Next Article in Journal
Merging the Sol–Gel Technique with the Pulsed Microplasma Cluster Source Deposition to Improve Control over the Memristive Response of TiO2 Thin Films
Next Article in Special Issue
Investigation of Color Reproduction on Linen Fabrics when Printing with Mimaki TX400-1800D Inkjet with Pigment TP250 Dyes
Previous Article in Journal
Shoulder Implant Manufacturer Detection by Using Deep Learning: Proposed Channel Selection Layer
Previous Article in Special Issue
The Interface Characterization of 2-Mercapto-1-methylimidazole Corrosion Inhibitor on Brass
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Water-Resistant Latex Coatings: Tuning of Properties by Polymerizable Surfactant, Covalent Crosslinking and Nanostructured ZnO Additive

1
Institute of Chemistry and Technology of Macromolecular Materials, Faculty of Chemical Technology, University of Pardubice, Studentská 573, 532 10 Pardubice, Czech Republic
2
Institute of Theoretical and Applied Mechanics of the Czech Academy of Sciences, Prosecká 809/76, 190 00 Prague, Czech Republic
3
Center of Materials and Nanotechnologies, Faculty of Chemical Technology, University of Pardubice, Studentská 573, 532 10 Pardubice, Czech Republic
4
Centre of Polymer Systems, Tomas Bata University in Zlín, Třída Tomáše Bati 5678, 760 01 Zlín, Czech Republic
*
Author to whom correspondence should be addressed.
Coatings 2021, 11(3), 347; https://doi.org/10.3390/coatings11030347
Submission received: 1 March 2021 / Revised: 11 March 2021 / Accepted: 15 March 2021 / Published: 18 March 2021
(This article belongs to the Collection Coatings: 10th Anniversary)

Abstract

:
This paper deals with the development of acrylic latexes providing high-performance water-resistant coatings. For this purpose, mutual effects of anionic surfactant type (ordinary and polymerizable), covalent intra- and/or interparticle crosslinking (introduced by allyl methacrylate copolymerization and keto-hydrazide reaction, respectively) and ionic crosslinking (provided by nanostructured ZnO additive) were investigated. The latexes were prepared by the standard emulsion polymerization of methyl methacrylate, butyl acrylate and methacrylic acid as the main monomers. The addition of surface-untreated powdered nanostructured ZnO was performed during latex synthesis, resulting in stable latexes comprising dispersed nanosized additive in the content of ca 0.9−1.0 wt.% (based on solids). The coating performance with emphasis on water resistance was evaluated. It was determined that the application of the polymerizable surfactant improved coating adhesion and water-resistance, but it wasn′t able to ensure high water-resistance of coatings. Highly water-resistant coatings were obtained provided that covalent intra- and interparticle crosslinking together with ionic crosslinking were employed in the coating composition, forming densely crosslinked latex films. Moreover, coatings comprising nanostructured ZnO additive displayed a significant antibacterial activity and improved solvent resistance.

1. Introduction

Development of ecological coatings has attracted particular interest in coating and paint industry, where acrylic latex binders are of considerable interest due to their favorable properties, such as low cost, resistance to atmospheric conditions, ease of preparation, modification and pigmentation. However, they have the disadvantage of water sensitivity of coatings, which often limits their practical application in high-performance products. The improvement of water resistance of acrylic latex-based coating films is therefore highly challenging and should be addressed preferably already in the synthesis phase of an acrylic emulsion polymer.
One approach to enhance the properties of latex coatings including the water resistance is the employment of intraparticle and interparticle covalent (chemical) crosslinking chemistry in emulsion polymers [1,2]. The intraparticle covalent crosslinking, usually introduced by copolymerizing multifunctional monomers such as divinylbenzene, ethylene glycol dimethacrylate or allyl methacrylate (ALMA), often provides rigid latex particles and results in deteriorated film formation [3]. The interparticle covalent crosslinking refers to a chemical reaction designed to occur between adjacent latex particles in the course of film formation. As reflected by number of publications and patents, there has been particular interest in one-pack acrylic coating compositions curable at room temperature [4,5,6]. Among these materials, self-crosslinking systems based on keto-hydrazide crosslinking reaction have been shown to be particularly effective [7,8,9]. Diacetone-acrylamide (DAAM) with adipic acid dihydrazide (ADH) is the most favored combination used in the self-crosslinking latexes [10,11,12]. The most remarkable feature of the keto-hydrazide crosslinking reaction is fast curing at room temperature, providing minimal consumption of energy during processing and easy applicability in case of recoating substrates.
Apart from the covalent crosslinking, the improved protective properties of latex coatings can be achieved by ionic (physical) interparticle crosslinking which occurs in the interfacial zones between adjacent latex particles, through ionic dipolar interactions [2]. In the case of carboxylated latexes, the ionic crosslinks are generated between carboxylic groups on the surface of latex particles in the presence of salts and oxides of multivalent metals [13,14,15]. The ionic bonds are believed to have the nature of ion multiplets or even larger clusters formed by associated ionized carboxylic groups and metal counterions [16,17]. Among the suitable ionic crosslinking additives, ZnO, typically in the form of water dispersion of small particles, has been reported in the relevant literature as an effective crosslinker of latex-based rubbers and coatings [18,19]. As ZnO has a limited water solubility, it neutralizes (ionizes) carboxylated latexes in the wet state and cures them upon drying. Recently, nanosized ZnO has been presented in the connection with ionic crosslinking of latex coatings leading to a pronounced improvement of coating protective properties, such as water and solvent resistance, anticorrosion protection and mechanical resistance [20,21,22]. Nanostructured ZnO has also met with considerable enthusiasm in the field of latex coatings due to the introduction of UV-shielding and antibacterial properties [23,24,25].
Another tool, being recognized to be effective in the improvement of water resistance of latex coatings, is the utilization of polymerizable surfactants in the process of emulsion polymerization [26,27,28,29]. In the case of the coating films made of latexes synthesized using conventional surfactants, the surfactant molecules, adsorbed on the surface of latex particles, can desorb easily and migrate through the film forming aggregates, which leads to increased penetration of water [30]. The small and mobile surfactant species may also migrate to the air-film interface affecting gloss of the coating film or to the film-substrate interface deteriorating adhesion [31]. By using the polymerizable surfactants, which become covalently attached to the latex particles, the surfactant desorption and migration problems were avoided or at least minimized [32,33]. Positive effects were obtained both on the latex stability and on the properties of the dried coating films [34,35,36,37].
The present paper deals with the preparation and evaluation of acrylic latexes suitable for water-resistant high-performance coating applications. While studying the optimal and highly effective latex composition, we concentrated on mutual effects of various unconventional strategies, namely, (i) the emulsion polymerization performed using a polymerizable surfactant, (ii) employing the covalent interparticle crosslinking based on the keto-hydrazide crosslinking reaction, (iii) introducing the covalent intraparticle crosslinking into latex polymer by copolymerizing ALMA and (iv) introducing the ionic crosslinking by means of a nanostructured ZnO additive, being incorporated in the course of the polymerization process. The latex coating performance with emphasis on water resistance was evaluated and mutually compared.

2. Materials and Methods

2.1. Materials

Latexes were prepared of methyl methacrylate (MMA, CAS: 80-62-6), butyl acrylate (BA, CAS: 141-32-2), methacrylic acid (MAA, CAS: 79-41-4), allyl methacrylate (ALMA, CAS: 96-05-9) and diacetone acrylamide (DAAM, CAS: 2873-97-4). All the monomers were purchased from Sigma-Aldrich (Prague, Czech Republic). Disponil FES 993 (BASF, Prague, Czech Republic) was used as the ordinary anionic surfactant. HITENOL AR-10 was used as the polymerizable anionic surfactant and was kindly supplied by DKS Co. Ltd. (Tokyo, Japan). Table 1 presents the main characteristics of the surfactants used in this work. Ammonium persulfate (CAS: 7727-54-0, Penta, Prague, Czech Republic) was utilized as the initiator of the polymerization reaction. Adipic acid dihydrazide (ADH, CAS: 1071-93-8) was utilized as the covalent crosslinking agent and was purchased from TCI Europe (Zwijndrecht, Belgium). Surface-untreated nanostructured ZnO (NanoGard®, CAS: 1314-13-2) having the average particle size in the range of 40–100 nm was used as the ionic crosslinking additive and was purchased from Alfa Aesar (Kandel, Germany). All the chemicals were utilized as received without any further purification.

2.2. Preparation and Characterization of Latexes

Four series of acrylic latexes differing in the covalent crosslinking strategy were synthesized by two-step nonseeded emulsion polymerization comprising a variable content of acrylic monomers (see Table 2). The detailed recipe for emulsion polymerization is presented in Table 3. The proportions of the acrylic monomers forming latex particles were chosen to achieve a calculated Tg (using the Fox equation [38]) of approximately 5 °C, so that film-formation was also kept in the case of the densely crosslinked latex films. Every series contained four latex samples which were synthesized using the same types and proportions of the acrylic monomers; two samples always being prepared with an ordinary (nonpolymerizable) surfactant (samples labelled D) and two samples being synthesized using a polymerizable surfactant (samples labelled H) that contains reactive double bonds, which can copolymerize with the acrylic monomers. The amounts of the particular surfactants were designed to maintain the same surface-active matter weight concentration in both types of latex samples. Moreover, one of the D-labelled samples and one of the H-labelled samples were synthesized by a standard synthetic route, while the rest two samples (labelled D_ZnO and H_ZnO) comprised nanostructured ZnO as the multifunctional additive, being incorporated in the course of the polymerization process of the second step polymer. The content of nanostructured ZnO was 1 wt.% with respect to the total monomer feeds. The way and experimental procedure of the nano-ZnO addition were based on our recent experience [20,21,25]. We have demonstrated that powdered surface-untreated nanostructured ZnO can be successfully added into a latex in the course of emulsion polymerization, providing a long-term stable coating binder.
In the Series 1, no covalent crosslinking was introduced into latex polymers. In the Series 2, the composition of the latex samples was designed to provide covalent interparticle crosslinking based on the keto-hydrazide crosslinking reaction. For this purpose, a constant amount of DAAM (5 wt.% with respect to the second step monomer feeds) was incorporated into the second step polymer to introduce ketone carbonyl groups into latex polymers for the subsequent interfacial covalent crosslinking by the reaction with ADH, added during latex formulation. The Series 3 was represented by latexes employing covalent intraparticle crosslinking, which was achieved by introducing a constant amount of ALMA (1 wt.% with respect to the first step monomer feeds) in the first step polymer. In the Series 4, the latexes were designed to provide both covalent inter- and intraparticle crosslinking by copolymerizing DAAM and ALMA in the same manner as described above. In the case of all series, carboxylic functionalities (in a relatively high content, considering standard acrylic latex coating compositions) were incorporated into the structure of the first and second step polymers by copolymerizing MAA (3 wt.% based on total monomer feeds) for three reasons: (i) stabilization of latex particles, (ii) acid catalysis of the keto-hydrazide covalent interparticle self-crosslinking reaction, (iii) enabling the ionic crosslinking via Zn2+ cations.
The latexes were produced in a 500 mL glass reaction flask under a nitrogen atmosphere at a polymerization temperature of 85 °C. The reaction flask charge (consisting of distilled water, initiator and the respective surfactant, see Table 3) was placed into the reaction flask and heated to the polymerization temperature. The monomer emulsion was consequently dosed into the stirred reaction flask at a dosing rate of about 2.5 mL/min in two steps (first polymer synthesis, second polymer synthesis), while a 15 min-long period between the two dosing steps was kept. The polymerization was then completed, during the 2 h of the hold period.
The synthesis of latexes added with nanostructured ZnO was conducted modifying the above described procedure. First, a finely dispersed ZnO aqueous dispersion was prepared by adding nanostructured ZnO to water that was designed for the preparation of the second step monomer emulsion. To facilitate the fragmentation of aggregates formed by ZnO primary nanoparticles, a proper dispersion process using a SilentCrusher M disperser (Heidolph, Schwabach, Germany) at 14,000 rpm was carried out for 20 min followed by 1 h-long ultrasonication. The fine nanostructured ZnO aqueous dispersion was then gently mixed with the monomers, surfactant and initiator designed for the preparation of the second step monomer emulsion (using a stirrer at low speed for 3 min). In the end, the resulting second step monomer emulsion with nanostructured ZnO was dosed to the reaction flask immediately at a dosing rate of about 2.5 mL/min.
After the synthesis, the latexes were filtered and the coagulum, consisting of filterable solids and solids deposited on the stirrer and the reaction flask walls, was collected. The solids content, coagulum content and monomer conversion were measured by gravimetric method and were calculated according Equations (1)–(3) [39].
solids   content   wt .   %   =   m 3     m 2   m 1   ×   100
coagulum   content   wt .   % =   m C   solids   content   ×   m L 100 +   m C   ×   100
monomer   conversion   wt .   %   =     solids   content   ×   m T 100     m I   +   m S   +   m Z n O m M   × 100
where m1 is the certain quantity of a latex sample cast onto a Petri dish; m2 is the weight of the Petri dish; m3 is the weight of the latex sample and the Petri dish dried to a constant weight at 110 °C; mC is the weight of the dried coagulum; mL is the weight of the filtered latex; mT is the total weight of all the materials put in the reaction flask; mI is the weight of the initiator; mS is the weight of the surfactant (active matter); mZnO is the weight of the nanostructured ZnO; and mM is the weight of the total monomers.
The latex pH was consequently adjusted to 8.5 with ammonia solution. Finally, an aqueous solution of ADH consisting of 1.25 g ADH and 11.3 g water was added to the latex samples of Series 2 and 4 with agitation. The prepared latexes were evaluated from the point of view of their minimum film-forming temperature (MFFT) and particle size. The minimum film-forming temperature (MFFT) was determined according to ISO 2115 [40], using the MFFT-60 instrument (Rhopoint Instruments, St Leonards, UK). The average particle sizes of polymer particles in the water phase were obtained from dynamic light scattering (DLS) experiments performed using a Litesizer 500 instrument (Anton Paar, Graz, Austria). All the DLS measurements were conducted at 25 °C.

2.3. Testing of Latex Stability

The stability of prepared latexes was evaluated using several testing methods. The heat storage stability testing was carried out by storing latexes for 60 days in an oven with a set temperature of 50 °C. The mechanical stability of latexes was evaluated by centrifuging 10 mL of latex in a sealed centrifugal tube for 15 min at a speed of 4000 rpm in a centrifugal machine. The freeze-thaw stability testing was carried out by storing 10 mL of latex in a plastic can at −5, −10 and −18 °C, respectively, for about 17 h and then allowing to thaw for 48 h at RT. The stability tests mentioned above were evaluated by observing whether there was no coagulation produced after the particular latex stability test and also, whether the tested latex provided smooth coating films without a significant loss of their gloss (below 10 rel.%). In that case the latex was considered to display the particular stability. A calcium ion stability test was carried out by dropping a small amount (two to three drops) of the latex to CaCl2 solution of a specific concentration (1, 1.5, 2 and 5 wt.%). If there was no precipitation, the latex was considered to have calcium ion stability at the given CaCl2 concentration.

2.4. Preparation and Characterization of Free-Standing Films

The free-standing films were prepared by pouring and drying the latexes in silicone molds. The samples were first air-dried at room temperature (RT, 23 ± 1 °C) for a month and then vacuum-dried at 30 °C for two weeks. The wet thickness of the free-standing films was approximately 0.7 mm. The free-standing films were used for the investigation of latex polymer structure, namely the chemical composition, the glass transition temperature (Tg) and the degree of crosslinking. In the case of the films based on the latexes comprising the nanostructured ZnO additive, the distribution and the actual content of ZnO-based nanoparticles inside the coating films were also determined. In addition, the free-standing films were used for the testing of water absorption and antibacterial activity of the latex films.
The chemical structure of the prepared latex polymers was studied employing a Fourier transform infrared (FT-IR) spectroscopy on a secondary module iZ10 of Nicolet iN10 instrument (Thermo Fisher Scientific, Waltham, MA, USA). The infrared spectra were obtained by attenuated total reflectance (ATR) on a built-in all-reflective diamond crystal in a range from 4000 cm−1 to 525 cm−1 using the resolution of 4 cm−1 (64 scans per spectrum). Tg of dried latex polymers was measured by means of differential scanning calorimetry (DSC) using a Pyris 1 DSC instrument (Perkin-Elmer, Waltham, MA, USA). The measurements were performed under N2 atmosphere at a heating rate of 10 °C/min from −50 to 120 °C and the second heating curve was used for Tg determination.
The degree of crosslinking introduced into latex polymers was evaluated according to gel content and crosslink density. The gel content was determined by the extraction in a Soxhlet extractor with tetrahydrofuran (THF) for 24 h according to CSN EN ISO 6427 [41]. The crosslink density was evaluated from swelling experiments performed on dry gel polymer samples (around 0.2 g) which were immersed in toluene at 50 °C for one week. A swelling time of one week was chosen on the basis of the test results on several samples, which manifested no significant changes after one week of immersion in toluene. At the end of the immersion period, the sample was removed, rapidly blotted with tissue and transferred to a weighing bottle to obtain the swollen weight of the sample. Equations (4)–(7) [42], employing the theory of Flory and Rehner [43], were used to calculate the average molecular weight between crosslinks (Mc) and the crosslink density (expressed as moles of crosslinks per cm3 of a polymer network), as given in the following:
M c   =   V 1 ρ p [ ϕ 1 3     ϕ / 2 ] l n 1 ϕ +   ϕ +   χ ϕ 2
ϕ   =   W p ρ s W p ρ s   +   W s ρ p
χ   =   0.34   +   V 1 RT δ 1     δ 2 2
Crosslink density = ρp/Mc
where V1 is the molar volume of toluene (106.3 cm3/mol); ρp is the density of polymer that was calculated to be 1.103 g/cm3 for the BA/MMA/MAA (60/37/3 by weight) copolymer from 1.06, 1.18 and 1.015 g/cm3 for poly(BA), poly(MMA) and poly(MAA), respectively; φ is the volume fraction of the gel polymer in the swollen gel; Wp and Ws are the weight fractions of the gel polymer and solvent (toluene) in the swollen gel, respectively; ρs is the density of solvent (0.8669 g/cm3); χ is the polymer and solvent interaction parameter; δ1 is the solubility parameter of polymer that was calculated to be 9.135 (cal/cm3)1/2 for the BA/MMA/MAA (60/37/3 by weight) copolymer from 9.0, 9.3 and 9.8 (cal/cm3)1/2 for poly(BA), poly(MMA) and poly(MAA), respectively [44,45]; and δ2 is the solubility parameter of toluene, 8.9 (cal/cm3)1/2.
The actual content of the nanostructured ZnO additive, embedded in the dried latex coating films, was determined by means of an inductively coupled plasma optical emission spectrometry (ICP-OES) using a spectrometer INTEGRA XL 2 (GBC, Dandenong, Australia), equipped with a concentric nebulizer and a glass cyclonic spray chamber (both Glass Expansion, Dangenong, Australia). The ZnO additive concentration was calculated using the simplified assumption that all the determined zinc in the coating film was only in the form of ZnO.
The distribution of ZnO additive in the coating film was investigated by scanning electron microscopy (SEM) and micro-Raman spectroscopy 2D mapping. The SEM measurements were performed using a LYRA 3 scanning electron microscope (Tescan, Brno, Czech Republic). Observations were conducted on fractured surfaces at the accelerating voltage of 5 kV. Samples were placed on carbon tape and carbon coated with a 20 nm thick layer. The 2D mapping on coating surfaces was performed using a micro-Raman spectrometer Nicolet DXR3xi (Thermo Fisher Scientific, Waltham, MA, USA) equipped with EMCCD detector. The laser beam (532 nm) was focused with a 100× objective (N.A = 0.90) and laser power of 5 mW. Maps were collected over an area of 50 × 100 µm using a 0.5 µm step size in a spectral range from 3400 to 50 cm1. Five exposures for each spectrum were recorded with the exposure time of 0.2 s. The analysis of collected data was done by a multivariate curve resolution.
The water absorption expressed in terms of water uptake by the latex films was tested by immersing specimens of the approximate dimensions 20 × 20 × 0.7 mm3 in distilled water at RT for 30 days. The soaked film was consequently removed from the water and the surface of the film was carefully dried by touching the polymer with filter paper. The water absorption (A) was calculated according Equation (8).
A = 100(wtw0)/w0
where w0 is the initial weight of a specimen before distilled water exposure and wt is the weight of the specimen after performing the immersion test.
The antibacterial activity of the coating films of the Series 4 (the samples D4, H4, D4_ZnO and H4_ZnO) was evaluated and compared using the modified ISO 22,196 method [46]. Four bacterial strains were used for the antimicrobial tests, namely, Staphylococcus aureus (S. Aureus, CCM 4516), Escherichia coli (E. coli, CCM 4517), Enterococcus faecalis (E. faecalis, CCM 3956) and Klebsiella pneumoniae (K. pneumoniae, CCM 4425). All the microbial strains were provided by the Czech Collection of Microorganisms (CCM, Brno, Czech Republic). The coating films were cut into squares of 25 × 25 mm2, sterilized by UV radiation and then placed in sterile Petri dishes. This was followed by inoculation of the samples using 0.1 mL of a particular standardized bacteria suspension, namely S. aureus (6.3 × 106 cfu/mL), E. coli (7.5 × 106 cfu/mL), E. faecalis (5.1 × 106 cfu/mL) and K. pneumoniae (1.8 × 106 cfu/mL). The samples were covered by polypropylene foil (20 × 20 mm2) disinfected with 70% ethanol. Incubation of the inoculated samples was performed at 95% of relative humidity at 35 °C for 24 h. The polypropylene foil was then removed, and the coating samples were imprinted on Plate Count Agar (HIMEDIA Laboratories Pvt., Mumbai, India). Each coating sample was imprinted three times on different areas of the Plate Count Agar, and incubated at 35 °C for 24 h. The results were then read and the increase in the number of bacterial colonies was evaluated based on scaling from 0 to 5, where 0 represents the best antimicrobial effect, with no growth of bacteria colonies. All of these analyses were performed using three different coating samples to ensure reliable antibacterial efficiency results.

2.5. Preparation and Characterization of Coatings Cast on Glass Substrates

Liquid latexes were applied onto glass panels using a blade applicator. The thickness of the wet coatings was 120 µm. No coalescing agents were used. The coatings were air-dried at RT and relative humidity of 45 ± 5% for seven days. The coating films were evaluated for their gloss, transparency, solvent resistance, adhesion and water whitening. The dry coating thickness was measured using a three-point instrument (BYK-Gardner, Geretsried, Germany). The gloss of coatings was evaluated by a micro-TRI-gloss µ instrument (BYK-Gardner, Geretsried, Germany) using a gloss-measuring geometry at 60°. Coatings cast on glass panels coated with black matte paint (RAL 9005) were used for the gloss measurements. The transparency of the coatings was evaluated by light transmission (measuring the transmittance at the wavelength 500 nm) using a ColorQuest XE Spectrometer (Hunterlab, Reston, VA, USA). The solvent resistance was tested by methyl ethyl ketone (MEK) rubbing following ASTM D 4752. The adhesion of coatings was evaluated by means of the pull-off test according to ISO 4624 using an Elcometer 510 Automatic Adhesion Tester (Elcometer Instruments, Aalen, Germany). All the above-mentioned experiments were carried out at RT.
The water whitening of coatings was evaluated by measuring the change in transmittance at a fixed wavelength (500 nm, near the green light, which is most sensitive to the human eye) using a ColorQuest XE Spectrometer (Hunterlab, Reston, VA, USA). The coatings were immersed in distilled water at RT for 24 h, followed by the immediate measurement of the transmittance of the exposed coating film area. The extent of water whitening (W) was calculated according Equation (9).
W = 100(T0Tt)/T0
where T0 is the coating sample transmittance before immersion in distilled water and Tt is the sample transmittance after immersion in distilled water for 24 h.

3. Results and Discussion

3.1. Characterization of Latexes

The characteristic properties of the prepared latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition, are listed in Table 4. In all the polymerizations, regardless of the type of the used surfactant and polymer composition, stable latexes with low amounts of coagulum (<1 wt.%) were obtained. This fact indicates a sufficient latex colloidal stability during the synthesis in the case of using the ordinary or the polymerizable type of surfactant. The addition of nanostructured ZnO, carried out during the process of emulsion polymerization, was also shown not to increase significantly the coagulum content in the respective latexes, suggesting no markedly negative effect of the in situ ZnO addition on latex stability during the synthesis.
When comparing the conversion and particle size for the latexes of the same composition but differing in the type of surfactant, a higher monomer conversion and simultaneously a lower final particle size were achieved in the case of using the polymerizable surfactant. This phenomenon has already been reported in the relevant literature [47,48,49,50]. It has been ascribed to the increased number of micelles, leading to a higher polymerization rate and a higher number of smaller polymer particles generated during the polymerization process using the polymerizable surfactant.
In addition, the mutual effects of covalent crosslinking strategy, surfactant type, and addition of nanostructured ZnO on MFFT of latexes were also investigated (see Table 4). As expected, a perceptible increase in MFFT values appeared in the Series 3 and 4, both employing intraparticle covalent crosslinking via ALMA copolymerization, which is generally regarded to introduce rigidity into latex polymer particles [51]. It was also found that the latexes added with nanostructured ZnO exhibited increased MFFT values in comparison with the corresponding latexes without the ZnO additive. This effect may be ascribed to the occurrence of ionic interparticle crosslinking, proceeding in the course of the film-formation via carboxylic acid functional groups of emulsion polymer and dissolved Zn2+ ions. When comparing the MFFT values for the corresponding latexes differing in the surfactant type, no remarkable difference was determined.

3.2. Stability of Latexes

From a practical point of view, a favorable coating binder has to exhibit a satisfactory stability under various conditions. For that reason, we also tested the stability of the prepared liquid latexes. We used several methods, simulating risks during storing and application, to evaluate the impact of latex composition, including surfactant type and ZnO addition on the colloidal stability of latexes. In the case of the heat storage stability and the mechanical stability (see Table 5), all the latexes apart from the sample D4_ZnO were found to be stable. The results of freeze-thaw tests (Table 5) based on the freezing cycle at −5 °C indicated no coagulation and no significant change in the coating gloss for all the prepared latexes, whereas the freezing cycles at lower temperatures (−10 and −18 °C) were found to destabilize the latex samples polymerized with the ordinary surfactant and those added with nanostructured ZnO. This fact suggests a thin hydration surface layer around polymer particles of these latexes, which caused interparticle coalescence resulting in coagulation. It was also found that the presence of keto-hydrazide chemistry improved the freeze-thaw stability significantly. (Notice all the samples of the Series 2 which manifested an excellent freeze-thaw stability.) This phenomenon can be explained firstly by thickening of the hydration layer (which remains unfrozen) around the particles due to the surface-bound hydrophilic groups of copolymerized DAAM; secondly by the presence of dissolved ADH that could decrease the freezing point of water.
The calcium ion stability test (see Table 6) revealed a higher resistance to the electrolyte in the case of the ZnO added latexes. This effect was observed in all the series, where the D1_ZnO and D3_ZnO samples even exhibited the stability at the highest CaCl2 concentration. It was also shown that there was no significant difference in the electrolyte stability between the D-labelled and H-labelled samples. Increased calcium ion stability in the case of the ZnO added latexes was in all probability caused by a more negatively charged electric layer on the surface of the latex particles. As ZnO exhibits a limited water solubility, resulting in dissociated Zn2+ and OH ions, a certain number of the carboxylic groups on the surface of latex particles was ionized, which increased the negative charge of the electric layer surrounding the latex particles.

3.3. Characterization of Coating Structure

The chemical structure of the coating materials prepared from latexes, differing in the covalent crosslinking strategy, surfactant type and addition of nanostructured ZnO, was analyzed using FT-IR spectroscopy. Representative absorption spectra of coating materials of the Series 1 (providing no crosslinking) and Series 2 (providing interparticle keto-hydrazide crosslinking) are presented in Figure 1 and Figure 2. It can be seen that there are no significant differences between the corresponding samples varying in the surfactant type. All the spectra display a typical pattern of acrylic polymers with a characteristic strong absorption band of the C=O bond at 1728 cm−1 assigned to the carboxylic acid ester group. The characteristic absorption band around 1530–1540 cm−1 appearing in the spectra of all the materials of the Series 2 can be attributed to bending vibration of the N–H bonds and proves that DAAM was copolymerized with the acrylic monomers. All the polymers of the Series 2 further manifested a weak absorption band around 1640 cm−1 corresponding in all probability to the N=C vibration, which indicates that the keto-hydrazide interparticle self-crosslinking reaction proceeded in the latex polymers [52]. In addition, a broad absorption band of the O–H bond around 3250 cm−1 appeared in all the latex samples, being more significant in the case of the ZnO addition. This absorption band can be assigned to the protonated carboxylic group (from MAA) and hydrogen bonded water. A higher intensity of this band in the case of the samples comprising the ZnO additive may be attributed to a low solubility of ZnO in water leading to dissolved Zn(OH)2, which caused a higher degree of ionization of the carboxylic groups. The ionized carboxylic groups were thus responsible for imparting a higher amount of molecularly bonded (hydroplasticizing) water in latex copolymers [53,54]. Moreover, a weak absorption band corresponding to the carboxylate anion νa(COO) appeared around 1600–1610 cm−1 in the spectra of the D_ZnO− and H_ZnO− labelled samples, which proves again the presence of carboxylic acid salts and also suggests the existence of ionic crosslinking between the carboxylic groups and dissociated Zn2+ cations. The samples of both series also exhibited a weak absorption band of the sulfate anion νa(SO42−) around 620 cm−1 which corresponds to the presence of the sulfate-based emulsifiers and also to the persulfate initiator producing sulfuric acid as the result of the reaction with water. A higher intensity of this band was manifested in the case of the D_ZnO− and H_ZnO− labelled samples of both series, which suggests that the original ZnO nanoparticles were converted into zinc sulfate hydroxide (ZnSO4·3Zn(OH)2) to some extent, after being subject to the polymerization conditions, as has been evidenced in our previous work [21]. The chemical conversion of nanostructured ZnO in the strongly acidic aqueous environment of the polymerization system (caused by sulfuric acid generated from the ammonium persulfate initiator) can be represented by Equation (10) [21,55].
ZnO + H2SO4 + 2H2O = ZnSO4 + 3Zn(OH)2
The effects of crosslinking strategy, surfactant type and addition of nanostructured ZnO on the structure of coating materials were also studied from the point of view of polymer Tg (see Table 7). As expected, the presence of the covalent crosslinking was connected with the increase of Tg of latex polymers. This effect was the most pronounced in the Series 4, where both the intra- and interparticle covalent crosslinking strategies were introduced. Focusing on the surfactant type, it can be observed that the H-labelled samples exhibited a higher Tg in contrast to the corresponding D-labelled samples, which may be related to the higher molar mass of the latex polymers synthesized with the polymerizable surfactant. This assumption can be deduced from the results of particle size and conversion (see Table 4) showing that the latexes polymerized using the polymerizable surfactant exhibited smaller particles and a higher monomer conversion in comparison with the corresponding latex samples based on the ordinary surfactant. A similar effect has been described for classic emulsion polymerization, where increased concentrations of the ordinary surfactant at a fixed level of initiator were applied. In these cases, the overall molar mass increased because more particles shared the same or even fewer number of radicals (due to the higher polymerization rate) [56]. With regard to the ZnO addition, the coating samples added with nanostructured ZnO displayed increased Tg values in comparison with the corresponding blank polymer samples (without ZnO additive), which indicates again the occurrence of ionic crosslinking.
Additionally, the structure of the coating films in terms of the degree of crosslinking introduced into latex materials was evaluated. The results of gel content, Mc (listed in Table 7) and crosslink density (depicted in Figure 3) confirmed that the most densely crosslinked polymer network was formed in the case of latex polymers of the Series 4, employing both inter- and intraparticle covalent crosslinking. A high level of crosslinking was also determined for the samples of the Series 3, which were crosslinked by ALMA copolymerization. However, the polymers of the Series 1 (without any proposed crosslinking strategy) also provided a certain level of crosslinking; minor in the case of the sample D1 and significant in the case of the sample H1. The presence of crosslinked structures in these samples can be explained by intermolecular chain transfer reactions to polymer as a result of BA copolymerization [57]. A comparison of the corresponding samples differing in the surfactant type revealed that a higher degree of crosslinking was always detected for the H-labelled polymers, which suggests again a higher molar mass of polymers prepared using the polymerizable surfactant. When comparing the ZnO-added and the blank samples of the same composition, it is evident that additional crosslinks were introduced into latex polymers due to the ZnO addition. We assume that carboxylic functionalities on the particle surface and also those buried in the particle interior were probably involved in ionic bonding, providing ionic inter- and intraparticle crosslinks. At this stage, it should be mentioned that the ionic intraparticle crosslinks are believed to have proceeded already during the polymer synthesis after dosing the nanostructured ZnO into the reactor.
The structure of latex coatings was also examined from the point of view of the actual content and distribution of the nanostructured ZnO additive in the coating films. The content of the inorganic additive in the respective coating films was determined by the ICP-OES measurements of the zinc concentration (see Table 7). The analysis revealed that the procedure of nanostructured ZnO incorporation, performed during the process of latex synthesis, provided a comparable portion of the inorganic additive in all the ZnO-based latexes (0.9–1.0 wt.% in the dried coating films). In addition, the morphology and location of the ZnO additive inside the representative coating films D4_ZnO and H4_ZnO was investigated using SEM (see Figure 4). For better resolution, SEM micrographs of cryofractures of the coating films were taken in both secondary electron (SE) mode (illustrating a topographical view) and backscattered electron (BSE) mode (showing an elementary contrast of the surface). The inorganic additive was shown to be present in the form of distinct nanosized spherical particles (apparently having the predominant chemical nature of ZnSO4·3Zn(OH)2, which were distributed distinctly and regularly inside the polymer matrix, which predetermines the transparent and glossy nature of the resulting coatings. Concurrently, the presence and distribution of the ZnO additive on the coating surface of the samples D4_ZnO and H4_ZnO was monitored using micro-Raman spectroscopy 2D mapping. Both samples displayed the same character and the representative illustration of the D4_ZnO coating surface is displayed in Figure 5. Two-dimensional Raman mapping revealed clearly that the surface of latex coatings comprising the ZnO-based additive was homogeneous without micrometer-sized inorganic aggregates.

3.4. Coating Properties

The properties of coatings differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition are presented in Table 8. The thickness of the dried coatings drawn down on glass substrates was about 50 µm. All the prepared coatings exhibited a high gloss and transparency without any pronounced difference. We can therefore assume that no significant desorption, migration and aggregation of surfactant molecules and ionic species at the film-air interface occurred (the risk considered especially for the coatings comprising the ordinary surfactant). When comparing MEK resistance of coatings, the results indicated that the solvent resistance wasn’t increased markedly by covalent intra- or/and interparticle crosslinking. In contrast, ZnO additive introducing ionic crosslinks was shown to be a very effective tool for obtaining a high solvent resistance. The coating samples D2_ZnO and D4_ZnO even passed the MEK test without failure, which was in all probability caused by the combination of ionic and covalent interparticle crosslinking. The results of the MEK test also revealed that the D_ZnO− labelled coating samples exhibited a higher MEK resistance in comparison with the corresponding H_ZnO− labelled coating samples. This finding may be related to the enhanced molar mass and rigidity of the H-labelled latex polymers (prepared by the polymerizable surfactant). As the ionic crosslinks between the carboxylic groups and zinc ions are supposed to be formed to some extent already during the synthesis of the emulsion polymer, we can assume that more rigid latex particles composed of longer polymer chains linked to each other by ionic bonds (intraparticle crosslinks) were formed in the case of the H_ZnO− labelled latexes (resembling the covalent intraparticle crosslinking by ALMA copolymerization). As the result, a porous film structure was apparently formed due to the deteriorated particle deformation and suppressed polymer chain interdiffusion, allowing the solvent molecules to penetrate more easily inside the coating. This effect can be observed more distinctly in the case of a poor MEK resistance for coatings of the Series 3, composed of ALMA-crosslinked latex particles.
In addition, the adhesion of the coatings to a glass substrate was evaluated (see Figure 6). Focusing on the surfactant type, the results confirmed that the use of the polymerizable surfactant provided improved adhesion of coatings, as has been expected. This effect can be attributed to the covalent attachment of surfactant molecules on the surface of latex particles, disabling their desorption and migration at the substrate–film interface. When comparing the coating adhesion in terms of the covalent crosslinking strategy, the best adhesion to the glass substrate was detected for the coatings of the Series 2 (inter-particularly crosslinked by the keto-hydrazide reaction). The reason is in all probability related to the presence of the additional polar functionalities incorporated into the molecular structure of the latex coatings by the keto-hydrazide crosslinks which together with the pendant carboxylic groups on polymer backbone provided secondary interactions between the acrylic polymer chains and the glass substrate. This effect was also manifested in the case of coatings of the Series 4 and explains their enhanced adhesion in comparison with the corresponding coatings of the Series 3. The weaker coating adhesion of the Series 4 in contrast to Series 2 can be attributed to the fixation of polymer chains by ALMA crosslinking, which hindered the orientation of polar functionalities towards the film–substrate interface during film-formation. Focusing on the effect of ZnO addition causing further polymer chain fixation by ionic crosslinks, the coatings added with nanostructured ZnO were surprisingly found to provide significantly improved adhesion in comparison with the corresponding coating samples without the ZnO additive. The improvement of coating adhesion can be explained by ionization of the carboxylic groups by dissolved zinc hydroxide, which resulted in stronger secondary interactions and bonding to glass substrate.
Water resistance of the latex coating films was evaluated in terms of water absorption (i.e., water uptake) and water whitening (symbolized by a decrease in transmittance). It was evidenced that water resistance of latex coatings can be enhanced markedly by introducing crosslinking [1,2,3,4,21,58]. The improved water resistance is usually attributed to the enhanced stiffness of crosslinked polymer, which restricts the influx of water and does not allow water domains to grow. In the case of water whitening, it is believed that only water domains exceeding a certain size within the film are responsible for the water whitening effect (light of longer wavelengths becomes more strongly scattered as the domains grow larger) [54,59]. The latex films composed of densely crosslinked polymers, after being exposed in water, usually exhibit low water uptake and contain smaller water domains, therefore they are observed less cloudy to the human eye. In contrast, the noncrosslinked or slightly crosslinked latex films generally suffer from high water swelling and pronounced water whitening.
The results of water absorption and water whitening measurements are demonstrated in Figure 7 and Figure 8, respectively. It is obvious that similar trends can be observed for both experiments. Focusing on the covalent crosslinking strategy, the assumptions were fulfilled; the higher the crosslink density (see the results presented in Figure 3), the increased water resistance of the coating films. As expected, the coatings of the Series 1 without any covalent crosslinking were the most water-sensitive, while the most densely crosslinked coatings of the Series 4, combining the intra- and interparticle covalent crosslinking, were shown to be the most water-resistant in terms of both water uptake and water whitening. When comparing the results from the point of view of the surfactant type, the coatings comprising the polymerizable surfactant always provided higher water resistance, which is in all probability related to the smaller interstitial areas (filled only with initiator-based salts) in contrast to the larger interstices (filled also with desorbed surfactant molecules) in the case of the ordinary surfactant-based coatings, leading to decreased osmotic pressure (as the main water penetration driving force) and suppressed influx of water. With regard to the ZnO addition, it can be stated that the coating films added with nanostructured ZnO displayed increased water resistance in comparison with the corresponding blank coatings (without ZnO additive), which corresponds to the increased crosslink density due to the contribution of ionic bonds. However, this claim was contradicted by the results of water absorption of the H_ ZnO− labelled samples of the Series 1 and 2 that exhibited higher water absorption in comparison with the blank samples of the same composition. The conflict between the results can be explained by the dissolution of a certain portion of nanostructured ZnO, providing a higher amount of ionically charged inorganic components, trapped in interstices between coalesced latex particles. It should be noted at this point that the latexes of the Series 4, namely the samples H4, D4_ZnO and H4_ZnO were found to provide highly water whitening resistant coatings, which was caused by the occurrence of intra- and intercovalent and ionic crosslinking, resulting in a highly dense polymer network.
Nanostructured ZnO has been successfully employed as the antibacterial agent in organic coatings [21,60,61,62,63]. Its antibacterial effect is still not fully understood, although several mechanisms have been proposed, namely (i) direct contact of nanoparticles with cell walls, resulting in destructing bacterial cell integrity [64], (ii) liberation of antimicrobial Zn2+ ions [65], and (iii) formation of reactive oxygen species including hydrogen peroxide [66,67]. The antibacterial efficiency of the latex coating films of the Series 4, providing the best water resistance, was tested and compared against major food and hospital pathogens, namely S. aureus, E. faecalis (gram-positive bacteria) and E. coli, K. pneumoniae (Gram-negative bacteria). Testing revealed (see Table 9) the strong antibacterial activity of all the ZnO-added coatings against S. aureus, E. coli and K. pneumoniae and moderate antibacterial efficiency against E. faecalis, the latter effect being probably related to the exceptional antioxidative mechanisms in E. faecalis [68]. Considering almost the same content of the ZnO additive in the respective coatings, it can be supposed that similar results would be also obtained in the case of the Series 1–3. It can be concluded that the addition of nanostructured ZnO provided a significant antibacterial activity of coatings and, therefore, the nanostructured ZnO can be considered not only as an effective ionic crosslinking additive but also as an antibacterial agent.

4. Conclusions

The present study is devoted to the development of latex-based water-resistant acrylic coatings suitable for high-performance applications. To decrease the water sensitivity of coating films, we investigated the combined effects of covalent crosslinking (ALMA-based intraparticle and/or keto-hydrazide-based interparticle), surfactant type (nonpolymerizable and polymerizable), and ionic crosslinking (nanostructured ZnO additive). The addition of nanostructured ZnO, carried out during the process emulsion polymerization, proved to be an effective tool for introducing ionic crosslinks into latex polymers without deteriorating the coating gloss and transparency. Inter- and intraparticle ionic crosslinks were in all probability formed via surface and interior carboxylic functionalities and Zn2+ ions. The results of water resistance testing (focused on water absorption and water whitening) confirmed that the coatings comprising the polymerizable surfactant exhibited improved water resistance in contrast to the ordinary surfactant-based coatings, in all probability due to the presence of smaller interstitial areas (filled only with initiator-based salts), leading to decreased osmotic pressure and suppressed influx of water. It was further proved that introducing of the covalent crosslinking provided increased water-resistance of coatings, the higher the crosslink density, the more water-resistant coating. In addition, the ZnO additive was also shown to introduce improved water resistance of coatings, which corresponds to the increased crosslink density due to the contribution of ionic bonds. Highly water-resistant coatings were formed regardless of the surfactant type in case of a concurrent employment of three crosslinking strategies, namely keto-hydrazide crosslinking, ALMA copolymerization and ZnO-based ionic crosslinking. Nevertheless, the application of the polymerizable surfactant was shown to improve significantly coating adhesion, apparently due to the covalent attachment of surfactant molecules to latex particles, disabling their desorption and migration at the substrate–film interface. Moreover, ZnO addition was shown to provide pronouncedly improved MEK resistance and significant antibacterial activity of the coating films. It can be concluded that environmentally friendly coating binders using nanostructured ZnO in the role of an ionic and antibacterial additive were developed, providing water- and solvent-resistant hygienic coatings, suitable for the protection of various materials.

Author Contributions

Conceptualization, J.M.; methodology, J.M., A.K., and J.V.; validation, J.M.; investigation, J.M., P.M., and D.S.; writing—original draft preparation, J.M.; writing—review and editing, J.Š.; visualization, D.S., P.M., and S.Š.; supervision, J.M.; project administration, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Education, Youth and Sports of the Czech Republic (project LM2018103) and the Czech Academy of Sciences, Institute of Theoretical and Applied Mechanics (project RVO 68378297). One of us, J.V., acknowledges the support of an internal grant from TBU in Zlín (IGA/CPS/2020/001) financed from funds of specific academic research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Roman Svoboda (University of Pardubice, Faculty of Chemical Technology, Department of Physical Chemistry) for the DSC measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Winnik, M.A. Interdiffusion and crosslinking in thermoset latex films. J. Coat. Technol. 2002, 74, 49–63. [Google Scholar] [CrossRef]
  2. Taylor, J.W.; Winnik, M.A. Functional latex and thermoset latex films. Jct Res. 2004, 1, 163–190. [Google Scholar] [CrossRef]
  3. Machotová, J.; Černošková, E.; Honzíček, J.; Šňupárek, J. Water sensitivity of fluorine-containing polyacrylate latex coatings: Effects of crosslinking and ambient drying conditions. Prog. Org. Coat. 2018, 120, 266–273. [Google Scholar] [CrossRef]
  4. Ruckerova, A.; Machotova, J.; Svoboda, R.; Pukova, K.; Bohacik, P.; Valka, R. Ambient temperature self-crosslinking latices using low generation PAMAM dendrimers as inter-particle crosslinking agents. Prog. Org. Coat. 2019, 119, 91–98. [Google Scholar] [CrossRef]
  5. Tillet, G.; Boutevin, B.; Ameduri, B. Chemical reactions of polymer crosslinking and post-crosslinking at room temperature. Prog. Polym. Sci. 2011, 36, 191–217. [Google Scholar] [CrossRef]
  6. Gonzáles, I.; Asua, J.M.; Leiza, J.R. Crosslinking in acetoacetoxy functional waterborne crosslinkable latexes. Macromol. Symp. 2006, 243, 53–62. [Google Scholar] [CrossRef]
  7. Nakayama, Y. Development of novel aqueous coatings which meet the requirements of ecology-conscious society: Novel cross-linking system based on the carbonyl-hydrazide reaction and its applications. Prog. Org. Coat. 2004, 51, 280–299. [Google Scholar] [CrossRef]
  8. Koukiotis, C.G.; Karabela, M.M.; Sideridou, I.D. Mechanical properties of films of latexes based on copolymers BA/MMA/DAAM and BA/MMA/VEOVA-10/DAAM and the corresponding self-crosslinked copolymers using the adipic acid dihydrazide as crosslinking agent. Prog. Org. Coat. 2012, 75, 106–115. [Google Scholar] [CrossRef]
  9. Koukiotis, C.; Sideridou, I.D. Synthesis and characterization of latexes based on copolymers BA/MMA/DAAM and BA/MMA/VEOVA-10/DAAM and the corresponding 1K crosslinkable binder using the adipic acid dihydrazide as crosslinking agent. Prog. Org. Coat. 2010, 69, 504–509. [Google Scholar] [CrossRef]
  10. Zhang, X.; Liu, Y.; Huang, H.; Li, Y.; Chen, H. The diacetone acrylamide crosslinking reaction and its control of core–shell polyacrylate lattices at ambient temperature. J. Appl. Polym. Sci. 2012, 123, 1822–1832. [Google Scholar] [CrossRef]
  11. Li, H.; Kan, C.; Du, Y.; Liu, D. Effects of the amount of diacetone acrylamide on the properties of styrene-acrylic copolymer latexes and their films. Polym. Prep. 2002, 43, 413–414. [Google Scholar]
  12. Kessel, N.; Illsley, D.R.; Keddie, J.L. The diacetone acrylamide crosslinking reaction and its influence on the film formation of an acrylic latex. J. Coat. Technol. Res. 2008, 5, 285–297. [Google Scholar] [CrossRef] [Green Version]
  13. Matsuda, H. Ionic crosslinking of carboxyl-terminated liquid rubbers with metal oxides. J. Appl. Polym. Sci. 1979, 24, 811–826. [Google Scholar] [CrossRef]
  14. Zosel, A. Mechanical properties of films from polymer lattices. Polym. Adv. Techol. 2003, 6, 263–269. [Google Scholar] [CrossRef]
  15. Kan, C.S.; Blackson, J.H. Effect of Ionomeric behavior on the viscoelastic properties and morphology of carboxylated latex films. Macromolecules 1996, 29, 6853–6864. [Google Scholar] [CrossRef]
  16. Schuman, T.; Wikström, M.; Rigdahl, M. Dispersion coating with carboxylated and cross-linked styrene–butadiene lattices. 1. Effect of some polymer characteristics on film properties. Prog. Org. Coat. 2004, 51, 220–227. [Google Scholar] [CrossRef]
  17. Richard, J.; Maquet, J. Dynamic micromechanical investigations into particle/particle interfaces in latex films. Polymer 1992, 33, 4164–4173. [Google Scholar] [CrossRef]
  18. Lee, D.I. The effects of latex coalescence and interfacial crosslinking on the mechanical properties of latex films. Polymer 2005, 46, 1287–1293. [Google Scholar] [CrossRef]
  19. Pinprayoon, O.; Saiani, A.; Groves, R.; Saunders, B.R. Particulate ionomer films prepared from dispersions of crosslinked polymer colloids: A structure–property study. J. Colloid Interface Sci. 2009, 336, 73–81. [Google Scholar] [CrossRef]
  20. Machotova, J.; Ruckerova, A.; Bohacik, P.; Pukova, K.; Kalendova, A.; Palarcik, J. High-performance one-pack ambient cross-linking latex binders containing low-generation PAMAM dendrimers and ZnO nanoparticles. J. Coat. Technol. Res. 2018, 15, 1167–1179. [Google Scholar] [CrossRef]
  21. Machotová, J.; Kalendová, A.; Zlámaná, B.; Šňupárek, J.; Palarčík, J.; Svoboda, R. Waterborne coating binders based on self-crosslinking acrylic latex with embedded inorganic nanoparticles: A comparison of nanostructured ZnO and MgO as crosslink density enhancing agents. Coatings 2020, 10, 339. [Google Scholar] [CrossRef] [Green Version]
  22. Dankova, M.; Kalendova, A.; Machotova, J. Waterborne coatings based on acrylic latex containing nanostructured ZnO as an active additive. J. Coat. Technol. Res. 2020, 17, 517–529. [Google Scholar] [CrossRef]
  23. Morsi, R.E.; Labena, A.; Khamis, E.A. Core/shell (ZnO/polyacrylamide) nanocomposite: In-situ emulsion polymerization, corrosion inhibition, anti-microbial and anti-biofilm characteristics. J. Taiwan Inst. Chem. Eng. 2016, 63, 512–522. [Google Scholar] [CrossRef]
  24. Tang, E.; Cheng, G.; Pang, X.; Ma, X.; Xing, F. Synthesis of nano-ZnO/poly(methyl methacrylate) composite microsphere through emulsion polymerization and its UV-shielding property. Colloid Polym. Sci. 2006, 284, 422–428. [Google Scholar] [CrossRef]
  25. Machotová, J.; Kalendová, A.; Voleská, M.; Steinerová, D.; Pejchalová, M.; Knotek, P.; Zárybnická, L. Waterborne hygienic coatings based on self-crosslinking acrylic latex with embedded inorganic nanoparticles: A comparison of nanostructured ZnO and MgO as antibacterial additives. Prog. Org. Coat. 2020, 147, 105704. [Google Scholar] [CrossRef]
  26. Kang, K.; Kan, C.Y.; Du, Y.; Liu, D.S. Synthesis and properties of soap-free poly(methyl methacrylate-ethyl acrylate-methacrylic acid) latex particles prepared by seeded emulsion polymerization. Eur. Polym. J. 2005, 41, 439–445. [Google Scholar] [CrossRef]
  27. Xu, G.; Deng, L.; Wen, X.; Pi, P.; Zheng, D.; Cheng, J.; Yang, Z. Synthesis and characterization of fluorine-containing poly-styrene-acrylate latex with core–shell structure using a reactive surfactant. J. Coat. Technol. Res. 2011, 8, 401–407. [Google Scholar] [CrossRef]
  28. Aramendia, E.; Barandiaran, M.J.; Grade, J.; Blease, T.; Asua, J.M. Polymerization of high-solids-content acrylic latexes using a nonionic polymerizable surfactant. J. Polym. Sci. Part A Polym. Chem. 2002, 40, 1552–1559. [Google Scholar] [CrossRef]
  29. Yang, S.F.; Xiong, P.T.; Gong, T.; Lu, D.P.; Guan, R. St-BA copolymer emulsion prepared by using novel cationic maleic dialkyl polymerizable emulsifier. Eur. Polym. J. 2005, 41, 2973–2979. [Google Scholar] [CrossRef]
  30. Filet, A.; Guillot, J.; Hamaide, T.; Guyot, A. Emulsion copolymerization of styrene with a nonionic styrenic polymerizable surfactant. Polym. Adv. Techol. 1995, 6, 465–472. [Google Scholar] [CrossRef]
  31. Chern, C.S.; Chen, Y.C. Semibatch emulsion polymerization of butyl acrylate stabilized by a polymerizable surfactant. Polym. J. 1996, 28, 627–632. [Google Scholar] [CrossRef] [Green Version]
  32. Schoonbrood, H.A.S.; Asua, J.M. Reactive surfactants in heterophase polymerization. 9. Optimum surfmer behaviour in emulsion polymerization. Macromolecules 1997, 30, 6034–6041. [Google Scholar] [CrossRef]
  33. Guyot, A. Advances in reactive surfactants. Adv. Colloid Interface Sci. 2004, 108–109, 3–22. [Google Scholar] [CrossRef] [PubMed]
  34. Hellgren, A.C.; Weissenborn, P.; Holmberg, K. Surfactants in water-borne paints. Prog. Org. Coat. 1999, 35, 79–87. [Google Scholar] [CrossRef]
  35. Chern, C.S.; Chen, Y.C. Stability of the polymerizable surfactant stabilized latex particles during semibatch emulsion polymerization. Colloid Polym. Sci. 1997, 275, 124–130. [Google Scholar] [CrossRef]
  36. Xiao, X.; Wang, Y. Emulsion polymerization of fluorinated acrylate in the presence of a polymerizable emulsifier. Colloids Surf. A Physicochem. Eng. Asp. 2009, 348, 151–156. [Google Scholar] [CrossRef]
  37. Wu, Z.; Zhong, C.Z.; Song, Y.; Zhang, Y.Q. Preparation and characterization of novel acrylate emulsiones with ketone-hydrazide crosslinking structure based on application of reactive emulsifier. Adv. Mat. Res. 2013, 833, 335–338. [Google Scholar] [CrossRef]
  38. Fox, T.G.; Flory, P.J. 2nd-Order transition temperatures and related properties of polystyrene. 1. Influence of molecular weight. J. Appl. Phys. 1950, 21, 581–591. [Google Scholar] [CrossRef]
  39. Ling, H.; Junyan, L. Synthesis, modification and characterization of core–shell fluoroacrylate copolymer latexes. J. Fluor. Chem. 2008, 129, 590–597. [Google Scholar] [CrossRef]
  40. Kumthekar, V.; Kolekar, S. Attributes of the latex emulsion processing and its role in morphology and performance in paints. Prog. Org. Coat. 2011, 72, 380–386. [Google Scholar] [CrossRef]
  41. Kan, C.Y.; Liu, D.S.; Kong, X.Z.; Zhu, X.L. Study on the preparation and properties of styrene–butyl acrylate–silicone copolymer lattices. J. Appl. Polym. Sci. 2001, 82, 3194–3200. [Google Scholar] [CrossRef]
  42. Sperling, L.H. Introduction to Physical Polymer Science, 4th ed.; John Wiley & Sons: Philadelphia, PA, USA, 2005; pp. 427–473. [Google Scholar]
  43. Flory, P.J.; Rehner, J. Statistical mechanics of cross-linked polymer networks II. Swelling. J. Chem. Phys. 1943, 11, 521–526. [Google Scholar] [CrossRef]
  44. Tobing, S.; Klein, A. Molecular parameters and their relation to the adhesive performance of acrylic pressure-sensitive adhesives. J. Appl. Polym. Sci. 2001, 79, 2230–2244. [Google Scholar] [CrossRef]
  45. Vandenburg, H.J.; Clifford, A.A.; Bartle, K.D.; Carlson, R.E.; Caroll, J.; Newton, I.D. A simple solvent selection method accelerated solvent extraction of additives from polymers. Analyst 1999, 124, 1707–1710. [Google Scholar] [CrossRef]
  46. Habib, S.; Lehocky, M.; Vesela, D.; Humpolíček, P.; Krupa, I.; Popelka, A. Preparation of progressive antibacterial LDPE surface via active biomolecule deposition approach. Polymers 2019, 11, 1704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Aguirreurreta, Z.; Cal, J.C.; Leiza, J.R. Anionic polymerizable surfactants and stabilizers in emulsion polymerization: A comparative study. Macromol. React. Eng. 2017, 11, 1600033. [Google Scholar] [CrossRef] [Green Version]
  48. Liu, J.; Chew, C.H.; Gan, L.M.; Teo, W.K.; Gan, L.H. Synthesis of monodisperse polystyrene microlatexes by emulsion polymerization using a polymerizable surfactant. Langmuir 1997, 13, 4988–4994. [Google Scholar] [CrossRef]
  49. Mekki, S.; Saïdi-Besbes, S.; Elaïssari, A.; Valour, J.P.; Derdour, A. Novel polymerizable surfactants: Synthesis and application in the emulsion polymerization of styrene. Polym. J. 2010, 42, 401–405. [Google Scholar] [CrossRef]
  50. Urquiola, M.B.; Dimonie, V.L.; Sudol, E.D.; El-Aasser, M.S. Emulsion polymerization of vinyl acetate using a polymerizable surfactant. I. Kinetic studies. J. Appl. Polym. Sci. Part A Polym. Chem. 1992, 30, 2619–2629. [Google Scholar] [CrossRef]
  51. Matsumoto, A.; Kodama, K.; Aota, H.; Capek, I. Kinetics of emulsion crosslinking polymerization and copolymerization of allyl methacrylate. Eur. Polym. J. 1999, 35, 1509–1517. [Google Scholar] [CrossRef]
  52. Pi, P.; Wang, W.; Wen, X.; Xu, S.; Cheng, J. Synthesis and characterization of low-temperature self-crosslinkable acrylic emulsion for PE film ink. Prog. Org. Coat. 2015, 81, 66–71. [Google Scholar] [CrossRef]
  53. Tsavalas, J.G.; Sundberg, D.C. Hydroplasticization of polymers: Model predictions and application to emulsion polymers. Langmuir 2010, 26, 6960–6966. [Google Scholar] [CrossRef]
  54. Jiang, B.; Tsavalas, J.G.; Sundberg, D.C. Water whitening of polymer films: Mechanistic studies and comparison between water and solvent borne films. Prog. Org. Coat. 2017, 105, 56–66. [Google Scholar] [CrossRef]
  55. Leitner, J.; Sedmidubský, D. Preparation, properties and utilization of nanostructured ZnO. Chem. Listy 2017, 110, 406–417. [Google Scholar]
  56. El-Aasser, M.S.; Tang, J.; Wang, X.; Daniels, E.S.; Dimonie, V.; Sudol, E.D. Advances in emulsion polymerization for coatings applications: Latex blends and reactive surfactants. J. Coat. Technol. 2001, 73, 51–63. [Google Scholar] [CrossRef]
  57. Podzimek, S.; Machotova, J.; Snuparek, J.; Vecera, M.; Prokupek, L. Characterization of molecular structure of acrylic copolymers prepared via emulsion polymerization using A4F-MALS technique. J. Appl. Polym. Sci. 2014, 131, 11178–11185. [Google Scholar] [CrossRef]
  58. Tian, Y.; Du, E.; Abdelmola, F.; Qiang, Y.; Carlsson, L.A. Rapid Characterization of water diffusion in polymer specimens using a droplet-based method. Langmuir 2020, 36, 7309–7314. [Google Scholar] [CrossRef]
  59. Liu, Y.; Gajawicz, A.M.; Rodin, V.; Soer, W.; Scheerder, J.; Satgurunathan, G.; McDonald, P.J.; Keddie, J.L. Explanations for water whitening in secondary dispersion and emulsion polymer films. J. Polym. Sci. Part B Polym. Phys. 2016, 54, 1658–1674. [Google Scholar] [CrossRef] [Green Version]
  60. Schwartz, V.B.; Thétiot, F.; Ritz, S.; Pütz, S.; Choritz, L.; Lappas, A.; Förch, R.; Landfester, K.; Jonas, U. Antibacterial surface coatings from zinc oxide nanoparticles embedded in poly(n-isopropylacrylamide) hydrogel surface layers. Adv. Funct. Mater. 2012, 22, 2376–2386. [Google Scholar] [CrossRef]
  61. Li, J.H.; Hong, R.Y.; Li, M.Y.; Li, H.Z.; Zheng, Y.; Ding, J. Effects of ZnO nanoparticles on the mechanical and antibacterial properties of polyurethane coatings. Prog. Org. Coat. 2009, 64, 504–509. [Google Scholar] [CrossRef]
  62. Seo, J.; Jeon, G.; Jang, E.S.; Khan, S.B.; Han, H. Preparation and properties of poly(propylene carbonate) and nanosized ZnO composite films for packaging applications. J. Appl. Polym. Sci. 2011, 122, 1101–1108. [Google Scholar] [CrossRef]
  63. Chaurasia, V.; Chand, N.; Bajpai, S.K. Water sorption properties and antimicrobial action of zinc oxide nanoparticles-loaded cellulose acetate films. J. Macromol. Sci. Pure Appl. Chem. 2010, 47, 309–317. [Google Scholar] [CrossRef]
  64. Brayner, R.; Ferrari-Iliou, R.; Brivois, N.; Djediat, S.; Benedetti, M.F.; Fievet, F. Toxicological impact studies based on Escherichia coli bacteria in ultrafine ZnO nanoparticles colloidal medium. Nano Lett. 2006, 6, 866–870. [Google Scholar] [CrossRef]
  65. Pasquet, J.; Chevalier, Y.; Pelletier, J.; Couval, E.; Bouvier, D.; Bolzinger, M.A. The contribution of zinc ions to the antimicrobial activity of zinc oxide. Colloid Surf. A Phys. Eng. Asp. 2014, 457, 263–274. [Google Scholar] [CrossRef]
  66. Sawai, J.; Shoji, S.; Igarashi, H.; Hashimoto, A.; Kokugan, T.; Shimizu, M.; Kojima, H. Hydrogen peroxide as an antibacterial factor in zinc oxide powder slurry. J. Ferment. Bioeng. 1998, 86, 521–552. [Google Scholar] [CrossRef]
  67. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on zinc oxide nanoparticles: Antibacterial activity and toxicity mechanism. Nano Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Szemes, T.; Vlkova, B.; Minarik, G.; Tothova, L.; Drahovska, H.; Turna, J.; Celec, P. On the origin of reactive oxygen species and antioxidative mechanisms in Enterococcus faecalis. Redox Rep. 2010, 15, 202–206. [Google Scholar] [CrossRef] [Green Version]
Figure 1. FT-IR spectra of D-labelled latex coatings of Series 1 and Series 2.
Figure 1. FT-IR spectra of D-labelled latex coatings of Series 1 and Series 2.
Coatings 11 00347 g001
Figure 2. FT-IR spectra of H-labelled latex coatings of Series 1 and Series 2.
Figure 2. FT-IR spectra of H-labelled latex coatings of Series 1 and Series 2.
Coatings 11 00347 g002
Figure 3. Crosslink density (expressed as moles of crosslinks per cm3 of a polymer network) of the coating films based on latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition.
Figure 3. Crosslink density (expressed as moles of crosslinks per cm3 of a polymer network) of the coating films based on latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition.
Coatings 11 00347 g003
Figure 4. SEM photographs of cryofractures of latex coatings taken in secondary electron (left) and backscattered electron mode (right): (a) blank sample D4; (b) sample D4_ZnO; (c) sample H4_ZnO. The scalebar applies to both SE and BSE images.
Figure 4. SEM photographs of cryofractures of latex coatings taken in secondary electron (left) and backscattered electron mode (right): (a) blank sample D4; (b) sample D4_ZnO; (c) sample H4_ZnO. The scalebar applies to both SE and BSE images.
Coatings 11 00347 g004
Figure 5. Two-dimensional Raman mapping of the coating surface of the sample D4_ZnO. Green color corresponds to the spectrum of the studied latex polymer. A vibration band of sulfate anion at 975 cm−1 was not detected.
Figure 5. Two-dimensional Raman mapping of the coating surface of the sample D4_ZnO. Green color corresponds to the spectrum of the studied latex polymer. A vibration band of sulfate anion at 975 cm−1 was not detected.
Coatings 11 00347 g005
Figure 6. Adhesion (expressed as pull strength) of coatings based on latexes, differing in the covalent crosslinking strategy, surfactant type and addition of nanostructured ZnO.
Figure 6. Adhesion (expressed as pull strength) of coatings based on latexes, differing in the covalent crosslinking strategy, surfactant type and addition of nanostructured ZnO.
Coatings 11 00347 g006
Figure 7. Water absorption into coatings based on latexes, differing in the covalent crosslinking strategy, surfactant type and addition of nanostructured ZnO. The water absorption was measured after 30-day immersion in distilled water at RT.
Figure 7. Water absorption into coatings based on latexes, differing in the covalent crosslinking strategy, surfactant type and addition of nanostructured ZnO. The water absorption was measured after 30-day immersion in distilled water at RT.
Coatings 11 00347 g007
Figure 8. Water whitening of coatings, expressed in terms of transmittance decrease. The water whitening was evaluated after a one-day exposition of coatings cast on glass panels in distilled water. Transmittance values were measured at 500 nm.
Figure 8. Water whitening of coatings, expressed in terms of transmittance decrease. The water whitening was evaluated after a one-day exposition of coatings cast on glass panels in distilled water. Transmittance values were measured at 500 nm.
Coatings 11 00347 g008
Table 1. Main characteristics of the used surfactants.
Table 1. Main characteristics of the used surfactants.
SurfactantTypeChemical CompositionActive Matter (wt.%)Critical Micelle Concentration (mg/L) a
Disponil FES 993NonpolymerizableSodium salt of ethoxylated fatty alcohol ethersulfate29.7200
HITENOL AR-10PolymerizableAmmonium salt of ethoxylated styrenated propenyl phenol ethersulfate99.1250
a Data given by the supplier.
Table 2. Monomeric compositions of latexes differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
Table 2. Monomeric compositions of latexes differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
SampleComposition of Monomer Feeds (g)Surfactant TypeNanostructured ZnO Additive
MMA/BA/MAA/DAAM/ALMA
First StepSecond Step
Series 1: No covalent crosslinking
D118.5/30/1.5/0/018.5/30/1.5/0/0nonpolymerizableno
H118.5/30/1.5/0/018.5/30/1.5/0/0polymerizableno
D1_ZnO18.5/30/1.5/0/018.5/30/1.5/0/0nonpolymerizableyes
H1_ZnO18.5/30/1.5/0/018.5/30/1.5/0/0polymerizableyes
Series 2: Interparticle covalent crosslinking
D218.5/30/1.5/0/016.5/29.5/1.5/2.5/0nonpolymerizableno
H218.5/30/1.5/0/016.5/29.5/1.5/2.5/0polymerizableno
D2_ZnO18.5/30/1.5/0/016.5/29.5/1.5/2.5/0nonpolymerizableyes
H2_ZnO18.5/30/1.5/0/016.5/29.5/1.5/2.5/0polymerizableyes
Series 3: Intraparticle covalent crosslinking
D318/30/1.5/0/0.518.5/30/1.5/0/0nonpolymerizableno
H318/30/1.5/0/0.518.5/30/1.5/0/0polymerizableno
D3_ZnO18/30/1.5/0/0.518.5/30/1.5/0/0nonpolymerizableyes
H3_ZnO18/30/1.5/0/0.518.5/30/1.5/0/0polymerizableyes
Series 4: Intra- and interparticle covalent crosslinking
D418/30/1.5/0/0.516.5/29.5/1.5/2.5/0nonpolymerizableno
H418/30/1.5/0/0.516.5/29.5/1.5/2.5/0polymerizableno
D4_ZnO18/30/1.5/0/0.516.5/29.5/1.5/2.5/0nonpolymerizableyes
H4_ZnO18/30/1.5/0/0.516.5/29.5/1.5/2.5/0polymerizableyes
Table 3. Recipe for emulsion polymerization by using different surfactants and nanostructured ZnO additive.
Table 3. Recipe for emulsion polymerization by using different surfactants and nanostructured ZnO additive.
ComponentReaction Flask Charge (g)First Step Monomer Emulsion (g)Second Step Monomer Emulsion (g)
Water 35.047.572.5
Disponil FES 993 a 0.243.73.7
HITENOL AR-10 b0.071.11.1
Ammonium Persulfate 0.20.20.2
Monomers 50.050.0
Nanostructured ZnO c1.0
a Nonpolymerizable surfactant used for the preparation of the D-labelled latex samples. b Polymerizable surfactant used for the preparation of the H-labelled latex samples. c Ionic crosslinking additive incorporated in the course of the preparation of the D_ZnO− and H_ZnO− labelled latex samples.
Table 4. Characteristics of latexes, differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
Table 4. Characteristics of latexes, differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
SampleCoagulum Content (wt.%)Solids (wt.%)Conversion (wt. %)Particle Size (nm)MFFT (°C)
Series 1: No covalent crosslinking
D1~0 a38.1 ± 0.195.3 ± 0.3123.4 ± 0.8<0
H1~0 a39.7 ± 0.199.5 ± 0.3114.3 ± 0.9<0
D1_ZnO~0 a37.8 ± 0.194.0 ± 0.2118.3 ± 1.83.6 ± 0.2
H1_ZnO0.239.2 ± 0.197.8 ± 0.2104.9 ± 1.54.5 ± 0.1
Series 2: Interparticle covalent crosslinking
D20.438.3 ± 0.295.9 ± 0.4120.4 ± 1.0<0
H2~0 a39.8 ± 0.199.7 ± 0.2112.3± 1.0<0
D2_ZnO0.637.0 ± 0.291.9 ± 0.4117.9 ± 1.64.1 ± 0.1
H2_ZnO~0 a39.3 ± 0.197.8 ± 0.2100.6 ± 1.34.8 ± 0.2
Series 3: Intraparticle covalent crosslinking
D3~038.3 ± 0.195.9 ± 0.3123.7 ± 1.00.6 ± 0.2
H30.339.4 ± 0.298.7 ± 0.4114.7 ± 0.7<0
D3_ZnO0.137.2 ± 0.292.4 ± 0.5124.7 ± 1.13.8 ± 0.2
H3_ZnO0.838.8 ± 0.196.6 ± 0.3112.7 ± 1.75.3 ± 0.3
Series 4: Intra- and interparticle covalent crosslinking
D40.238.6 ± 0.096.7 ± 0.1118.3 ± 1.21.0 ± 0.3
H4~0 a39.7 ± 0.299.5 ± 0.5105.7 ± 1.1<0
D4_ZnO~0 a37.6 ± 0.193.4 ± 0.2121.1 ± 1.94.8 ± 0.1
H4_ZnO0.738.9 ± 0.196.8 ± 0.2110.8 ± 1.25.9 ± 0.2
a No coagulum was present in latex.
Table 5. Stability testing results of latexes, differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
Table 5. Stability testing results of latexes, differing in the covalent crosslinking strategy, surfactant type and the application of nanostructured ZnO additive.
SampleHeat Storage Stability aMechanical Stability aFreeze-Thaw Stability a
Freezing at −5 °CFreezing at −10 °CFreezing at −18 °C
Series 1: No covalent crosslinking
D1
H1
D1_ZnO
H1_ZnO
Series 2: Interparticle covalent crosslinking
D2
H2
D2_ZnO
H2_ZnO
Series 3: Intraparticle covalent crosslinking
D3
H3
D3_ZnO
H3_ZnO––
Series 4: Intra- and interparticle covalent crosslinking
D4
H4
D4_ZnO
H4_ZnO
a Stability results expressed by symbols: “√” means no visible coagulation and the tested latex also provides coatings with a retained gloss (a decrease in gloss < 10 rel.%); “–” means visible coagulation and/or the tested latex provides coatings with a deteriorated gloss (a decrease in gloss > 10 rel.%).
Table 6. Calcium ion stability results of latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition.
Table 6. Calcium ion stability results of latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition.
SampleCa2+ Stability a
CaCl2 Concentration (wt.%)
1.01.52.05.0
Series 1: No covalent crosslinking
D1
H1
D1_ZnO
H1_ZnO
Series 2: Interparticle covalent crosslinking
D2
H2
D2_ZnO
H2_ZnO
Series 3: Intraparticle covalent crosslinking
D3
H3
D3_ZnO
H3_ZnO
Series 4: Intra- and interparticle covalent crosslinking
D4
H4
D4_ZnO
H4_ZnO
a Stability results expressed by symbols: “√” means no visible precipitation of the tested latex in CaCl2 solution; “–” means precipitation of the tested latex in CaCl2 solution.
Table 7. Characteristics of coating materials prepared from latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition, in terms of Tg, gel content and crosslink density, expressed as the average molecular weight between crosslinks (Mc), and Zinc content.
Table 7. Characteristics of coating materials prepared from latexes, differing in the covalent crosslinking strategy, surfactant type and nanostructured ZnO addition, in terms of Tg, gel content and crosslink density, expressed as the average molecular weight between crosslinks (Mc), and Zinc content.
SampleTg (°C)Gel Content (wt.%)Mc (g/mol)Zinc Content (mg/kg)Actual Content of ZnO (wt.%) b
Series 1: No covalent crosslinking
D11.2 ± 0.23.2 ± 1.8a00
H12.4 ± 0.154.0 ± 0.9199671 ± 779200
D1_ZnO2.6 ± 0.360.0 ± 1.8148201 ± 14987241 ± 650.90 ± 0.008
H1_ZnO3.8 ± 0.373.1 ± 1.550726 ± 24787420 ± 300.92 ± 0.004
Series 2: Interparticle covalent crosslinking
D23.2 ± 0.380.5 ± 1.645612 ± 547900
H27.1 ± 0.291.2 ± 1.138254 ± 537300
D2_ZnO4.1 ± 0.493.0 ± 2.316263 ± 1917968 ± 300.99 ± 0.003
H2_ZnO4.5 ± 0.294.1 ± 1.611884 ± 1057309 ± 100.91 ± 0.001
Series 3: Intraparticle covalent crosslinking
D32.9 ± 0.491.6 ± 0.718273 ± 15000
H34.1 ± 0.795.8 ± 1.05576 ± 71600
D3_ZnO3.6 ± 0.596.0 ± 1.74050 ± 1217967 ± 500.99 ± 0.005
H3_ZnO5.3 ± 0.598.6 ± 2.12957 ± 1057740 ± 490.96 ± 0.005
Series 4: Intra- and interparticle covalent crosslinking
D46.3 ± 0.695.6 ± 2.15579 ± 14100
H47.3 ± 0.497.7 ± 1.23427 ± 18800
D4_ZnO3.6 ± 0.298.9 ± 1.22720 ± 1377336 ± 290.92 ± 0.003
H4_ZnO6.4 ± 0.499.0 ± 1.41827 ± 417254 ± 680.90 ± 0.008
a Value was not measured. b Percentage of ZnO-based nanoparticles in a dried coating film calculated from the inductively coupled plasma optical emission spectrometry (ICP-OES) results using the simplified assumption that all the determined Zn in the polymer film was only in the form of ZnO.
Table 8. Properties of coatings cast on glass substrates.
Table 8. Properties of coatings cast on glass substrates.
SampleThickness (µm)Gloss 60° (GU)Transparency (%) 1MEK Resistance (Number of Strikes)
Series 1: No covalent crosslinking
D158.4 ± 8.180.5 ± 0.290.1 ± 0.310.3 ± 0.6
H163.2 ± 7.385.1 ± 0.190.2 ± 0.312.3 ± 0.6
D1_ZnO61.5 ± 6.285.5 ± 0.189.6 ± 0.4210.4 ± 19.2
H1_ZnO48.8 ± 5.485.2 ± 0.290.0 ± 0.3133.2 ± 15.1
Series 2: Interparticle covalent crosslinking
D251.8 ± 7.885.0 ± 0.391.0 ± 0.211.8 ± 1.5
H258.6 ± 6.385.2 ± 0.190.2 ± 0.418.7 ± 0.6
D2_ZnO55.2 ± 6.786.0 ± 0.189.9 ± 0.2above 300 2
H2_ZnO60.5 ± 7.085.4 ± 0.390.7 ± 0.3213.7 ± 18.2
Series 3: Intraparticle covalent crosslinking
D361.9 ± 5.584.3 ± 0.291.1 ± 0.38.0 ± 1.0
H358.5 ± 7.384.0 ± 0.190.6 ± 0.410.7 ± 0.6
D3_ZnO65.0 ± 6.781.5 ± 0.490.5 ± 0.2102.3 ± 2.4
H3_ZnO53.5 ± 4.984.3 ± 0.190.6 ± 0.323.1 ± 4.7
Series 4: Intra- and interparticle covalent crosslinking
D457.5 ± 6.885.2 ± 0.191.1 ± 0.229.4 ± 4.6
H448.1 ± 8.585.5 ± 0.290.2 ± 0.324.6 ± 6.0
D4_ZnO45.7 ± 9.785.6 ± 0.189.8 ± 0.3above 300
H4_ZnO56.5 ± 6.184.9 ± 0.190.6 ± 0.3236 ± 14
1 Transmittance measured at 500 nm. 2 Maximum evaluative value (representing the best property).
Table 9. Results of antibacterial activity of the coating films of the Series 4.
Table 9. Results of antibacterial activity of the coating films of the Series 4.
SampleGrowth of Bacterial Colonies a
S. aureusE. coliE. faecalisK. pneumoniae
D45,5,55,5,54,5,45,5,5
H45,5,55,5,55,4,55,5,5
D4_ZnO0,0,00,0,03,3,30,0,0
H4_ZnO0,0,00,0,03,3,30,0,0
a The scale for assessing the growth of bacterial colonies: 0—without growth; 1—detectable amount (single colony); 2—detectable amount (combined colony); 3—second imprint, distinguishable colonies, third imprint can be detected; 4—third imprint, distinguishable colonies; 5—overgrown, continuous growth.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Machotová, J.; Kalendová, A.; Steinerová, D.; Mácová, P.; Šlang, S.; Šňupárek, J.; Vajdák, J. Water-Resistant Latex Coatings: Tuning of Properties by Polymerizable Surfactant, Covalent Crosslinking and Nanostructured ZnO Additive. Coatings 2021, 11, 347. https://doi.org/10.3390/coatings11030347

AMA Style

Machotová J, Kalendová A, Steinerová D, Mácová P, Šlang S, Šňupárek J, Vajdák J. Water-Resistant Latex Coatings: Tuning of Properties by Polymerizable Surfactant, Covalent Crosslinking and Nanostructured ZnO Additive. Coatings. 2021; 11(3):347. https://doi.org/10.3390/coatings11030347

Chicago/Turabian Style

Machotová, Jana, Andréa Kalendová, Denisa Steinerová, Petra Mácová, Stanislav Šlang, Jaromír Šňupárek, and Jan Vajdák. 2021. "Water-Resistant Latex Coatings: Tuning of Properties by Polymerizable Surfactant, Covalent Crosslinking and Nanostructured ZnO Additive" Coatings 11, no. 3: 347. https://doi.org/10.3390/coatings11030347

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