Next Article in Journal
Fluoride Exposure and the Effect of Tobacco Smoking on Urinary Fluoride Levels in Primary Aluminum Workers
Next Article in Special Issue
Nanostructured Black Nickel Coating as Replacement for Black Cr(VI) Finish
Previous Article in Journal
Morphological Study of Bacillus thuringiensis Crystals and Spores
Previous Article in Special Issue
Can Encapsulation of the Biocide DCOIT Affect the Anti-Fouling Efficacy and Toxicity on Tropical Bivalves?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Silver Nanoparticles with Gemini Surfactants as Efficient Capping and Stabilizing Agents

Department of Bioactive Products, Faculty of Chemistry, Adam Mickiewicz University Poznan, 61-614 Poznan, Poland
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(1), 154; https://doi.org/10.3390/app11010154
Submission received: 27 November 2020 / Revised: 14 December 2020 / Accepted: 23 December 2020 / Published: 26 December 2020

Abstract

:
The scientific community has paid special attention to silver nanoparticles (AgNPs) in recent years due to their huge technological capacities, particularly in biomedical applications, such as antimicrobials, drug-delivery carriers, device coatings, imaging probes, diagnostic, and optoelectronic platforms. The most popular method of obtaining silver nanoparticles as a colloidal dispersion in aqueous solution is chemical reduction. The choice of the capping agent is particularly important in order to obtain the desired size distribution, shape, and dispersion rate of AgNPs. Gemini alkylammonium salts are named as multifunctional surfactants, and possess a wide variety of applications, which include their use as capping agents for metal nanoparticles synthesis. Because of the high antimicrobial activity of gemini surfactants, AgNPs stabilized by this kind of surfactant may possess unique and strengthened biocidal properties. The present paper presents the synthesis of AgNPs stabilized by gemini surfactants with hexadecyl substituent and variable structure of spacer, obtained via ecofriendly synthesis. UV-Vis spectroscopy and dynamic light scattering were used as analyzing tools in order to confirm physicochemical characterization of the AgNPs (characteristic UV-Vis bands, hydrodynamic diameter of NPs, polydispersity index (PDI)).

1. Introduction

Over the past twenty years, silver nanoparticles (AgNPs) have become increasingly popular due to their special physical, chemical, and biological properties. It is characteristic that, depending on the surface to volume ratio, the properties of nanosilver change, and they have been exploited for different purposes [1,2]. AgNPs offer promising prospects for a wide range of applications, such as antimicrobials against bacteria [3,4,5,6,7,8,9,10,11,12], fungi [10,13,14,15,16,17], and viruses [9,18,19,20,21,22,23]. Silver nanoparticles can also be helpful in managing the ongoing pandemic of COVID-19 (Coronavrus Disease 2019) caused by the SARS-CoV-2 (severe acute respiratory syndrome coronavirus 2) [24,25]. AgNPs have been tested as biomedical device coatings [12,26,27,28], combating multidrug-resistant cancer [12,29], drug-delivery carriers [12,30], and imaging probes in ultrasensitive analysis [26,31,32,33,34]. They may also find applications in sensing [27,31,35,36] and chemical catalysis [32,35,37,38,39]. Due to the great interest of AgNPs, different synthetic methods have been developed, such as physical, chemical, and biological approaches [26,40,41]. Among them, chemical reduction is the most frequently used synthesis method of silver nanoparticles, with high productivity and low costs [39,41]. In this approach, stable colloids in water or non-aqueous solvents can be obtained. Generally, synthesis of AgNPs in solution requires three particular components: metal precursors (silver nitrate, perchlorate, chloride, trifluoroacetate or (PPh3)3AgNO3), reducers, and stabilizers [42]. Several reductants are commonly used for the preparation of the AgNPs: sodium borohydride, sodium citrate, glucose, N,N-dimethylformamide, hydrazine, polyols, (e.g., ethylene glycol, diethylene glycol), formaldehyde [42]. One important factor is to apply capping agents to stabilize dispersive NPs, thus protect them against their agglomeration [41]. In general, the shape and size distribution of the synthesized nanoparticles are controlled by altering the method and conditions of the synthesis, reducing and stabilizing agents, their concentrations, and molar ratios [43]. Recently, different types of surfactants started to be used as stabilizing agents for AgNPs. We should mention that a proper stabilizer and/or reaction time is crucial for the obtaining of stable and small size Ag nanoparticles [44].
Dimeric quaternary alkylammonium salts, named by Menger and Littau as gemini, are modern type of surfactants [45]. These compounds consist of two long alkyl substituents and two cationic headgroups connected by a linker. The connector may have various structures: short, long, flexible, stiff, polar (e.g., functionalized by heteroatoms), and nonpolar (e.g., hydrocarbon chains or rings) [46,47,48,49,50]. Spacer and hydrophobic chain are very important units in gemini surfactant architecture. They are a keystone in the adsorption of dimeric surfactants on surfaces and interfaces, in creation of micelles of different shapes [48,49,51,52,53,54]. Considering the structure of gemini surfactants, they exhibit a higher efficacy in contrast to monomeric (conventional) surfactants, such as decreased critical micelle concentration (CMC), greater ability to reduce surface tension, or promote antimicrobial activity [48,50,55,56,57]. From an ecological point of view, dimeric alkylammonium salts are considerably less toxic to marine organisms than quaternary ammonium salts [58,59], and can be biodegradable [48,60,61,62]. Considering the above properties, and the fact that the use of gemini surfactant provides the desired effect at a lower concentration in comparison to monomeric analogues, the use of dimeric alkylammonium salts is correlated with the greenolution idea in chemistry [63,64]. Gemini surfactants, due to their antimicrobial properties against bacteria [48,55,56,65,66,67,68,69,70] and fungi [47,48,56,66,71], can be used as microbicides or biofilm eradication agents in many areas of life and industries [66,72,73,74].
In reference to the growing interest in silver nanoparticles and gemini surfactants, we focused on the different procedures for obtaining silver nanoparticles capped with gemini surfactants, having a different structure in spacer unit, obtained in easy, low-cost, and green synthesis. Detailed spectroscopic analysis and surface activity of the gemini surfactants, as well as analysis of AgNPs based on a UV-Vis, and dynamic light scattering (DLS) study were presented.

2. Materials and Methods

2.1. Material Used

N,N-Dimethyl-N-hexadecylamine, bis(2-bromoethyl)ether, 1,6-dibromohexane, silver nitrate, and sodium borohydride, were obtained from Sigma-Aldrich (Poznan, Poland). Acetonitrile and diphosphorus pentoxide were purchased from VWR Chemicals (Gdansk, Poland). Reagent purity was at least 95%. All materials were used without prior purification.

2.2. Synthesis

2.2.1. Synthesis of Gemini Surfactants

The 16-6-16 (1,6-hexamethylene-bis(N-hexadecyl-N,N-dimethylammonium bromide)): the N,N-dimethyl-N-hexadecylamine (10 g; 37 mmol) and 1,6-dibromohexane (4.5 g; 18.8 mmol) were placed in a 250 mL round-bottom flask. Reaction mixture was stirred without the solvent at room temperature for 2 h. Then, the reaction mixture was left to solidify completely. The crude product was purified by crystallization from acetonitrile, filtered, and dried in a vacuum desiccator over P2O5 and in an oven at 60 oC. Reaction yield was over 98%.
The 16-O-16 (3-oxa-1,5-pentane-bis(N-hexadecyl-N,N-dimethylammonium bromide)): the N,N-dimethyl-N-hexadecylamine (10 g; 37 mmol) and bis(2-bromoethyl)ether (4.36 g; 18.5 mmol) were placed in a 250 mL round-bottom flask. Reaction mixture was stirred without the solvent at room temperature for 2 h. Then, the reaction mixture was left to solidify completely. The crude product was purified by crystallization from acetonitrile, filtered, and dried in a vacuum desiccator over P2O5 and in an oven at 60 oC. Reaction yield was over 95%.

2.2.2. Synthesis of Silver Nanoparticles

Experiments 1–6: 40 mL of 0.75 mM silver nitrate was mixed with 10 mL of specific amount of gemini surfactant (16-6-16 or 16-O-16) corresponding to three different molar ratio nAg/nGemini = 2.5; 5; 10. The mixture was stirred for 5 min, and then 5 mL of 1 mM sodium borohydride cold solution was added dropwise. Color changed to intense orange. After adding all of the solution, stirring occurred for another 20 min.
Experiment 7: 15 mL of 4 mM NaBH4 and 10 mL of 0.2 mM 16-6-16 solutions were placed in a 100 mL Erlenmeyer flask and stirred for 5 min. Then 5 mL of 2 mM AgNO3 was added dropwise to the flask. Stirring continued for another 20 min.
Experiment 8: 5 mL of 12 mM NaBH4 was placed in a 100 mL Erlenmeyer flask with 20 mL of water. Solution of 10 mL of 1 mM AgNO3 and 10 mL of 0.2 mM 16-6-16 was prepared separately and added dropwise into NaBH4. After adding all of the solution, stirring occurred for another 20 min.
All of the experiments were repeated three times in order to confirm the reproducibility of the synthetic procedures. The AgNPs stabilized by gemini surfactants were stable for at least three months, during which no changes in the UV-vis and DLS were observed.

2.3. Experimental Methods

2.3.1. Analysis of Dimeric Alkylammonium Salts

The melting point of the synthesized gemini surfactants was measured on Stuart SMP30 apparatus (Staffordshire, UK), using capillary with one sealed side. The measurements give the values with resolution of 1.0 °C. The elemental analysis measurements were performed on a FLASH 2000 elemental analyzer (Delft, The Netherlands) with a thermal conductivity detector. The nuclear magnetic resonance (NMR) spectra were measured with a Varian VNMR-S 400 MHz (Oxford, UK), operating at 403 and 101 MHz for 1H and 13C respectively, attendant with software Vnmr VERSION 2.3 REVISION A (Varian, Oxford, UK). The chemical shifts were performed in CDCl3 relative to an internal standard—TMS (tetramethylsilane). The Fourier Transform Infrared Spectroscopy (FTIR) spectra were performed on FT-IR Bruker IFS 66v/S (Poznan, Poland) apparatus. All of the samples were tested in solid state in the form of tablets with potassium bromide.
The critical micelle concentration values were determined by conductivity analysis using a Conductivity Meter Elmetron CC-505 (Zabrze, Poland). The apparatus was calibrated by a standard (147 μS/cm in 298.15 K). All samples were obtained using deionized water. Conductivity measurements were carried out at 25 °C. The titration was repeated at least three times for each compound.

2.3.2. Analysis of Silver Nanoparticles

UV-Vis absorption spectra of all the AgNPs solutions capped with dimeric ammonium salts were recorded on Varian Cary 50 Scan UV-Vis spectrophotometer (Varian, Oxford, UK) at 25 °C, operated with Cary WinUV software (Varian, Oxford, UK) with quartz cuvettes of 1 cm path length. In all cases, samples were diluted 5-fold with deionized water, to decrease the absorbance to the range suitable for UV-Vis measurements.
Dynamic light scattering measurements and polydispersity index (PDI) were performed at 25 °C using Zetasizer, Nano-ZS Malvern Instruments (Malvern, UK) with a He-Ne laser (633 nm, 4 mW) equipped with a built-in termo-controller. Ten repeated measurements were conducted for each sample.

3. Results and Discussion

3.1. Synthesis and Spectroscopic Characterization of Gemini Surfactants

Gemini surfactants tested in this paper as capping agents for AgNPs were received by alkylation of N,N-dimethyl-N-hexadecylamine with 1,6-dibromohexane or bis(2-bromoethyl)ether, for 16-6-16, and 16-O-16 respectively (Figure 1). A synthetic procedure developed in our laboratory assumes carrying out the reaction without a solvent at room temperature [50,75]. These reactions proceed according to the nucleophilic substitution mechanism SN2. The rate of these reactions depends on the substrates concentrations. In case of the reaction without a solvent, the concentration of the reactants is the highest possible. Such reactions take place with high efficiency without waste. These reactions are in accordance with the greenolution approach, because they reduce the use of solvent to a minimum and minimize costs.
The structure and purity of gemini surfactants obtained were confirmed by 1H NMR, 13C NMR, FTIR, and elemental analysis.
The 16-6-16: mp 222–223 °C; elemental analysis (%) calcd. for C42H90Br2N2: C 64.43, H 11.58, N 3.58; found C 64.72, H 11.68, N 3.40; 1H NMR (403 MHz, CDCl3) 3.73 (bs, 4H, N+CH2(spacer)), 3.49 (m, 4H, -N+CH2-), 3.39 (s, 12H, -N+(CH3)2), 2,01 (bs, 4H, -N+CH2CH2-(spacer)), 1.72 (bs, 4H, -N+CH2CH2-),1.57 (bs, 4H, -N+CH2CH2-CH2(spacer)), 1.36-1.25 (m, 52H, -CH2(CH2)13CH3), 0.88 (t, 6H, -CH2CH3); 13C NMR (101 MHz, CDCl3): 65.9 (-N+CH2spacer), 64.06 (-N+CH2-), 50.98 (-N+CH3), 31.89 (-CH2CH2CH3), 29.61, 29.59, 29.56, 29.34, 29.27, 29.45, 29.37, 29.12 (-N+(CH2)2(CH2)11-), 26.86 (-N+CH2CH2-), 24.37 –(N+CH2CH2-(spacer)), 22.6 (-N+CH2CH2-CH2(spacer)), 21.61 (-CH2CH3), 14.11 (-CH2CH3).
The 16-O-16: mp 240-242 oC; elemental analysis (%) calcd. for C40H86Br2N2O: C 62.32, H 11.24, N 3.63; found C 61.41, H 11.68, N 3.20; 1H NMR (403 MHz, CDCl3) 4.36 (bs, 4H, ­OCH2), 4.05 (bs, 4H, N+CH2CH2O-), 3.67-3.62 (m, 4H, -N+CH2-), 3.46 (s, 12H, -N+(CH3)2), 1.73 (bs, 4H, -N+CH2CH2-), 1.47-1.10 (m, 52H, -CH2(CH2)13CH3), 0.88 (t, 6H, -CH2CH3); 13C NMR (101 MHz, CDCl3): 65.92(-N+CH2CH2O-), 64.59 (-OCH2-), 64.03 (-N+CH2-), 51.57 (-N+CH3), 31.82 (-CH2CH2CH3), 29.61, 29.58, 29.56, 29.54, 29.45, 29.37, 29.25, 26.25 (-N+(CH2)2(CH2)11-), 22.83 (-N+CH2CH2-), 22.58 (-CH2CH3), 14.00 (-CH2CH3).
FTIR spectra for tested dimeric alkylammonium salts (Figure 2) show regular bands for the asymmetric (νas) and symmetric (νs) stretching vibrations of methyl and methylene groups at 2916–2846 cm−1, as well as bands for the deformation vibration (δ) of methyl groups at 1472–1464 cm−1. The 16-O-16 spectra shows usual peaks of C-O stretching vibration at 1146 cm−1. Due to hydrophilicity (ability to absorb water) of gemini surfactants the strong absorption at the 3500–3350 cm−1 region from the -OH stretching vibrations of water molecule is detected in FTIR spectra.

3.2. Surface Properties of Gemini Surfactants

The basic ability of surface active agents is their tendency to be adsorbed at interfaces. Mechanism of surface action of dimeric surfactants based on the adsorption of ammonium cations into a polar phase, and hydrocarbon chains in a nonpolar phase [48]. The keystone of all surfactants research is determination of their critical micelle concentration [49], the lowest concentration at which particles rapidly aggregate into micelles. The fact that dimeric alkylammonium salts have much lower CMC values than monomeric ones, creates great application possibilities for them [52,63,76,77,78,79]. The determination of CMC is also important, because toxicity of gemini rises when their concentration surpasses CMC [80].
The CMC was determined using a conductometric titration, creating dependency graphs of the characteristic conductivity in water of the obtained compounds as a function of the concentration. This relationship generates two lines with different slopes. The straight line with higher inclination illustrates behavior before micellization, whereas the line with a smaller slope indicates the process of micellization (Figure 3). The intersection of lines formed as a result of linear regression defines CMC [50,59].
The values for the slope ratios of linear regression in pre- and post-micellization regions allow estimating the degree of ionization (α) and the degree of counterion binding parameter (β). These parameters demonstrate the capability of the counterion binding on the micelles [50,59]. Knowing the values of CMC and β allow for calculation of Gibbs free energy of micellization (ΔGomic) [81]. The experimental values of CMC, α, β and ΔGomic for the investigated gemini surfactants are given in Table 1.
Critical micelle concentration of dimeric surfactants depends on many structural elements: length and structure of spacer and substituents and type of anion. It is well known that gemini surfactants with long (hexadecyl, octadecyl) substituents possess lower CMC values than analogues with shorter hydrocarbon chain [48,49,50,77,82,83,84,85,86]. CMCs of dimeric alkylammonium salts tested in this work are over thirty times lower, than CMC of 12-6-12 (1,6-hexamethylene-bis(N-hexadecyl-N,N-dimethylammonium bromide)) which is 0.98 (mM) [57]. Surface activity of our gemini surfactants, although they have different spacer structure, is comparable. This result arein good correlation to our previous result—that an element of the structure, which essentially affects the surface activity, is the length of the substituent, not the structure of the linker [50].
The 16-O-16 possesses a lower degree of ionization value than 16-6-16. The small α is caused by the stronger binding of the counterion to the aggregates, suggesting better packing of the head groups and higher surface charge density at the interface [50,51]. Both obtained gemini surfactants have negative ΔGomic, indicating that the micellization process is spontaneous [50,55]. Surfactant with oxygen-functionalized spacer (16-O-16) shows greater tendency to form micelles, because in this case ΔGomic is lower.

3.3. Preparation and Characterization of Silver Nanoparticles

AgNPs were prepared by chemical synthetic procedure under different conditions. We used: silver salt (AgNO3) as a metal precursor, sodium borohydride (NaBH4) as a reductant and gemini surfactants (16-6-16 in experiments 1–3 and 7–8, 16-O-16 in experiments 4–6) as stabilizing agents (Figure 4). Experiments 1–6 were performed according to preparation described by Pisárčik [87]. Different molar ratios of AgNO3 and gemini surfactants were applied to study the formation of AgNPs (Table 2). In these cases, nAgNO3 was taken in excess of 6-fold to nNaBH4. Experiments 7 and 8 were carried out by another procedure, with 16-6-16 as stabilizing agent in silver-gemini surfactants molar ratio 5 (Table 2). In these cases, nNaBH4 was taken in excess of 6-fold to nAgNO3.
The color of AgNPs solution depends on the degree of dilution of the colloids, starting from yellow and going to intense orange with the increase of colloids concentration. Gemini surfactant allows for keeping the nanodispersion homogeneous, with a color depending on the nanodispersion dilution. In our work, in all experiments, the intense orange color was formed, showing that the reaction was followed by the formation of nanoparticles [87,88,89,90]. The color of AgNPs solution is suggested to appear due to localized surface plasmon resonance (LSPRs). LSPR is a collective excitation of the free electrons in the conduction band around the surface of the nanoparticles. The electrons are specified to particular vibration modes, which depend on particle size and shape. Therefore, AgNPs could be detected and characterized by UV-Vis. With increasing of particle size, the absorption wavelength shifts to longer wavelengths [91].
UV-Vis spectroscopic analysis is an essential technique for observation of AgNPs formation and stability in solution. The absorbance peak of silver nanoparticles usually occurs at wavelength range of 350–450 nm, and moves to longer wavelengths with progressive particle size [91]. In Figure 5a, the UV-Vis spectra of AgNPs, which were stabilized with 16-6-16 is presented. The spectra were measured for three various Ag-to-gemini surfactant molar ratios: nAg/n16-6-16 = 2.5; 5; 10. We can see that from the highest gemini concentration (experiment 1) to the lowest (experiment 3), the peak of absorbance increases gradually: 0.286, 0.332 and 0.359. As it is known, the absorbance is directly connected with the concentration, thus, since the difference is not significant, it can be concluded that 16-6-16 has a strong stabilizing effect on AgNPs in the extensive range of nAg/n16-6-16 molar ratio values. However, the nAg/n16-6-16 molar ratio at 2.5 indicates a slightly weak stabilizing effect since the concentration of AgNPs is not high.
Figure 5b shows the UV-Vis absorption spectra of nanoparticles stabilized with 16-O-16. These AgNPs were synthesized similar to the previous ones, but 16-O-16 was taken as capping agent (experiment 4–6). We observe that the peak of absorbance gradually increases with the decrease of gemini surfactant concentration as well, as in the case of experiments 1–3. The peaks of absorbance for experiments 4, 5, and 6 are 0.185, 0.313, and 0.397, respectively. The stabilizing effect of 16-O-16 on AgNPs is stronger for molar ratios of nAg/n16-O-16 = 5 and 10, whereas it has weaker effect for reaction with nAg/n16-O-16 molar ratio of 2.5. Same trend was observed for of 16-6-16.
UV–Vis absorption spectra of AgNPs obtained in experiments 7 and 8 are shown in Figure 6. Strong absorption peaks at approximately 410 and 419 nm come from the surface plasmon absorption of nanosized silver particles. These spectra stand out; the good symmetric absorption peaks with a nearly unaltered width, suggesting that the size of the nanoparticles is very homogenous [89,92].
All UV–Visible absorption spectra exhibit the absorbance peaks in the range of 410–435 nm (Table 3), which is typical feature for colloidal silver nanoparticles [93], confirming our statement that gemini surfactants in the Ag nanodispersion ensures a stabilizing effect on the nanoparticles. These results are in good correlation with what was described previously [87,94,95].
Consequently, the colloidal stability is also confirmed by the polydispersity index (PDI), which is given in DLS analysis and should be less or about 0.3. If the PDI is equal or higher than 1, the observable precipitation may occur [96]. The PDI values were between 0.2 and 0.4 in experiments 1–3 and 7–8, illustrating good colloidal features, and PDI values around 0.5 were found to be for the experiments 4–6. Higher PDI values in experiments 4–6 supports relatively broad and slightly distorted width of the UV-Vis spectra (Figure 5b) as well as the visual minor nanodispersions, which were of white color.

3.4. Effect of Spacer Structure on the Size of Silver Nanoparticles

The size of AgNPs capped with gemini and their particle size distributions were obtained using DLS measurements. The Figure 7 shows hydrodynamic diameter of nanoparticles stabilized with gemini surfactants16-6-16 and 16-O-16, which is plotted as a function of the nAg/nGemini molar ratio. The nanoparticle diameter values illustrate an average value of the particle size distribution.
For experiments 4–6 we can observe, that with decreasing of 16-O-16 concentration the silver nanoparticles diameter size increase. However, in the case of 16-6-16 (experiments 1–3) the opposite trend is noticed—we monitor the almost constant diameter size in the range of all molar ratios. Construction of the spacer of gemini surfactants have a considerable impact on nanoparticles size. Compounds with more hydrophobic linker stabilizes more efficiently and gives smaller AgNPs than analog with spacer functionalized by ether group.
Hydrodynamic diameter of silver nanoparticles, stabilized with 16-6-16 from synthetic procedure with excess of reductor, are 110 and 220 nm for experiments 7 and 8, respectively. These results clearly show that synthetic procedure applied in experiment 7 gives the smallest AgNPs with the best stabilizing effect of gemini bilayer. Figure 8 represents the particle size distribution histograms of AgNPs synthesized in the experiments 7 and 8.
In fact, the particle sizes of AgNPs measured by DLS is the whole conjugate size, with countslayer of capping agents too [88]. This explains the relatively big diameters. He et al. explained that the gemini cationic surfactants can be adsorbed on the particle surface due to electrostatic interactions between surfactant positive charge head group and counterions attached onto the nanoparticle surface. This causes the formation of coating layers, which lowers the aggregation of particles [90]. Pisárčik et al. demonstrates that gemini surfactants with dodecyl substituent and different spacer length from two to twelve carbon atoms are able to efficiently stabilize silver nanoparticles [87]. Our results indicate that gemini surfactants with longer substituent can also be applied as efficient stabilizer for AgNPs. Introduction of oxygen atom to spacer results in an increase of hydrophilicity of surfactant molecule. This caused increasing of formed silver nanoparticles size, indicating that the stabilizing effect of gemini bilayer is less efficient. We also demonstrate that, not only the kind of stabilizing agent is crucial, but also synthetic procedures and silver to gemini surfactant molar ratios are very important factors in AgNPs synthesis. These results are in good correlation with what was described previously [87].

4. Conclusions

The 1,6-hexamethylene-bis(N-hexadecyl-N,N-dimethylammonium bromide) and 3-oxa-1,5-pentane-bis(N-hexadecyl-N,N-dimethylammonium) bromide) were obtained via solvent-free green synthetic procedure. The products structures have been confirmed by 1H and 13C NMR, FITR, and elemental analysis. The critical micelle concentrations were specified by conductivity measurement. Silver nanoparticles were obtained via chemical reduction, using gemini surfactants as stabilizing agent. Physicochemical investigations were done by using UV-Vis spectroscopy and DLS measurements. We demonstrated that the structure of the gemini surfactant and conditions of the silver nanoparticles preparation determine their shape and size distribution. It was shown that the hydrophilicity of the dimeric alkylammonium salts can affect the size distribution of AgNPs. The best stabilization, which is represented by a small nanoparticle size, was received in experiment with 16-6-16 as a stabilizing agent, molar ratio nAg/nGemini =5, and with excess of reductant. Results of our research show that application of gemini surfactants is a prospective field of study, which may provide an easy-to-follow green method for the synthesis of nanoparticles.

Author Contributions

Conceptualization, B.B. and A.S.; methodology, A.S. and M.B.; software, M.B.; validation, B.B., A.S., and M.B.; formal analysis, M.B.; investigation, A.S.; resources, B.B.; data curation, A.S.; writing—original draft preparation, A.S.; writing—review and editing, B.B. and M.B.; visualization, M.B.; supervision, B.B.; project administration, B.B.; funding acquisition, B.B. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was funded from the Faculty of Chemistry, Adam Mickiewicz University, subvention for maintaining research potential (B.B.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharma, V.K.; Yngard, R.A.; Lin, Y. Silver nanoparticles: Green synthesis and their antimicrobial activities. Adv. Colloid Interface Sci. 2009, 145, 83–96. [Google Scholar] [CrossRef] [PubMed]
  2. Helmlinger, J.; Sengstock, C.; Groß-Heitfeld, C.; Mayer, C.; Schildhauer, T.A.; Köller, M.; Epple, M. Silver nanoparticles with different size and shape: Equal cytotoxicity, but different antibacterial effects. RSC Adv. 2016, 6, 18490–18501. [Google Scholar] [CrossRef] [Green Version]
  3. Ulubayram, K.; Calamak, S.; Shahbazi, R.; Eroglu, I. Nanofibers Based Antibacterial Drug Design, Delivery and Applications. Curr. Pharm. Des. 2015, 21, 1930–1943. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmed, S.; Ikram, S. Silver Nanoparticles: One Pot Green Synthesis Using Terminalia arjuna Extract for Biological Application. J. Nanomed. Nanotechnol. 2015, 6. [Google Scholar] [CrossRef] [Green Version]
  5. Yaqoob, S.B.; Adnan, R.; Rameez Khan, R.M.; Rashid, M. Gold, Silver, and Palladium Nanoparticles: A Chemical Tool for Biomedical Applications. Front. Chem. 2020, 8, 376. [Google Scholar] [CrossRef]
  6. Panáček, A.; Kvítek, L.; Prucek, R.; Kolář, M.; Večeřová, R.; Pizúrová, N.; Sharma, V.K.; Nevěčná, T.; Zbořil, R. Silver Colloid Nanoparticles: Synthesis, Characterization, and Their Antibacterial Activity. J. Phys. Chem. B 2006, 110, 16248–16253. [Google Scholar] [CrossRef]
  7. Quang, D.V.; Sarawade, P.B.; Hilonga, A.; Kim, J.-K.; Chai, Y.G.; Kim, S.H.; Ryu, J.-Y.; Kim, H.T. Preparation of silver nanoparticle containing silica micro beads and investigation of their antibacterial activity. Appl. Surf. Sci. 2011, 257, 6963–6970. [Google Scholar] [CrossRef]
  8. Martínez-Gutierrez, F.; Thi, E.P.; Silverman, J.M.; de Oliveira, C.C.; Svensson, S.L.; Hoek, A.V.; Sánchez, E.M.; Reiner, N.E.; Gaynor, E.C.; Pryzdial, E.L.G.; et al. Antibacterial activity, inflammatory response, coagulation and cytotoxicity effects of silver nanoparticles. Nanomedicine 2012, 8, 328–336. [Google Scholar] [CrossRef]
  9. Salleh, A.; Naomi, R.; Utami, N.D.; Mohammad, A.W.; Mahmoudi, E.; Mustafa, N.; Fauzi, M.B. The Potential of Silver Nanoparticles for Antiviral and Antibacterial Applications: A Mechanism of Action. Nanomaterials 2020, 10, 1566. [Google Scholar] [CrossRef] [PubMed]
  10. Oluwaniyi, O.O.; Adegoke, H.I.; Adesuji, E.T.; Alabi, A.B.; Bodede, S.O.; Labulo, A.H.; Oseghale, C.O. Biosynthesis of silver nanoparticles using aqueous leaf extract of Thevetia peruviana Juss and its antimicrobial activities. Appl. Nanosci. 2016, 6, 903–912. [Google Scholar] [CrossRef]
  11. Youssef, F.S.; El-Banna, H.A.; Elzorba, H.Y.; Galal, A.M. Application of some nanoparticles in the field of veterinary medicine. Int. J. Vet. Sci. Med. 2019, 7, 78–93. [Google Scholar] [CrossRef] [PubMed]
  12. Burdușel, A.-C.; Gherasim, O.; Grumezescu, A.M.; Mogoantă, L.; Ficai, A.; Andronescu, E. Biomedical Applications of Silver Nanoparticles: An Up-to-Date Overview. Nanomaterials 2018, 8, 681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Vankar, P.S.; Shukla, D. Biosynthesis of silver nanoparticles using lemon leaves extract and its application for antimicrobial finish on fabric. Appl. Nanosci. 2012, 2, 163–168. [Google Scholar] [CrossRef] [Green Version]
  14. Xia, Z.-K.; Ma, Q.-H.; Li, S.-Y.; Zhang, D.-Q.; Cong, L.; Tian, Y.-L.; Yang, R.-Y. The antifungal effect of silver nanoparticles on Trichosporon asahii. J. Microbiol. Immunol. Infect. 2016, 49, 182–188. [Google Scholar] [CrossRef] [Green Version]
  15. Panáček, A.; Kolář, M.; Večeřová, R.; Prucek, R.; Soukupová, J.; Kryštof, V.; Hamal, P.; Zbořil, R.; Kvítek, L. Antifungal activity of silver nanoparticles against Candida spp. Biomaterials 2009, 30, 6333–6340. [Google Scholar] [CrossRef]
  16. Elgorban, A.M.; El-Samawaty, A.E.-R.M.; Yassin, M.A.; Sayed, S.R.; Adil, S.F.; Elhindi, K.M.; Bakri, M.; Khan, M. Antifungal silver nanoparticles: Synthesis, characterization and biological evaluation. Biotechnol. Biotechnol. Equip. 2016, 30, 56–62. [Google Scholar] [CrossRef] [Green Version]
  17. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Cheema, S.A.; Rehman, H.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 721, 137778. [Google Scholar] [CrossRef]
  18. Galdiero, S.; Falanga, A.; Vitiello, M.; Cantisani, M.; Marra, V.; Galdiero, M. Silver Nanoparticles as Potential Antiviral Agents. Molecules 2011, 16, 8894–8918. [Google Scholar] [CrossRef] [Green Version]
  19. Thi Ngoc Dung, T.; Nang Nam, V.; Thi Nhan, T.; Ngoc, T.T.B.; Minh, L.Q.; Nga, B.T.T.; Phan Le, V.; Viet Quang, D. Silver nanoparticles as potential antiviral agents against African swine fever virus. Mater. Res. Express 2020, 6, 1250–1259. [Google Scholar] [CrossRef]
  20. Mori, Y.; Ono, T.; Miyahira, Y.; Nguyen, V.Q.; Matsui, T.; Ishihara, M. Antiviral activity of silver nanoparticle/chitosan composites against H1N1 influenza A virus. Nanoscale Res. Lett. 2013, 8, 93. [Google Scholar] [CrossRef] [Green Version]
  21. Bekele, A.Z.; Gokulan, K.; Williams, K.M.; Khare, S. Dose and Size-Dependent Antiviral Effects of Silver Nanoparticles on Feline Calicivirus, a Human Norovirus Surrogate. Foodborne Pathog. Dis. 2016, 13, 239–244. [Google Scholar] [CrossRef] [PubMed]
  22. Singh, L.; Kruger, H.G.; Maguire, G.E.M.; Govender, T.; Parboosing, R. The role of nanotechnology in the treatment of viral infections. Ther. Adv. Infect. Dis. 2017, 4, 105–131. [Google Scholar] [CrossRef] [PubMed]
  23. Seino, S.; Imoto, Y.; Kosaka, T.; Nishida, T.; Nakagawa, T.; Yamamoto, T.A. Antiviral Activity of Silver Nanoparticles Immobilized onto Textile Fabrics Synthesized by Radiochemical Process. MRS Adv. 2016, 1, 705–710. [Google Scholar] [CrossRef]
  24. Balagna, C.; Perero, S.; Percivalle, E.; Nepita, E.V.; Ferraris, M. Virucidal effect against coronavirus SARS-CoV-2 of a silver nanocluster/silica composite sputtered coating. Open Ceram. 2020, 1, 100006. [Google Scholar] [CrossRef]
  25. Nikaeen, G.; Abbaszadeh, S.; Yousefinejad, S. Application of nanomaterials in treatment, anti-infection and detection of coronaviruses. Nanomedicine 2020, 15, 1501–1512. [Google Scholar] [CrossRef] [PubMed]
  26. Lee, S.; Jun, B.-H. Silver Nanoparticles: Synthesis and Application for Nanomedicine. Int. J. Mol. Sci. 2019, 20, 865. [Google Scholar] [CrossRef] [Green Version]
  27. Sotiriou, G.A.; Pratsinis, S.E. Engineering nanosilver as an antibacterial, biosensor and bioimaging material. Curr. Opin. Chem. Eng. 2011, 1, 3–10. [Google Scholar] [CrossRef] [Green Version]
  28. Marassi, V.; Cristo, L.D.; Smith, G.J.; Ortelli, S.; Blosi, M.; Costa, A.L.; Reschiglian, P.; Volkov, Y.; Prina-Mello, A. Silver nanoparticles as a medical device in healthcare settings: A five-step approach for candidate screening of coating agents. R. Soc. Open Sci. 2018, 5, 171113. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, J.; Zhao, Y.; Guo, Q.; Wang, Z.; Wang, H.; Yang, Y.; Huang, Y. TAT-modified nanosilver for combating multidrug-resistant cancer. Biomaterials 2012, 33, 6155–6161. [Google Scholar] [CrossRef]
  30. Ardhendu, K.M. Silver Nanoparticles as Drug Delivery Vehicle. Glob. J. Nanomed. 2017, 3, 555607. [Google Scholar] [CrossRef]
  31. Lee, K.-S.; El-Sayed, M.A. Gold and Silver Nanoparticles in Sensing and Imaging: Sensitivity of Plasmon Response to Size, Shape, and Metal Composition. J. Phys. Chem. B 2006, 110, 19220–19225. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, L.; Sun, Y.; Che, G.; Li, Z. Self-assembled silver nanoparticle films at an air–liquid interface and their applications in SERS and electrochemistry. App. Surf. Sci. 2011, 257, 7150–7155. [Google Scholar] [CrossRef]
  33. Alvarez-Puebla, R.A.; Aroca, R.F. Synthesis of Silver Nanoparticles with Controllable Surface Charge and Their Application to Surface-Enhanced Raman Scattering. Anal. Chem. 2009, 81, 2280–2285. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, F.; Wang, Q.; Gu, Z.; Fang, K.; Marriott, G.; Gu, N. Silver Nanoparticle-Embedded Microbubble as a Dual-Mode Ultrasound and Optical Imaging Probe. ACS Appl. Mater. Interfaces 2013, 5, 9217–9223. [Google Scholar] [CrossRef] [PubMed]
  35. Varghese Alex, K.; Tamil Pavai, P.; Rugmini, R.; Shiva Prasad, M.; Kamakshi, K.; Sekhar, K.C. Green Synthesized Ag Nanoparticles for Bio-Sensing and Photocatalytic Applications. ACS Omega 2020, 5, 13123–13129. [Google Scholar] [CrossRef]
  36. Manno, D.; Filippo, E.; Di Giulio, M.; Serra, A. Synthesis and characterization of starch-stabilized Ag nanostructures for sensors applications. J. Non-Cryst. Solids 2008, 354, 5515–5520. [Google Scholar] [CrossRef]
  37. Dong, X.-Y.; Gao, Z.-W.; Yang, K.-F.; Zhang, W.-Q.; Xu, L.-W. Nanosilver as a new generation of silver catalysts in organic transformations for efficient synthesis of fine chemicals. Catal. Sci. Technol. 2015, 5, 2554–2574. [Google Scholar] [CrossRef]
  38. Treshchalov, A.; Erikson, H.; Puust, L.; Tsarenko, S.; Saar, R.; Vanetsev, A.; Tammeveski, K.; Sildos, I. Stabilizer-free silver nanoparticles as efficient catalysts for electrochemical reduction of oxygen. J. Colloid Interface Sci. 2017, 491, 358–366. [Google Scholar] [CrossRef]
  39. Yaqoob, A.A.; Umar, K.; Ibrahim, M.N.M. Silver nanoparticles: Various methods of synthesis, size affecting factors and their potential applications—A review. Appl. Nanosci. 2020, 10, 1369–1378. [Google Scholar] [CrossRef]
  40. Zhang, X.-F.; Liu, Z.-G.; Shen, W.; Gurunathan, S. Silver Nanoparticles: Synthesis, Characterization, Properties, Applications, and Therapeutic Approaches. Int. J. Mol. Sci. 2016, 17, 1534. [Google Scholar] [CrossRef]
  41. Abou El-Nour, K.M.M.; Eftaiha, A.; Al-Warthan, A.; Ammar, R.A.A. Synthesis and applications of silver nanoparticles. Arab. J. Chem. 2010, 3, 135–140. [Google Scholar] [CrossRef] [Green Version]
  42. Chouhan, N. Silver Nanoparticles: Synthesis, Characterization and Applications. In Silver Nanoparticles—Fabrication, Characterization and Applications; Maaz, K., Ed.; InTech: Rijeka, Croatia, 2018; ISBN 978-1-78923-478-7. [Google Scholar]
  43. Khodashenas, B.; Ghorbani, H.R. Synthesis of silver nanoparticles with different shapes. Arab. J. Chem. 2019, 12, 1823–1838. [Google Scholar] [CrossRef] [Green Version]
  44. AL-Thabaiti, S.A.; Al-Nowaiser, F.M.; Obaid, A.Y.; Al-Youbi, A.O.; Khan, Z. Formation and characterization of surfactant stabilized silver nanoparticles: A kinetic study. Colloids Surf. B 2008, 67, 230–237. [Google Scholar] [CrossRef] [PubMed]
  45. Menger, F.M.; Littau, C.A. Gemini-surfactants: Synthesis and properties. J. Am. Chem. Soc. 1991, 113, 1451–1452. [Google Scholar] [CrossRef]
  46. Alami, E.; Beinert, G.; Marie, P.; Zana, R. Alkanediyl-α,ω-bis(dimethylalkylammonium bromide) surfactants. 3. Behavior at the air-water interface. Langmuir 1993, 9, 1465–1467. [Google Scholar] [CrossRef]
  47. Brycki, B.; Kowalczyk, I.; Kozirog, A. Synthesis, Molecular Structure, Spectral Properties and Antifungal Activity of Polymethylene-α,ω-bis(N,N- dimethyl-N-dodecyloammonium Bromides). Molecules 2011, 16, 319–335. [Google Scholar] [CrossRef] [Green Version]
  48. Brycki, B.E.; Kowalczyk, I.H.; Szulc, A.; Kaczerewska, O.; Pakiet, M. Multifunctional Gemini Surfactants: Structure, Synthesis, Properties and Applications. In Application and Characterization of Surfactants; Najjar, R., Ed.; InTech: Rijeka, Croatia, 2017; ISBN 978-953-51-3325-4. [Google Scholar]
  49. Menger, F.M.; Keiper, J.S. Gemini Surfactants. Angew. Chem. Int. Ed. 2000, 39, 1906–1920. [Google Scholar] [CrossRef]
  50. Brycki, B.; Szulc, A.; Koenig, H.; Kowalczyk, I.; Pospieszny, T.; Górka, S. Effect of the alkyl chain length on micelle formation for bis(N-alkyl-N,N-dimethylethylammonium)ether dibromides. C. R. Chim. 2019, 22, 386–392. [Google Scholar] [CrossRef]
  51. Chlebicki, J.; Węgrzyńska, J.; Wilk, K.A. Surface-active, micellar, and antielectrostatic properties of bis-ammonium salts. J. Colloid Interface Sci. 2008, 323, 372–378. [Google Scholar] [CrossRef]
  52. Das, S.; Mukherjee, I.; Paul, B.K.; Ghosh, S. Physicochemical Behaviors of Cationic Gemini Surfactant (14-4-14) Based Microheterogeneous Assemblies. Langmuir 2014, 30, 12483–12493. [Google Scholar] [CrossRef]
  53. Pisárčik, M.; Devínsky, F. Surface tension study of cationic gemini surfactants binding to DNA. Cent. Eur. J. Chem. 2014, 12, 577–585. [Google Scholar] [CrossRef]
  54. Devínsky, F.; Masárová, Ľ.; Lacko, I. Surface activity and micelle formation of some new bisquaternary ammonium salts. J. Colloid Interface Sci. 1985, 105, 235–239. [Google Scholar] [CrossRef]
  55. Kowalczyk, I.; Pakiet, M.; Szulc, A.; Koziróg, A. Antimicrobial Activity of Gemini Surfactants with Azapolymethylene Spacer. Molecules 2020, 25, 4054. [Google Scholar] [CrossRef] [PubMed]
  56. Koziróg, A.; Brycki, B. Monomeric and gemini surfactants as antimicrobial agents—Influence on environmental and reference strains. Acta Biochim. Pol. 2015, 62, 879–883. [Google Scholar] [CrossRef] [PubMed]
  57. Kaczerewska, O.; Leiva-Garcia, R.; Akid, R.; Brycki, B.; Kowalczyk, I.; Pospieszny, T. Heteroatoms and π electrons as favorable factors for efficient corrosion protection. Mater. Corros. 2019, 70, 1099–1110. [Google Scholar] [CrossRef]
  58. Garcia, M.T.; Kaczerewska, O.; Ribosa, I.; Brycki, B.; Materna, P.; Drgas, M. Biodegradability and aquatic toxicity of quaternary ammonium-based gemini surfactants: Effect of the spacer on their ecological properties. Chemosphere 2016, 154, 155–160. [Google Scholar] [CrossRef] [Green Version]
  59. Kaczerewska, O.; Brycki, B.; Ribosa, I.; Comelles, F.; Garcia, M.T. Cationic gemini surfactants containing an O-substituted spacer and hydroxyethyl moiety in the polar heads: Self-assembly, biodegradability and aquatic toxicity. J. Ind. Eng. Chem. 2018, 59, 141–148. [Google Scholar] [CrossRef]
  60. Brycki, B.; Waligórska, M.; Szulc, A. The biodegradation of monomeric and dimeric alkylammonium surfactants. J. Hazard. Mater. 2014, 280, 797–815. [Google Scholar] [CrossRef]
  61. Akram, M.; Anwar, S.; Ansari, F.; Bhat, I.A.; Kabir-ud-Din, K.-D. Bio-physicochemical analysis of ethylene oxide-linked diester-functionalized green cationic gemini surfactants. RSC Adv. 2016, 6, 21697–21705. [Google Scholar] [CrossRef]
  62. Bergero, M.F.; Liffourrena, A.S.; Opizzo, B.A.; Fochesatto, A.S.; Lucchesi, G.I. Immobilization of a microbial consortium on Ca-alginate enhances degradation of cationic surfactants in flasks and bioreactor. Int. Biodeterior. Biodegrad. 2017, 117, 39–44. [Google Scholar] [CrossRef]
  63. Mondal, M.H.; Roy, A.; Malik, S.; Ghosh, A.; Saha, B. Review on chemically bonded geminis with cationic heads: Second-generation interfactants. Res. Chem. Intermed. 2016, 42, 1913–1928. [Google Scholar] [CrossRef]
  64. Mondal, M.H.; Malik, S.; Roy, A.; Saha, R.; Saha, B. Modernization of surfactant chemistry in the age of gemini and bio-surfactants: A review. RSC Adv. 2015, 5, 92707–92718. [Google Scholar] [CrossRef]
  65. Martín, V.I.; de la Haba, R.R.; Ventosa, A.; Congiu, E.; Ortega-Calvo, J.J.; Moyá, M.L. Colloidal and biological properties of cationic single-chain and dimeric surfactants. Colloids Surf. B 2014, 114, 247–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Brycki, B.; Szulc, A. Gemini Alkyldeoxy-D-Glucitolammonium Salts as Modern Surfactants and Microbiocides: Synthesis, Antimicrobial and Surface Activity, Biodegradation. PLoS ONE 2014, 9, e84936. [Google Scholar] [CrossRef] [Green Version]
  67. Dani, U.; Bahadur, A.; Kuperkar, K. Micellization, Antimicrobial Activity and Curcumin Solubilization in Gemini Surfactants: Influence of Spacer and Non-Polar Tail. Colloids Interface Sci. Commun. 2018, 25, 22–30. [Google Scholar] [CrossRef]
  68. Jennings, M.C.; Buttaro, B.A.; Minbiole, K.P.C.; Wuest, W.M. Bioorganic Investigation of Multicationic Antimicrobials to Combat QAC-Resistant Staphylococcus aureus. ACS Infect. Dis. 2015, 1, 304–309. [Google Scholar] [CrossRef]
  69. Zhang, S.; Ding, S.; Yu, J.; Chen, X.; Lei, Q.; Fang, W. Antibacterial Activity, in Vitro Cytotoxicity, and Cell Cycle Arrest of Gemini Quaternary Ammonium Surfactants. Langmuir 2015, 31, 12161–12169. [Google Scholar] [CrossRef]
  70. Kuperkar, K.; Modi, J.; Patel, K. Surface-Active Properties and Antimicrobial Study of Conventional Cationic and Synthesized Symmetrical Gemini Surfactants. J. Surfactants Deterg. 2012, 15, 107–115. [Google Scholar] [CrossRef]
  71. Obłąk, E.; Piecuch, A.; Krasowska, A.; Łuczyński, J. Antifungal activity of gemini quaternary ammonium salts. Microbiol. Res. 2013, 168, 630–638. [Google Scholar] [CrossRef]
  72. Koziróg, A.; Kręgiel, D.; Brycki, B. Action of Monomeric/Gemini Surfactants on Free Cells and Biofilm of Asaia lannensis. Molecules 2017, 22, 2036. [Google Scholar] [CrossRef] [Green Version]
  73. Labena, A.; Hegazy, M.A.; Sami, R.M.; Hozzein, W.N. Multiple Applications of a Novel Cationic Gemini Surfactant: Anti-Microbial, Anti-Biofilm, Biocide, Salinity Corrosion Inhibitor, and Biofilm Dispersion (Part II). Molecules 2020, 25, 1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kumar, N.; Tyagi, R. Industrial Applications of Dimeric Surfactants: A Review. J. Dispers. Sci. Technol. 2014, 35, 205–214. [Google Scholar] [CrossRef]
  75. Brycki, B.; Drgas, M.; Bielawska, M.; Zdziennicka, A.; Jańczuk, B. Synthesis, spectroscopic studies, aggregation and surface behavior of hexamethylene-1,6-bis(N,N-dimethyl-N-dodecylammonium bromide). J. Mol. Liq. 2016, 221, 1086–1096. [Google Scholar] [CrossRef]
  76. 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]
  77. Zana, R. Dimeric (Gemini) Surfactants: Effect of the Spacer Group on the Association Behavior in Aqueous Solution. J. Colloid Interface Sci. 2002, 248, 203–220. [Google Scholar] [CrossRef]
  78. Singh, V.; Tyagi, R. Unique Micellization and CMC Aspects of Gemini Surfactant: An Overview. J. Dispers. Sci. Technol. 2014, 35, 1774–1792. [Google Scholar] [CrossRef]
  79. ud din Parray, M.; Maurya, N.; Ahmad Wani, F.; Borse, M.S.; Arfin, N.; Ahmad Malik, M.; Patel, R. Comparative effect of cationic gemini surfactant and its monomeric counterpart on the conformational stability of phospholipase A2. J. Mol. Struct. 2019, 1175, 49–55. [Google Scholar] [CrossRef]
  80. Egorova, E.M.; Kaba, S.I. The effect of surfactant micellization on the cytotoxicity of silver nanoparticles stabilized with aerosol-OT. Toxicol. In Vitro 2019, 57, 244–254. [Google Scholar] [CrossRef]
  81. Zana, R. Critical Micellization Concentration of Surfactants in Aqueous Solution and Free Energy of Micellization. Langmuir 1996, 12, 1208–1211. [Google Scholar] [CrossRef]
  82. Menger, F.M.; Littau, C.A. Gemini surfactants: A new class of self-assembling molecules. J. Am. Chem. Soc. 1993, 115, 10083–10090. [Google Scholar] [CrossRef]
  83. Zana, R.; Levy, H. Alkanediyl-α,ω-bis(dimethylalkylammonium bromide) surfactants (dimeric surfactants) Part 6. CMC of the ethanediyl- 1,2-bis(dimethylalkylammonium bromide) series. Colloids Surf. A 1997, 127, 229–232. [Google Scholar] [CrossRef]
  84. Dam, T.; Engberts, J.B.F.N.; Karthäuser, J.; Karaborni, S.; van Os, N.M. Synthesis, surface properties and oil solubilisation capacity of cationic gemini surfactants. Colloids Surf. A 1996, 118, 41–49. [Google Scholar] [CrossRef] [Green Version]
  85. Oda, R.; Candau, S.J.; Oda, R.; Huc, I. Gemini surfactants, the effect of hydrophobic chain length and dissymmetry. Chem. Commun. 1997, 2105–2106. [Google Scholar] [CrossRef]
  86. Chang, H.; Cui, Y.; Wang, Y.; Li, G.; Gao, W.; Li, X.; Zhao, X.; Wei, W. Wettability and adsorption of PTFE and paraffin surfaces by aqueous solutions of biquaternary ammonium salt Gemini surfactants with hydroxyl. Colloids Surf. A 2016, 506, 416–424. [Google Scholar] [CrossRef]
  87. Pisárčik, M.; Jampílek, J.; Lukáč, M.; Horáková, R.; Devínsky, F.; Bukovský, M.; Kalina, M.; Tkacz, J.; Opravil, T. Silver Nanoparticles Stabilised by Cationic Gemini Surfactants with Variable Spacer Length. Molecules 2017, 22, 1794. [Google Scholar] [CrossRef] [Green Version]
  88. Siddiq, A.M.; Parandhaman, T.; Begam, A.F.; Das, S.K.; Alam, M.S. Effect of gemini surfactant (16-6-16) on the synthesis of silver nanoparticles: A facile approach for antibacterial application. Enzym. Microb. Technol. 2016, 95, 118–127. [Google Scholar] [CrossRef]
  89. Xu, J.; Han, X.; Liu, H.; Hu, Y. Synthesis and optical properties of silver nanoparticles stabilized by gemini surfactant. Colloids Surf. A 2006, 273, 179–183. [Google Scholar] [CrossRef]
  90. He, S.; Chen, H.; Guo, Z.; Wang, B.; Tang, C.; Feng, Y. High-concentration silver colloid stabilized by a cationic gemini surfactant. Colloids Surf. A 2013, 429, 98–105. [Google Scholar] [CrossRef]
  91. Guzman, M.; Dille, J.; Godet, S. Synthesis and antibacterial activity of silver nanoparticles against gram-positive and gram-negative bacteria. Nanomedicine 2012, 8, 37–45. [Google Scholar] [CrossRef]
  92. He, S.; Yao, J.; Jiang, P.; Shi, D.; Zhang, H.; Xie, S.; Pang, S.; Gao, H. Formation of Silver Nanoparticles and Self-Assembled Two-Dimensional Ordered Superlattice. Langmuir 2001, 17, 1571–1575. [Google Scholar] [CrossRef]
  93. Njagi, E.C.; Huang, H.; Stafford, L.; Genuino, H.; Galindo, H.M.; Collins, J.B.; Hoag, G.E.; Suib, S.L. Biosynthesis of Iron and Silver Nanoparticles at Room Temperature Using Aqueous Sorghum Bran Extracts. Langmuir 2011, 27, 264–271. [Google Scholar] [CrossRef] [PubMed]
  94. Li, D.; Fang, W.; Feng, Y.; Geng, Q.; Song, M. Stability properties of water-based gold and silver nanofluids stabilized by cationic gemini surfactants. J. Taiwan Inst. Chem. Eng. 2019, 97, 458–465. [Google Scholar] [CrossRef]
  95. Siddiq, A.M.; Thangam, R.; Madhan, B.; Alam, M.S. Counterion coupled (COCO) gemini surfactant capped Ag/Au alloy and core–shell nanoparticles for cancer therapy. RSC Adv. 2019, 9, 37830–37845. [Google Scholar] [CrossRef]
  96. Haque, M.N.; Kwon, S.; Cho, D. Formation and stability study of silver nano-particles in aqueous and organic medium. Korean J. Chem. Eng. 2017, 34, 2072–2078. [Google Scholar] [CrossRef]
Figure 1. Synthesis of 16-6-16 and 16-O-16.
Figure 1. Synthesis of 16-6-16 and 16-O-16.
Applsci 11 00154 g001
Figure 2. FTIR spectra for 16-6-16 and 16-O-16.
Figure 2. FTIR spectra for 16-6-16 and 16-O-16.
Applsci 11 00154 g002
Figure 3. Specific conductivity versus surfactant concentration in water for 16-O-16.
Figure 3. Specific conductivity versus surfactant concentration in water for 16-O-16.
Applsci 11 00154 g003
Figure 4. Schematic representation of the formation of layer of gemini surfactants, surrounding the Ag nanoparticles.
Figure 4. Schematic representation of the formation of layer of gemini surfactants, surrounding the Ag nanoparticles.
Applsci 11 00154 g004
Figure 5. UV-Vis spectra of silver nanoparticles stabilized with gemini surfactants (a) 16-6-16; (b) 16-O-16 at different nAg/nGemini values.
Figure 5. UV-Vis spectra of silver nanoparticles stabilized with gemini surfactants (a) 16-6-16; (b) 16-O-16 at different nAg/nGemini values.
Applsci 11 00154 g005
Figure 6. UV-Vis spectra of silver nanoparticles stabilized with 16-6-16 obtained in procedure with excess of reductant.
Figure 6. UV-Vis spectra of silver nanoparticles stabilized with 16-6-16 obtained in procedure with excess of reductant.
Applsci 11 00154 g006
Figure 7. Size of silver nanoparticles capped with 16-O-16 and 16-6-16 as a function of nAg/nGemini molar ratio.
Figure 7. Size of silver nanoparticles capped with 16-O-16 and 16-6-16 as a function of nAg/nGemini molar ratio.
Applsci 11 00154 g007
Figure 8. Particle size distribution histograms of AgNPs obtained by DLS measurements.
Figure 8. Particle size distribution histograms of AgNPs obtained by DLS measurements.
Applsci 11 00154 g008
Table 1. Critical micelle concentration, degree of ionization, counterion binding parameter, and Gibbs free energy of micellization in 25 °C.
Table 1. Critical micelle concentration, degree of ionization, counterion binding parameter, and Gibbs free energy of micellization in 25 °C.
SurfactantCMC (mM)αβΔGomic (kJ/mol)
16-6-160.0340.550.45−50.16
16-O-160.0310.400.60−58.32
Table 2. Experimental data of different procedures for AgNPs obtaining.
Table 2. Experimental data of different procedures for AgNPs obtaining.
ExperimentGemini SurfactantMolar Ratio nAg/nGeminiConcentration of AgNO3 [mM]
116-6-162.50.75
216-6-1650.75
316-6-16100.75
416-O-162.50.75
516-O-1650.75
616-O-16100.75
716-6-1652
816-6-1651
Table 3. The wavelength and absorbance data obtained from UV-Vis spectra.
Table 3. The wavelength and absorbance data obtained from UV-Vis spectra.
ExperimentWavelength [nm]Absorbance
14200.286
24350.332
34250.359
44200.185
54200.313
64150.397
74100.690
84190.596
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Brycki, B.; Szulc, A.; Babkova, M. Synthesis of Silver Nanoparticles with Gemini Surfactants as Efficient Capping and Stabilizing Agents. Appl. Sci. 2021, 11, 154. https://doi.org/10.3390/app11010154

AMA Style

Brycki B, Szulc A, Babkova M. Synthesis of Silver Nanoparticles with Gemini Surfactants as Efficient Capping and Stabilizing Agents. Applied Sciences. 2021; 11(1):154. https://doi.org/10.3390/app11010154

Chicago/Turabian Style

Brycki, Bogumił, Adrianna Szulc, and Mariia Babkova. 2021. "Synthesis of Silver Nanoparticles with Gemini Surfactants as Efficient Capping and Stabilizing Agents" Applied Sciences 11, no. 1: 154. https://doi.org/10.3390/app11010154

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