and such receive as DNA, protection proteins and from biodegrada- sotion. on [ 7From,21–23 this]. This point is of of importance view, cationic in gene surfactants therapy [24,25 are] because among the the application most prospective of carriers naked[30–34 DNA], including molecules is those complicated bearing due an to theirimidazolium huge size, the moiety electrostatic [35– repulsion38]. Imidazolium am- between nucleotides and negatively charged cell membranes and the immune response phiphiles and DNA were documented to form sustainable nontoxic complexes that pro- observed [26]. Therefore, efficient transfection can be achieved using optimized carriers of polynucleotides,tect from enzymatic which allow degradation compacting large and DNA provide molecules a andcharge recharging density them. sufficient This for DNA allowstransport genetic through material the to be cell recognized membrane by the [35]. cell [27 Nonviral,28], pass throughvectors the based on the of byand the imidazolium endocytosis mechanism amphiphiles [29] and were receive reported protection and from were biodegradation. less effective From than this commercially pointavailable of view, lipofectamine cationic surfactants 2000 are; however among the, they most prospectivehad lower carriers toxicity [30 [36].–34], includ- ing those bearing an imidazolium moiety [35–38]. Imidazolium amphiphiles and DNA were documentedA brief literature to form sustainableoverview nontoxicmakes it complexes obvious thatthat protect the fabrication from enzymatic of complexes based degradationon amphiphiles and provide and polyelectrolytes a charge density sufficient of various for DNA types transport is urgent. through Nevertheless, the cell despite the membranelong-term [35 study]. Nonviral of these vectors systems, based onthere the mixtureis little of understanding lipids and imidazolium of the factors am- influencing phiphilestheir functional were reported activity. and were Much less effectiveattention than is commercially paid to the available interaction lipofectamine of cationic surfactants 2000; however, they had lower toxicity [36]. with anionic synthetic polyelectrolytes because they could simulate complexation be- A brief literature overview makes it obvious that the fabrication of complexes based ontween amphiphiles DNA and and polyelectrolytesamphiphilic compounds. of various types Unlike is urgent. strong Nevertheless, polyelectrolytes, despite the weak polyelec- long-termtrolytes studydemonstrate of these systems, complicated there is and little tunable understanding solution of the behavior, factors influencing which is controlled by theirthe charge functional density activity. of Much a polymer attention chain, is paid which to the, interactionin turn, is of determined cationic surfactants by the nature of mac- withromolecules, anionic synthetic concentration polyelectrolytes of charged because theyadditives, could simulate solution complexation pH, etc. For between cationic surfactant– DNA and amphiphilic compounds. Unlike strong polyelectrolytes, weak polyelectrolytes demonstratepolyacrylic complicated acid (PAA) and binary tunable solutionsystems behavior,, the critical which isionization controlled bydegree the charge αc should be taken densityinto account. of a polymer The chain, theoretical which, invalue turn, of is determinedαc = 0.35 can by thebe naturecalculated of macromolecules, based on Manning’s coun- concentrationterion condensation of charged additives, theory [39]. solution Typically, pH, etc. For electrostatic cationic surfactant–polyacrylic interactions prevail above αc, acidwhile (PAA) the binary hydrophobic systems, the effect critical in ionizationcombination degree withαc should hydrogen be taken binding into account. contributes to com- The theoretical value of αc = 0.35 can be calculated based on Manning’s counterion con- plexation below αc. In accordance with the literature data [40], for surfactant–PAA sys- densation theory [39]. Typically, electrostatic interactions prevail above αc, while the hydrophobictems, this threshold effect in combination occurs at with a pH hydrogen of ca. 4.2, binding beyond contributes which to the complexation ionization degree mark- belowedly αincreasesc. In accordance and electrostatically with the literature driven data [40 ], complexes for surfactant–PAA are formed. systems, Based this on this infor- thresholdmation, occursit can at be a pHexpected of ca. 4.2, that beyond these which complexes the ionization show degree stimuli markedly-responsive increases properties regu- andlated electrostatically by the solution driven pH, complexes concentration are formed. of Based components, on this information, ionic force, it can etc. be ex- This allows con- pected that these complexes show stimuli-responsive properties regulated by the solution pH,trol concentrationling the aggregation of components, capacity ionic force, and etc. functional This allows ac controllingtivity of thethe aggregation systems based on PAA. capacityTherefore, and functionalthis study activity is devoted of the systemsto the basedelucidation on PAA. of Therefore, the self- thisassembling study is behavior and devotedpractical to thepotentiality elucidation of of the self-assembling systems based behavior on imidazolium and practical amphiphiles potentiality of thebearing a dodecyl systemshydrophobic based on tail imidazolium (IA-12, Figure amphiphiles 1) and bearing two polyanions a dodecyl hydrophobic: (i) synthetic tail polyelectrolyte (IA-12, (poly- Figure1) and two polyanions: (i) synthetic polyelectrolyte (polyacrylic acid, PAA, Figure1 ) andacrylic (ii) a DNA acid, decamer PAA, (oligonucleotide). Figure 1) and For (ii) amphiphile/polyelectrolyte, a DNA decamer (oligonucleotide). systems their For am- aggregationphiphile/polyelectrolyte characteristics in, aqueoussystems solutions their aggregation were quantified characteristics under different insurfac- aqueous solutions tantwere quantified and under fixed different polyelectrolyte surfactant concentrations concentrations in accordance and with fixed an earlier polyelectrolyte con- protocolcentrations [41]. in accordance with an earlier protocol [41].

FigureFigure 1. 1.Chemical Chemical formulas formulas of IA-12 of and IA- PAA12 and studied PAA in thisstudied report. in this report.

Molecules 2021, 26, 2363 3 of 17

2. Results and Discussion 2.1. Mixed Systems Based on IA-12 and Polyacrylic Acid Binary IA-12/PAA systems were studied by the variation of the amphiphile concen- tration and three fixed concentrations of polyelectrolyte (1, 3 and 5 mM). In accordance with our previous reports [41,42], a spontaneous pH of ~4 occurs under these conditions, which corresponds to an ~8% degree of polyelectrolyte ionization. This means that PAA can be specified as a weak polyelectrolyte bearing a negative charge. The first stage of the investigation was dedicated to the study of the aggregation properties of polymer–colloid systems. For this purpose, the surface tension technique was successfully used (Figure2). Inflection points in surface-tension isotherms are usually taken as the critical concentration (CMC) or the critical association concentration (CAC). Usually, there are more inflection points in the case of surfactant/polyelectrolyte systems in comparison with individual amphiphile solutions. This depends on various factors, including the nature and concentration of components, molecular weight of the polymer, geometry of the molecules, solution pH, polyelectrolyte ionization degree, ionic strength, etc. It was shown that two inflection points were observed for the IA-12/PAA binary systems (Figure2) regardless of the PAA concentration. The first inflection point (CAC1, see arrows in Figure2) reflects the IA-12 concentration, at which the formation of amphiphile/polyelectrolyte complexes begins. The second inflection point (CAC2, see arrows in Figure2) corresponds to the saturation of macromolecules by amphiphile aggregates, after which individual IA-12 aggregates are formed. The CAC1 and CAC2 values obtained are compared with the CMC value for the individual IA-12 solutions [43] (Table1). The addition of polyelectrolyte results in the formation of aggregates at a ca. 200 times lower amphiphile concentration (0.046 mM) than in the case of individual amphiphile solutions (9 mM). This effect is mostly expressed in the case of the lowest Molecules 2021, 26, x PAA concentration (1 mM) and diminishes with an increase in the PAA concentration 4 of 17 (Table1). This fact is probably due to the distribution of amphiphiles between different macromolecules with an increase in their amount in solution.

FigureFigure 2. 2.Surface Surface tension tension isotherms isotherms for the for IA-12/PAA the IA-12/PAA binary systems binary at systems three fixed at PAA three concentra- fixed PAA concen- tions;trations pH =; pH 4, 25 =◦ 4,C. 25 °С.

Table 1. Aggregation thresholds of the IA-12/PAA binary systems obtained by different physico- chemical techniques.

Aggregation Thresholds (mM) CPAA, mM Tensiometry Conductometry Fluorimetry CAC1 CAC2 0 9 [43] - 10 [43] 9 [43] 1 0.05 10.0 5 0.05 3 0.08 9.0 4 0.05 5 0.10 8.0 3.5 0.07

The size characteristics of the IA-12/PAA binary systems were evaluated using the dynamic light scattering technique for amphiphile concentrations beyond CAC1 (Figure 3). This method allows characterizing the hydrodynamic diameter DH of mixed aggregates and proposes their morphology in aqueous solutions. As can be seen, large particles with a DH of 100–150 nm are formed, which can increase to 350 nm with the amphiphile con- centration in some cases. Earlier, we reported that the hydrodynamic diameter of individ- ual amphiphile aggregates is 4–6 nm, and they most likely have a spherical shape [43]. Taking into account the small size of the particles in the individual PAA solutions (DH = ~2 nm), it should be suggested that aggregates with a size of ≥100 nm are formed in mixed IA-12/PAA systems due to electrostatic attraction forces. Analogous trends in the varia- tion of DH with the amphiphile concentration are observed for all three fixed PAA con- centrations. First, the aggregate sizes increase with the surfactant concentration, whereaf- ter a decrease in the sizes occurs (Figure 3). The largest size of complexes probably reflects the agglomeration of IA-12 and PAA macromolecules close to the isoelectric point. The comparison of the number-averaged size distributions with intensity-averaged data re- veals that good agreement occurred in the majority of systems (Figure 3). However, bi- modal intensity-averaged size distributions could be seen in some cases, indicating an increase in the polydispersity of the system.

Molecules 2021, 26, 2363 4 of 17

Table 1. Aggregation thresholds of the IA-12/PAA binary systems obtained by different physico- chemical techniques.

Aggregation Thresholds (mM) Tensiometry CPAA, mM Conductometry Fluorimetry CAC1 CAC2 0 9 [43] - 10 [43] 9 [43] 1 0.05 10.0 5 0.05 3 0.08 9.0 4 0.05 5 0.10 8.0 3.5 0.07

To obtain reliable information on the aggregation behavior of IA-12/PAA binary sys- tems, they were examined by two additional techniques: conductometry and fluorescence spectroscopy using pyrene as a probe (Figures S1 and S2, Supplementary Materials). This is motivated by the fact that these methods are based on different physical principles responsible for their accuracy while providing specific advantages and limitations. Ten- siometry evaluates the structural behavior of the systems at the surface layer rather than in the bulk phase. Meanwhile, the interfacial characteristics of polyelectrolyte–surfactant systems can be modulated by different factors that are not related to the aggregation. At the same time, conductometry provides a valuable instrument for the quantification of the aggregation behavior of charged species. The fluorimetry technique is very sensitive to changes in micropolarity that result from the surfactant aggregation. Data of the two alternative techniques are consistent with those obtained by tensiometry (Table1). The size characteristics of the IA-12/PAA binary systems were evaluated using the dynamic light scattering technique for amphiphile concentrations beyond CAC1 (Figure3). This method allows characterizing the hydrodynamic diameter DH of mixed aggregates and proposes their morphology in aqueous solutions. As can be seen, large particles with a DH of 100–150 nm are formed, which can increase to 350 nm with the amphiphile concen- tration in some cases. Earlier, we reported that the hydrodynamic diameter of individual amphiphile aggregates is 4–6 nm, and they most likely have a spherical shape [43]. Taking into account the small size of the particles in the individual PAA solutions (DH = ~2 nm), it should be suggested that aggregates with a size of ≥100 nm are formed in mixed IA- 12/PAA systems due to electrostatic attraction forces. Analogous trends in the variation of DH with the amphiphile concentration are observed for all three fixed PAA concentra- tions. First, the aggregate sizes increase with the surfactant concentration, whereafter a decrease in the sizes occurs (Figure3). The largest size of complexes probably reflects the agglomeration of IA-12 and PAA macromolecules close to the isoelectric point. The comparison of the number-averaged size distributions with intensity-averaged data reveals that good agreement occurred in the majority of systems (Figure3). However, bimodal intensity-averaged size distributions could be seen in some cases, indicating an increase in the polydispersity of the system. Additional information can be obtained through the analysis of correlation functions (Figures S3–S5). Aggregates of a smaller size show faster diffusion; hence, decay of the correlation function to the baseline occurs for a shorter period of time. In the case of the larger aggregates, the time of decay increased. For the IA-12/PAA 1 mM binary system (Figure S3), the maximum time of the decay of the correlation function (about 2000 µs) was found at 0.5 and 2 mM amphiphile concentrations. This corresponds to the largest aggregates in this system (Figure3a). Moreover, the shape of the correlation function provides evidence of the existence of only one type of aggregate in the polymer– colloid solutions. This trend is observed for all the IA-12/PAA binary systems studied (Figures S3–S5). Molecules 2021, 26, 2363 5 of 17 Molecules 2021, 26, x 5 of 17

(a)

(b)

(c)

Figure 3. Number- and intensity-averaged size distribution for the IA-12/PAA binary systems at ◦ various fixed PAA concentrations: (a) 1 mM; (b) 3 mM; (c) 5 mM; 25 C. Molecules 2021, 26, 2363 6 of 17

The charge characteristics of the binary IA-12/PAA systems were evaluated using the electrophoretic light scattering technique. The data shown in Figure4 and Figure S6 testify that an increase in the surfactant concentration is accompanied by a compensation of the negative charge of the PAA anion up to a zero value, after which positively charged systems appeared. Unlike the data available [42], the isoelectric point corresponding to the zero zeta potential of binary systems is practically unaffected by the PAA concentration. This can be due to several factors. First, PAA is a weak polyelectrolyte, and its charge behavior is strongly determined by the solution pH. Under spontaneous conditions, the solution pH is close to four (Figure S7). A slight decrease is observed with an increase in the surfactant concentration. The latter is probably caused by the shift of the pKa value induced by . Under these conditions, only~8% of carboxylic groups are ionized [41], which is probably responsible for the weak influence of the polyelectrolyte concentration on electrostatic interactions. Second, the charge behavior of the IA-12/PAA binary systems can be guided by a fine balance between several factors, including the variation in the solution pH as a function of the polymer and surfactant concentration and a shift in pKa values. Therefore, no simple correlation is observed in this case. Third, additional information can be obtained from the turbidity data, demonstrating that binary systems tend to become inhomogenous close to an equimolar ratio (Figure5); therefore, some discrepancy of data in this concentration range can occur.

Molecules 2021, 26, x 7 of 17

This phenomenon could be caused by the fact that uncharged colloid species tend to ag- gregate, thereby inducing inhomogeneity. On the contrary, charged particles are more stable due to electrostatic repulsion preventing their aggregation and sedimentation. As can be seen, the maximum of absorbance shifts to the right with an increase in the PAA concentration, thereby indicating that the charge behavior of the binary systems is influ- enced by the polyelectrolyte concentration, although this is unobvious from the zeta po- Figuretential 4. dataZeta potentialin Figure versus 4. IA-12 concentration plot for the IA-12/PAA binary systems; 25 ◦C.

FigureFigure 5. 5.Optical Optical density density at λat= λ 500 = 500 nm versusnm versus IA-12 IA concentration-12 concentration plot for plot the IA-12/PAAfor the IA-12/PAA binary binary systems; CPAA = 1, 3, 5 mM, 25◦ °C. systems; CPAA = 1, 3, 5 mM, 25 C.

2.2. Amphiphile/DNA Decamer Interactions The next step of the study was dedicated to the interactions between the imidazo- lium-containing amphiphile and DNA decamer (oligonucleotide (oNu)). These investiga- tions allow simulating the binding of IA-12 with nucleotide units of DNA. Initially, the size and charge characteristics of IA-12/DNA decamer complexes were evaluated, and the binding parameters of the components were quantified. To this end, methods of dynamic and electrophoretic light scattering and fluorescence spectroscopy were used. Dynamic light scattering was applied for the determination of the hydrodynamic diameter of par- ticles in the IA-12/oNu binary system (Figure 6). As can be seen, binding of the compo- nents responsible for the complex formation is observed. This is evident from the enlarge- ment of the DH value of particles from 2 nm for an individual oNu to 50 nm and higher sizes (up to 250 nm) for the IA-12/oNu binary systems. It is noteworthy that these sizes are optimal for biotechnological purposes because they allow the nanoparticles to circu- late in the bloodstream for a long time [44].

Figure 6. Number-averaged size distribution for the IA-12/oNu binary system at various am- phiphile/oNu molar ratios: 1—individual oNu, 2—0.02, 3—0.045, 4—0.094, 5—0.194, 6—0.394, 7— 0.79, 8—1.2; 9—3.9; 25 °C.

Molecules 2021, 26, x 7 of 17

This phenomenon could be caused by the fact that uncharged colloid species tend to ag- gregate, thereby inducing inhomogeneity. On the contrary, charged particles are more stable due to electrostatic repulsion preventing their aggregation and sedimentation. As can be seen, the maximum of absorbance shifts to the right with an increase in the PAA concentration, thereby indicating that the charge behavior of the binary systems is influ- enced by the polyelectrolyte concentration, although this is unobvious from the zeta po- tential data in Figure 4.

Molecules 2021, 26, 2363 7 of 17

In all cases, an increase in the IA-12 concentration results in the enhancement of the opalescence in the system followed by its disappearance, with a further increase in the micelle concentration (Figure5). It is noteworthy that the turbidimetric data are in good accordance with the above variation in zeta potential in the systems. The highest absorbance was detected in the area of the isoelectric points of the systems, while the solutions were more transparent in the region of the existence of positive or negative zeta potential. This phenomenon could be caused by the fact that uncharged colloid species tendFigure to aggregate,5. Optical therebydensity inducing at λ = 500 inhomogeneity. nm versus IA On-12 the concentration contrary, charged plot particlesfor the IA are-12/PAA binary moresystems; stable CPAA due = to1, electrostatic3, 5 mM, 25 repulsion °C. preventing their aggregation and sedimentation. As can be seen, the maximum of absorbance shifts to the right with an increase in the PAA2.2. Amphiphile/DNA concentration, thereby Decamer indicating Interactions that the charge behavior of the binary systems is influenced by the polyelectrolyte concentration, although this is unobvious from the zeta potentialThe data next in Figurestep of4. the study was dedicated to the interactions between the imidazo- lium-containing amphiphile and DNA decamer (oligonucleotide (oNu)). These investiga- 2.2. Amphiphile/DNA Decamer Interactions tions allow simulating the binding of IA-12 with nucleotide units of DNA. Initially, the The next step of the study was dedicated to the interactions between the imidazolium- containingsize and charge amphiphile characteristics and DNA decamer of IA- (oligonucleotide12/DNA decamer (oNu)). complexes These investigations were evaluated, and the allowbinding simulating parameters the binding of the of IA-12components with nucleotide were quantified. units of DNA. To Initially, this end, the size methods and of dynamic chargeand electrophoretic characteristics of IA-12/DNAlight scattering decamer and complexes fluorescence were evaluated, spectroscopy and the bindingwere used. Dynamic parameterslight scattering of the componentswas applied were for quantified.the determination To this end, of methods the hydrodynamic of dynamic and diameter of par- electrophoretic light scattering and fluorescence spectroscopy were used. Dynamic light scatteringticles in the was IA applied-12/oNu for the binary determination system of(Figure the hydrodynamic 6). As can diameter be seen, of binding particles of the compo- innents the IA-12/oNuresponsible binary for the system complex (Figure formation6). As can be is seen, observed. binding This of the is components evident from the enlarge- responsiblement of the for D theH value complex of formation particles is from observed. 2 nm This for is an evident individual from the oNu enlargement to 50 nm and higher ofsizes the D(upH value to 250 of particlesnm) for from the 2IA nm-12/oNu for an individual binary systems. oNu to 50 It nm is and noteworthy higher sizes that these sizes (upare tooptimal 250 nm) for for biotechnological the IA-12/oNu binary purposes systems. because It is noteworthy they allow that these the nanoparticles sizes are to circu- optimal for biotechnological purposes because they allow the nanoparticles to circulate in thelate bloodstream in the bloodstream for a long time for [a44 long]. time [44].

Figure 6. 6.Number-averaged Number-averaged size size distribution distribution for the for IA-12/oNu the IA-12/oNu binary systembinary at system various at am- various am- phiphile/oNu molar molar ratios: ratios: 1—individual 1—individual oNu, oNu, 2—0.02, 2— 3—0.045,0.02, 3— 4—0.094,0.045, 4 5—0.194,—0.094, 6—0.394,5—0.194, 6—0.394, 7— ◦ 7—0.79,0.79, 8— 8—1.2;1.2; 9 9—3.9;—3.9; 2525 C.°C. Another important criterion for the efficacy of potential nonviral vectors is their ability to neutralize the negative charges of nucleotide units, which was examined by the electrophoretic light scattering method (Figure7). It was demonstrated that IA-12 has a weak ability of oNu charge compensation (∆ξ does not exceed 20 mV); the isoelectric point was not achieved even with the excess of amphiphile. This trend has earlier been reported for the higher homologs of IA-12 [45], which was attributed to the fact that the delocalized positive charge of the imidazolium group cannot compensate for the negative charge of the backbones. Molecules 2021, 26, x 8 of 17

Molecules 2021, 26, x 8 of 17 Another important criterion for the efficacy of potential nonviral vectors is their abil- ity to neutralize the negative charges of nucleotide units, which was examined by the elec- trophoretic light scattering method (Figure 7). It was demonstrated that IA-12 has a weak abilityAnother of oNu important charge compensation criterion for (Δξthe efficacydoes not of exceed potential 20 mV); nonviral the isoelectric vectors is theirpoint abil- was itynot to achieved neutralize even the with negative the excess charges of ofamphiphile. nucleotide Thisunits, trend which has was earlier examined been reported by the elec- for trophoreticthe higher homologslight scattering of IA -method12 [45], (Figurewhich was7). It attributed was demonstrated to the fact that that IA the-12 delocalizedhas a weak abilitypositive of charge oNu charge of the compensation imidazolium (Δξgroup does cannot not exceed compensate 20 mV); for the the isoelectric negative pointcharge was of notthe phosphateachieved even backbones. with the excess of amphiphile. This trend has earlier been reported for the higher homologs of IA-12 [45], which was attributed to the fact that the delocalized Molecules 2021, 26, 2363 positive charge of the imidazolium group cannot compensate for the8 ofnegative 17 charge of the phosphate backbones.

Figure 7. Electrokinetic potential versus amphiphile/oNu molar ratio for the IA-12/oNu, IA- 14/oNu [45], IA-16/oNu [45] and IA-18/oNu [45] binary systems; 25 °C.

FigureFigureTo 7. 7. Electrokinetictest Electrokinetic whether potential a potential DNA versus oligomer versus amphiphile/oNu amphiphile/ can be used molaroNu for ratio molar the for reliable ratio the IA-12/oNu, for prediction the IA- IA-12/oNu of the, IA binding- 14/oNu14/behavioroNu [45 [45]], IA-16/oNuof, IA macromolecular-16/oNu [45] and[45] IA-18/oNu and analogs,IA-18/ [45oNu] binary electrophoretic [45] systems; binary 25 systems;◦C. analysis 25 °C. involving pDNA was car- ried out (Figure 8), which demonstrated that IA-12 does not bind with pDNA in a wide To test whether a DNA oligomer can be used for the reliable prediction of the binding behaviorconcentrationTo of test macromolecular whether range aof DNA analogs,the amphiphile. oligomer electrophoretic can Some be analysis used retardation for involving the reliable of pDNA both wasprediction supercoiled carried of andthe binding relaxed outbehaviorforms (Figure of8 pDNA),of which macromolecular demonstratedwas, however, thatanalogs, observed IA-12 does electrophoretic notat increased bind with pDNA analysisN/P inratios a wideinvolving from concen- 38:1 pDNA to 300:1, was indi- car- trationriedcating out range a weak(Figure of the interaction amphiphile.8), which ofdemonstrated Some the retardationcationic surfactantthat of both IA supercoiled-12 withdoes pDNA.not and bind relaxed T hewith formstypical pDNA cationic in a wide sur- ofconcentrationfactant pDNA CTAB was, however, rangeused as observedof a the reference amphiphile. at increased compound N/PSome ratios wasretardation from shown 38:1 ofto 300:1,bothcompletely indicatingsupercoiled bind pDNAand relaxed at an aforms weak interactionof pDNA of was, the cationic however, surfactant observed with pDNA.at increased The typical N/P cationic ratios surfactantfrom 38:1 to 300:1, indi- CTABN/P usedratio as of a reference14:1 (Figure compound S9). wasThe shown DNA to-binding completely ability bind pDNA of CTAB at an N/Pwas ratioinferior to that ob- ofcatingserved 14:1 (Figure a for weak cationic S9). interaction The polymers DNA-binding of the such abilitycationic as cationic of CTABsurfactant polyaspartamides was inferior with topDNA. that observed withThe tanypical for N/P cationic ratio as sur-low cationicfactantas 1:1 [46]. polymers CTAB These used such dat asasa cationic aare reference in polyaspartamides good compound agreement with was with an shown N/Pthe ratioabove to ascompletely lowzeta as potential 1:1 [46 bind]. resultspDNA of at thean TheseN/PIA-12/oNu ratio data are of complexes,in 14:1 good (Figure agreement testifying S9). with The the thatDNA above electrostatic-binding zeta potential ability interactions results of ofCTAB the IA-12/oNu play was a inferior minor roleto that in ob- the complexes, testifying that electrostatic interactions play a minor role in the complexation. servedcomplexation. for cationic Unlike polymers such surfactants,as cationic polyaspartamides imidazolium head with groups an N/Pwith ratio delocalized as low Unlike ammonium surfactants, imidazolium head groups with delocalized cationic charge giveascationic 1:1 an insufficient[46]. charg Thesee give contribution dat aan are insufficient in to phosphategood agreement contribution anion neutralization. with to phosphatethe above zetaanion potential neutralization. results of the IA-12/oNu complexes, testifying that electrostatic interactions play a minor role in the complexation. Unlike ammonium surfactants, imidazolium head groups with delocalized cationic charge give an insufficient contribution to phosphate anion neutralization.

Figure 8. Electrophoretic mobility of pDNA in a complex with IA-12. Ctrl shows pure pDNA. N/P ratios are indicated above the corresponding wells. A DNA ladder from 100 to 10,000 bp was used.

Molecules 2021, 26, x 9 of 17

Figure 8. Electrophoretic mobility of pDNA in a complex with IA-12. Ctrl shows pure pDNA. N/P ratios are indicated above the corresponding wells. A DNA ladder from 100 to 10,000 bp was used.

DNA–surfactant complexes were characterized by the DLS technique at an N/P ratio of 10:1 (Figures S10–S12). CTAB formed with pDNA well-defined nanosized complexes Molecules 2021, 26, 2363 (DH 267 nm, PDI 0.234) with a highly positive zeta potential, indicating9 of the 17 condensation and cationization of pDNA by the surfactant. Under the same conditions, IA-12 neither packed pDNA nor induced its cationization, further showing the lack of its strong elec- trostaticDNA–surfactant interaction complexes with pDNA were characterized and, therefore, by the the DLS different technique atDNA an N/P-bindi rationg properties of ofIA 10:1-12 (Figures compared S10–S12). to CTAB. CTAB formed with pDNA well-defined nanosized complexes (DH 267 nm,The PDI a 0.234)dditional with aquantification highly positive zetaof IA potential,-12/oNu indicating binding the was condensation carried out and using the fluo- cationization of pDNA by the surfactant. Under the same conditions, IA-12 neither packed pDNArescence nor technique induced its cationization,and ethidium further bromide showing (EB) the as lack an of intercalation its strong electrostatic probe. This approach interactionis based on with the pDNA exclusion and, therefore, of EB from the differentthe complex DNA-bindinges by surfactant properties molecules, of IA-12 which is ac- comparedcompanied to CTAB. by the quenching of EB fluorescence. Emission spectra of EB/oNu complexes at variousThe additional IA-12 quantificationconcentrations of IA-12/oNu (Figure S13) binding demonstrate was carried that out using the higher the fluo- the amphiphile rescenceconcentration, technique the and lower ethidium the bromideintensity (EB) of asthe an corresponding intercalation probe. band. This Fluorescence approach spectra for is based on the exclusion of EB from the complexes by surfactant molecules, which is accompaniedvarious IA- by12/oNu the quenching molar of ratios EB fluorescence. were used Emission for the spectra calculation of EB/oNu of complexesthe IA-12/oNu binding atdegree various β IA-12 (Figure concentrations 9) using (FigureEquation S13) ( demonstrate1). An increase that the in higher the amphiphile the amphiphile concentration is concentration,shown to induce the lower an increase the intensity in ofthe the β correspondingvalue up to 75% band. under Fluorescence IA-12 spectra excess, for while no charge variousneutralization IA-12/oNu was molar observed ratios were in this used range for the (Figure calculation 7). of It theis n IA-12/oNuoteworthy binding that earlier reported degree β (Figure9) using Equation (1). An increase in the amphiphile concentration is showndata for to induce higher an homologs increase in the of βIAvalue-12 up(tetradecyl, to 75% under hexadecyl IA-12 excess, and while octadecyl no charge derivatives) ex- neutralizationhibited an almost was observed quantit inative this range binding (Figure of7 ).oNu It is [45]. noteworthy This emphasizes that earlier reported the crucial role of the datahydrophobic for higher homologs effect and of IA-12 cooperative (tetradecyl, interactions hexadecyl and in octadecyl the binding derivatives) efficacy exhib- of imidazolium itedsurfactants an almost toward quantitative nucleotides. binding of I oNut can [45 be]. proposed This emphasizes that an the alternative crucial role of intercalation the mech- hydrophobic effect and cooperative interactions in the binding efficacy of imidazolium anism of binding probably occurs apart from a weak electrostatic contribution. This is surfactants toward nucleotides. It can be proposed that an alternative intercalation mecha- nismbased of bindingon the probablyfact that occurs amphiphiles apart from awith weak a electrostatic planar aromatic contribution. imidazolium This is based head group are oncapable the fact of that locat amphiphilesing nucleobases. with a planar Importantly, aromatic imidazolium the alkyl head tail group length are capable of imidazolium am- ofphiphiles locating nucleobases. strongly determines Importantly, theirthe alkyl intercalation tail length of ability. imidazolium In particular, amphiphiles homologs with stronglyshorter determinesalkyl tails their cannot intercalation interact ability. with Inpolynucleotides particular, homologs in withsuch shorter a way alkyl [47]. Comparative tails cannot interact with polynucleotides in such a way [47]. Comparative analysis of data obtainedanalysis for of imidazolium data obtained homological for imidazolium series allows ushomological to conclude thatseries the superpositionallows us to conclude that ofthe two superposition factors controls theof two IA-12 factors complexation controls with the a DNA IA- decamer,12 complexation namely hydrophobic with a DNA decamer, bindingnamely and hydrophobic the intercalation binding mechanism. and the intercalation mechanism.

FigureFigure 9. 9.Binding Binding degree degree of IA-12 of IA with-12 with oNu versusoNu versus IA-n concentration IA-n concentration plot for the plot IA-12/oNu, for the IA-12/oNu, IA-14/oNuIA-14/oNu [45 [45],], IA-16/oNu IA-16/oNu [45] and[45] IA-18/oNu and IA-18/oNu [45] mixtures; [45] mixtures; 25 ◦C. 25 °C.

To test whether the IA-12/oNu complex could be transfected through the cell barrier to its interior,To test awhether fluorescence the microscopyIA-12/oNu technique complex was could used. be Figuretransfected 10 demonstrates through the cell barrier HeLato its cancer interior cells, staineda fluorescence with Hoechst microscopy 33342 in the technique absence and was in the used. presence Figure of the 10 demonstrates IA-12/oNuHeLa cancer complexes. cells stained This dye with is capable Hoechst of binding 33342 in with the both absence DNA ofand the in nucleus the presence of of the IA- HeLa12/oNu cells complexes. (Figure 10a), This and thedye added is capable oNu marked of bind theming with with blue. both In DNA Figure of 10 b,the the nucleus of HeLa blue nimbus corresponding to oNu localization is clearly seen close to the stained nucleus cells (Figure 10a), and the added oNu marked them with blue. In Figure 10b, the blue nimbus corresponding to oNu localization is clearly seen close to the stained nucleus of

Molecules 2021, 26, x 10 of 17

Molecules 2021, 26, x 10 of 17

Molecules 2021, 26, 2363 the HeLa cell, which indicates that successful penetration of the complexes10 of 17 through the cell barrier occurred. the HeLa cell, which indicates that successful penetration of the complexes through the of the HeLa cell, which indicates that successful penetration of the complexes through the cell barrierbarrier occurred. occurred.

(a) (b)

Figure 10. Fluorescent micrographs of HeLa cells in the absence (a) and in the presence of the IA- (a) (b) 12/oNu complex (b); CoNu = 1 mМ, CIA-12 = 0.5 mМ. The arrow indicates the presence of oligonucle- otideFigure in 10. theFluorescent cytoplasm micrographs of cells. of HeLa cells in the absence (a) and in the presence of the Figure 10. Fluorescent micrographs of HeLa cells in the absence (a) and in the presence of the IA- IA-12/oNu complex (b); CoNu = 1 mM, CIA-12 = 0.5 mM. The arrow indicates the presence of 12/oligonucleotideoNu complex in the (b cytoplasm); CoNu = of1 mМ, cells. CIA-12 = 0.5 mМ. The arrow indicates the presence of oligonucle- otide Thein the cellular cytoplasm uptake of cells. of the IA-12/oNu complexes was further testified using flow cy- tometryThe cellular(Figure uptake 11). Untreated of the IA-12/oNu cells were complexes used wasas negative further testified controls. using After flow treatment with substances,cytometryThe (Figurecellular M- HeLa11 ).uptake Untreated cells of were cellsthe IA werefixed-12/oNu used and as stained negativecomplexes controls.with was Hoechst After further treatment 33342 testified with(blue), using which flow was cy- usedsubstances, as a fluorescent M-HeLa cells probe. were fixed The and dye stained allows with one Hoechst to mark 33342 cell (blue), nuclei which an wasd oNu penetrated tometryused as a fluorescent(Figure 11). probe. Untreated The dye allowscells were one to used mark as cell negative nuclei and controls. oNu penetrated After treatment with intosubstances,into cellscells in in the Mthe form-HeLa form of bluecells of spots.blue were spots. It fixed was shownIt and was stained that shown upon with treatmentthat Hoechstupon with treatment IA-12/oNu33342 (blue), with IAwhich-12/oNu was complexesusedcomplexes, as a thefluorescent, the fluorescence fluorescence probe. intensity intensityThe was dye higher wasallows than higher one in the thanto control, mark in the whichcell control, nuclei indicates anwhich thatd oNu indicates penetrated that theintothe penetrationpenetration cells in the of theof form the test complexestestof blue complexes spots. into M-HeLa Itinto was M cells -shownHeLa occurred. cells that occurrupon ed.treatment with IA-12/oNu complexes, the fluorescence intensity was higher than in the control, which indicates that the penetration of the test complexes into M-HeLa cells occurred.

FigureFigure 11. 11.Cellular Cellular uptake uptake study study of the of IA-12/oNu the IA-12/ complex.oNu complex 1—Intact. M-HeLa1—Intact cells; M- 2—M-HeLaHeLa cells; 2—M-HeLa cellscells inin the the presence presence of the of IA-12/oNuthe IA-12/ complex.oNu complex. FigureFrom 11. the Cellular viewpoint uptake of a practicalstudy of application,the IA-12/oNu cytotoxicity complex is. an 1— essentialIntact M characteristic-HeLa cells; 2—M-HeLa cellsof the inFrom systems. the presencethe Table viewpoint2 of summarizes the IA of- 12/a practicaloNu the cytotoxic complex application, effect. of the cytotoxicity single IA-12 solutionis an essential and characteris- ticIA-12/oNu of the systems. complexes Tabl towarde 2 summarizes M-HeLa and Changthe cytotoxic liver cell effect lines. Cationicof the single surfactants IA-12 solution and IAare-12/oNu typicallyFrom the characterizedcomplexes viewpoint toward by of a rathera practical M- highHeLa cytotoxicity, application,and Chang which livercytotoxicity is confirmedcell lines. is byCationican the essential IC50 surfactants characteris- are values of IA-12 given in Table2, with a higher cytotoxic effect occurring against Chang typically characterized by a rather high cytotoxicity, which is confirmed by the IC50 values ticliver of cells. the However,systems. theTabl transitione 2 summarizes to the binary the IA-12/oNu cytotoxic complex effect of is the accompanied single IA by-12 solution and of IA-12 given in Table 2, with a higher cytotoxic effect occurring against Chang liver cells. IAthe- appearance12/oNu complexes of selectivity toward in cytotoxic M-HeLa action, and namely Chang a sharp liver decrease cell lines. in IC50 Cationicvalues (to surfactants are However,1.1typicallyµM), which characterized the occurs transition with by cancer to a the rather cells binary along high IA with cytotoxicity,-12/oNu an increase complex which in the isviability is accompanied confirmed of normal by by the the IC 50appear- values anceofcells. IA Analogousof-12 selectivity given changesin Table in cytotoxic in 2, cytotoxic with aaction, activityhigher namely arecytotoxic documented a sharp effect in decreaseoccurring [48], where in anagainst IC MTT50 values test Chang (to liver 1.1 µMcells.), revealed a decrease in the cytotoxicity of complexes compared to single cationic surfactants. whichHowever, occurs the with transition cancer to cells the alongbinary with IA- 12/oNuan increase complex in the is viability accompanied of normal by thecells. appear- Anal- ogousance of changes selectivity in cytotoxic in cytotoxic activity action, are namely documented a sharp in decrease[48], where in ICan50 MTT values test (to revealed 1.1 µM a), decreasewhich occurs in the with cytotoxicity cancer cells of complexesalong with comparedan increase t oin single the viability cationic of surfactants. normal cells. Anal- ogous changes in cytotoxic activity are documented in [48], where an MTT test revealed a Tabledecrease 2. The in cytotoxicthe cytotoxicity effect of ofIA -complexes12 in the individual compared system to single and in cationic the binary surfactants. system with oNu toward normal and tumor human cell lines. Table 2. The cytotoxic effect of IA-12 in the individual system and in the binary system with oNu IC50 µM toward normal and tumor human cell lines.

IC50 µM

Molecules 2021, 26, x 11 of 17 Molecules 2021, 26, 2363 11 of 17

Table 2. The cytotoxic effectM-HeLa of IA-12 in the individual system and in the binaryChang system withLiver oNu toward normalIA-12 and tumor human cellIA lines.-12/oNu IA-12 IA-12/oNu

39 ± 0.1 1.1 ± 0.08IC50 µM 17.0 ± 0.1 21.3 ± 1.8 M-HeLa Chang Liver 2.3. MembranotropicIA-12 Properties IA-12/oNu and Cell Penetration IA-12 IA-12/oNu Another39 ± 0.1 important feature 1.1 ± 0.08 of drug and 17.0 gene± 0.1 delivery systems 21.3 ± 1.8based on amphiphilic compounds is their ability to pass through cell membranes composed of bilayers. With2.3. Membranotropic this in mind Properties, the ability and Cell of PenetrationIA-12 to integrate with dipalmitoylphosphatidylcholine (DPPC)Another liposomes important was feature examined. of drug and This gene property delivery systemswas studied based onusing amphiphilic the turbidimetric ti- compounds is their ability to pass through cell membranes composed of lipid bilayers. With tration of DPPC liposome dispersion by IA-12 solution to determine the temperature of this in mind, the ability of IA-12 to integrate with dipalmitoylphosphatidylcholine (DPPC) theliposomes gel/liquid was examined.crystal phase This propertytransition. was In studied the absence using the of turbidimetric foreign compounds, titration of the tempera- tureDPPC of liposome the DPPC dispersion main byphase IA-12 transition solution to determine (TPT) equals the temperature 41 ± 0.1 ° ofС the(Figure gel/liquid S14). Therein, TPT iscrystal taken phase as the transition. inflection In the point absence in of reversed foreign compounds, S-shape plots. the temperature Surfactants of the and DPPC other additives ◦ main phase transition (TPT) equals 41 ± 0.1 C (Figure S14). Therein, TPT is taken as the are capable of changing the TPT value due to ordering or disturbing the (Fig- inflection point in reversed S-shape plots. Surfactants and other additives are capable ure S14). Figure 12 visualizes turbidimetric data in terms of the temperature of the main of changing the TPT value due to ordering or disturbing the lipid bilayer (Figure S14). phaseFigure transition12 visualizes versus turbidimetric the IA- data12/DPPC in terms molar of the ratio. temperature The initial of the decrease main phase in the TPT value withtransition the IA versus-12 concentration the IA-12/DPPC indicates molar ratio. the The integration initial decrease of amphiphile in the TPT value molecules with in the lipid bilayerthe IA-12 that concentration resulted in indicates the disordering the integration of DPPC of amphiphile alkyl ch moleculesains. This in thepromotes lipid an increase bilayer that resulted in the disordering of DPPC alkyl chains. This promotes an increase inin the the permeabilitypermeability of theof the lipid lipid bilayer bilayer that could that be could used for be biotechnological used for biotechnological tasks. The tasks. The IA-12 DPPC slopeslope ofof dependence dependence became became higher higher at CIA-12 at/C CDPPC ≥/C0.1, which ≥ 0.1, can which be attributed can be to theattributed to the solubilizationsolubilization of of liposomes liposomes in IA-12 in IA aggregates.-12 aggregates.

FigureFigure 12.12.DPPC DPPC main main phase phase transition transition temperature temperature versus the versus amphiphile/lipid the amphiphile/lipid molar ratio plot formolar ratio plot forthe the IA-12/DPPC IA-12/DPPC binary binary system. system. 3. Materials and Methods 3.3.1. Materials Chemicals and Methods 3.1. ChemicalsImidazolium amphiphile IA-12 was synthesized in accordance with the standard experimental procedure [45,49]. Commercially available double-stranded oligonucleotide (oNu)Imidazolium with 10 nucleotide amphiphile units in IA each-12 strand was synthesized (GCGTTAACGC, in accordance molecular weight with the is standard ex- perimental3028 g/mol, Jointprocedure Stock Company [45,49]. Syntol,Commercially Moscow, Russia)available was double used. Plasmid-stranded DNA oligonucleotide ((pDNA)oNu) with pVax1-DsRed 10 nucleotide (3665 bp)units was in kindly each giftedstrand by (GCGTTAACGC, Dilara Gatina (Omics molecular Technology weight is 3028 g/mol,Laboratory, Joint Kazan Stock Federal Company University, Syntol Kazan,, Moscow, Russia). Russia Agarose) was for molecularused. Plasmid biology, DNA (pDNA) ethidium bromide and hexadecyltrimethylammonium bromide (≥99.0%) were purchased pVax1-DsRed (3665 bp) was kindly gifted by Dilara Gatina (Omics Technology Labora- tory, Kazan Federal University, Kazan, Russia). Agarose for molecular biology, ethidium bromide and hexadecyltrimethylammonium bromide (≥99.0%) were purchased from Sigma-Aldrich. As probes for fluorometric measurements, ethidium bromide (Sigma-Al- drich, St. Louis, MO, USA, 95%) and pyrene (Sigma-Aldrich, St. Louis, Mo, USA, 99%) were used. Lipoid PC 16:0/16:0 (1,2-dipalmitoyl-sn-glycero-3-phosphocholine, DPPC) was a gift from Lipoid GmbH (Ludwigshafen, Germany). Hoechst 33342 (Sigma-Aldrich, St. Louis, MO, USA, 99%) was applied for the fluorescence microscopy assay. Polyacrylic acid with an average molecular weight of 1800 g/mol (Sigma-Aldrich, St. Louis, MO, USA,

Molecules 2021, 26, 2363 12 of 17

from Sigma-Aldrich. As probes for fluorometric measurements, ethidium bromide (Sigma- Aldrich, St. Louis, MO, USA, 95%) and pyrene (Sigma-Aldrich, St. Louis, MO, USA, 99%) were used. Lipoid PC 16:0/16:0 (1,2-dipalmitoyl-sn-glycero-3-phosphocholine, DPPC) was a gift from Lipoid GmbH (Ludwigshafen, Germany). Hoechst 33342 (Sigma-Aldrich, St. Louis, MO, USA, 99%) was applied for the fluorescence microscopy assay. Polyacrylic acid with an average molecular weight of 1800 g/mol (Sigma-Aldrich, St. Louis, MO, USA, 99%) was used. Polymer concentrations were given as a on a monomer basis. Purified water (18.2 MΩ cm resistivity at 25 ◦C) from Direct-Q 5 UV equipment (Millipore S.A.S. 67120 Molsheim-France) was used for all solution preparation.

3.2. Preparation of Samples Liposome preparation was carried out by the thin lipid hydration technique as fol- lows [49–51]: a specific amount of DPPC (5.4 mg) was solved in 100 µL of chloroform. The was removed by evaporation overnight at room temperature. The obtained lipid film was suspended by 1 mL of water and thermostated at 55–60 ◦C in a water bath for 30 min. The fabricated rude was frozen and thawed 5 times using liquid nitrogen and a water bath (60 ◦C), respectively. After that, the dispersion was extruded 21 times by using a syringe extruder (LiposoFast Basic, Avestin, Ottawa, ON, Canada) through a polycarbonate filter with 100 nm pores. A stock solution of a DNA decamer was prepared by the dissolvation of commercial crystalline oligonucleotide in 4 mM Tris–HCl buffer (pH = 8.0) and its down to 10 mM (concentration per nucleotide unit). The mixture was heated at 95 ◦C for 5 min and immediately chilled in an ice bath.

3.3. Methods 3.3.1. Tensiometry A Kr˝ussK06 tensiometer (Hamburg, Germany, Du Nouy ring detachment method) was used for measurements of the surface tension of the amphiphile/PAA binary systems. The volume of amphiphiles solutions was 10 mL. The sample was thermostated for 5 min at 25 ◦C before measurement. Detachment of the ring for each sample was repeated at least 3 times, and measurements with a deviation of ≤0.1 mN/m were taken into account.

3.3.2. Conductometry Specific conductivity was measured using an Inolab Cond 720 conductometer at 25 ◦C. The sample was thermostated for 5 min at 25 ◦C before measurement. Specific conductivity values with a deviation of no more than ±1 µSm/cm were taken into account.

3.3.3. Fluorescence Spectroscopy A survey of the emission fluorescence spectra of pyrene was performed using a Cary Eclipse G9800A fluorescence spectrophotometer (Agilent Technologies, Santa Clara, CA, USA) at 25 ◦C. The excitation wavelength was 335 nm. Emission spectra were registered in the range of 350–500 nm with a constant scanning speed (120 nm/min). A cuvette with a 1 cm width was used in all measurements. Fluorescence intensities of the first (II) and third (IIII) vibrational peaks of pyrene at 373 nm and 384 nm, respectively, were extracted from the collected spectra. Fluorescence spectra of the oligonucleotide–ethidium bromide complexes were recorded in the range from 500 to 700 nm at an excitation wavelength of 480 nm. A 0.8 mL sam- ple containing 0.5 µM of EB and a 10 µM concentration of oligonucleotide (counting on 1 nucleotide unit) in 4 mM Tris–HCl buffer (pH = 8.0) was equilibrated for 10 min at 25 ◦C. After that, a specific aliquot of the amphiphile solution was added to the mixture, and the fluorescence emission spectrum was registered [52,53]. Each concentration dependence was performed 2 times, and the standard deviation of the results was less than 2%. The Molecules 2021, 26, 2363 13 of 17

following equation was used for the quantitative evaluation of the binding degree of oNu to amphiphile: β = (Ibound − Iobs)/(Ibound − Ifree) (1)

where Ifree and Ibound are the intensities of fluorescence of free EB and probe bound to oNu, respectively; and Iobs is the fluorescence intensity observed during titration measurements at a specific amount of amphiphile added.

3.3.4. Dynamic and Electrophoretic Light Scattering Measurements of size and charge characteristics of samples were carried out using dynamic and electrophoretic light scattering techniques at Malvern ZetaSizer Nano ZS apparatus (Malvern Instruments Ltd., Malvern, UK). A source of irradiation was a He-Ne gas laser with 4 MW power (λoperating = 633 nm). Collected signals were treated in terms of frequency and phase analysis of scattered light using software attached to the device. All measurements were carried out at a 173◦ scattering angle. Computation of particle size was performed in accordance with the Stokes–Einstein equation [54]:

D = kT/(6πηR) (2)

where k is the Boltzmann’s constant, T is the absolute temperature, η is the solvent’s viscosity and R is the hydrodynamic radius. The zeta potential calculation was based on three measurements of electrophoretic mobilities using the Smoluchowski relationship:

ζ = µη/ε (3)

where ζ is the zeta potential, η is the dynamic viscosity of the solution, µ is the mobility of the particle and ε is the dielectric constant [55]. The hydrodynamic diameter and zeta potential of pDNA complexes with IA-12 and CTAB were measured on a ZetaSizer Nano ZS analyzer (Malvern Instruments) in a U- cuvette in 50 mM HEPES buffer (pH = 7). The final concentration of DNA and the N/P ratio were 10 µg/mL and 10:1, respectively, for both complexes.

3.3.5. Electrophoretic Analysis To obtain complexes, pDNA at a final concentration of 10 µg/mL was gently mixed with IA-12 (90–3000 µg/mL) or CTAB (5–160 µg/mL) in phosphate-buffered saline (PBS). The mixture was incubated at ambient temperature for 20 min. Electrophoretic analysis of pDNA–surfactant complexes was performed in 1% agarose gel in Tris-acetate–EDTA buffer using electrophoresis apparatus (Bio-Rad Laboratories, Hercules, CA, USA). Electrophore- sis conditions were detailed previously [46]. Briefly, pDNA (200 ng per well) was separated at a voltage of 8 V/cm for 60 min and then stained with ethidium bromide (0.5 µg/mL). The gels were visualized using a ChemiDoc™ XRS Plus gel documentation system (Bio- Rad Laboratories, Hercules, CA, USA). O’Gene Ruler DNA Ladder Mix (100–10,000 bp) (Fermentas, Waltham, MA, USA) was used.

3.3.6. Turbidimetry The phase behavior of amphiphile/PAA binary mixtures was studied using a Specord 250 PLUS spectrophotometer (Analytik Jena, Jena, Germany). An aliquot of polyelectrolyte solution was added to IA-12 solution of a specific concentration. The mixture was kept for 2 min, and the optical density of the solution was measured at a 500 nm wavelength. Phase transitions in amphiphile/DPPC mixtures were studied using a Specord 250 PLUS spectrophotometer (Analytik Jena, Jena, Germany). An optical density at a 350 nm wave- length as a function of the temperature was registered for the establishment of the lipid main phase transition in the IA-12/DPPC binary system. The temperature varied in the range from 35 to 45 ◦C. Primarily, a turbidimetric plot for individual DPPC liposomes (0.7 mM) was obtained. Further, an aliquot of 5 mM amphiphile solution was added to the Molecules 2021, 26, 2363 14 of 17

liposome dispersion, and an analogous turbidimetric plot was registered. These operations were repeated for several added aliquots of amphiphile solution. Turbidimetric plots were fitted in accordance with the van ’t Hoff two-state model, which assumes that the inflection point in the turbidimetric plot corresponds to the main phase transition of DPPC.

3.3.7. Fluorescence Microscopy Hoechst 33342 dye (0.1 mg/mL) was used as a contrast agent to localize the oligonu- cleotide. This dye binds to double-stranded DNA with a preference for domains enriched in adenine and thymine. Experiments were carried out using a Nikon Eclipse Ci-S fluores- cence microscope (Nikon, Tokyo, Japan) at a magnification of 1000× and by means of the multifunctional Cytell Cell Imaging system (GE Health Care Life Science, Umeå, Sweden) using Automated Imaging BioApp at a magnification of 1000×.

3.3.8. Flow Cytometry Assay M-HeLa cells at 1 × 106 cells/well were seeded in six-well plates. After 24 h of incubation, the IA-12/oNu complex was added to the test wells at a concentration of 0.5 µM. The uptake of the IA-12/oNu complex cells was analyzed using flow cytometry (Guava easyCyte 8HT, Luminex Corporation, Austin, TX, USA). Untreated cells were used as negative controls.

3.3.9. In Vitro Cytotoxicity Assay The evaluation of the toxicity of the IA-12 and IA-12/oNu systems was estimated by means of the multifunctional Cytell Cell Imaging system (GE Health Care Life Science, Sweden) using the Cell Viability Bio App. Cell lines M-HeLa (epithelial carcinoma of the cervix, HeLa subline, M-HeLa clone) were acquired from the Type Culture Collection of the Institute of Cytology (Russian Academy of Sciences, Saint Petersburg, Russia) and Chang liver normal cells from the N.F. Gamaleya Research Center of Epidemiology and Microbiology (Moscow, Russia). The cells were seeded in a concentration of 105 cells/mL in 96-well plates (Eppendorf, Hamburg, Germany), then 150 µL of standard Eagle’s medium (PanEco, Moscow, Russia) was added per well, and incubation proceeded with CO2 at 37 ◦C. The medium was then supplemented with 10% fetal calf serum and 1% nonessential amino acids. Twenty-four hours after seeding, the systems under study were added at preset dilutions, 150 µL to each well. The dilutions were prepared immediately in a nutrient medium. The experiments were repeated three times. Intact cells cultured in parallel with experimental cells were used as a control.

4. Conclusions In this study, complexation in mixed systems based on imidazolium amphiphile IA-12 and poly- or oligomeric anions of synthetic and natural origin was demonstrated. The addition of polyacrylic acid to a single surfactant system markedly reduces the aggregation thresholds of the system by ca. 200 times and leads to the electrostatically driven forma- tion of stable polymer–colloid complexes with a hydrodynamic diameter of 100–150 nm. Unlike synthetic polyelectrolytes, the complexation of IA-12 with oligonucleotide is not electrostatically mediated, as evidenced by weak charge neutralization demonstrated by the zeta potential study and gel electrophoresis experiments. At the same time, high IA- 12/oNu complexation efficacy is testified by ethidium bromide quenching data. Turbidity measurements demonstrated the membrane tropic ability of the imidazolium surfactant, after which the cellular uptake of the IA-12/oNu complex was proved by fluorescence microscopy and flow cytometry data. Selective cytotoxicity of the complexes was observed toward M-HeLa cells with IC50 = 1.1 µM, while a markedly higher value of IC50 = 23.1 µM was determined for Chang liver cells.

Supplementary Materials: The following are available online. Figure S1: Specific conductivity versus amphiphile concentration plot for individual IA-12 solutions and IA-12/PAA binary systems at fixed PAA concentrations; 25 ◦C. Figure S2: Intensity ratio of the first and third vibronic peaks of Molecules 2021, 26, 2363 15 of 17

pyrene vs. amphiphile concentration for the IA-12/PAA binary systems at fixed PAA concentrations; 25 ◦C. Figure S3: DLS correlation functions for the IA-12/PAA binary system at various amphiphile ◦ concentrations and a constant PAA concentration; CPAA = 1 mM; 25 C. Figure S4. DLS correlation functions for the IA-12/PAA binary system at various amphiphile concentrations and a constant PAA ◦ concentration; CPAA = 3 mM; 25 C. Figure S5: DLS correlation functions for the IA-12/PAA binary ◦ system at various amphiphile concentrations and a constant PAA concentration; CPAA = 5 mM; 25 C. Figure S6: Zeta potential versus IA-12 concentration plot for the IA-12/PAA binary systems; 25 ◦C (the linear X-axis). Figure S7: pH changes in IA-12/PAA aqueous solutions; 25 ◦C. Figure S8: Molecular weight marker O’Gene Ruler DNA Ladder Mix. Figure S9: Electrophoretic mobility of pDNA in complex with CTAB. Ctrl shows pure pDNA, N/P ratios are indicated above the corresponding wells. A DNA ladder from 100 to 10,000 bp was used. Figure S10: Intensity-averaged size distribution for the IA-12/pDNA system; N/P ratio of 10:1; 25 ◦C. Figure S11: Intensity-averaged size distribution for CTAB/pDNA system; N/P ratio of 10:1; 25 ◦C. Figure S12: Zeta potential for IA-12/pDNA (green) and CTAB/pDNA (red) complexes; N/P ratio of 10:1; 25 ◦C. Figure S13. EB/oNu emission fluorescence spectra in the presence of various amounts of IA-12 (the arrow indicates the direction of the increase in the amphiphile concentration); 25 ◦C. Figure S14: Typical turbidimetric plot for DPPC liposomes in the absence and in the presence of IA-12 (the amphiphile/lipid molar ratio is 1:5), CDPPC = 0.7 mM [56]. Author Contributions: Conceptualization, methodology, supervision, L.Y.Z.; experiments, D.A.K., D.M.K., S.S.L., V.M.Z., A.S.S., S.K.A., A.P.L., A.D.V. and D.V.S.; formal analysis, validation, data curation, visualization, writing—original draft preparation, D.A.K.; writing—review and editing, L.Y.Z. and D.R.G. All authors have read and agreed to the published version of the manuscript. Funding: This research was funded by a government assignment for the FRC Kazan Scientific Center of RAS. D.A.K., D.R.G., D.M.K., S.S.L., A.S.S., S.K.A., A.P.L., A.D.V., L.Y.Z. and D.V.S. acknowledge the Kazan Federal University Strategic Academic Leadership Program. Institutional Review Board Statement: Not applicable. Informed Consent Statement: Not applicable. Data Availability Statement: Data are contained within the article. Conflicts of Interest: The authors declare no conflict of interest. Sample Availability: Samples of the compounds are not available from the authors.

References 1. Varga, I.; Campbell, R.A. General physical description of the behavior of oppositely charged polyelectrolyte/surfactant mixtures at the air/water interface. Langmuir 2017, 33, 5915–5924. [CrossRef][PubMed] 2. Lele, B.J.; Tilton, R.D. Colloidal depletion and structural force synergism or antagonism in solutions of mutually repelling polyelectrolytes and ionic surfactants. Langmuir 2019, 35, 15937–15947. [CrossRef][PubMed] 3. Gradzielski, M.; Hoffmann, I. Polyelectrolyte-surfactant complexes (PESCs) composed of oppositely charged components. Curr. Opin. Colloid Interface Sci. 2018, 35, 124–141. [CrossRef] 4. Guzmán, E.; Fernández-Peña, L.; Ortega, F.; Rubio, R.G. Equilibrium and kinetically trapped aggregates in polyelectrolyte– oppositely charged surfactant mixtures. Curr. Opin. Colloid Interface Sci. 2020, 48, 91–108. [CrossRef] 5. Guzmán, E.; Llamas, S.; Maestro, A.; Fernández-Peña, L.; Akanno, A.; Miller, R.; Ortega, F.; Rubi, R.G. Polymer–surfactant systems in bulk and at fluid interfaces. Adv. Colloid Interface Sci. 2016, 233, 38–64. [CrossRef] 6. Kluˇcáková, M.; Jarábková, S.; Velcer, T.; Kalina, M.; Pekaˇr,M. Transport of a model diffusion probe in polyelectrolyte-surfactant hydrogels. Surf. A 2019, 573, 73–79. [CrossRef] 7. Vasilieva, E.A.; Samarkina, D.A.; Gaynanova, G.A.; Lukashenko, S.S.; Gabdrakhmanov, D.R.; Zakharov, V.M.; Vasileva, L.A.; Zakharova, L.Y. Self-assembly of the mixed systems based on cationic surfactants and different types of polyanions: The influence of structural and concentration factors. J. Mol. Liq. 2018, 272, 892–901. [CrossRef] 8. Tseng, H.-W.; Chen, P.-C.; Tsui, H.-W.; Wang, C.-H.; Hu, T.-Y.; Chen, L.-J. Effect of molecular weight of poly(acrylic acid) on the interaction of oppositely charged ionic surfactant–polyelectrolyte mixtures. J. Taiwan Inst. Chem. Eng. 2018, 92, 50–57. [CrossRef] 9. Asadov, Z.H.; Nasibova, S.M.; Ahmadova, G.A.; Zubkov, F.I.; Rahimov, R.A. Head-group effect of surfactants of cationic type in interaction with propoxylated sodium salt of polyacrylic acid in . Colloids Surf. A. 2017, 527, 95–100. [CrossRef] 10. Yan, P.; Jin, C.; Wang, C.; Ye, J.P.; Xiao, J.X. Effect of surfactant head group size on polyelectrolyte–surfactant interactions: Steady-state and time-resolved fluorescence study. J. Colloid Interface Sci. 2005, 282, 188–192. [CrossRef] 11. Langevin, D. Complexation of oppositely charged polyelectrolytes and surfactants in aqueous solutions. A review. Adv. Colloid Interface Sci. 2009, 147−148, 170–177. [CrossRef] Molecules 2021, 26, 2363 16 of 17

12. Wang, C.; Tam, K.C. Interaction between polyelectrolyte and oppositely charged surfactant: Effect of charge density. J. Phys. Chem. B. 2004, 108, 8976–8982. [CrossRef] 13. Schulze-Zachau, F.; Braunschweig, B. CnTAB/polystyrene sulfonate mixtures at air–water interfaces: Effects of alkyl chain length on surface activity and charging state. Phys. Chem. Chem. Phys. 2019, 21, 7847–7856. [CrossRef][PubMed] 14. Tripathy, S.K.; Kumar, J.; Nalwa, H.S. Handbook of Polyelectrolytes and Their Applications; American Scientific Publishers: Stevenson Ranch, CA, USA, 2002. 15. Ray, D.; Das, B. Micellization of ionic liquid surfactants induced by sodium polystyrenesulfonate in aqueous solutions. J. Solut. Chem. 2019, 48, 1576–1590. [CrossRef] 16. Koolivand-Salooki, M.; Javadi, A.; Bahramian, A.; Abdollahi, M. Dynamic interfacial properties and foamability of polyelectrolyte- surfactant mixtures. Colloids Surf. A 2019, 562, 345–353. [CrossRef] 17. Gabdrakhmanov, D.R.; Valeeva, F.G.; Samarkina, D.A.; Lukashenko, S.S.; Mirgorodskaya, A.B.; Zakharova, L.Y. The first representative of cationic amphiphiles bearing three unsaturated moieties: Self-assembly and interaction with polypeptide. Colloids Surf. A. 2018, 558, 463–469. [CrossRef] 18. Jain, N.; Trabelsi, S.; Guillot, S.; McLoughlin, D.; Langevin, D.; Letellier, P.; Turmine, M. Citical aggregation concentration in mixed solutions of anionic polyelectrolytes and cationic surfactants. Langmuir 2004, 20, 8496–8503. [CrossRef] 19. Taylor, D.J.F.; Thomas, R.K.; Li, P.X. Adsorption of oppositely charged polyelectrolyte/surfactant mixtures. neutron reflection from alkyl trimethylammonium bromides and sodium poly(styrenesulfonate) at the air/water interface: The effect of surfactant chain length. Langmuir 2003, 19, 3712–3719. [CrossRef] 20. Kuznetsova, D.A.; Gabdrakhmanov, D.R.; Vasilieva, E.A.; Lukashenko, S.S.; Ahtamyanova, L.R.; Siraev, I.S.; Zakharova, L.Y. Supramolecular catalytic systems based on a cationic amphiphile and sodium polystyrene sulfonate for decomposition of organophosphorus pollutants. Russ. J. Org. Chem. 2019, 55, 11–16. [CrossRef] 21. Zhang, Y.; Chan, H.F.; Leong, K.W. Advanced materials and processing for drug delivery: The past and the future. Adv. Drug Delivery Rev. 2013, 65, 104–120. [CrossRef][PubMed] 22. Kuznetsova, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Faizullin, D.A.; Zuev, Y.F.; Nizameev, I.R.; Kadirov, M.K.; Kuznetsov, D.M.; Zakharova, L.Y. Interaction of bovine serum albumin with cationic imidazoliumcontaining amphiphiles bearing urethane fragment: Effect of hydrophobic tail length. J. Mol. Liq. 2020, 307, 113001. [CrossRef] 23. Samarkina, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Nizameev, I.R.; Kadirov, M.K.; Zakharova, L.Y. Homologous series of amphiphiles bearing imidazolium head group: Complexation with bovine serum albumin. J. Mol. Liq. 2019, 275, 232–240. [CrossRef] 24. Kim, H.J.; Kim, A.; Miyata, K.; Kataoka, K. Recent progress in development of siRNA delivery vehicles for cancer therapy. Adv. Drug Deliv. Rev. 2016, 104, 61–77. [CrossRef][PubMed] 25. Wittrup, A.; Lieberman, J. Knocking down disease: A progress report on siRNA therapeutics. Nat. Rev. Genet. 2015, 16, 543–552. [CrossRef][PubMed] 26. Wettig, S.D.; Verrall, R.E.; Foldvari, M. Gemini surfactants: A new family of building blocks for non-viral gene delivery systems. Curr. Gene Ther. 2008, 8, 9–23. [CrossRef][PubMed] 27. Xu, L.; Feng, L.; Dong, R.; Hao, J.; Dong, S. Transfection efficiency of DNA enhanced by association with salt-free catanionic vesicles. Biomacromolecules 2013, 14, 2781–2789. [CrossRef][PubMed] 28. Andrzejewska, W.; Wilkowska, M.; Peplinska, B.; Skrzypczak, A.; Kozak, M. Structural characterization of transfection nanosys- tems based on tricationic surfactants and short double stranded oligonucleotides. Biochem. Biophys. Res. Commun. 2019, 518, 706–711. [CrossRef] 29. Szala, S. Terapia Genowa; Wydawnictwo Naukowe PWN: Warsaw, Poland, 2003. 30. Leitner, S.; Grijalvo, S.; Solans, C.; Eritja, R.; García-Celma, M.J.; Calderó, G. Ethylcellulose nanoparticles as a new “in vitro” transfection tool for antisense oligonucleotide delivery. Carbohydr. Polym. 2020, 229, 115451. [CrossRef] 31. López-López, M.; López-Cornejo, P.; Martín, V.I.; Ostos, F.J.; Checa-Rodríguez, C.; Prados-Carvajal, R.; Lebrón, J.A.; Huertas, P.; Moyá, M.L. Importance of hydrophobic interactions in the single-chained cationic surfactant-DNA complexation. J. Colloid Interface Sci. 2018, 521, 197–205. [CrossRef] 32. Limeresa, M.J.; Suñé-Poua, M.; Prieto-Sánchez, S.; Moreno-Castro, C.; Nusblat, A.D.; Hernández-Munain, C.; Castro, G.R.; Suñé, C.; Suñé-Negre, J.M.; Cuetsas, M.L. Development and characterization of an improved formulation of cholesteryl oleate-loaded cationic solid-lipid nanoparticles as an efficient non-viral gene delivery system. Colloids Surf. B. 2019, 184, 110533. [CrossRef] 33. Mashal, M.; Attia, N.; Soto-Sánchez, C.; Martínez-Navarrete, G.; Fernández, E.; Puras, G.; Pedraz, J.L. Non-viral vectors based on cationic niosomes as efficient gene delivery vehicles to central nervous system cells into the brain. Int. J. Pharm. 2018, 552, 48–55. [CrossRef] 34. Zhang, J.; Wang, Z.; Lin, W.; Chen, S. Gene transfection in complex media using PCBMAEE-PCBMA copolymer with both hydrolytic and zwitterionic blocks. Biomaterials 2014, 35, 7909–7918. [CrossRef] 35. Soni, S.; Sarkar, S.; Mirzadeh, N.; Selvakannan, P.; Bhargava, S. Self-assembled functional nanostructure of plasmid DNA with ionic liquid [Bmim][PF6]: Enhanced efficiency in bacterial gene transformation. Langmuir 2015, 31, 4722–4732. [CrossRef] 36. Huang, Q.; Ou, W.; Chen, H.; Feng, Z.; Wang, J.-Y.; Zhang, J.; Zhu, W.; Yu, X.-Q. Novel cationic lipids possessing protonated cyclen and imidazolium salt for gene delivery. Eur. J. Pharm. Biopharm. 2011, 78, 326–335. [CrossRef][PubMed] Molecules 2021, 26, 2363 17 of 17

37. Zhou, T.; Llizo, A.; Li, P.; Wang, C.; Guo, Y.; Ao, M.; Bai, L. High transfection efficiency of homogeneous DNA nanoparticles induced by imidazolium gemini surfactant as nonviral vector. J. Phys. Chem. C 2013, 117, 26573–26581. [CrossRef] 38. Zhou, T.; Xu, G.; Ao, M.; Yang, Y.; Wang, C. DNA compaction to multi-molecular DNA condensation induced by cationic imidazolium gemini surfactants. Colloids Surf. A. 2012, 414, 33–40. [CrossRef] 39. Manning, G.S. Limiting Laws and Counterion Condensation in Polyelectrolyte Solutions I. Colligative Properties. J. Chem. Phys. 1969, 51, 924–938. [CrossRef] 40. Yoshida, K.; Dubin, P.L. Complex formation between polyacrylic acid and cationic:nonionic mixed micelles: Effect of pH on electrostatic interaction and hydrogen bonding. Colloids Surf. A. 1999, 147, 161–167. [CrossRef] 41. Vasilieva, E.A.; Lukashenko, S.S.; Vasileva, L.A.; Pavlov, R.V.; Gaynanova, G.A.; Zakharova, L.Y. Aggregation behavior of the surfactant bearing pyrrolidinium head group in the presence of polyacrylic acid. Russ. Chem. Bull. 2019, 68, 341–346. [CrossRef] 42. Vasilieva, E.A.; Ibragimova, A.R.; Lukashenko, S.S.; Konovalov, A.I.; Zakharova, L.Y. Mixed self-assembly of polyacrylic acid and oppositely charged gemini surfactants differing in the structure of head group. Fluid Phase Equilib. 2014, 376, 172–180. [CrossRef] 43. Samarkina, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Khamatgalimov, A.R.; Zakharova, L.Y. Aggregation capacity and complexation properties of a system based on an imidazole-containing amphiphile and bovine serum albumin. Russ. J. Gen. Chem. 2017, 87, 2826–2831. [CrossRef] 44. Hunt, K.K.; Vorburger, S.A.; Swisher, S.G. Gene Therapy for Cancer; Humana Press: Totowa, NJ, USA, 2007. 45. Samarkina, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Khamatgalimov, A.R.; Kovalenko, V.I.; Zakharova, L.Y. Cationic amphiphiles bearing imidazole fragment: From aggregation properties to potential in biotechnologies. Colloids Surf. A. 2017, 529, 990–997. [CrossRef] 46. Salakhieva, D.; Shevchenko, V.; Németh, C.; Gyarmati, B.; Szilágyi, A.; Abdullin, T. Structure-biocompatibility and transfection activity relationships of cationic polyaspartamides with (dialkylamino)alkyl and alkyl or hydroxyalkyl side groups. Int. J. Pharm. 2017, 517, 234–246. [CrossRef] 47. Jumbri, K.; Ahmad, H.; Abdulmalek, E.; Basyaruddin, M.; Rahman, A. Binding energy and biophysical properties of ionic liquid-DNA complex: Understanding the role of hydrophobic interactions. J. Mol. Liq. 2016, 223, 1197–1203. [CrossRef] 48. Grigoriev, I.V.; Korobeynikov, V.A.; Cheresiz, S.V.; Pokrovskiy, A.G.; Zakharova, L.Y.; Voronin, M.A.; Lukashenko, S.S.; Konovalov, A.I.; Zuev, Y.F. Cationic gemini surfactants as new agents for plasmid DNA delivery into cells. Dokl. Biochem. Biophys. 2012, 445, 197–199. [CrossRef][PubMed] 49. Kuznetsova, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Voloshina, A.D.; Sapunova, A.S.; Kashapov, R.R.; Zakharova, L.Y. Self-assembled systems based on novel hydroxyethylated imidazolium-containing amphiphiles: Interaction with DNA decamer, and lipid. Chem. Phys. Lipids. 2019, 223, 104791. [CrossRef] 50. Gabdrakhmanov, D.R.; Vasilieva, E.A.; Voronin, M.A.; Kuznetsova, D.A.; Valeeva, F.G.; Mirgorodskaya, A.B.; Lukashenko, S.S.; Zakharov, V.M.; Mukhitov, A.R.; Faizullin, D.A.; et al. Soft nanocontainers based on hydroxyethylated geminis: Role of spacer in self-assembling, solubilization, and complexation with oligonucleotide. J. Phys. Chem. C 2020, 124, 2178–2192. [CrossRef] 51. Kuznetsova, D.A.; Gabdrakhmanov, D.R.; Lukashenko, S.S.; Ahtamyanova, L.R.; Nizameev, I.R.; Kadirov, M.K.; Zakharova, L.Y. Novel hybrid liposomal formulations based on imidazolium-containing amphiphiles for drug encapsulation. Colloids Surf. B 2019, 178, 352–357. [CrossRef][PubMed] 52. Zakharova, L.; Voronin, M.; Semenov, V.; Gabdrakhmanov, D.; Syakaev, V.; Gogolev, Y.; Giniyatullin, R.; Lukashenko, S.; Reznik, V.; Latypov, S.; et al. Supramolecular systems based on novel mono- and dicationic pyrimidinic amphiphiles and oligonucleotides: A self-organization and complexation study. Chem. Phys. Chem. 2012, 13, 788–796. [CrossRef][PubMed] 53. Gabdrakhmanov, D.; Samarkina, D.; Semenov, V.; Syakaev, V.; Giniyatullin, R.; Gogoleva, N.; Zakharova, L. Novel dicationic pyrimidinic surfactant: Self-assembly and DNA complexation. Colloids Surf. A 2015, 480, 113–121. [CrossRef] 54. Kuznetsova, D.A.; Gaynanova, G.A.; Vasileva, L.A.; Sibgatullina, G.V.; Samigullin, D.V.; Sapunova, A.S.; Voloshina, A.D.; Galkina, I.V.; Petrov, K.A.; Zakharova, L.Y. Mitochondria-targeted cationic liposomes modified with alkyltriphenylphosphonium bromides loaded with hydrophilic drugs: Preparation, cytotoxicity and colocalization assay. J. Mater. Chem. B 2019, 7, 7351–7362. [CrossRef] [PubMed] 55. Von Smoluchowski, M. Contribution to the theory of electro-osmosis and related phenomena. Bull. Int. Acad. Sci. Cracovie 1903, 3, 184–199. 56. Aguiar, J.; Carpena, P.; Molina-Bolívar, J.A.; Carnero Ruiz, C. On the determination of the critical micelle concentration by the pyrene 1:3 ratio method. J. Colloid Interface Sci. 2003, 258, 116–122. [CrossRef]