Next Article in Journal
Advances in Cellulose-Based Composites for Energy Applications
Previous Article in Journal
Validating the Use of Gaussian Process Regression for Adaptive Mapping of Residual Stress Fields
Previous Article in Special Issue
Biomechanical Evaluation of Patient-Specific Polymethylmethacrylate Cranial Implants for Virtual Surgical Planning: An In-Vitro Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cytotoxicity and Microbiological Properties of Copolymers Comprising Quaternary Ammonium Urethane-Dimethacrylates with Bisphenol A Glycerolate Dimethacrylate and Triethylene Glycol Dimethacrylate

by
Marta W. Chrószcz-Porębska
1,
Izabela M. Barszczewska-Rybarek
1,*,
Alicja Kazek-Kęsik
2,3 and
Izabella Ślęzak-Prochazka
3,4
1
Department of Physical Chemistry and Technology of Polymers, Faculty of Chemistry, Silesian University of Technology, Strzody 9 Str., 44-100 Gliwice, Poland
2
Department of Inorganic Chemistry, Analytical Chemistry and Electrochemistry, Faculty of Chemistry, Silesian University of Technology, Krzywoustego 6 Str., 44-100 Gliwice, Poland
3
Biotechnology Centre, Silesian University of Technology, Krzywoustego 8 Str., 44-100 Gliwice, Poland
4
Department of Systems Biology and Engineering, Faculty of Automatic Control, Electronics and Computer Science, Silesian University of Technology, Akademicka 16 Str., 44-100 Gliwice, Poland
*
Author to whom correspondence should be addressed.
Materials 2023, 16(10), 3855; https://doi.org/10.3390/ma16103855
Submission received: 12 April 2023 / Revised: 15 May 2023 / Accepted: 18 May 2023 / Published: 20 May 2023

Abstract

:
Using dental composite restorative materials with a copolymeric matrix chemically modified towards bioactive properties can help fight secondary caries. In this study, copolymers of 40 wt.% bisphenol A glycerolate dimethacrylate, 40 wt.% quaternary ammonium urethane-dimethacrylates (QAUDMA-m, where m represents 8, 10, 12, 14, 16 and 18 carbon atoms in the N-alkyl substituent), and 20 wt.% triethylene glycol dimethacrylate (BG:QAm:TEGs) were tested for (i) cytotoxicity on the L929 mouse fibroblast cell line; (ii) fungal adhesion, fungal growth inhibition zone, and fungicidal activity against C. albicans; and (iii) bactericidal activity against S. aureus and E. coli. BG:QAm:TEGs had no cytotoxic effects on L929 mouse fibroblasts because the reduction of cell viability was less than 30% compared to the control. BG:QAm:TEGs also showed antifungal activity. The number of fungal colonies on their surfaces depended on the water contact angle (WCA). The higher the WCA, the greater the scale of fungal adhesion. The fungal growth inhibition zone depended on the concentration of QA groups (xQA). The lower the xQA, the lower the inhibition zone. In addition, 25 mg/mL BG:QAm:TEGs suspensions in culture media showed fungicidal and bactericidal effects. In conclusion, BG:QAm:TEGs can be recognized as antimicrobial biomaterials with negligible biological patient risk.

1. Introduction

Materials with methacrylate matrices are the most commonly used in dentistry [1]. They have good utility properties, low cytotoxicity, and negligible microbial activity [2,3,4]. Although dentistry offers a wide range of excellent materials, dental caries are observed on a large scale [5,6]. A large part of this problem concerns children [7,8]. Materials with antimicrobial properties could help to fight this disease [9].
Methacrylate monomers with quaternary ammonium (QA) groups are suitable components that, when chemically incorporated into the structure of the organic matrix of dental composite restorative materials (DCRM), give them anti-microbiological activity [10,11,12]. This type of modification involves the free radical copolymerization of QA monomers with common dental dimethacrylates such as bisphenol A glycerolate dimethacrylate (Bis-GMA), urethane-dimethacrylate (UDMA), and triethylene glycol dimethacrylate (TEGDMA), which leads to the formation of polymer networks with high crosslink density [13,14].
Methacrylates with QA groups include monomethacrylates (mono-QAMs) [15,16,17,18,19,20,21] and dimethacrylates (di-QAMs) [22,23,24,25,26,27,28,29,30]. Mono-QAMs such as 12-methacryloyloxydodecylpyridinium bromide [15], and N,N-dimethylaminoethyl methacrylate (DMAEMA) derivatives with bromide [16,17], iodide [18,19,20], or chloride [21] counter ions were the first to be obtained. DCRMs modified with the above mono-QAMs showed high antibacterial activity against many strains of bacteria, such as Streptococcus mutans, Actinomyces viscosus, and Lactobacillus casei [15,16,17,19,20,21]. However, it turned out that they caused deterioration of the physical and mechanical properties of the DCRM matrix, mainly by increasing its elasticity, water absorption, and free monomer leachability [17,18,19,22].
In the hope of eliminating this flow, di-QAMs were developed. The presence of two methacrylate groups in the monomer molecule theoretically gives it a chance to build onto the structure of the polymer network at both ends [22]. Di-QAMs include monomers obtained via the reaction of DMAEMA with dibromo alkanes [23] and urethane-dimethacrylates obtained via the reaction of quaternized N-methyldiethanolamine with 2-(methacryloyloxy)ethyl isocyanate [24,25], or 2-hydroxyethyl methacrylate and 5-isocyanato-1-(isocyanatomethyl)-1,3,3-trimethylcyclohexane [26,27]. Due to the aliphatic structures of these di-QAMs, polymers with no satisfactory properties were obtained. Therefore, the quaternized Bis-GMA derivative was developed. It was obtained from quaternized bisphenol A glycerolate diethylamine and methacryloyl chloride. The polymers modified with quaternized Bis-GMA derivative showed increased values of selected physical properties (water sorption and solubility) [28]. However, high antibacterial activity was the standard and essential feature of dimethacrylate copolymers containing di-QAMs and related materials. It was observed against the following bacteria strains: S. mutans, Escherichia coli, Streptococcus aureus, Pseudomonas aeruginosa, and Bacillus subtillis [23,24,25,26,27,28].
On the other hand, published literature shows that QA monomers may increase the cytotoxicity of methacrylate copolymers. Li et al. [30] reported that the addition of 10 wt.% DMAEMA dimethacrylate derivative with one QA group did not change the Bis-GMA:HEMA cytotoxicity determined on a human gingival fibroblast cell line. Antonucci et al. [29] used the same monomer to test its influence on the cytotoxicity of Bis-GMA:TEGDMA. Any difference in cell viability (RAW 264.7 macrophage-like cell line) was observed between the copolymer modified with 10 wt. % QA monomer and the reference. When the QA monomer concentration increased to 20 and 30 wt.%, reductions of no more than 30% were observed in cell viability. It was also found that monomers with two QA groups increased copolymer cytotoxicity at a lower concentration. Manouchehri et al. [23] reported that 1 wt.% DMAEMA dimethacrylate derivatives with two QA groups caused a one-quarter decrease in cell viability (human foreskin fibroblast) to the reference. Makvandi et al. [28] tested the cytotoxicity of Bis-GMA:TEGDMA modified with 5 to 15 wt.% Bis-GMA derivative containing two QA groups. They observed that the introduction of QA monomer caused a decrease in cell viability (L929 mouse fibroblast) from 10 to 40%.
The bioactivity of compounds containing QA groups, including dimethacrylates, encouraged further research towards developing novel monomer chemical structures and their compositions that result in polymeric materials combining high antibacterial activity, low cytotoxicity, and appropriate utility properties.
Our research group has been working on new di-QAMs and their copolymers. These monomers include a series of urethane-dimethacrylates with two quaternary ammonium groups (QAUDMA-m, m being related to the length of the N-alkyl substituent and corresponding to 8, 10, 12, 14, 16, and 18 carbon atoms) (Figure 1) [31]. QAUDMA-m monomers are similar in chemical structure to the so-called urethane dimethacrylate (UDMA) monomer due to the following structural elements: (i) the 1,6-diisocyanato-2,4,4-trimethylhexylene (TMDI) core, (ii) two arms attached to the core by carbamate linkages, and (iii) methacrylate group at the end of each arm. The presence of the QA group in each arm distinguishes QAUDMA-m from UDMA.
The results of the initial study showed that the QAUDMA-m monomers were highly viscous resins, had appropriate transparency, polymerized to high degrees of conversion (DC), and were characterized by low polymerization shrinkage (Se) [31]. These promising characteristics of QAUDMA-m monomers qualified them for further research to assess their influence on the properties of copolymers with common dental dimethacrylates.
Copolymers of 60 wt.% QAUDMA-m with 40 wt.% TEGDMA (QAm:TEGs) were first investigated by analogy to the 60 wt.% Bis-GMA with 40 wt.% TEGDMA copolymer (BG:TEG), representing standard DCRM matrix [32,33]. QAm:TEGs revealed significantly higher antibacterial activity against S. aureus and E. coli than BG:TEG, which depended on the number of carbon atoms in the N-alkyl substituent (Cm). In the case of S. aureus, QA14:TEG, on whose surface no bacteria colonies were observed, showed the highest antibacterial activity. On the other hand, QA8:TEG had the weakest antibacterial action. In the case of E. coli, QAm:TEGs from C8 to C14, on whose surfaces no bacteria colonies were observed, showed the highest antibacterial activity. On the other hand, QA16:TEG had the weakest antibacterial action [32]. However, the physicomechanical analysis showed that none of the QAm:TEGs had sufficiently high flexural properties or hardness. Their water sorption and residual monomer leachability were also excessively high [33].
Due to the detected intense antibacterial activity of QAm:TEGs, another series of copolymers was developed. They consisted of 40 wt.% Bis-GMA, 40 wt.% QAUDMA-m, and 20 wt.% TEGDMA (BG:QAm:TEG), analogous to the composition of another standard DCRM matrix, consisting of 40 wt.% Bis-GMA, 40 wt.% UDMA and 20 wt.% TEGDMA (BG:UD:TEG). They were characterized in terms of physical properties [34]. As can be seen from Table 1, they had low polymerization shrinkage, high glass transition temperature, and hydrophilic surfaces—except for BG:QA16:TEG and BG:QA18:TEG, which were hydrophobic. They also had sufficiently low water sorption and solubility, except BG:QA8:TEG and BG:QA10:TEG. These latter absorbed greater water quantity than is specified in ISO 4049 standard (40 µg/mm3) [35].
This work aimed to further characterize BG:QAm:TEGs for cytotoxicity on the L929 mouse fibroblast cell line and antimicrobial activity against C. albicans, S. aureus, and E. coli. The null hypothesis was that the BG:QAm:TEGs have antimicrobial activity and do not have cytotoxic effects on cells chosen for testing. Bioactive DCRM matrices, such as our proposed BG:QAm:TEGs, are highly desirable because infections caused by oral microflora are observed among adults and children [7,8]. These include tooth decay caused by bacteria and candidiasis caused by fungi. The available dental materials have an insufficient biocidal effect, and, in addition, growing antibiotic resistance is also a problem. This article may have cognitive and utility dimensions for developing potential novel DCRM systems with antimicrobial activity.

2. Materials and Methods

2.1. Monomer Synthesis and Crosslinking Copolymerization

The QAUDMA-m monomers were synthesized using the previously described procedure [31]. First, N,N-(2-hydroxyethyl)methylaminoethyl methacrylate (HAMA) was obtained using methyl methacrylate (MMA; Acros Organics, Geel, Belgium) and 2,2′-methyliminodiethanol (MDEA; Acros Organics, Geel, Belgium). Next, it was N-alkylated with alkyl 1-bromooctane,1-bromodecane, 1-bromododecane, 1-bromotetradecane, 1-bromohexadecane, and 1-bromooctadecane (Acros Organics, Geel, Belgium) to obtain 2-(methacryloyloxy)ethyl-2-hydroxyethylmethylalkylammonium bromides (QAHAMAs). Finally, the addition reaction of QAHAMAs to 1,6-diisocyanato-2,4,4-trimethylhexylene (TMDI; Tokyo Chemical Industry, Tokyo, Japan) was performed, and the QAUDMA-m monomers were obtained.
Next, 40 wt.% Bis-GMA (Sigma-Aldrich, St. Louis, MO, USA), 40 wt.% QAUDMA-m, and 20 wt.% TEGDMA (Sigma-Aldrich, St. Louis, MO, USA) were stirred at 60 °C to obtain homogenous compositions. As 0.4 wt.% camphorquinone (CQ; Sigma-Aldrich, St. Louis, MO, USA) and 1 wt.% 2-dimethylaminoethyl methacrylate (DMAEMA; Sigma-Aldrich, St. Louis, MO, USA), respectively the photoinitiator and accelerator, were added, the compositions were subjected to irradiation by utilizing a UV-Vis lamp (Ultra Vitalux 300, Osram, Munich, Germany) with a wavelength range of 280 to 780 nm for 60 min at room temperature (bulk polymerization). The PET film covering the photopolymerizing system prevented oxidation inhibition. Disc-like casts (diameter × thickness = 10 mm × 3 mm) were obtained. Once polished with sandpaper, they were used for the following tests: cytotoxicity, fungal adhesion, and fungal growth inhibition zone. The fungicidal and bactericidal tests employed powdered copolymers with a grain size of less than 25 µm. They were used to prepare suspensions in culture media.
The names of copolymer samples were abbreviated as BG:QAm:TEGs (m—number of carbon atoms in N-alkyl chain in QAUDMA-m equal to 8, 10, 12, 14, 16, or 18). The copolymer of 40 wt.% Bis-GMA, 40 wt.% urethane-dimethacrylate (UDMA; Sigma-Aldrich, St. Louis, MO, USA), and 20 wt.% TEGDMA (BG:UD:TEG) was prepared as a reference (Figure 2).

2.2. Cytotoxicity

Cytotoxicity was determined according to the ISO 10993-5 standard [36] using the L929 mouse fibroblast cell line (ATCC-CCL-1). The L929 cell line was obtained from ATCC-CCL-1, NCTC clone 929 Areolar Fibroblast Mouse, delivered by LGC Standards Sp. z o.o. ul. Ogrodowa 27/29, Kiełpin 05-092 Łomianki.
Copolymer specimens of disc-like casts of 10 mm × 3 mm (diameter × thickness) were sanded clean and immersed in a culture medium (1 mL per 3 cm2 of the specimen surface) for 72 h at 37 °C with shaking at 80 rpm. The achieved extracts were mixed with fresh culture medium in a volume ratio of 1:1 and placed in a 96-well culture plate. Then, cell suspensions in a culture medium containing 5000 L929 mouse fibroblast cells were added to each well in the culture plate. Samples were incubated at 5% CO2 at 37 °C for 24 h (IncuSafe CO2 Incubator MCO-170AICUVD-PE PHCbi, Etten-Leur, The Netherlands). Next, 10 µL of Alamar Blue (G-Biosciences, St Louis, MO, USA) was added to each well, and samples were incubated at 37 °C for 4 h. Finally, fluorescence of aliquots was measured using a Varioskan LUX Multimode microplate reader (Thermo Scientific, Waltham, MA, USA).

2.3. Microbiological Properties

Microbiological properties were determined against a C. albicans (ATCC 2091) reference fungus strain as well as S. aureus (ATCC 25923) and E. coli (ATCC 25922) reference bacterium strains. Before testing, fungal and bacterial strains were cultured at 37 °C for 18 h (incubator POL-EKO, Wodzisław Śląski, Poland) in SDA (sabouraud dextrose lab-agar, Biomaxima, Lublin, Poland) and TSB (tryptic soy broth, Biomaxima, Lublin, Poland) medium, respectively.

2.3.1. Fungal Adhesion

The adhesion of fungi to copolymer surfaces was determined on specimens in the form of disc-like casts of 10 mm × 3 mm (diameter × thickness). Specimen surfaces were finished with sandpaper before experiments.
Specimens were submerged in 1 mL of fungal suspension in a culture medium (~5 × 106 CFU/mL) and incubated at 37 °C for 18 h. Next, the fungal suspension was removed with a Pasteur pipette, and samples were washed gently with 1 mL of sterile water. Washed specimens were placed in clean centrifuge tubes, soaked in 1 mL of sterile water, and vortexed (1 min, 3000 rpm) to remove fungi from sample surfaces. Next, 100 µL of the obtained fungal suspensions were diluted with 0.9% NaCl to achieve concentrations of 0.1, 0.01, 0.001, 0.0001, 0.00001, and 0.000001%. Then 100 µL of each solution was applied on agar plates (Sabourand agar, Diag-Med, Warszawa, Poland) and incubated at 37 °C for 18 h. Finally, the number of fungi colonies on plates illuminated by transmitted light was manually counted.

2.3.2. Fungal Growth Inhibition Zone

The fungal growth inhibition zone diameter was determined for extracts obtained by soaking disc-like copolymer casts of 10 mm × 3 mm (diameter × thickness) in sterile water for seven days at room temperature.
We applied 100 µL of the fungal suspension in a culture medium (~5 × 108 CFU/mL) to the agar plates. Next, 5 mm diameter holes were cut out. They were filled with 100 µL of the previously obtained copolymer extracts. Agar plates were incubated at 37 °C for 24 h. The inhibition zone diameter was measured.

2.3.3. Fungicidal Activity

Copolymer powders of a grain with a diameter of less than 25 µm were used to prepare 25 mg/mL suspensions in SDA. Next, 20 µL of C. albicans fungal suspension in culture medium (~5 × 108 CFU/mL) were introduced into the obtained copolymer suspensions, vortexed (10 s, 2000 rpm), and incubated at 37 °C for 18 h. Suspensions were vortexed again (10 s, 2000 rpm), and 100 µL was applied to agar plates. The agar plates were incubated at 37 °C for 18 h. The amount of fungal colonies was qualitatively assessed by visual comparison to the control.

2.3.4. Bactericidal Activity

Copolymer powders of a grain with a diameter of less than 25 µm were used to prepare 25 mg/mL suspensions in TSB. Next, 20 µL of S. aureus and E. coli bacterial suspensions in a culture medium (~5 × 108 CFU/mL) was introduced into the achieved copolymer suspensions, vortexed (10 s, 2000 rpm), and incubated at 37 °C for 18 h. Suspensions were vortexed again (10 s, 2000 rpm), and 100 µL was applied to agar plates (Müller-Hinton agar, Diag-Med, Warszawa, Poland). The agar plates were incubated at 37 °C for 18 h. The amount of bacterial colonies was qualitatively assessed by visual comparison to the control.

2.4. Statistical Analysis

Statistica 13.1 (TIBCO Software Inc., Palo Alto, CA, USA) software was used to perform statistical analysis. The experimental results represent the mean value of five measurements, accompanied by a standard deviation (SD). The statistical significance of the results was assessed by the non-parametric Wilcoxon test (level of significance (p) of 0.05).

3. Results

In this work, a set of dimethacrylate BG:QAm:TEGs were obtained by photopolymerization and characterized for cytotoxicity and microbiological properties. The BG:UD:TEG was used for comparison purposes.

3.1. Cytotoxicity

Figure 3 shows the results obtained for the Alamar Blue percent reduction. They correspond to the viability of the L929 cell line.
The cell viability determined for BG:QAm:TEGs was from 63.81 to 76.77%. Alamar Blue percent reduction values were statistically similar (p > 0.05), except for BG:QA8:TEG. The cell viability of the latter was lower (p ≤ 0.05) than that of BG:QA12:TEG and BG:QA18:TEG. Compared to the BG:UD:TEG, cell viability was reduced in all BG:QAm:TEGs (p ≤ 0.05).
The viability of L929 mouse fibroblast cells in BG:QAm:TEGs was related to the control. The results of these calculations are shown in Figure 4. As can be seen, the reduction in cell viability observed for BG:QAm:TEGs ranged from 12.28 to 27.09%.

3.2. Microbiological Properties

In Figure 5, the results of the fungal C. albicans adhesion test are shown.
The number of C. albicans colonies observed on BG:QAm:TEGs’ surfaces were from 3.20 to 4.15 log(CFU/mL). The number of fungal colonies on BG:QAm:TEGs with Cm from C8 to C14 was statistically similar (p > 0.05). The further increase in Cm up to C18 caused a significant increase in the log(CFU/mL) value (p ≤ 0.05). Compared to the BG:UD:TEG, the number of fungal colonies on almost all BG:QAm:TEGs was lower (p ≤ 0.05), except the BG:QA18:TEG. Its log(CFU/mL) value was similar (p > 0.05) to that determined for BG:UD:TEG.
In Figure 6, the results for the fungal growth inhibition zone are shown.
The fungal growth inhibition zones were only observed for BG:QAm:TEGs with Cm lower than C16. Their diameters ranged from 13 to 7 mm and decreased as the Cm increased. BG:QA16:TEG and BG:QA18:TEG showed no inhibition zone, similar (p > 0.05) to the BG:UD:TEG.
In Figure 7, the results from the microbial activity test are shown.
Antimicrobial activity tests were performed by qualitatively assessing the number of microbial colonies on agar plates overlaid with 25 mg/mL copolymer suspensions in appropriate culture media, SDA, and TSB, respectively, for fungi and bacteria. As can be seen from Figure 7, no colonies of tested microbial strains (C. albicansFigure 7a, S. aureusFigure 7b, E. coliFigure 7c) were observed for all BG:QAm:TEGs. The number of fungal colonies on the agar plate with the BG:UD:TEG suspension was similar to that of fungal colonies observed on the agar plate without any copolymer (control). On the other hand, the number of colonies of both bacterial strains on the agar plates with the BG:UD:TEG suspension was noticeably lower than the number of bacterial colonies observed on the agar plate without any copolymer (control).

4. Discussion

A set of six QAUDMA-m monomers with two QA groups substituted with N-alkyl chains of various lengths (Cm = C8, C10, C12, C14, C16, and C18) (Figure 1) were synthesized to obtain novel copolymers for dental applications. The concept assumed the total replacement of UDMA with QAUDMA-m in a typical dental BG:UD:TEG dimethacrylate copolymer.
The cytotoxic effects of DCRM are closely related to sol fraction leaching [37,38,39]. The scale of this phenomenon depends on the combination of several factors, such as the matrix chemical structure, the degree of conversion [37,38], and oxygen inhibition on the DCRM surface [39]. The eluted substances can penetrate bodily tissues [38]. When they are cytotoxic, they cause inflammation, inhibition of cell division, or even necrosis (Figure 8) [40,41,42,43,44,45]. Therefore, the leachable fraction of dental material is standardized to 7.5 µg/mm3 [35].
The polymers with QA units have higher cytotoxicity than typical dental copolymers due to the presence of quaternary nitrogen atoms [46,47]. Therefore, analysis of their cytotoxicity is fundamental in dental materials studies.
This study evaluated the cytotoxicity of BG:QAm:TEGs compared to the BG:UD:TEG reference on the L929 mouse fibroblast cell line. In accordance with ISO 10993-5 standard [36], the results of cell viability allowed us to identify all tested BG:QAm:TEGs as noncytotoxic. This standard states that a reduction in cell viability by more than 30% to the control is considered a cytotoxic effect, whereas the decrease in cell viability of BG:QAm:TEGs was 15.32 to 27.09% (Figure 4).
The biofilm formed on the tooth and dental filling surfaces consists of agglomerates of bacteria and fungi [48,49]. The acidic products of the bacterial metabolism cause enamel acidolysis and further secondary caries formation [50]. On the other hand, fungi cause candidiasis, considered the primary fungal infection of the oral cavity (Figure 9) [51]. The proliferation of bacteria and fungi is promoted by insufficient oral hygiene, primary and secondary caries, and untreated inflammation of the oral cavity [52,53]. All those factors indicate that the antifungal properties of dental materials are as crucial as antibacterial activity. Unfortunately, the antifungal activity of dimethacrylate copolymers containing QA groups has rarely been investigated.
In this study, the antifungal activity of BG:QAm:TEGs was examined against C. albicans (ATCC 2091) by evaluating the number of fungal colonies on the copolymer surfaces, inhibition zone diameters, and fungicidal effect of copolymer suspensions.
All BG:QAm:TEGs and BG:UD:TEG showed antifungal activity against C. albicans. The BG:QAm:TEGs with Cm lower than C16 showed similar and the highest antifungal activity. The number of fungus colonies found on their surfaces was 103 orders of magnitude greater. Further increase in the Cm resulted in a significant increase in the log(CFU/mL) values. The BG:QA18:TEG had the highest log(CFU/mL) value, which simultaneously was almost the same as that for BG:UD:TEG.
The literature shows that surface character is crucial in Candida biofilm formation [48]. Therefore, the results obtained for antifungal activity were related to those obtained for water contact angle (WCA) [34] (Figure 10). It can be seen that the greater the WCA values, the greater the number of fungus colonies on the BG:QAm:TEG surfaces. It can be concluded that the greater the surface hydrophobicity, the lower the antifungal activity of BG:QAm:TEGs.
Antifungal activity was also characterized by measuring inhibition zone diameter. The results showed that the greater the Cm, the smaller the inhibition zone diameter. However, the fungal growth-inhibitory effect was observed for BG:QAm:TEGs with Cm lower than C16. BG:QA16:TEG and BG:QA18:TEG did not show a fungal growth inhibition zone.
The results of the inhibition zone diameter were related to the concentration of QA groups (xQA) in the BG:QAm:TEG repeating units (Figure 11). As can be seen, the fungal growth inhibitory effects depend on the xQA. The lower its value, the lower the inhibition zone diameter. It can also be assumed that xQA of 3.63 mol/kg is the minimum below which no fungal growth inhibitory effect of the BG:QAm:TEG is observed.
The qualitative assessment of the amount of microbial colonies observed on agar plates overlaid with BG:QAm:TEGs suspensions indicated that the concentration of 25 mg/mL is sufficient to kill all tested microbial strains (C. albicansFigure 7a, S. aureusFigure 7b, E. coliFigure 7c). On the other hand, the concentration of 25 mg/mL of BG:UD:TEG reference copolymer can be recognized only as sufficient to inhibit the growth of tested bacteria strains. An inhibitory effect was not observed for the C. albicans strain, as the visual assessment of the colony number was similar for BG:UD:TEG and the control. The results for BG:UD:TEG reference sample showed that this typical dental copolymer had antibacterial activity but not antifungal activity.
Data obtained for BG:QAm:TEGs suggests their antimicrobial activity and lack of cytotoxicity. However, a more comprehensive survey will be done to assess the possible therapeutic advantage of these copolymers. Microbiological characterization of BG:QAm:TEGs including the number of bacterial colonies on the sample surface and bacterial growth inhibition zone as well as mechanical properties will be the subject of another publication. Regarding cytotoxicity, as it was determined using only the L929 mouse fibroblast cell line, it should be checked using more cell lines, including human ones. Another area for development is that the results presented are limited to copolymers. If they were to be used as DCRMs’ matrices, testing on the model and commercial composites modified with QAUDMA-m should be carried out.

5. Conclusions

The modification of BG:UD:TEG by replacement 40 wt.% UDMA with the same amount of QAUDMA-m resulted in copolymers of low cytotoxicity and high antimicrobial activity against C. albicans, S. aureus, and E. coli.
BG:QAm:TEGs did not show cytotoxic effects on the L929 mouse fibroblast cell line, as the cell viability decreased by no more than 30% compared to the control. The Cm did not reveal the influence on cytotoxicity.
The antifungal effect against C. albicans was demonstrated by the reduction in the number of fungal colonies on copolymer surfaces, the occurrence of a growth inhibition zone, and the fungicidal effect of the 25 mg/mL copolymer suspensions. The antifungal effect depended on the Cm only from C14 to C18. It was revealed in correlations of a number of C. albicans colonies on copolymer surfaces with water contact angle (WCA) and fungal growth inhibition zone with a concentration of QA groups (xQA). The higher the WCA (lower surface hydrophilicity), the greater the number of C. albicans colonies on BG:QAm:TEG surfaces. The lower the xQA, the lower the inhibition zone diameter.
Antibacterial activity of BG:QAm:TEGs against S. aureus and E. coli was demonstrated by the bactericidal effect of the 25 mg/mL copolymer suspensions.
In summary, the null hypothesis has been verified, and BG:QAm:TEGs were found noncytotoxic and antimicrobial. However, further analysis should be done to demonstrate that they are safe bioactive biomaterials and do not pose a biological risk to patients.

Author Contributions

Conceptualization, I.M.B.-R. and M.W.C.-P.; methodology, I.M.B.-R., M.W.C.-P. and A.K.-K.; investigation, M.W.C.-P., A.K.-K. and I.Ś.-P.; statistical analysis, M.W.C.-P.; resources, I.M.B.-R. and M.W.C.-P.; data curation, M.W.C.-P., A.K.-K. and I.Ś.-P.; writing—original draft preparation, I.M.B.-R. and M.W.C.-P.; writing—review and editing, I.M.B.-R.; visualization, M.W.C.-P.; supervision, I.M.B.-R.; project administration, I.M.B.-R.; funding acquisition, I.M.B.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Rector’s grant for scientific research and development activities at the Silesian University of Technology, grant number: 04/040/RGJ23/0243.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data supporting the reported results are available from the authors upon reasonable request.

Conflicts of Interest

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

References

  1. Dursun, E.; Fron-Chabouis, H.; Attal, J.-P.; Raskin, A. Bisphenol A Release: Survey of the Composition of Dental Composite Resins. Open Dent. J. 2016, 10, 446–453. [Google Scholar] [CrossRef] [PubMed]
  2. Spencer, P.; Ye, Q.; Misra, A.; Goncalves, S.E.P.; Laurence, J.S. Proteins, Pathogens, and Failure at the Composite-Tooth Interface. J. Dent. Res. 2014, 93, 1243–1249. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, N.; Melo, M.A.S.; Weir, M.D.; Reynolds, M.A.; Bai, Y.; Xu, H.H.K. Do Dental Resin Composites Accumulate More Oral Biofilms and Plaque than Amalgam and Glass Ionomer Materials? Materials 2016, 9, 888. [Google Scholar] [CrossRef] [PubMed]
  4. Pagano, S.; Lombardo, G.; Costanzi, E.; Balloni, S.; Bruscoli, S.; Flamini, S.; Coniglio, M.; Valenti, C.; Cianetti, S.; Marinucci, L. Morpho-Functional Effects of Different Universal Dental Adhesives on Human Gingival Fibroblasts: An in Vitro Study. Odontology 2021, 109, 524–539. [Google Scholar] [CrossRef] [PubMed]
  5. Janakiram, C.; Dye, B.A. A Public Health Approach for Prevention of Periodontal Disease. Periodontology 2000 2020, 84, 202. [Google Scholar] [CrossRef] [PubMed]
  6. James, S.L.; Abate, D.; Abate, K.H.; Abay, S.M.; Abbafati, C.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; et al. Global, Regional, and National Incidence, Prevalence, and Years Lived with Disability for 354 Diseases and Injuries for 195 Countries and Territories, 1990-2017: A Systematic Analysis for the Global Burden of Disease Study 2017. Lancet 2018, 392, 1789–1858. [Google Scholar] [CrossRef]
  7. Carli, E.; Pasini, M.; Pardossi, F.; Capotosti, I.; Narzisi, A.; Lardani, L. Oral Health Preventive Program in Patients with Autism Spectrum Disorder. Children 2022, 9, 535. [Google Scholar] [CrossRef] [PubMed]
  8. World Heath Organisation. Global Oral Health Status Report: Towards Universal Health Coverage for Oral Health by 2030. Available online: https://www.who.int/publications/i/item/9789240061484 (accessed on 27 January 2023).
  9. Ramburrun, P.; Pringle, N.A.; Dube, A.; Adam, R.Z.; D’souza, S.; Aucamp, M. Recent Advances in the Development of Antimicrobial and Antifouling Biocompatible Materials for Dental Applications. Materials 2021, 14, 3167. [Google Scholar] [CrossRef]
  10. Zhang, Y.; Chen, Y.; Hu, Y.; Huang, F.; Xiao, Y. Quaternary Ammonium Compounds in Dental Restorative Materials. Dent. Mater. J. 2018, 37, 183–191. [Google Scholar] [CrossRef]
  11. Imazato, S.; Chen, J.-h.; Ma, S.; Izutani, N.; Li, F. Antibacterial Resin Monomers Based on Quaternary Ammonium and Their Benefits in Restorative Dentistry. Jpn. Dent. Sci. Rev. 2012, 48, 115–125. [Google Scholar] [CrossRef]
  12. Makvandi, P.; Jamaledin, R.; Jabbari, M.; Nikfarjam, N.; Borzacchiello, A. Antibacterial Quaternary Ammonium Compounds in Dental Materials: A Systematic Review. Dent. Mater. 2018, 34, 851–867. [Google Scholar] [CrossRef] [PubMed]
  13. Hild, G. Model Networks Based on ‘endlinking’ Processes: Synthesis, Structure and Properties. Prog. Polym. Sci. 1998, 23, 1019–1149. [Google Scholar] [CrossRef]
  14. Morgan, D.R.; Kalachandra, S.; Shobha, H.K.; Gunduz, N.; Stejskal, E.O. Analysis of a Dimethacrylate Copolymer (Bis-GMA and TEGDMA) Network by DSC and 13C Solution and Solid-State NMR Spectroscopy. Biomaterials 2000, 21, 1897–1903. [Google Scholar] [CrossRef] [PubMed]
  15. Imazato, S.; Torii, M.; Tsuchitani, Y.; Mccabe, J.F.; Russell, R.R.B. Incorporation of Bacterial Inhibitor into Resin Composite. J. Dent. Res. 1994, 73, 1437–1443. [Google Scholar] [CrossRef] [PubMed]
  16. Li, F.; Weir, M.D.; Xu, H.H.K. Effects of Quaternary Ammonium Chain Length on Antibacterial Bonding Agents. J. Dent. Res. 2013, 92, 932–938. [Google Scholar] [CrossRef]
  17. Vidal, M.L.; Rego, G.F.; Viana, G.M.; Cabral, L.M.; Souza, J.P.B.; Silikas, N.; Schneider, L.F.; Cavalcante, L.M. Physical and Chemical Properties of Model Composites Containing Quaternary Ammonium Methacrylates. Dent. Mater. 2018, 34, 143–151. [Google Scholar] [CrossRef]
  18. Cherchali, F.Z.; Attik, N.; Mouzali, M.; Tommasino, J.B.; Abouelleil, H.; Decoret, D.; Seux, D.; Grosgogeat, B. Structural Stability of DHMAI Antibacterial Dental Composite Following in Vitro Biological Aging. Dent. Mater. 2020, 36, 1161–1169. [Google Scholar] [CrossRef]
  19. Cherchali, F.Z.; Mouzali, M.; Tommasino, J.B.; Decoret, D.; Attik, N.; Aboulleil, H.; Seux, D.; Grosgogeat, B. Effectiveness of the DHMAI Monomer in the Development of an Antibacterial Dental Composite. Dent. Mater. 2017, 33, 1381–1391. [Google Scholar] [CrossRef]
  20. He, J.; Söderling, E.; Österblad, M.; Vallittu, P.K.; Lassila, L.V.J. Synthesis of Methacrylate Monomers with Antibacterial Effects against S. Mutans. Molecules 2011, 16, 9755–9763. [Google Scholar] [CrossRef]
  21. Li, F.; Li, F.; Wu, D.; Ma, S.; Gao, J.; Li, Y.; Xiao, Y.; Chen, J. The Effect of an Antibacterial Monomer on the Antibacterial Activity and Mechanical Properties of a Pit-and-Fissure Sealant. J. Am. Dent. Assoc. 2011, 142, 184–193. [Google Scholar] [CrossRef]
  22. Huang, L.; Yu, F.; Sun, X.; Dong, Y.; Lin, P.T.; Yu, H.H.; Xiao, Y.H.; Chai, Z.G.; Xing, X.D.; Chen, J.H. Antibacterial Activity of a Modified Unfilled Resin Containing a Novel Polymerizable Quaternary Ammonium Salt MAE-HB. Sci. Rep. 2016, 6, 33858. [Google Scholar] [CrossRef] [PubMed]
  23. Manouchehri, F.; Sadeghi, B.; Najafi, F.; Mosslemin, M.H.; Niakan, M. Synthesis and Characterization of Novel Polymerizable Bis-Quaternary Ammonium Dimethacrylate Monomers with Antibacterial Activity as an Efficient Adhesive System for Dental Restoration. Polym. Bull. 2019, 76, 1295–1315. [Google Scholar] [CrossRef]
  24. Liang, X.; Söderling, E.; Liu, F.; He, J.; Lassila, L.V.J.; Vallittu, P.K. Optimizing the Concentration of Quaternary Ammonium Dimethacrylate Monomer in Bis-GMA/TEGDMA Dental Resin System for Antibacterial Activity and Mechanical Properties. J. Mater. Sci. Mater. Med. 2014, 25, 1387–1393. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, Q.T.; He, J.W.; Lin, Z.M.; Liu, F.; Lassila, L.V.J.; Vallittu, P.K. Physical and Chemical Properties of an Antimicrobial Bis-GMA Free Dental Resin with Quaternary Ammonium Dimethacrylate Monomer. J. Mech. Behav. Biomed. Mater. 2016, 56, 68–76. [Google Scholar] [CrossRef]
  26. Huang, Q.; Lin, Z.; Liang, X.; Liu, F.; He, J. Preparation and Characterization of Antibacterial Dental Resin with UDMQA-12. Adv. Polym. Technol. 2014, 33, 21395. [Google Scholar] [CrossRef]
  27. Liang, X.; Huang, Q.; Liu, F.; He, J.; Lin, Z. Synthesis of Novel Antibacterial Monomers (UDMQA) and Their Potential Application in Dental Resin. J. Appl. Polym. Sci. 2013, 129, 3373–3381. [Google Scholar] [CrossRef]
  28. Makvandi, P.; Ghaemy, M.; Mohseni, M. Synthesis and Characterization of Photo-Curable Bis-Quaternary Ammonium Dimethacrylate with Antimicrobial Activity for Dental Restoration Materials. Eur. Polym. J. 2016, 74, 81–90. [Google Scholar] [CrossRef]
  29. Antonucci, J.M.; Zeiger, D.N.; Tang, K.; Lin-Gibson, S.; Fowler, B.O.; Lin, N.J. Synthesis and Characterization of Dimethacrylates Containing Quaternary Ammonium Functionalities for Dental Applications. Dent. Mater. 2012, 28, 219–228. [Google Scholar] [CrossRef]
  30. Li, F.; Weir, M.D.; Chen, J.; Xu, H.H.K. Comparison of Quaternary Ammonium-Containing with Nano-Silver-Containing Adhesive in Antibacterial Properties and Cytotoxicity. Dent. Mater. 2013, 29, 450–461. [Google Scholar] [CrossRef]
  31. Chrószcz, M.W.; Barszczewska-Rybarek, I.M. Synthesis and Characterization of Novel Quaternary Ammonium Urethane-Dimethacrylate Monomers—A Pilot Study. Int. J. Mol. Sci. 2021, 22, 8842. [Google Scholar] [CrossRef]
  32. Chrószcz, M.W.; Barszczewska-Rybarek, I.M.; Kazek-K, A. Novel Antibacterial Copolymers Based on Quaternary Ammonium Urethane-Dimethacrylate Analogues and Triethylene Glycol Dimethacrylate. Int. J. Mol. Sci. 2022, 23, 4954. [Google Scholar] [CrossRef] [PubMed]
  33. Chrószcz-Porębska, M.W.; Barszczewska-Rybarek, I.M.; Chladek, G. Characterization of the Mechanical Properties, Water Sorption, and Solubility of Antibacterial Copolymers of Quaternary Ammonium Urethane-Dimethacrylates and Triethylene Glycol Dimethacrylate. Materials 2022, 15, 5530. [Google Scholar] [CrossRef] [PubMed]
  34. Chrószcz-Porębska, M.W.; Barszczewska-Rybarek, I.M.; Chladek, G. Physicochemical Properties of Novel Copolymers of Quaternary Ammonium UDMA Analogues, Bis-GMA, and TEGDMA. Int. J. Mol. Sci. 2023, 24, 1400. [Google Scholar] [CrossRef]
  35. ISO 4049:2019; Dentistry—Polymer Based Restorative Materials. International Standrad Organisation: London, UK, 2019.
  36. ISO 10993-5:2009; Biological Evaluation of Medical Devices—Part 5: Tests for In Vivo Cytotoxicity. International Standard Organisation: London, UK, 2009.
  37. Lee, M.J.; Kim, M.J.; Kwon, J.S.; Lee, S.B.; Kim, K.M. Cytotoxicity of Light-Cured Dental Materials According to Different Sample Preparation Methods. Materials 2017, 10, 288. [Google Scholar] [CrossRef] [PubMed]
  38. Furche, S.; Hickel, R.; Reichl, F.X.; Van Landuyt, K.; Shehata, M.; Durner, J. Quantification of Elutable Substances from Methacrylate Based Sealers and Their Cytotoxicity Effect on with Human Gingival Fibroblasts. Dent. Mater. 2013, 29, 618–625. [Google Scholar] [CrossRef] [PubMed]
  39. Gociu, M.; Pǎtroi, D.; Prejmerean, C.; Pǎstrǎv, O.; Boboia, S.; Prodan, D.; Moldovan, M. Biology and Cytotoxicity of Dental Materials: An in Vitro Study. Rom. J. Morphol. Embryol. 2013, 54, 261–265. [Google Scholar] [PubMed]
  40. Hampe, T.; Wiessner, A.; Frauendorf, H.; Alhussein, M.; Karlovsky, P.; Bürgers, R.; Krohn, S. Monomer Release from Dental Resins: The Current Status on Study Setup, Detection and Quantification for In Vitro Testing. Polymers 2022, 14, 1790. [Google Scholar] [CrossRef]
  41. Polydorou, O.; Trittler, R.; Hellwig, E.; Kümmerer, K. Elution of Monomers from Two Conventional Dental Composite Materials. Dent. Mater. 2007, 23, 1535–1541. [Google Scholar] [CrossRef]
  42. Volk, J.; Ziemann, C.; Leyhausen, G.; Geurtsen, W. Non-Irradiated Campherquinone Induces DNA Damage in Human Gingival Fibroblasts. Dent. Mater. 2009, 25, 1556–1563. [Google Scholar] [CrossRef]
  43. Durner, J.; Dȩbiak, M.; Bürkle, A.; Hickel, R.; Reichl, F.X. Induction of DNA Strand Breaks by Dental Composite Components Compared to X-ray Exposure in Human Gingival Fibroblasts. Arch. Toxicol. 2011, 85, 143–148. [Google Scholar] [CrossRef]
  44. Lodiene, G.; Morisbak, E.; Bruzell, E.; Ørstavik, D. Toxicity Evaluation of Root Canal Sealers in Vitro. Int. Endod. J. 2008, 41, 72–77. [Google Scholar] [CrossRef] [PubMed]
  45. Demirci, M.; Hiller, K.A.; Bosl, C.; Galler, K.; Schmalz, G.; Schweikl, H. The Induction of Oxidative Stress, Cytotoxicity, and Genotoxicity by Dental Adhesives. Dent. Mater. 2008, 24, 362–371. [Google Scholar] [CrossRef] [PubMed]
  46. Hrubec, T.C.; Seguin, R.P.; Xu, L.; Cortopassi, G.A.; Datta, S.; Hanlon, A.L.; Lozano, A.J.; McDonald, V.A.; Healy, C.A.; Anderson, T.C.; et al. Altered Toxicological Endpoints in Humans from Common Quaternary Ammonium Compound Disinfectant Exposure. Toxicol. Rep. 2021, 8, 646–656. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, C.; Cui, F.; Zeng, G.-M.; Jiang, M.; Yang, Z.-Z.; Yu, Z.-G.; Zhu, M.-Y.; Shen, L.-Q. Quaternary Ammonium Compounds (QACs): A Review on Occurrence, Fate and Toxicity in the Environment. Sci. Total Environ. 2015, 518–519, 352–362. [Google Scholar] [CrossRef]
  48. Ramage, G.; Saville, S.P.; Thomas, D.P.; López-Ribot, J.L. Candida Biofilms: An Update. Eukaryot. Cell 2005, 4, 633–638. [Google Scholar] [CrossRef] [PubMed]
  49. Bertolini, M.; Costa, R.C.; Barão, V.A.R.; Cunha Villar, C.; Retamal-Valdes, B.; Feres, M.; Silva Souza, J.G. Oral Microorganisms and Biofilms: New Insights to Defeat the Main Etiologic Factor of Oral Diseases. Microorganisms 2022, 10, 2413. [Google Scholar] [CrossRef] [PubMed]
  50. Chen, F.; Wang, D. Novel Technologies for the Prevention and Treatment of Dental Caries: A Patent Survey. Expert Opin. Ther. Pat. 2010, 20, 681. [Google Scholar] [CrossRef] [PubMed]
  51. Zupancic Cepic, L.; Dvorak, G.; Piehslinger, E.; Georgopoulos, A. In Vitro Adherence of Candida Albicans to Zirconia Surfaces. Oral Dis. 2020, 26, 1072–1080. [Google Scholar] [CrossRef]
  52. Patil, S.; Rao, R.S.; Majumdar, B.; Anil, S. Clinical Appearance of Oral Candida Infection and Therapeutic Strategies. Front. Microbiol. 2015, 6, 1391. [Google Scholar] [CrossRef]
  53. Garcia-Cuesta, C.; Sarrion-Pérez, M.G.; Bagán, J.V. Current Treatment of Oral Candidiasis: A Literature Review. J. Clin. Exp. Dent. 2014, 6, e576–e582. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of QAUDMA-m monomers.
Figure 1. Chemical structure of QAUDMA-m monomers.
Materials 16 03855 g001
Figure 2. Monomer composition preparation and subsequent crosslinking copolymerization.
Figure 2. Monomer composition preparation and subsequent crosslinking copolymerization.
Materials 16 03855 g002
Figure 3. Viability of L929 mouse fibroblast cells after 24 h of incubation. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Figure 3. Viability of L929 mouse fibroblast cells after 24 h of incubation. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Materials 16 03855 g003
Figure 4. Reduction in cell viability of BG:QAm:TEGs relative to the control.
Figure 4. Reduction in cell viability of BG:QAm:TEGs relative to the control.
Materials 16 03855 g004
Figure 5. The results from the fungal C. albicans (ATCC 2091) adhesion test. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Figure 5. The results from the fungal C. albicans (ATCC 2091) adhesion test. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Materials 16 03855 g005
Figure 6. The fungal C. albicans (ATCC 2091) growth inhibition zone diameter. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Figure 6. The fungal C. albicans (ATCC 2091) growth inhibition zone diameter. Lowercase letters indicate statistically insignificant differences (p > 0.05) (non-parametric Wilcoxon test).
Materials 16 03855 g006
Figure 7. Results of microbial activity tests: (a) C. albicans (ATCC 2091), (b) S. aureus (ATCC 25923), and (c) E. coli (ATCC 25922).
Figure 7. Results of microbial activity tests: (a) C. albicans (ATCC 2091), (b) S. aureus (ATCC 25923), and (c) E. coli (ATCC 25922).
Materials 16 03855 g007
Figure 8. The origin of the DCRMs’ cytotoxic effect.
Figure 8. The origin of the DCRMs’ cytotoxic effect.
Materials 16 03855 g008
Figure 9. The action of bacteria and fungi comprising oral biofilm.
Figure 9. The action of bacteria and fungi comprising oral biofilm.
Materials 16 03855 g009
Figure 10. Comparing the water contact angle (WCA) values [28] with the number of fungi colonies on BG:QAm:TEG surfaces.
Figure 10. Comparing the water contact angle (WCA) values [28] with the number of fungi colonies on BG:QAm:TEG surfaces.
Materials 16 03855 g010
Figure 11. Comparing the C. albicans (ATCC 2091) growth inhibition zones with QA groups concentration (xQA) in BG:QAm:TEG compositions.
Figure 11. Comparing the C. albicans (ATCC 2091) growth inhibition zones with QA groups concentration (xQA) in BG:QAm:TEG compositions.
Materials 16 03855 g011
Table 1. Polymerization shrinkage (Se), glass transition temperature (Tgp), water contact angle (WCA), water sorption (WS), and solubility (SL) of BG:QAm:TEGs [34].
Table 1. Polymerization shrinkage (Se), glass transition temperature (Tgp), water contact angle (WCA), water sorption (WS), and solubility (SL) of BG:QAm:TEGs [34].
Se (%)Tgp (°C)WCA (°)WS
(µg/mm3)
SL
(µg/mm3)
BG:QA8:TEG5.0842.2181.4168.275.15
BG:QA10:TEG5.4845.8184.6848.425.18
BG:QA12:TEG6.0746.6386.3235.545.22
BG:QA14:TEG6.1447.8385.5234.435.58
BG:QA16:TEG6.2450.4191.0532.675.42
BG:QA18:TEG6.4050.8199.5325.945.54
BG:UD:TEG8.3555.9080.7611.711.12
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chrószcz-Porębska, M.W.; Barszczewska-Rybarek, I.M.; Kazek-Kęsik, A.; Ślęzak-Prochazka, I. Cytotoxicity and Microbiological Properties of Copolymers Comprising Quaternary Ammonium Urethane-Dimethacrylates with Bisphenol A Glycerolate Dimethacrylate and Triethylene Glycol Dimethacrylate. Materials 2023, 16, 3855. https://doi.org/10.3390/ma16103855

AMA Style

Chrószcz-Porębska MW, Barszczewska-Rybarek IM, Kazek-Kęsik A, Ślęzak-Prochazka I. Cytotoxicity and Microbiological Properties of Copolymers Comprising Quaternary Ammonium Urethane-Dimethacrylates with Bisphenol A Glycerolate Dimethacrylate and Triethylene Glycol Dimethacrylate. Materials. 2023; 16(10):3855. https://doi.org/10.3390/ma16103855

Chicago/Turabian Style

Chrószcz-Porębska, Marta W., Izabela M. Barszczewska-Rybarek, Alicja Kazek-Kęsik, and Izabella Ślęzak-Prochazka. 2023. "Cytotoxicity and Microbiological Properties of Copolymers Comprising Quaternary Ammonium Urethane-Dimethacrylates with Bisphenol A Glycerolate Dimethacrylate and Triethylene Glycol Dimethacrylate" Materials 16, no. 10: 3855. https://doi.org/10.3390/ma16103855

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