Next Article in Journal
Efficient Photocatalytic Luminous Textile for Simulated Real Water Purification: Advancing Economical and Compact Reactors
Previous Article in Journal
Thermodynamic Guidelines for the Mechanosynthesis or Solid-State Synthesis of MnFe2O4 at Relatively Low Temperatures
Previous Article in Special Issue
Medical-Radiation-Shielding Film Fabricated by Imitating the Layered Structure Pattern of Abalone Shell and Verification of Its Shielding Effect
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical Modification of Dental Dimethacrylate Copolymer with Tetramethylxylylene Diisocyanate-Based Quaternary Ammonium Urethane-Dimethacrylates—Physicochemical, Mechanical, and Antibacterial Properties

by
Patryk Drejka
1,
Marta Chrószcz-Porębska
1,
Alicja Kazek-Kęsik
2,3,
Grzegorz Chladek
4 and
Izabela Barszczewska-Rybarek
1,*
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 Engineering Materials and Biomaterials, Faculty of Mechanical Engineering, Silesian University of Technology, Konarskiego 18A Str., 44-100 Gliwice, Poland
*
Author to whom correspondence should be addressed.
Materials 2024, 17(2), 298; https://doi.org/10.3390/ma17020298
Submission received: 18 December 2023 / Revised: 1 January 2024 / Accepted: 5 January 2024 / Published: 7 January 2024

Abstract

:
In this study, two novel quaternary ammonium urethane-dimethacrylates (QAUDMAs) were designed for potential use as comonomers in antibacterial dental composite restorative materials. QAUDMAs were synthesized via the reaction of 1,3-bis(1-isocyanato-1-methylethyl)benzene with 2-(methacryloyloxy)ethyl-2-decylhydroxyethylmethylammonium bromide (QA10+TMXDI) and 2-(methacryloyloxy)ethyl-2-dodecylhydroxyethylmethylammonium bromide (QA12+TMXDI). Their compositions with common dental dimethacrylates comprising QAUDMA 20 wt.%, urethane-dimethacrylate monomer (UDMA) 20 wt.%, bisphenol A glycerolate dimethacrylate (Bis-GMA) 40 wt.%, and triethylene glycol dimethacrylate (TEGDMA) 20 wt.%, were photocured. The achieved copolymers were characterized for their physicochemical and mechanical properties, including their degree of conversion (DC), glass transition temperature (Tg), polymerization shrinkage (S), water contact angle (WCA), flexural modulus (E), flexural strength (FS), hardness (HB), water sorption (WS), and water leachability (WL). The antibacterial activity of the copolymers was characterized by the minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC) against Staphylococcus aureus and Escherichia coli. The achieved results were compared to the properties of a typical dental copolymer comprising UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. The introduction of QAUDMAs did not deteriorate physicochemical and mechanical properties. The WS and WL increased; however, they were still satisfactory. The copolymer comprising QA10+TMXDI showed a higher antibacterial effect than that comprising QA12+TMXDI and that of the reference copolymer.

Graphical Abstract

1. Introduction

Faced with the increase in the dental caries scale [1,2], it is important to find effective solutions for curing this disease. Therefore, in recent years, an upward trend in the development of antimicrobial dental composite restorative materials (DCRMs) is observed [3,4,5,6,7,8,9,10,11,12,13,14,15]. The elementary concept of providing DCRMs with antimicrobial properties assumes physically dispersing bioactive substances (such as antibiotics, antimicrobial enzymes, chlorohexidine, triclosan, metals, metal oxides, and the nanoparticles of quaternized polyethyleneimine) in available dental materials [5,6,7,8,9,10,11,12,14]. As these types of biocides can easily elute from a material, the material achieved shows only a short-term antimicrobial effect. It quickly loses its mechanical and antimicrobial properties and causes cytotoxic effects [16,17].
Another concept of providing DCRMs with antimicrobial properties is based on the covalent incorporation of a monomeric biocide into the polymer network structure, constituting the DCRM matrix. It involves the copolymerization of universal dental dimethacrylates (DMAs), such as bisphenol A glycerolate dimethacrylate (Bis-GMA), its ethoxylated derivative (Bis-EMA), urethane-dimethacrylate monomer (UDMA), and triethylene glycol dimethacrylate (TEGDMA), with dimethacrylates having bioactive groups, of which dimethacrylates with quaternary ammonium (QA) groups (QADMAs) are the best known [18,19,20,21,22]. The QADMA content in a copolymer ranges from a few to several dozen percent. This alternative offers the stable, non-leaching, and long-term antimicrobial action of a DCRM [23,24]. The huge advantages of monomeric antimicrobials include lower cytotoxic effects in comparison to physically dispersed antibacterial agents [16,25] and an alternative to antibiotics, whose application should be limited due to the increasing resistance of microorganisms [26,27,28]. The aforementioned features of QADMAs explain the great interest among scientists in designing novel monomer structures containing QA groups and the development of related materials.
The following series of QADMAs are described in the literature:
(i)
dimethacrylate derivatives of N-methyldiethanolamine (MDEA) with one central quaternary ammonium group (QAn+MDEA, where n is the number of carbon atoms in the N-alkyl substituent of the QA group). QA12+MDEA and QA16+MDEA were studied [29,30];
(ii)
dimethacrylate derivatives of N,N-dimethylaminoethyl methacrylate (DMAEMA) with two QA groups separated with a central aliphatic chain (QAn+DMAEMA, where n is the number of carbon atoms in the chain separating QA groups). QA groups in these monomers were substituted with two methyl groups. QA4+DMAEMA and QA6+DMAEMA were studied [31];
(iii)
the Bis-GMA quaternary ammonium derivative with two QA groups (QAn+bis-GMA, where n is the number of carbon atoms in the N-alkyl substituent of the QA group). QA6+bis-GMA was studied [32];
(iv)
fully aliphatic urethane-dimethacrylates with one central QA group, based on N-methyldiethanolamine (MDEA) and 2-isocyanatoethyl methacrylate (IEM) (QAn+MDEA+IEM, where n is the number of carbon atoms in the N-alkyl substituents of the QA group). QAn+MDEA+IEMs with n of 12, 14, 16, and 18 were studied [33,34];
(v)
urethane-dimethacrylates with one central QA group, based on N-methyldiethanolamine (MDEA) and cycloaliphatic isophorone diisocyanate (IPDI) (QAn+MDEA+IPDI, where n is the number of carbon atoms in the N-alkyl substituent of the QA group). QAn+MDEA+IPDIs with n of 12, 14, 16, and 18 were studied [35,36];
(vi)
quaternary ammonium analogues of the urethane-dimethacrylate monomer with two QA groups (QAn+TMDI, where n is the number of carbon atoms in the N-alkyl substituent of the QA group). QAn+TMDI with n of 8, 10, 12, 14, 16, and 18 were studied [37,38,39,40,41,42].
QAn+MDEAs showed high antibacterial activity against many bacteria strains, including S. aureus, Streptococcus mutans, Actinomyces viscosus, Lactobacillus acidophilus, Streptococcus sanguinis, Porphyromonas gingivalis, Prevotella melaninogenica, and Enterococcus faecalis. In addition, these monomers showed cytotoxicity lower than Bis-GMA [29]. The copolymer of QAn+MDEA 10 wt.% with Bis-GMA/TEGDMA 90 wt.% showed antibacterial activity against S. mutans [30]. QAn+DMAEMA 1 wt.% was used to modify a commercial adhesive resin (Tetric N-Bond) and the resulting material also showed antibacterial activity against S. mutans [31]. The introduction of QAn+Bis-GMA 5 wt.% into the Bis-GMA/TEGDMA formulation resulted in a copolymer with antibacterial activity against E. coli, S. aureus, S. mutans, and Bacillus subtillis. The antifungal activity of the modified copolymer against Candida albicans was observed too. As the QAn+Bis-GMA concentration increased, the antimicrobial activity increased; however, it was accompanied by an increase in cytotoxicity and the deterioration of mechanical properties [32]. The modification of the same Bis-GMA/TEGDMA formulation with QAn+MDEA+IEM 5 wt.% resulted in an antibacterial copolymer only for QA16+MDEA+IEM. It showed antibacterial activity against S. mutans. All other copolymers gained antibacterial activity as the QAn+MDEA+IEMs concentration increased to 10 wt.%, but this was accompanied by the deterioration in mechanical properties [33,34]. The QAn+MDEA+IPDI 50 wt.% and TEGDMA 50 wt.% copolymers also showed high antibacterial activity against S. mutans. However, the disadvantages of these copolymers included a low flexural strength, low modulus of elasticity, and high water sorption [35,36]. By reducing the QAn+MDEA+IPDI content to 30 wt.% and using a formulation of Bis-GMA/TEGDMA 70 wt.%, an improvement in mechanical properties was achieved [36]. QAn+TMDIs and their copolymers were the most extensively studied [37,38,39,40,41,42]. Copolymers of QAn+TMDI 60 wt.% and TEGDMA 40 wt.% were characterized by very high antibacterial activity against S. aureus and E. coli. However, they had insufficiently good mechanical parameters, excessively high water sorption, and water leachability [38,39]. Copolymers of QAn+TMDI 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.% also showed high antimicrobial activity against S. aureus and E. coli [40]. In addition, antifungal activity against C. albicans was observed for these copolymers [41]. Many of their physicochemical, mechanical, and biological parameters were good. However, the detailed analysis showed that an increase in the N-alkyl substituent length resulted in a decrease in the antibacterial activity, and the deterioration of mechanical properties [40,41,42]. This analysis led to the conclusion that copolymers based on QA10+TMDI and QA12+TMDI contain the optimum combination of properties [40].
The promising results obtained for copolymers based on QA10+TMDI and QA12+TMDI motivated us to replace the TMDI core with a 1,3-bis(1-isocyanate-1-methylethyl)benzene (TMXDI) core to obtain new urethane-dimethacrylate monomers with two QA groups located in the wings and 1,3-phenylene ring located in the core. We formulated a hypothesis that the copolymerization of the achieved quaternary ammonium urethane-dimethacrylates (QAUDMAs) with UDMA, Bis-GMA, and TEGDMA results in an antibacterial copolymer with good physicochemical and mechanical properties. Therefore, the goal of this work was to synthesize two novel urethane-dimethacrylates with the TMXDI core and two methacrylate-ended wings comprising the QA group in the middle. The QA groups were substituted with an N-alkyl chain of 10 and 12 carbon atoms (QA10+TMXDI and QA12+TMXDI, respectively). The two novel monomers were compounded with common dental dimethacrylates in the following weight fractions: QAUDMA 20 wt.%, UDMA 20 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. The two novel monomers were also photocured. The novel monomers, as well as their compositions with dental dimethacrylates, were characterized for molecular weight (MW), concentration of double bonds (xDB), viscosity (η), refractive index (RI), and density (dm). The degree of conversion (DC), density (dp), polymerization shrinkage (S), glass transition temperature (Tg), water contact angle (WCA), water sorption (WS), water solubility (SL), hardness (HB), flexural strength (FS), flexural modulus (E), minimum bactericidal concentration (MBC), and minimum inhibitory concentration (MIC) were all measured for the resulting copolymers (the first comprising QA10+TMXDI (20(QA10+TMXDI)p) and the second comprising QA12+TMXDI (20(QA12+TMXDI)p). The results were subjected to a comparative analysis of the properties of the standard dental copolymer consisting of UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.% (40(UDMA)p).
The novelty of this work combines the development of two new urethane-dimethacrylate monomers with quaternary ammonium groups (QAUDMAs) and the complex physicochemical, mechanical, and antibacterial characteristics of their copolymers with common dimethacrylates used in dentistry. So far, QAUDMAs synthesized from 1,3-bis(1-isocyanato-1-methylethyl)benzene have not been described in the literature.
The development of such materials is very important for dental science and may contribute to reducing the scale of dental carries in the future.

2. Materials and Methods

2.1. Materials

Bisphenol A-glycidyl dimethacrylate (Bis-GMA), triethylene glycol dimethacrylate (TEGDMA), urethane-dimethacrylate (UDMA), 1,3-bis(1-isocyanato-1-methylethyl)benzene (TMXDI), phenothiazine (PTZ), camphorquinone (CQ), N,N-dimethylaminoethyl methacrylate (DMAEMA), potassium bromide (KBr, FT-IR grade), and tetramethylsilane (TMS) were purchased from Sigma-Aldrich, St. Louis, MO, USA. Methyl methacrylate (MMA), N-methyldiethanolamine (MDEA), decyl bromide (DB), and dodecyl bromide (DDB) were purchased from Acros Organics, Geel, Belgium. Dibutyltin dilaurate (DBTDL) was purchased from Fluka, Charlotte, NC, USA. Deuterated chloroform (CDCl3) and deuterated dichloromethane (CD2Cl2) were purchased from Deutero GMBH, Kastellaun, Germany. Potassium carbonate (K2CO3), and magnesium sulfate (MgSO4) were purchased from Chempur, Piekary Śląskie, Poland. Toluene, chloroform, and dichloromethane were purchased from Stanlab, Lublin, Poland. All chemicals were used as received.

2.2. Chemical Syntheses

2.2.1. N,N-(2-Hydroxyethyl)methylaminoethyl Methacrylate (HAMA)

The transesterification reaction reagents—MMA 1.00 mol (100.12 g), and MDEA 0.67 mol (79.85 g), the catalyst—K2CO3 8 wt.% (14.40 g), the polymerization inhibitor—PTZ 0.05% (0.09 g) and the solvent—toluene 400 cm3 were introduced into a single-necked round-bottom flask equipped with a Vigreux column and distillation head. The reaction mixture was brought to a boil. The equilibrium of the reaction was moved in favor of HAMA by the continuous collection of a distillate consisting of an azeotropic mixture of methanol (condensation reaction by-product), MMA (used in excess), and toluene (solvent). The reaction was carried out until the temperature at the column head reached 100 °C, which was achieved after 2.5 h. The cooled reaction mixture was filtered and washed three times with distilled water in a 1:2 volume ratio (HAMA and MDEA were soluble in water). The aqueous fractions were combined and extracted three times with chloroform in a 1:3 volume ratio (only HAMA was soluble in chloroform; as MDEA was water-insoluble, it remained in a water fraction). The chloroform fractions, containing only HAMA, were also combined, dried with MgSO4 overnight, and chloroform was evaporated on a rotary evaporator under reduced pressure (30 and then 3 mbar). The obtained raw product was subjected to vacuum distillation (3 mbar), collecting a boiling fraction from 110 to 130 °C. The process yielded HAMA 14%.

2.2.2. 2-(Methacryloyloxy)ethyl-2-hydroxyethylmethylalkylammonium Bromides (QAHAMA-n, Where n Is the Number of Carbon Atoms in the N-Alkyl Substituent)

The Menshutkin reaction reagents—HAMA 0.107 mol (20.00 g), and alkyl bromide 0.107 mol (DB 23.66 g and DDB 26.67 g), and the polymerization inhibitor—PTZ 0.05 wt.% (0.022 and 0.023 g, respectively, in the reaction with DB and DDB) were introduced into a three-necked round-bottom flask equipped with a mechanical stirrer, reflux condenser, and thermometer. The reaction was carried out at 80 °C for 90 h. The reaction yielded 100% of 2-(methacryloyloxy)ethyl-2-decylhydroxyethylmethylammonium bromide (QAHAMA-10) and 2-(methacryloyloxy)ethyl-2-dodecylhydroxyethylmethylammonium bromide (QAHAMA-12).

2.2.3. Quaternary Ammonium Urethane-Dimethacrylates (QAUDMA)

A 50% solution of an addition reaction reagent—QAHAMA-n 0.070 mol (QAHAMA-10 28.50 g and QAHAMA-12 28.6 g), a catalyst—DBTDL 0.035 wt.% (0.013 g) and a solvent—dichloromethane 21 cm3 were introduced into a three-necked round-bottom flask equipped with a reflux condenser, thermometer, and dropping funnel. A 50% solution of the second reagent, composed of TMXDI 0.035 mol (8.55 g) and dichloromethane 6.5 cm3, was placed in a dropping funnel. The flask content was brought to a boil (approximately 42 °C), and then a TMXDI solution was added dropwise for 1.5 h. The reaction was carried out for 5 h, maintaining the reaction mixture at the boiling point. Dichloromethane was evaporated on a rotary evaporator under reduced pressure (30 and then 3 mbar). A light-yellow, viscous, liquid substance remained in the flask, constituting the product. The flask with the product was placed in the laboratory dryer (SLW 53 STD, POL-EKO, Wodzisław Śląski, Poland) and heated at 42 °C for 24 h. The reaction yielded 100% both QAUDMAs, i.e., the addition reaction of TMXDI and QAHAMA-10 yielded QA10+TMXDI, and that of QAHAMA-12 yielded QA12+TMXDI.

2.3. Curing Procedure and Sample Preparation

Two experimental monomer compositions were prepared using mechanical stirring at 50 °C. They consisted of QAUDMA (QA10+TMXDI and QA12+TMXDI) 20 wt.%, UDMA 20 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. The reference composition consisted of UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. The components of the initiating system, comprising CQ 0.4 wt.% and DMAEMA 1 wt.%, were introduced into homogeneous monomer mixtures and stirring was continued at the same temperature until CQ dissolved. The achieved compositions were poured into the molds, covered with the PET foil, and irradiated with the UV–VIS lamp (Ultra Vitalux 300, Osram, Munich, Germany) at room temperature for 1 h. The lamp emitted radiation in the range of 280 to 780 nm and radiation exitance of 2400 mW/cm2. The achieved casts were processed following the guidelines for specimen preparation for testing particular properties. Each specimen was polished with fine sandpaper before experiments.

2.4. Nuclear Magnetic Resonance Spectroscopy (NMR)

The NMR 300 MHz spectrometer (UNITY/INOVA, Varian, Palo Alto, CA, USA) was used to collect 256 scan 1H spectra and 40,000 scan 13C NMR spectra of products achieved. CD2Cl2 and CDCl3 were used as solvents and TMS was used as an internal standard.

2.5. Fourier Transform Infrared Spectroscopy (FTIR)

The FTIR spectrometer (Spectrum Two, Perkin-Elmer, Waltham, MA, USA) was used to collect spectra (128 scans, 1 cm−1 resolution).
Monomers were tested as a very thin layer between two KBr pellets. Copolymers were powdered, sieved to a diameter of less than 25 µm, and dispersed in KBr pellets.
The degree of conversion (DC) in copolymers was calculated from the following equation:
D C = 1 A C = C A A r p o l y m e r A C = C A A r m o n o m e r
The variables in the equation are defined as follows:
  • AC=C—the absorbance of the stretching vibration band of the carbon–carbon double bond in the methacrylate group at 1636 cm−1;
  • AAr—the absorbance of the aromatic stretching skeletal vibrations band at 1608 cm−1.

2.6. Viscosity

The rotating viscometer (Visco Star Plus L, Brookfield Fungilab Viscometer, Barcelona, Spain) was used to determine the viscosity (η). The procedure was carried out at 25 °C according to the ISO 2555-2018 standard [43].

2.7. Refractive Index

The digital refractometer (DR 6100T, Krüss Optronic, Hamburg, Germany) was used to determine the refractive index (RI). The procedure was carried out according to the ISO 489:2022 standard [44].

2.8. Density and Polymerization Shrinkage

The density of monomers (dm) was determined with the 1 mL pycnometer. The density of copolymers (dp) was determined with the analytical balance (XP Balance, Mettler Toledo, Greifensee, Switzerland) analytical balance equipped with the density determination kit.
The polymerization shrinkage (S) was calculated from the following equation:
S % = 1 d m d p × 100
The variables in the equation are defined as follows:
  • dm—the monomer density;
  • dp—the polymer density.

2.9. Differential Scanning Calorimetry (DSC)

The differential scanning calorimeter (DSC 3, Mettler Toledo, Greifensee, Switzerland) was used in the measurements, utilizing powdered polymer samples and standard aluminum crucibles. The 10 K/min heating rate, 0 to 200 °C temperature range, and air atmosphere were used in all experiments.
The glass transition temperature (Tg) of copolymers was read off the DSC curve as the midpoint of the transition region, following the 11357-2:2020 standard [45].

2.10. Water Contact Angle

The goniometer (OCA 15EC, Data Physics, Filderstadt, Germany) was used to determine the water contact angle (WCA) of copolymer surfaces. The analysis was performed on disc-like specimens with dimensions of 15 mm × 1.5 mm (diameter × thickness) utilizing the sessile drop method and 4 µL of deionized water. The measurement was performed immediately after the drop was placed on the sample surface.

2.11. Water Sorption and Solubility

The water sorption (WS) and solubility (SL) of polymers were determined according to the guidelines of the ISO 4049:2019 standard [46] on disc-like specimens with dimensions of 15 mm × 1.5 mm (diameter × thickness). The analytical balance (XP Balance, Mettler Toledo, Greifensee, Switzerland) was used in experiments.
The specimens were dried to constant weight (m0) in a laboratory dryer (SLW 53 STD, POL-EKO, Wodzisław Śląski, Poland) and immersed in deionized water at room temperature for 7 days. After that, samples were removed from the water, dried with blotting paper, weighed (m1), and dried again to a constant weight (m2).
WS and SL were calculated using the following equations:
W S   μ g m m 3 = m 1 m 0 V ,
S L   μ g m m 3 = m 0 m 2 V
The variables in the equation are defined as follows:
  • m0—the initial mass of the dried samples;
  • m1—the mass of the swollen samples;
  • m2—the mass of the dried samples after immersion in water;
  • V—the initial volume of the dried samples.

2.12. Mechanical Properties

2.12.1. Hardness

The hardness tester (VEB, Werkstoffprűfmaschinen, Leipzig, Germany) was used to determine the ball indentation hardness (HB). The procedure was carried out according to the ISO 2039-1:2001 standard [47] on disc-like specimens with dimensions of 40 mm × 4 mm (diameter × thickness).
HB was calculated using the following equation:
H B   M P a = F m 0.21 h h r + 0.21 π d h r
The variables in the equation are defined as follows:
  • Fm—the test load;
  • d—the ball intender diameter (d = 5 mm);
  • h—the immersion depth;
  • hr—the reduced immersion depth (hr = 0.25 mm).

2.12.2. Flexural Properties

The universal testing machine (Z020, Zwick, Ulm, Germany) was used to determine the flexural strength (FS) and flexural modulus (E). The procedure was carried out according to the ISO 178:2019 standard [48] on bar specimens with dimensions of 64 mm × 10 mm × 3.3 mm (length × width × thickness).
FS and E were calculated using the following equations:
F S   M P a = 3 P l 2 b d 2 ,
E   M P a = P 1 l 3 4 b d 3 δ
The variables in the equation are defined as follows:
  • P1—the load at the selected point of the elastic region of the stress–strain plot;
  • P—the maximum load;
  • l—the support span;
  • b—the sample width;
  • d—the sample thickness;
  • δ—the deflection of the sample at P1.

2.13. Antibacterial Properties

The antibacterial properties were characterized by the minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC) against S. aureus (ATCC 25923) and E. coli (ATCC 25922). The tryptic soy broth (TSB) nutritious medium was used for the bacteria cultivation, which was carried out at 37 °C for 18 h in an incubator (ILW 53, POL-EKO, Wodzisław Śląski, Poland). The experiments were carried out on polymer powders sieved to a grain diameter of less than 25 µm.
A series of six copolymer suspensions in TSB were prepared with concentrations of 50, 25, 12.5, 6.25, 3.125, and 1.563 mg/mL. Next, 20 µL of S. aureus and E. coli bacterial suspensions in TSB (~5 × 108 CFU/mL) were introduced into the copolymer suspensions, vortexed (10 s, 2000 rpm), and incubated at 37 °C for 18 h. Suspensions were then vortexed again (10 s, 2000 rpm), and 100 µL of each was applied on an agar plate (Müller-Hinton agar, Diag-Med, Raszyn, Poland). The agar plates were incubated at 37 °C for 18 h. The number of bacterial colonies was qualitatively assessed by visual comparison to the control. The lowest concentration at which no bacterial colonies were observed on agar plates was taken as the MBC. The lowest concentration at which the bacteria growth was reduced was taken as the MIC.

2.14. Statistical Analysis

The software (Statistica 13.3, TIBCO Software Inc., Palo Alto, CA, USA) was used to perform the non-parametric Wilcoxon test with a significance level (p) of 0.05. The statistical significance of the results was determined for five samples. The results were expressed as the average value (AV) and corresponding standard deviation (SD).

3. Results

3.1. Monomer Synthesis and Characterization

As part of this work, two QAUDMA monomers, QA10+TMXDI and QA12+TMXDI, were obtained. To achieve this goal, a procedure developed for the QAn+TMDI monomer synthesis was adopted [37].
The synthesis route of QA10+TMXDI and QA12+TMXDI included three stages (Figure 1):
  • The transesterification of MMA with MDEA that led to HAMA formation. It was conducted with the MMA excess and K2CO3 catalyst;
  • The Menshutkin reaction by the reaction of HAMA with DB and DDB resulted in a conversion of the HAMA tertiary amine group into a quaternary ammonium group in QAHAMA-n. QAHAMA-10 and QAHAMA-12 were achieved, respectively;
  • The addition reaction of QAHAMA-n and TMXDI was conducted with the DBTDL catalyst and resulted in the formation of QAUDMAs: QA10+TMXDI and QA12+TMXDI;
  • The chemical structure of the obtained monomers was confirmed using spectroscopic methods, including 1H NMR, 13C NMR, and FTIR.
Figure 2 shows the 1H NMR spectra of the QA10+TMXDI and QA12+TMXDI monomers. Table 1 lists the proton signals present in the spectra with their assignment to chemical groups.
Figure 3 shows the 13C NMR spectra of the QA10+TMXDI and QA12+TMXDI monomers. Table 2 lists carbon atom signals present in the spectra with their assignment to chemical groups.
Figure 4 shows the FTIR spectra of the QA10+TMXDI and QA12+TMXDI monomers. Table 3 lists of absorption bands present in the spectra with their assignment to chemical groups.
The QA10+TMXDI and QA12+TMXDI monomers were characterized by their molecular weight (MW), concentration of double bonds (xDB), refractive index (RI), and density (dm). Table 4 presents the results for the experimental monomers and three commercial dental dimethacrylate monomers (Bis-GMA, UDMA, and TEGDMA), for comparison.
The molecular weight (MW) of QA10+TMXDI was 1061 g/mol and QA12+TMXDI had a MW that was 5% higher. Compared to Bis-GMA, UDMA, and TEGDMA, QA10+TMXDI as well as QA12+TMXDI had a higher MW, on average by 53, 57, and 74%, respectively. The MW was converted to the concentration of double bonds (xDB). The xDB of QA10+TMXDI was 1.89 mol/kg and QA12+TMXDI had a xDB that was 5% lower. Compared to Bis-GMA, UDMA, and TEGDMA, QA10+TMXDI as well as QA12+TMXDI had a lower xDB, on average by 53, 57, and 74%, respectively.
The refractive index (RI) of QA10+TMXDI was 1.5230 and QA12+TMXDI had a RI that was 0.6% lower. This difference did not have statistical significance. Compared to Bis-GMA, QA10+TMXDI as well as QA12+TMXDI had a lower RI, on average by 2%. Compared to UDMA and TEGDMA, QA10+TMXDI as well as QA12+TMXDI had a higher RI, on average by 4 and 3%, respectively.
The density (dm) of QA10+TMXDI was 1.16 g/cm3 and QA12+TMXDI had a dm that was 15% higher. This difference was statistically significant. Compared to Bis-GMA, UDMA, and TEGDMA, QA10+TMXDI as well as QA12+TMXDI had a higher dm. In the case of QA10+TMXDI, these differences were 1, 7, and 9%, respectively, for Bis-GMA, UDMA, and TEGDMA. In the case of QA12+TMXDI, these differences were greater and equaled 16, 22, and 25%, respectively, for Bis-GMA, UDMA, and TEGDMA.

3.2. Characterization of Monomer Compositions

The obtained QA10+TMXDI and QA12+TMXDI monomers were used to prepare compositions with commercial DMA monomers. QAUDMA 20 wt.% was mixed with UDMA 20 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. For comparison purposes, the UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.% mixture was prepared. Table 5 shows the names, weight, and molar ratios of the prepared monomer compositions.
The 20(QA10+TMXDI)m and 20(QA12+TMXDI)m experimental monomer compositions and 40(UDMA)m reference monomer composition were characterized by their molecular weight (MW), concentration of double bonds (xDB), viscosity (η), refractive index (RI), and density (dm). Table 6 lists the values of these parameters.
The MW of 20(QA10+TMXDI)m was 476 g/mol and 20(QA12+TMXDI)m had a MW that was 1% higher. Compared to 40(UDMA)m, 20(QA10+TMXDI)m as well as 20(QA12+TMXDI)m had a higher MW, by 10 and 11%, respectively. The MW was converted to the concentration of double bonds in the monomer compositions. The xDB of 20(QA10+TMXDI)m was 4.20 mol/kg and 20(QA12+TMXDI)m had a xDB that was 1% lower. Compared to 40(UDMA)m, 20(QA10+TMXDI)m as well as 20(QA12+TMXDI)m had a lower xDB, by 10 and 11%, respectively.
The η of 20(QA10+TMXDI)m was 3.79 Pa∙s and 20(QA12+TMXDI)m had a 15% lower η. Compared to 40(UDMA)m, 20(QA10+TMXDI)m had a higher η, whereas 20(QA12+TMXDI)m had a lower η, by 7 and 9%, respectively. All these differences did not have statistical significance.
The RI of 20(QA10+TMXDI)m was 1.5101 and 20(QA12+TMXDI)m had a RI that was 0.05% higher. Compared to 40(UDMA)m, 20(QA10+TMXDI)m as well as 20(QA12+TMXDI)m had a higher RI, by 0.30 and 0.35%. All these differences did not have statistical significance.
The dm of 20(QA10+TMXDI)m was 1.13 g/cm3 and 20(QA12+TMXDI)m had a dm that was 2% lower. This difference was statistically significant. Compared to 40(UDMA)m, 20(QA10+TMXDI)m had a dm that was 2% lower, and this difference was statistically significant, whereas 20(QA12+TMXDI)m had the same dm, and this result did not have statistical significance.

3.3. Characterization of Copolymers

20(QA10+TMXDI)m, 20(QA12+TMXDI)m, and 40(UDMA)m were subjected to photopolymerization, utilizing an initiation system consisting of CQ and DMAEMA, which is typical for the DCRM photocuring. The structure of the obtained copolymers was characterized by the degree of conversion (DC). The physical properties were characterized by determining the polymerization shrinkage (S), glass transition temperature (Tg), water sorption (WS), water solubility (SL), and water contact angle (WCA). The mechanical properties were characterized by determining the hardness (HB), flexural modulus (E), and flexural strength (FS). The copolymers were also tested for antibacterial properties by determining the minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC) against S. aureus (ATCC 25923) and E. coli (ATCC 25922).

3.3.1. Degree of Conversion (DC)

The DC was determined utilizing the FTIR spectroscopy and internal standard method. As the tested systems comprised the TMXDI moiety, it was possible to use the aromatic skeletal vibration band as a standard (Figure 5). This is the most widely used method in the DC characterization of poly(dimethacrylate)s because this band does not change its specificity due to polymerization [49]. Table 7 shows the DC in the copolymers tested.
The DC in 20(QA10+TMXDI)p was 0.57 and 20(QA12+TMXDI)p had a DC that was 21% higher. This difference was statistically significant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p had a higher DC, by 6 and 28%, respectively. The difference for 20(QA10+TMXDI)p was statistically insignificant, whereas that for 20(QA12+TMXDI)p was statistically significant.

3.3.2. Physical Properties

Table 8 shows the results obtained for the dp, S, and Tg of the copolymers tested. The DSC thermograms of the copolymers are presented in Supplementary Materials in Figure S1.
The dp of 20(QA10+TMXDI)p was 1.21 g/cm3 and 20(QA12+TMXDI)p had a dp that was 1% lower. Compared to 40(UDMA)p, 20(QA10+TMXDI)p had a higher dp, by 1%, whereas 20(QA12+TMXDI)p had the same dp. All the results for the dp were statistically insignificant.
The S of 20(QA10+TMXDI)p was 7.04% and the S of 20(QA12+TMXDI)p was 10% higher. Compared to 40(UDMA)p, 20(QA10+TMXDI)p had a lower S, whereas 20(QA12+TMXDI)p had a higher S, by 5 and 4%, respectively. All these differences were statistically significant.
The Tg of 20(QA10+TMXDI)p was 53.46 °C and 20(QA12+TMXDI)p had a Tg that was 13% higher. This difference was statistically significant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p had a lower Tg, by 15 and 4%, respectively. The difference for 20(QA10+TMXDI)p was statistically significant. The other copolymer did not have statistical significance.
Table 9 shows the results obtained for the WS, SL, and WCA of the tested copolymers. The goniometry camera images of the deionized water droplets on the surfaces of the copolymers are presented in Supplementary Materials in Figure S2.
The WS of 20(QA10+TMXDI)p was 10.43 µg/mm3 and 20(QA12+TMXDI)p had a WS that was 1% higher. This difference was statistically insignificant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p had a higher WS, by 96 and 95%, respectively. These differences were statistically significant.
The SL of 20(QA10+TMXDI)p was 2.18 µg/mm3 and 20(QA12+TMXDI)p had a SL that was 13% higher. Compared to 40(UDMA)p, 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p had a higher SL, by 1182 and 1347%, respectively. All these differences were statistically significant.
The WCA of 20(QA10+TMXDI)p was 87.03° and 20(QA12+TMXDI)p had a WCA that was 5% higher. This difference was statistically insignificant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p had a higher WCA, by 17 and 23%, respectively. These differences were statistically significant.

3.3.3. Mechanical Properties

Table 10 shows the results obtained for the HB, E, and FS of the tested copolymers.
The HB of 20(QA10+TMXDI)p was 234.8 MPa and 20(QA12+TMXDI)p had a HB that was 17% lower. Compared to 40(UDMA)p, 20(QA10+TMXDI)p had a HB that was 4% higher, and 20(QA12+TMXDI)p had a HB that was 14% lower. All these differences were statistically significant.
The FS of 20(QA10+TMXDI)p was 100.63 MPa and 20(QA12+TMXDI)p had a FS that was 10% lower. This difference was not statistically significant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p and 20(QA12+TMXDI)p had a lower FS, by 20 and 28%, respectively. These differences were statistically significant.
The E of 20(QA10+TMXDI)p was 3244.9 MPa and 20(QA12+TMXDI)p had a E that was 7% lower. This difference was not statistically significant. Compared to 40(UDMA)p, 20(QA10+TMXDI)p and 20(QA12+TMXDI)p had a lower E, by 10 and 16%, respectively. These differences were statistically significant.

3.3.4. Antibacterial Properties

Table 11 shows the results obtained for the MIC and MBC of the tested copolymers concerning the S. aureus (ATCC 25923) and E. coli (ATCC 25922) bacteria strains. The results of the experiments are presented in Supplementary Materials in Figures S3–S8.
20(QA10+TMXDI)p had an MBC of 12.5 mg/mL, and an MIC of 6.25 mg/mL concerning both bacteria tested. 20(QA12+TMXDI)p had an MBC as well as MIC greater than 50 mg/mL concerning S. aureus. The same copolymer showed a lower MBC and MIC concerning E. coli, respectively, 25 and 12.5 mg/mL. 40(UDMA)p had an MBC as well as MIC greater than 50 mg/mL in each case, i.e., concerning both bacteria strains.

4. Discussion

As part of this study, two novel urethane-dimethacrylate monomers with two QA groups were obtained. For this purpose, the methodology that we developed to obtain an analogous series of urethane-dimethacrylate monomers with two QA groups and a core derived from 1,6-diisocyanato-2,2,4-trimethylhexane (TMDI) was adopted (QAn+TMDI, where n describes the number of carbon atoms in the N-alkyl substituent) [37]. Qan+TMDIs were obtained in a three-step process in which HAMA, QAHAMA-n, and Qan+TMDI were successively synthesized. In this study, we used TMXDI instead of TMDI. TMXDI is classified as an aliphatic isocyanate, but it has a central 1,3-phenylene ring [50]. The addition reaction of QAHAMA-n and TMXDI resulted in the formation of QA10+TMXDI and QA12+TMXDI (Figure 1).
The spectroscopic analysis confirmed the chemical structure of the obtained monomers. The 1H NMR (Figure 2, Table 1), as well as 13C NMR (Figure 3, Table 2) spectra of QA10+TMXDI and QA12+TMXDI, showed the presence of signals coming from the urethane linkage, TMXDI core, methacrylate group, MDEA moiety, and N-alkyl substituent. The signal locations of the -CH3 and -CH2- groups adjacent to the quaternary nitrogen were specific to the QA group neighborhood. Signals derived from the proton and carbon atoms of the -CH3 group neighboring to N+ were found from 3.30 to 3.77 ppm in the 1H NMR spectra and at 53 ppm in the 13C NMR spectra. In the 1H NMR spectrum of HAMA, that -CH3 group gives a singlet at a higher field, i.e., ca. 2.4 ppm. In the 13C NMR spectrum of HAMA, a signal derived from the -CH3 group adjacent to N+ can be also found at a higher field, i.e., ca. 42 ppm [37]). The formation of the urethane linkage was confirmed by the presence of signals of the -NHCOO- proton, present in the range from 7.10 to 7.80 ppm in the 1H NMR spectra, and the -NHCOO- carbon atom, present at 157 ppm in 13C NMR spectra. The quaternization and formation of the urethane linkage were additionally confirmed from the FTIR analysis. In the FTIR spectra, a band of the C-N+ bond vibrations was present at 943 cm−1, whereas the C-N bond vibrations of the -NHCOO- linkage resulted in two bands at 3215 and 1529 cm−1 (Figure 4, Table 3).
Physicochemical properties QA10+TMXDI and QA12+TMXDI depended on the N-alkyl substituent length (Table 4). Its extension with two methylene groups resulted in a 5% increase in the MW and xDB. The RI decreased by 0.6% and the dm increased by 15%.
QA10+TMXDI and QA12+TMXDI had a MW of 1061 and 1117 g/mol and a xDB of 1.89 and 1.79 mol/kg, respectively (Table 4). This showed that the extension of the N-alkyl substituent with two methylene groups resulted in a 5% increase in the MW and the same decrease in the xDB. The Ris of QA10+TMXDI and QA12+TMXDI were 1.5230 and 1.5138, respectively. This corresponded to the statistically insignificant decrease in the RI of 0.6% when the N-alkyl substituent length increased. The dm increased from 1.16 to 1.33 g/cm3 with the increase in the N-alkyl substituent length. This was a 15% change, and this had statistical significance. It can be seen that the physicochemical properties of experimental QAUDMAs depended on the N-alkyl substituent length. The more noticeable change was observed for density. The increase in density may suggest the tighter packing of the QA12+TMXDI molecule than that of QA10+TMXDI, regardless of whether it had longer N-alkyl substituents.
Compared to Bis-GMA, UDMA, and TEGDMA, QA10+TMXDI and QA12+TMXDI had a higher MW (Table 4). The MWs of QA10+TMXDI and QA12+TMXDI were approximately more than twice as much as Bis-GMA (512 g/mol) and UDMA (470 g/mol), and almost four times as much as TEGDMA (286 g/mol). This resulted in a lower xDB of QA10+TMXDI (1.89 mol/kg) and QA12+TMXDI (1.79 mol/kg) compared to dental DMAs (3.90, 4.25, and 6.99 mol/kg, respectively, for Bis-GMA, UDMA, and TEGDMA [37]).
The light-bending ability of QA10+TMXDI and QA12+TMXDI was adequate for a DCRM matrix. Their Ris (respectively, 1.5230 and 1.5138) were very close to that of dental DMAs (1.5493, 1.4614, and 1.4852, respectively, for Bis-GMA, UDMA, and TEGDMA [37]) and were within the range of 1.46 to 1.55, which is recommended for dental systems [51,52].
The densities of QA10+TMXDI (1.16 g/cm3) and QA12+TMXDI (1.33 g/cm3) were higher than those of Bis-GMA, UDMA, and TEGDMA (1.15, 1.09, and 1.07 g/cm3, respectively [37]). If compared to fully aliphatic UDMA and TEGDMA, it is understandable that the synthesized QAUDMAs had a higher density, as the TMXDI benzene ring is heavy and capable of tight packing [53]. The comparison of the dm of QA10+TMXDI and QA12+TMXDI to that of Bis-GMA showed that they had higher densities. Listing the important differences in the chemical structure of experimental QAUDMAs and Bis-GMA may be useful in explaining this observation:
  • Bis-GMA has a more spacious core comprising two 1,4-phenylene rings, whereas QA10+TMXDI and QA12+TMXDI has one central 1,3-phenylene ring;
  • Bis-GMA has shorter aliphatic wings than QA10+TMXDI and QA12+TMXDI;
  • Bis-GMA provides a hydroxyl proton donor to the hydrogen bonds, whereas QA10+TMXDI and QA12+TMXD create these bonds involving the urethane proton donor group;
  • QA10+TMXDI and QA12+TMXDI have two long N-alkyl substituents in the wings.
It can be concluded that the asymmetrical substitution in the central benzene ring and the long N-alkyl substituents do not disturb the tight packing of QA10+TMXDI and QA12+TMXDI.
The QA10+TMXDI and QA12+TMXDI monomers were introduced into the common DCRM dental dimethacrylate composition of UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. The idea was to replace half of UDMA with one of the QAUDMAs, i.e., QA10+TMXDI and QA12+TMXDI. Correspondingly, the novel formulations comprised QAUDMA 20 wt.%, UDMA 20 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.% (Table 5).
Only a negligible effect of the N-alkyl substituent on the 20(QA10+TMXDI)m and 20(QA12+TMXDI)m properties was observed (Table 6). The extension of the N-alkyl substituents by two methylene groups resulted in a 1% decrease in the xDB, a 15% decrease in the η, a 0.05% increase in the RI, and a 2% increase in the dm. These differences were statistically insignificant, except for the dm, which was statistically significant. The most notable change occurred in the viscosity. Its decrease can suggest the weakening of intermolecular interactions in 20(QA12+TMXDI)m caused by the N-dodecyl substituent [53].
Compared to 40(UDMA)m, the experimental compositions had about a 10% higher MW and 10% lower xDB. This suggests that the replacement of half of UDMA with one of QAUDMAs would have a minor impact on a decrease in the theoretical crosslink density of the resulting copolymer [54].
The comparison of 40(UDMA)m and the experimental compositions by the viscosity, refractive index, and density also led to satisfactory conclusions. 40(UDMA)m had the η of 3.53 Pa∙s, which was in the range limited by the η of 20(QA10+TMXDI)m and 20(QA12+TMXDI)m, i.e., 3.79 and 3.22 Pa∙s, respectively. In addition, differences in the η between 40(UDMA)m and 20(QA10+TMXDI)m or 20(QA12+TMXDI)m were only 7 and 9%, respectively, and they did not have statistical significance. This showed that the experimental formulations would have flowability similar to the reference formulation, which would be beneficial to the DCRM preparation. The transparency of 20(QA10+TMXDI)m and 20(QA12+TMXDI)m also was suitable for dental materials. Their Ris were of 1.5101 and 1.5094, respectively. This matches the RI range of 1.46 to 1.55, which is recommended for dental materials [51,52]. In addition, the difference in the RI of the experimental compositions and reference composition was statistically insignificant. Finally, the dm of 20(QA12+TMXDI)m and 40(UDMA)m had the same value of 1.11 g/cm3. 20(QA10+TMXDI)m had a dm that was 2% higher, and that difference was statistically significant.
20(QA10+TMXDI)m and 20(QA12+TMXDI)m were subjected to photopolymerization towards two experimental copolymers 20(QA10+TMXDI)p and 20(QA12+TMXDI)p. They were characterized by the DC as the fundamental parameter used to characterize the polymer network structure (Table 7). An appropriately high DC is responsible for the adequate physical and mechanical properties of the DCRM matrices and the effective functioning of a dental filling [55,56]. The DC of novel QAUDMAs increased statistically significantly with the extension of the N-alkyl substituent. 20(QA10+TMXDI)p had a DC of 0.57 and 20(QA12+TMXDI)p had a DC of 0.69. Their DCs also were higher than that in 40(UDMA)p., which was 0.54. In addition, The DCs of 20(QA10+TMXDI)p and 20(QA12+TMXDI)p were higher than 0.55, which is recommended for clinical applications [57]. Thus, we can conclude that polymerization of 20(QA10+TMXDI)m and 20(QA12+TMXDI)m exhibited high efficiency and was not negatively influenced by a central benzene ring (which increases system stiffness [54]) and N-alkyl substituents (which increase the distance between polymerizing species [58,59]) present in their molecules.
A significant effect of the N-alkyl substituent on the density, volumetric contraction, glass temperature, water sorption, and leachability of 20(QA10+TMXDI)p and 20(QA12+TMXDI)p was not observed (Table 8 and Table 9). 20(QA10+TMXDI)p had a dp of 1.21 g/cm3, S of 7.04%, Tg of 53.46 °C, WS of 10.43 µg/mm3, and SL of 2.18 µg/mm3. 20(QA12+TMXDI)p had a 1% lower dp, and this change was statistically insignificant. The S and Tg statistically significantly increased, and these changes were 10% and 13%, respectively. The WS decreased statistically insignificantly by 1%, whereas the SL increased by 13% and this change was statistically significant. Although the decrease in the WS was insignificant, the chemical character of the copolymer surface can explain this change. 20(QA12+TMXDI)p was characterized by a WCA of 91.30° (Table 9), which corresponds to a hydrophobic surface [60,61]. 20(QA12+TMXDI)p had a statistically insignificantly lower WS and WCA. However, its WCA was 91.30°, which corresponds to a hydrophobic surface [60,61]. An effect of the N-alkyl substituent on the SL can be attributed to a higher MW of QA12+TMXDI compared to that of QA10+TMXDI. It is possible that a smaller number of unreacted monomer molecules leached out from 20(QA12+TMXDI)p because of a higher DC compared to 20(QA10+TMXDI)p. A similar effect was also observed in other work on the copolymer of Qan+TMDI [39].
The reference copolymer, 40(UDMA)p, had a 7.42% S, which was within the range limited by the S of 20(QA10+TMXDI)p and 20(QA12+TMXDI)p (Table 8). Compared to 40(UDMA)p, 20(QA10+TMXDI)m had a 5% lower S and 20(QA12+TMXDI)m had a 4% higher S, and both changes were statistically significant. The small scale of these differences suggests that the experimental and reference formulations had similar volumetric contractions caused by polymerization. Therefore, the potential DCRM matrices modified with QA10+TMXDI and QA12+TMXDI would probably have a marginal gap similar to common dental dimethacrylate copolymers [62].
The thermal behavior of 20(QA10+TMXDI)p and 20(QA12+TMXDI)p also was appropriate for their application in the potential DCRM matrix. On one hand, the experimental copolymers had a Tg of 53.46 and 60.30 °C, respectively. These Tgs were lower than the Tg of 40(UDMA)p (62.97 °C). The difference in Tgs between the reference copolymer and 20(QA10+TMXDI)p was 15% and was statistically significant, whereas that for 20(QA12+TMXDI)p was 4% and was statistically insignificant (Table 8). However, the Tgs in both cases were higher than 50 °C. This is a highly satisfactory result because it means that the experimental copolymers occur in a glassy state up to a temperature of at least 50 °C. Therefore, we can assume that the experimental copolymers would maintain constant mechanical performance over the entire temperature range in the oral cavity [63].
Compared to 40(UDMA)p, which had a WS of 5.31 μg/mm3 and SL of 0.17 μg/mm3, the experimental copolymers had WSs almost twice as high and SLs approximately fourteen times higher (Table 9). These were statistically significant differences, but the values of the WS and SL were noticeably lower than those defined by ISO 4049, i.e., 40 µg/mm3 and 7.5 µg/mm3, respectively [46]. A higher WS (10.43 μg/mm3), which was determined for 20(QA10+TMXDI)p, was 26% of the maximum recommended by the ISO standard, and a higher SL, which was determined for 20(QA12+TMXDI)p (2.46 μg/mm3), was 33% of the maximum recommended by the ISO 4049 standard. This is a highly satisfactory outcome, as the hydrophilic QA groups usually cause an increase in a polymer’s WS [39]. The increased WS and SL of the experimental copolymers showed that the QA groups promote interactions with water molecules, even though the N-alkyl substituent decreased the hydrophilicity of the copolymer surface. This was confirmed by the result achieved for 40(UDMA)p, which was characterized by the lowest WCA (74.07°, Table 9), and thus its surface was the most hydrophilic [60].
The mechanical properties of 20(QA10+TMXDI)p and 20(QA12+TMXDI) p decreased as the N-alkyl substituent length increased. The E, FS, and HB for 20(QA10+TMXDI)p were 3244.9 Mpa, 100.6 Mpa, and 234.8 Mpa, respectively. They decreased by 7%, 10%, and 17% for 20(QA12+TMXDI)p (Table 10). Only the result for the HB was statistically significant. This led to the conclusion that hardness was the most sensitive to the N-alkyl chain extension, as its decline was the most pronounced.
40(UDMA)p exhibited the E, FS, and HB of 3597.7 Mpa, 125.4 Mpa, and 225.9 Mpa, respectively. Its modification by introducing QAUDMA resulted in statistically significant decreases of E, by 10 and 16%, and FS, by 20 and 28%, respectively, for 20(QA10+TMXDI)p and 20(QA12+TMXDI)p. The HB of 20(QA10+TMXDI)p increased by 4% and this difference was statistically significant. As the length of the N-alkyl substituent increased to twelve carbon atoms, the HB decreased statistically significantly by 14%. However, the overall mechanical performance of 20(QA10+TMXDI)p as well as 20(QA12+TMXDI)p was fully satisfactory, as the tested parameters were high.
Finally, the experimental copolymers revealed antibacterial activity against S. aureus and E. coli (Table 11). 20(QA10+TMXDI)p had a lower MIC and MBC than 20(QA12+TMXDI)p. This means that the extension of the N-alkyl substituent in the QAUDMA with two methylene groups decreased the copolymer’s antibacterial activity. If compared to 40(UDMA)p, both copolymers had higher antibacterial activity. In addition, intra-species differences were observed for 20(QA12+TMXDI)p. If compared to 40(UDMA)p, 20(QA12+TMXDI)p showed the same antibacterial activity concerning S. aureus, and greater antibacterial activity concerning E.coli. 20(QA12+TMXDI)p had a MBC 50% lower and MIC 75% lower than 40(UDMA)p in tests against E.coli. 20(QA10+TMXDI)p showed more significant antibacterial activity and differentiation according to bacterial strain was no longer observed. 20(QA10+TMXDI)p had a 75% lower MBC and 87.5% lower MIC than 40(UDMA)p in tests against both bacteria strains.
We can additionally characterize the experimental copolymers using a common classification of antimicrobials as bactericidal or bacteriostatic. This classification assumes that if the MBC to MIC ratio is lower than four, a biocide can be recognized bactericidal, and if the MBC to MIC ratio is higher than six, a biocide can be recognized bacteriostatic [64]. As this ratio for 20(QA10+TMXDI)p was two, one can assume this copolymer is bactericidal and it may be possible to administer the 20(QA10+TMXDI)p dosages to kill 99.9% of the bacteria.

5. Conclusions

The two novel quaternary ammonium urethane-dimethacrylate (QAUDMA) monomers were successfully synthesized from 1,3-bis(1-isocyanato-1-methylethyl)benzene TMXDI) and 2-(methacryloyloxy)ethyl-2-alkylhydroxyethylmethylammonium bromides with the decyl and dodecyl N-alkyl substituents, respectively, QA10+TMXDI and QA12+TMXDI. The TMXDI diisocyanate provided an aromatic central ring in the QAUDMA structure.
The achieved QAUDMAs were used as comonomers 20 wt.% in the copolymerization with common dental dimethacrylates, UDMA 20 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.%. For comparison, the UDMA 40 wt.%, Bis-GMA 40 wt.%, and TEGDMA 20 wt.% representative dental copolymer was also prepared.
The monomer compositions had adequate flowability and transparency required for dental materials.
A comparative analysis of the tested copolymers’ properties showed that their polymerization shrinkage, glass transition temperature, and mechanical properties, were satisfactory and were usually similar to the properties of the reference copolymer. The water sorption and water solubility mostly deteriorated; however, they were satisfactory. Their values were far below the limits pointed at the international standard ISO 4049.
The antibacterial properties of the copolymers were characterized against S. aureus (ATCC 25923) and E. coli (ATCC 25922). The copolymer comprising QA10+TMXDI showed a significantly higher antibacterial effect than both the copolymer comprising QA12+TMXDI and the reference copolymer, as it had a lower minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC).
This study demonstrated that the introduction of a benzene ring into the QAUDMA structure resulted in copolymers with very good physicochemical and mechanical characteristics and antibacterial activity. However, it was observed that the shorter the N-alkyl chain, the higher the antibacterial activity. Therefore, it is worth conducting further research with monomers that have shorter N-alkyl substituents. A more in-depth investigation of antibacterial properties should also be carried out for the tested copolymers, as the aim of this work was only to determine whether they have antibacterial activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17020298/s1, Figure S1: DSC thermograms of copolymers: (a) 20(QA10+TMXDI)p; (b) 20(QA12+TMXDI) p; (c) 40(UDMA) p, Figure S2: The goniometry camera images of deionized water droplets on the surfaces of copolymers: (a) 20(QA10+TMXDI)p; (b) 20(QA12+TMXDI) p; (c) 40(UDMA) p, Figure S3: The results of antibacterial activity tests of 20(QA10+TMXDI)p against S. aureus (ATCC 25923), Figure S4: The results of antibacterial activity tests of 20(QA12+TMXDI)p against S. aureus (ATCC 25923), Figure S5: The results of antibacterial activity tests of 40(UDMA)p against S. aureus (ATCC 25923), Figure S6: The results of antibacterial activity tests of 20(QA10+TMXDI)p against E. coli (ATCC 25922), Figure S7: The results of antibacterial activity tests of 20(QA12+TMXDI)p against E. coli (ATCC 25922), Figure S8: The results of antibacterial activity tests of 40(UDMA)p against E. coli (ATCC 25922).

Author Contributions

Conceptualization, P.D. and I.B.-R.; Methodology, P.D. and I.B.-R.; Validation, P.D.; Formal analysis, I.B.-R.; Investigation, P.D. and M.C.-P.; Resources, P.D. and I.B.-R.; Data curation, P.D., A.K.-K. and G.C.; Writing—original draft, P.D. and I.B.-R.; Writing—review & editing, I.B.-R.; Visualization, P.D.; Supervision, A.K.-K., G.C. and I.B.-R.; Project administration, I.B.-R.; Funding acquisition, P.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Polish Budget Funds for Scientific Research in 2023 as core funding for research and development activities at the Silesian University of Technology – funding for young scientists grant number 04/040/BKM23/0258.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and supplementary materials.

Conflicts of Interest

The authors declare no conflicts 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. WHO. Global Oral Health Status Report: Towards Universal Health Coverage for Oral Health by 2030; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  2. Wen, P.; Chen, M.; Zhong, Y.; Dong, Q.; Wong, H. Global Burden and Inequality of Dental Caries, 1990 to 2019. J. Dent. Res. 2022, 101, 392–399. [Google Scholar] [CrossRef]
  3. Chen, L.; Suh, B.; Yang, J. Antibacterial dental restorative materials: A review. Am. J. Dent. 2018, 31, 6B–12B. [Google Scholar] [PubMed]
  4. Beyth, N.; Farah, S.; Domb, A.; Weiss, E. Antibacterial dental resin composites. Reac. Funct. Polym. 2014, 75, 81–88. [Google Scholar] [CrossRef]
  5. Ferrando-Magraner, E.; Bellot-Arcís, C.; Paredes-Gallardo, V.; Almerich-Silla, J.M.; García-Sanz, V.; Fernández-Alonso, M.; Montiel-Company, J.M. Antibacterial Properties of Nanoparticles in Dental Restorative Materials. A Systematic Review and Meta-Analysis. Medicina 2020, 56, 55. [Google Scholar] [CrossRef] [PubMed]
  6. Domb, A.J.; Weiss, E.; Beyth, N.; Farber, I.; Perez-Davidi, M. Antimicrobial Nanoparticulate Additives Forming Non-Leachable Sustained Antimicrobial Polymeric Compositions. U.S. Patent 20080226728A1, 18 September 2008. [Google Scholar]
  7. Fernandes, J.; Menezes, V.; Albuquerque, A.; Oliveira, M.; Meira, K.; Menezes, R., Jr.; Sampaio, F.C. Improving Antimicrobial Activity of Dental Restorative Materials. In Emerging Trends in Oral Health Sciences and Dentistry; InTech: London, UK, 2015. [Google Scholar]
  8. Badr, S.; Abdulrahman, A.; Abdullah, A.; Essam, A.; Mohammad, Y.; Shahzeb, H. Effect of various antibacterial materials in dental composites: A systematic review. Ann. Dent. Spec. 2021, 9, 39–44. [Google Scholar]
  9. Mehdawi, I.; Young, A. Antibacterial composite restorative materials for dental applications. In Non-Metallic Biomaterials for Tooth Repair and Replacement; Elsevier: Amsterdam, The Netherlands, 2013; pp. 270–293. [Google Scholar]
  10. Zhou, X.; Huang, X.; Li, M.; Peng, X.; Wang, S.; Zhou, X.; Cheng, L. Development and status of resin composite as dental restorative materials. J. Appl. Polym. Sci. 2019, 136, 48180. [Google Scholar] [CrossRef]
  11. Chan, D.; Hu, W.; Chung, K.; Larsen, R.; Jensen, S.; Cao, D.; Gaviria, L.; Ong, J.; Whang, K.; Eiampongpaiboon, T. Reactions: Antibacterial and bioactive dental restorative materials: Do they really work? Am. J. Dent. 2018, 15, 32B–36B. [Google Scholar]
  12. Farrugia, C.; Camilleri, J. Antimicrobial properties of conventional restorative filling materials and advances in antimicrobial properties of composite resins and glass ionomer cements—A literature review. Dent. Mater. 2015, 31, e89–e99. [Google Scholar] [CrossRef]
  13. Ibrahim, M.; Garcia, I.; Kensara, A.; Balhaddad, A.; Collares, F.; Williams, M.; Ibrahim, A.; Lin, N.; Weir, M.; Xu, H.; et al. How we are assessing the developing antibacterial resin-based dental materials? A scoping review. J. Dent. 2020, 99, 103369. [Google Scholar] [CrossRef]
  14. Chen, Z.; Chu, Z.; Jiang, Y.; Xu, L.; Qian, H.; Wang, Y.; Wang, W. Recent advances on nanomaterials for antibacterial treatment of oral diseases. Mater. Today Bio. 2023, 20, 100635. [Google Scholar] [CrossRef]
  15. 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] [PubMed]
  16. Sun, Q.; Zhang, L.; Bai, R.; Zhuang, Z.; Zhang, Y.; Yu, T.; Peng, L.; Xin, T.; Chen, S.; Han, B. Recent Progress in Antimicrobial Strategies for Resin-Based Restoratives. Polymers 2021, 13, 1590. [Google Scholar] [CrossRef] [PubMed]
  17. Ramburrun, P.; Pringle, N.; Dube, A.; Adam, R.; 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]
  18. 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]
  19. Ge, Y.; Wang, S.; Zhou, X.; Wang, H.; Xu, H.H.K.; Cheng, L. The Use of Quaternary Ammonium to Combat Dental Caries. Materials 2015, 8, 3532–3549. [Google Scholar] [CrossRef] [PubMed]
  20. Imazato, S.; Chen, J.H.; Ma, S.; Izutani, N.; Li, F. Antibacterial Resin Monomers Based on Quaternary Ammonium and Their Benefits in Restorative Dentistry. Jap. Dent. Sci. Rev. 2012, 48, 115–125. [Google Scholar] [CrossRef]
  21. Featherstone, J. Dental restorative materials containing quaternary ammonium compounds have sustained antibacterial action. J. Am. Dent. Assoc. 2022, 153, 1114–1120. [Google Scholar] [CrossRef]
  22. Li, Y.; Li, B.; Guo, X.; Wang, H.; Cheng, L. Applications of quaternary ammonium compounds in the prevention and treatment of oral diseases: State-of-the-art and future directions. J. Dent. 2023, 137, 104678. [Google Scholar] [CrossRef]
  23. Jiao, Y.; Niua, L.; Maa, S.; Li, J.; Tayd, F.; Chena, J. Quaternary ammonium-based biomedical materials: State-of-the-art, toxicological aspects and antimicrobial resistance. Prog. Polym. Sci. 2017, 71, 53–90. [Google Scholar] [CrossRef]
  24. Mousavinasab, S. Biocompatibility of composite resins. Dent. Res. J. 2011, 8, S21–S29. [Google Scholar]
  25. Santos, M.R.E.; Fonseca, A.C.; Mendonça, P.V.; Branco, R.; Serra, A.C.; Morais, P.V.; Coelho, J.F.J. Recent Developments in Antimicrobial Polymers: A Review. Materials 2016, 9, 599. [Google Scholar] [CrossRef] [PubMed]
  26. Capita, R.; Alonso-Calleja, C. Antibiotic-resistant bacteria: A challenge for the food industry. Crit. Rev. Food Sci. Nutr. 2013, 53, 11–48. [Google Scholar] [CrossRef] [PubMed]
  27. Chinemerem, D.; Ugwu, M.; Oliseloke, A.; Al-Ouqaili, M.; Chinedu, I.; Victor, C.; Saki, M. Antibiotic resistance: The challenges and some emerging strategies for tackling a global menace. J. Clin. Lab. Anal. 2022, 36, e24655. [Google Scholar] [CrossRef]
  28. Saha, M.; Sarkar, A. Review on Multiple Facets of Drug Resistance: A Rising Challenge in the 21st Century. J. Xenobiot. 2021, 11, 197–214. [Google Scholar] [CrossRef] [PubMed]
  29. Huang, L.; Xiao, Y.H.; Xing, X.D.; Li, F.; Ma, S.; Qi, L.L.; Chen, J.H. Antibacterial Activity and Cytotoxicity of Two Novel Cross-Linking Antibacterial Monomers on Oral Pathogens. Arch. Oral Biol. 2011, 56, 367–373. [Google Scholar] [CrossRef] [PubMed]
  30. 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]
  31. 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]
  32. 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]
  33. 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]
  34. 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]
  35. 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]
  36. 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]
  37. 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]
  38. Chrószcz, M.W.; Barszczewska-Rybarek, I.M.; Kazek-Kęsik, 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]
  39. 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]
  40. Chrószcz-Porębska, M.; Kazek-Kęsik, A.; Chladek, G.; Barszczewska-Rybarek, I. Novel mechanically strong and antibacterial dimethacrylate copolymers based on quaternary ammonium urethane-dimethacrylate analogues. Dent. Mater. 2023, 39, 659–664. [Google Scholar] [CrossRef]
  41. 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. [Google Scholar] [CrossRef]
  42. 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]
  43. ISO 4049:2019; Plastics—Resins in the Liquid State or As Emulsions or Dispersions—Determination of Apparent Viscosity Using A Single Cylinder Type Rotational Viscometer Method. International Organization for Standardization: Geneva, Switzerland, 2019.
  44. ISO 489:2022; Plastics—Determination of Refractive Index. International Organization for Standardization: Geneva, Switzerland, 2022.
  45. ISO 11357-2:2020; Plastics—Differential Scanning Calorimetry (DSC)—Part 2: Determination of Glass Transition Temperature and Step Height. International Organization for Standardization: Geneva, Switzerland, 2020.
  46. ISO 4049:2019; Dentistry—Polymer-Based Restorative Materials. International Organization for Standardization: Geneva, Switzerland, 2019.
  47. ISO 2039-1:2001; Plastics—Determination of Hardness—Part 1: Ball Indentation Method. International Organization for Standardization: Geneva, Switzerland, 2001.
  48. ISO 178:2019; Plastics—Determination of Flexural Properties. International Organization for Standardization: Geneva, Switzerland, 2019.
  49. Barszczewska-Rybarek, I. Quantitative determination of degree of conversion in photocured poly(urethane-dimethacrylate)s by FTIR spectroscopy. J. Appl. Polym. Sci. 2012, 123, 1604–1611. [Google Scholar] [CrossRef]
  50. Six, C.; Richter, F. Isocyanates, Organic. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2003; Volume 20, pp. 63–82. [Google Scholar]
  51. Manappallil, J.J. Basic Dental Materials, 4th ed.; Jaypee Brothers Medical Publishers: New Delhi, India, 2015. [Google Scholar]
  52. Oivanen, M.; Keulemans, F.; Garoushi, S.; Vallittu, P.; Lassila, L. The effect of refractive index of fillers and polymer matrix on translucency and color matching of dental resin composite. Biomater. Investig. Dent. 2021, 8, 48–53. [Google Scholar] [CrossRef]
  53. Gonçalves, F.; Kawano, Y.; Pfeifer, C.; Stansbury, J.W.; Braga, R.R. Influence of BisGMA, TEGDMA, and BisEMA contents on viscosity, conversion, and flexural strength of experimental resins and composites. Eur. J. Oral Sci. 2009, 117, 442–446. [Google Scholar] [CrossRef]
  54. Stansbury, J.W. Dimethacrylate network formation and polymer property evolution as determined by the selection of monomers and curing conditions. Dent. Mater. 2012, 28, 13–22. [Google Scholar] [CrossRef]
  55. Du, M.; Zheng, Y. Degree of conversion and mechanical properties studies of UDMA based materials for producing dental posts. Polym. Compos. 2008, 29, 623–630. [Google Scholar] [CrossRef]
  56. Moldovan, M.; Balazsi, R.; Soanca, A.; Roman, A.; Sarosi, C.; Prodan, D.; Vlassa, M.; Cojocaru, I.; Saceleanu, V.; Cristescu, I. Evaluation of the Degree of Conversion, Residual Monomers and Mechanical Properties of Some Light-Cured Dental Resin Composites. Materials 2019, 12, 2109. [Google Scholar] [CrossRef]
  57. Alshali, R.Z.; Silikas, N.; Satterthwaite, J.D. Degree of conversion of bulk-fill compared to conventional resin-composites at two time intervals. Dent. Mater. 2013, 29, e213–e217. [Google Scholar] [CrossRef]
  58. Nuzhdina, A.V.; Morozov, A.S.; Kopitsyna, M.N.; Strukova, E.N.; Shlykova, D.S.; Bessonov, I.V.; Lobakova, E.S. Simple and Versatile Method for Creation of Non-Leaching Antimicrobial Surfaces Based on Cross-Linked Alkylated Polyethyleneimine Derivatives. Mater. Sci. Eng. C 2017, 70, 788–795. [Google Scholar] [CrossRef]
  59. Yudovin-Farber, I.; Beyth, N.; Weiss, E.I.; Domb, A.J. Antibacterial Effect of Composite Resins Containing Quaternary Ammonium Polyethyleneimine Nanoparticles. J. Nanopart Res. 2010, 12, 591–603. [Google Scholar] [CrossRef]
  60. Law, K.-Y. Definitions for Hydrophilicity, Hydrophobicity, and Superhydrophobicity: Getting the Basics Right. J. Phys. Chem. Lett. 2014, 5, 686–688. [Google Scholar] [CrossRef]
  61. Grainger, D.; Castner, D. 3.301—Surface Analysis and Biointerfaces: Vacuum and Ambient in Situ Techniques. In Comprehensive Biomaterials; Ducheyne, P., Ed.; Elsevier: Amsterdam, The Netherlands, 2011; pp. 1–22. [Google Scholar]
  62. Susila, A.V. Role of Composition on Polymerization Shrinkage and Shrinkage Stress in Dental Composites. J. Oper. Dent. Endod. 2021, 6, 31–44. [Google Scholar]
  63. Moraes, J.C.S.; Sostena, M.M.D.S.; Roberto, C. The Glass Transition Temperature in Dental Composites. In Metal, Ceramic and Polymeric Composites for Various Uses; Chapter 33; John Cuppoletti, J., Ed.; IntechOpen: London, UK, 2011. [Google Scholar]
  64. Davis, J.L. Pharmacologic Principles. In Equine Internal Medicine, 4th ed.; Reed, S.M., Bayly, W.M., Sellon, D.C., Eds.; Elsevier Inc.: Philadelphia, PA, USA, 2018; pp. 79–137. [Google Scholar]
Figure 1. The synthesis route of QAUDMA monomers, namely QA10+TMXDI and QA12+TMXDI.
Figure 1. The synthesis route of QAUDMA monomers, namely QA10+TMXDI and QA12+TMXDI.
Materials 17 00298 g001
Figure 2. 1H NMR spectra of monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Figure 2. 1H NMR spectra of monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Materials 17 00298 g002
Figure 3. 13C NMR spectra of monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Figure 3. 13C NMR spectra of monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Materials 17 00298 g003
Figure 4. FTIR spectra of the achieved QAUDMA monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Figure 4. FTIR spectra of the achieved QAUDMA monomers: (a) QA10+TMXDI; (b) QA12+TMXDI.
Materials 17 00298 g004
Figure 5. FTIR spectra showing the disappearance of the absorption band of the C=C stretching vibrations at 1636 cm−1 due to polymerization, and the absorption band of the aromatic stretching skeletal vibrations at 1608 cm−1, which was used as an internal standard: (a) 20(QA10+TMXDI)m; (b) 20(QA12+TMXDI)m; (c) 20(QA10+TMXDI)p; and (d) 20(QA12+TMXDI)p.
Figure 5. FTIR spectra showing the disappearance of the absorption band of the C=C stretching vibrations at 1636 cm−1 due to polymerization, and the absorption band of the aromatic stretching skeletal vibrations at 1608 cm−1, which was used as an internal standard: (a) 20(QA10+TMXDI)m; (b) 20(QA12+TMXDI)m; (c) 20(QA10+TMXDI)p; and (d) 20(QA12+TMXDI)p.
Materials 17 00298 g005
Table 1. 1H NMR signals of QA10+TMXDI and QA12+TMXDI monomers.
Table 1. 1H NMR signals of QA10+TMXDI and QA12+TMXDI monomers.
Signal SymbolHydrogen AtomMultiplicityNumber of ProtonsChemical Shift [ppm]
aCH3-C=s61.95
b=CH22 m45.70 and 6.14
c-O-CH2-CH2-N+-m44.33–4.91
d-O-CH2-CH2-N+-m44.00–4.26
e-N+-CH3s63.30–3.77
f-O-CH2-CH2-N+-m44.00–4.26
g-O-CH2-CH2-N+-m44.33–4.91
h-NH-C=Om27.10–7.80
i-N+-CH2-CH2-(CH2)7(or 9)-CH3 1m43.30–3.77
j-N+-CH2-CH2-(CH2)7(or 9)-CH3 1m41.49–1.88
k-N+-CH2-CH2-(CH2)7(or 9)-CH3 1m28/36 21.16–1.47
l-N+-CH2-CH2-(CH2)(7 or 9)-CH3 1m60.82–1.00
mCH3-C-CH3m121.49–1.88
n-CH- (Ar)m17.10–7.80
o-CH- (Ar)m27.10–7.80
p-CH- (Ar)m17.10–7.80
1 -N+-CH2-CH2-(CH2)7-CH3 corresponds to QA10+TMXDI, and -N+-CH2-CH2-(CH2)9-CH3 corresponds to QA12+TMXDI. 2 28 corresponds to QA10+TMXDI, and 36 corresponds to QA12+TMXDI.
Table 2. 13C NMR signals of QA10+TMXDI and QA12+TMXDI monomers.
Table 2. 13C NMR signals of QA10+TMXDI and QA12+TMXDI monomers.
Signal SymbolCarbon AtomChemical Shift [ppm]
aCH3-C=21
b=CH2130
cCH3-C=CH2138
d-O-C=O169
e-h-CH2-59–67
i-N+-CH353
j-N+-CH2-(CH2)8(or 10)-CH3 159–67
k-N+-CH2-(CH2)8(or 10)-CH3 126–35
l-N+-CH2-(CH2)8 (or 10)-CH3 117
m-NH-C=O157
n-NH-C-59–67
o>C-CH326–35
p-s-CH= (Ar)124–126
t>C= (Ar)
1 -N+-CH2-(CH2)8-CH3 corresponds to QA10+TMXDI, and -N+-CH2-(CH2)10-CH3 corresponds to QA12+TMXDI.
Table 3. The interpretation of FTIR spectra of QA10+TMXDI and QA12+TMXDI monomers.
Table 3. The interpretation of FTIR spectra of QA10+TMXDI and QA12+TMXDI monomers.
Chemical BondIntensity 1Wavenumber [cm−1]
N-Hw3215
CH3w3024
=CH2m2957
CH2, CH3m2925 and 2855
N=C=O 2s2257
C=Os1714
C=Cw1638
C=C (Ar)w1605
NH (urethane)m1529
CH2, CH3m1457
C-Nm1249 and 1158
C-O-Cm1092 and 1087
C-N+m943
1. The intensity of the absorption bands in the FTIR spectra was referred to as strong (s), medium (m), and weak (w). 2. Refers to the isocyanate group absorption band present in the FTIR spectra of the QAUDMA monomers before their thermal treatment.
Table 4. Properties of QA10+TMXDI and QA12+TMXDI related to properties of Bis-GMA, UDMA, and TEGDMA. The lowercase letter “a” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 4. Properties of QA10+TMXDI and QA12+TMXDI related to properties of Bis-GMA, UDMA, and TEGDMA. The lowercase letter “a” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Monomer NameMW (g/mol)xDB (mol/kg)RIdm (g/cm3)
AV 1AV 2
Experimental monomers
QA10+TMXDI10611.891.52301.16 a
QA12+TMXDI11171.791.51381.33 a
Reference monomers
Bis-GMA5123.901.5493 31.15 3
UDMA4704.251.4614 31.09 3
TEGDMA2866.991.4852 31.07 3
1 The SD for the RI values was 0.0001. 2 The SD for the dm values was 0.01. 3 Taken from [37].
Table 5. Compositions of QAUDMA, Bis-GMA, UDMA, and TEGDMA experimental mixtures related to the Bis-GMA, UDMA, and TEGDMA reference mixture.
Table 5. Compositions of QAUDMA, Bis-GMA, UDMA, and TEGDMA experimental mixtures related to the Bis-GMA, UDMA, and TEGDMA reference mixture.
Monomer Composition Sample Composition
QAUDMAUDMABis-GMATEGDMA
Mass
Fraction
Mole
Fraction
Mass
Fraction
Mole
Fraction
Mass
Fraction
Mole
Fraction
Mass
Fraction
Mole
Fraction
Experimental compositions
20(QA10+TMXDI)m0.200.090.200.200.400.370.200.34
20(QA12+TMXDI)m0.200.090.200.200.400.370.200.34
Reference composition
40(UDMA)m0.000.000.400.360.400.340.200.30
Table 6. Properties of 20(QA10+TMXDI)m and 20(QA12+TMXDI)m experimental monomer compositions related to properties of the 40(UDMA)m reference monomer composition. The lowercase letters “a” and “b” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 6. Properties of 20(QA10+TMXDI)m and 20(QA12+TMXDI)m experimental monomer compositions related to properties of the 40(UDMA)m reference monomer composition. The lowercase letters “a” and “b” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Monomer CompositionMW (g/mol)xDB (mol/kg)η (Pa∙s)RIdm (g/cm3)
AVSDAV 1AV 2
Experimental compositions
20(QA10+TMXDI)m4764.203.790.401.51011.13 a,b
20(QA12+TMXDI)m4814.163.220.301.50941.11 a
Reference composition
40(UDMA)m4294.663.530.401.50491.11 b
1 The SD for the RI values always was 0.0001. 2 The SD for the dm values was always 0.01.
Table 7. The DC in copolymers. The lowercase letters “a” and “b” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 7. The DC in copolymers. The lowercase letters “a” and “b” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
CopolymerDC
AVSD
Experimental copolymers
20(QA10+TMXDI)p0.57 a0.05
20(QA12+TMXDI)p0.69 a,b0.06
Reference copolymer
40(UDMA)p0.54 b0.04
Table 8. The physical properties of copolymers: density (dp), polymerization shrinkage (S), and glass transition temperature (Tg). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 8. The physical properties of copolymers: density (dp), polymerization shrinkage (S), and glass transition temperature (Tg). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Copolymerdp (g/cm3)S (%)Tg (°C)
AVSDAVSDAVSD
Experimental copolymers
20(QA10+TMXDI)p1.210.017.04 a,b0.8353.46 a,b3.27
20(QA12+TMXDI)p1.200.017.72 a,c1.1760.30 a4.34
Reference copolymer
40(UDMA)p1.200.017.42 b,c1.0862.97 b4.52
Table 9. The physical properties characterizing copolymers’ interaction with water: water sorption (WS), water leachability (SL), and water contact angle (WCA). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 9. The physical properties characterizing copolymers’ interaction with water: water sorption (WS), water leachability (SL), and water contact angle (WCA). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
CopolymerWS (μg/mm3)SL (μg/mm3)WCA (◦)
AVSDAVSDAVSD
Experimental copolymers
20(QA10+TMXDI)p10.43 a0.422.18 a,b0.0687.03 a 3.07
20(QA12+TMXDI)p10.35 b0.232.46 a,c0.2591.30 b 4.82
Reference copolymer
40(UDMA)p5.31 a,b0.360.17 b,c0.0974.07 a,b 2.24
Table 10. The mechanical properties of copolymers: hardness (HB), flexural modulus (E), and flexural strength (FS). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
Table 10. The mechanical properties of copolymers: hardness (HB), flexural modulus (E), and flexural strength (FS). The lowercase letters “a”–“c” in the superscripts indicate couples with statistically significant results (p ≤ 0.05). The remaining results did not have statistical significance (p > 0.05).
CopolymerHB (MPa)FS (MPa)E (MPa)
AVSDAVSDAVSD
Experimental copolymers
20(QA10+TMXDI)p234.8 a,b8.7100.6 a11.73244.9 a114.1
20(QA12+TMXDI)p194.6 a,c26.590.3 b6.13020.6 b198.9
Reference copolymer
40(UDMA)p225.9 b,c19.4125.4 a,b13.03597.7 a,b154.4
Table 11. The antibacterial properties of copolymers against S. aureus (ATCC 25923) and E. coli (ATCC 25922): minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC).
Table 11. The antibacterial properties of copolymers against S. aureus (ATCC 25923) and E. coli (ATCC 25922): minimum bactericidal concentration (MBC) and minimum inhibitory concentration (MIC).
Copolymer MBC [mg/mL]MIC [mg/mL]
S. aureus
(ATCC 25923)
E. coli
(ATCC 25922)
S. aureus
(ATCC 25923)
E. coli
(ATCC 25922)
Experimental copolymers
20(QA10+TMXDI)p12.512.56.256.25
20(QA12+TMXDI)p>5025>5012.5
Reference copolymer
40(UDMA)p>50>50>50>50
Control
Sterile Water---
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

Drejka, P.; Chrószcz-Porębska, M.; Kazek-Kęsik, A.; Chladek, G.; Barszczewska-Rybarek, I. Chemical Modification of Dental Dimethacrylate Copolymer with Tetramethylxylylene Diisocyanate-Based Quaternary Ammonium Urethane-Dimethacrylates—Physicochemical, Mechanical, and Antibacterial Properties. Materials 2024, 17, 298. https://doi.org/10.3390/ma17020298

AMA Style

Drejka P, Chrószcz-Porębska M, Kazek-Kęsik A, Chladek G, Barszczewska-Rybarek I. Chemical Modification of Dental Dimethacrylate Copolymer with Tetramethylxylylene Diisocyanate-Based Quaternary Ammonium Urethane-Dimethacrylates—Physicochemical, Mechanical, and Antibacterial Properties. Materials. 2024; 17(2):298. https://doi.org/10.3390/ma17020298

Chicago/Turabian Style

Drejka, Patryk, Marta Chrószcz-Porębska, Alicja Kazek-Kęsik, Grzegorz Chladek, and Izabela Barszczewska-Rybarek. 2024. "Chemical Modification of Dental Dimethacrylate Copolymer with Tetramethylxylylene Diisocyanate-Based Quaternary Ammonium Urethane-Dimethacrylates—Physicochemical, Mechanical, and Antibacterial Properties" Materials 17, no. 2: 298. https://doi.org/10.3390/ma17020298

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