Next Article in Journal
Application of Extension Sliding Mode Strategy to Maximum Power Point Tracking in Human Power Generation Systems
Previous Article in Journal
A Laboratory Test Expert System for Clinical Diagnosis Support in Primary Health Care
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Vinylcarbazole: As an Additive for Thermal Polymerization at Room Temperature with in situ Formation of Ag(0) Nanoparticules

Carrefour of Innovative Technology and Ecodesign (CITE), Faculty of Engineering, University of Sherbrooke, 2500 boul. de l'Université, Sherbrooke, QC J1K 2R1, Canada
*
Author to whom correspondence should be addressed.
Appl. Sci. 2015, 5(3), 241-258; https://doi.org/10.3390/app5030241
Submission received: 3 July 2015 / Revised: 20 July 2015 / Accepted: 28 July 2015 / Published: 5 August 2015
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
N-vinylcarbazole (NVK) is proposed as an additive for acrylates thermal free radical polymerization (FRP) and epoxy thermal ring opening polymerization (ROP) at room temperature. The new initiating systems are based on a silane/silver salt/N-vinylcarbazole interaction, which ensures good to excellent polymerization. Moreover, the polymerization is much more efficient under air than under argon. The effects of the N-vinylcarbazole, silane, silver salt and monomer structures are investigated. Interestingly, silver nanoparticles Ag(0) are formed in situ. The as-synthesized nanocomposite materials contained spherical nanoparticles homogeneously dispersed in the polymer matrices. Polymers and nanoparticles were characterized by Differential Scanning Calorimetry (DSC), Transmission Electron Microscopy (TEM), Energy-Dispersive X-ray Spectrometry (EDS), Fourier Transform Infrared Spectroscopy (FTIR), and UV-vis spectroscopy. A coherent picture of the involved chemical mechanisms is presented.

Graphical Abstract

1. Introduction

Nanocomposite materials containing noble metal and dispersed nanoparticles in a polymer matrix may exhibit peculiar physical and chemical properties that are of high scientific and technological importance [1]. These new composites are used for optical [2], electrical [3] and medical [4] applications as well as data storage [5]. In particular, silver nanoparticles are very important for their excellent electrical conductivity [6], anti-microbial effect [7] and optical properties [8]. In recent years, a whole bunch of synthetic methods for the preparation of metal nanoparticules have been developed: chemical, photochemical and thermal.
Metal-polymer nanocomposites are usually obtained via multistep methods. Dry silver nanoparticules (NPs) produced beforehand can be dispersed into a polymerizable formulation to obtain self-assembly functionalized structures. However, besides the specific hazards related to handling dry NPs, their size dispersity over a large scale is difficult to control, thus limiting the interest of this “ex situ” method [9,10].
The “in situ” approach that involves the generation of metal nanoparticules directly in a polymerizable medium through reduction of cationic precursors offers the advantages of better dispersion ability and facile chemical or photochemical reduction [11,12].
Thermally and photochemically induced free radical polymerizations (FRP) have been the subject of numerous studies [13,14,15,16,17,18]. In thermal FRPs, redox initiated processes are very attractive due e.g., to the absence of solvents and the use of quite low reaction temperatures. However, such FRPs suffer from a strong drawback concerned with the well-known oxygen inhibition [13,14,15,16,17,18,19,20]. Indeed, both the initiating (Scheme 1(r1)) and propagating (Scheme 1(r2)) radicals are scavenged by O2 and converted into stable peroxyl radicals, which cannot participate in any further polymerization initiation reactions and lead to a decrease of both polymerization rates and final conversions (Scheme 1) [21,22,23,24].
Scheme 1. Different polymerization reactions under oxygen (r1: initiating step and r2: propagation step).
Scheme 1. Different polymerization reactions under oxygen (r1: initiating step and r2: propagation step).
Applsci 05 00241 g010
In cationic polymerization (CP), such an inhibition does not exist and, interestingly, the use of less toxic monomers than the acrylates can be highly worthwhile. Redox systems being able to initiate a cationic polymerization have been already proposed [25,26,27,28]. Very recently, new bi-component combinations based on onium salt (diaryliodonium or triarylsulfonium salt) and organosilane (R3SiH) as the reducing agent were elegantly proposed: the redox reaction is catalyzed by noble metal complexes (platinum, palladium, rhodium) and leads to a silylium cation, which directly attacks the epoxide or/and reacts with water to form a Bronstedt acid [29].
The thermal FRP and/or CP at room temperature and under air remain an exciting research topic. This approach consists of producing a radical R (from an organosilane), to initiate FRP, that can be easily oxidized by a silver salt (e.g., AgSbF6) leading to R+, which corresponds to the ROP initiating structure (Scheme 2) [30,31,32].
Scheme 2. Schematic representation of the polymerization mechanisms.
Scheme 2. Schematic representation of the polymerization mechanisms.
Applsci 05 00241 g011
However, the nature of the primary radicals produced from the organosilane is crucial for getting a high polymerization efficiency, as they should be characterized by excellent oxidation properties; moreover, the resulting radicals and cations must be efficient structures for the FRP and CP reactions, respectively [33,34,35,36,37]. Even if a lot of systems have been developed in the last years [33,34,35,36,37,38,39,40,41,42], the search for novel additives being able to improve the radical and/or cationic polymerization processes still appears as a challenge.
In the present paper, N-vinylcarbazole (NVK) is used as an additive for thermal free and/or cationic polymerization of an acrylate and an epoxide at room temperature and under air (TMPTA, St and EPOX—see below, Scheme 3). This proposal is the result of three approaches connected with (i) the fascinating and versatile use and potential of organosilanes in radical and cationic photopolymerization under air recently outlined [43,44,45]; (ii) the general interest for the in situ incorporation of metallic particles into a polymer film (i.e., the insertion of Ag(0) into a matrix is highly worthwhile for the design of biocide or bactericide incorporating polymers); and (iii) the formation of an excellent film-forming properties and the high glass-transition-temperature (220 °C) due to the formation of polyvinylcarbazole (PVK) in the film of polymer [46,47].
Scheme 3. Investigated compounds for polymerization processes.
Scheme 3. Investigated compounds for polymerization processes.
Applsci 05 00241 g012

2. Experimental Section

2.1. Compounds

The compounds with the highest purity available used are presented in Scheme 3. Diphenylsilane (DPSi), N-Vinylcarbazole (NVK), Silver hexafluoroantimonate (AgSbF6), Trimethylol-propane triacrylate (TMPTA), Styrene (St) and (3,4-epoxycyclohexane) methyl 3,4-epoxycyclohexylcarboxylate (EPOX) were all purchased from Sigma-Aldrich and used as received.

2.2. Radical and Cationic Polymerization

Silver salt (1% w/w) was dissolved in the selected monomer (TMPTA, St or EPOX) and diphenylsilane (DPSi) was introduced into the formulation by a syring at room temperature. The polymerization of these monomers (TMPTA, St and EPOX) was carried out in pill-box for a sample of 1 g. The progress of the exothermic polymerization is followed by monitoring the sample temperature using a thermocouple connected to a DaqPro-5300 (resolution 0.1 °C, Spectris Canada, Inc; Dba Omega Environmental, Laval, Canada). A magnetic stirrer in contact with the pill-box was used to ensure a good homogeneity. The addition of the DPSi into the formulation corresponds to time t = 0 s.
Since most radical or cationic polymerizations rapidly take place with the release of a large amount of heat (exothermic reaction), it has been shown [29,48] that the increase of the temperature is directly proportional to the monomer conversion. However, as the set-up used here is not adiabatic, the relationship between the conversion and the temperature is only valid for the first steps of the polymerization process.

2.3. Interpenetrated Polymer Network Synthesis

For the polymerization of TMPTA/EPOX (50%/50% w/w) blends, silver salt (1% w/w) was dissolved into the mixture of monomers and DPSi was introduced into the formulation by a syring at RT. The polymerization of TMPTA/EPOX was carried out in pill-box for a sample of 1 g. The progress of the exothermic polymerization is followed by monitoring the sample temperature using a thermocouple connected to a DaqPro-5300 (resolution 0.1 °C). A magnetic stirrer in contact with the pill-box was used to ensure a good homogeneity. The addition of the DPSi into the formulation corresponds to time t = 0 s.

2.4. Characterization

UV measurements were carried out with a Spectra max plus 384 spectrometer (VWR International; Mont Royal, QC, Canada), whereas Fourier-transform infrared spectra (FTIR) were performed on an ABB Bomen Spectrometer (ABB Bomen, Quebec, QC, Canada).
Elemental analysis was obtained by energy-dispersive X-ray spectrometry (EDS) using an EDS spectrometer (Oxford instruments Austin, Inc., Austin, TX, USA).
Differential Scanning Calorimetry (DSC) analysis was performed on a DSC-Q2000 instrument (TA instruments Inc., Mississauga, ON, Canada) under a nitrogen atmosphere at a heating or cooling rate of 20 °C/min. The films were scanned in the range of 25 °C to 300 °C.
Transmission electron microscopy (TEM) analyses of composite materials were carried out at an accelerated voltage of 80 kV using a Hitachi H-7500 TEM (Hitachi High Technologies Canada, Inc., Toronto, ON, Canada).

3. Results and Discussion

3.1. Free Radical Polymerization (FRP)

The polymerization of TMPTA smoothly proceeds at room temperature using DPSi/AgSbF6 (2%/1% w/w) (Figure 1A); the stirring is stopped within 3 min due to the complete polymerization of the formulation. The polymerization does not occur using DPSi or AgSbF6 independently thereby demonstrating that this combination is necessary to initiate the process. Almost no inhibition is noted as a short period is already requested for the solubilization of the silane into the formulation. A complete polymerization is obtained after about 300 s. It can be noted that the polymerization is dramatically faster under air than under argon (tack free time ~3 min vs. 6 min, respectively, Figure 1A). This result shows that (i) no strong oxygen inhibition obviously occurs; and (ii) oxygen should positively participate in the initiation process (see below). Remarkably, a high Si-H consumption is observed during the polymerization (Figure 1B) confirming the hydrogen abstraction process (Scheme 2 and Scheme 3) (see below): This demonstrates that the Si-H conversion is directly related to the overall efficiency of the system. The formation of polyacrylate (PTMPTA) is well evidenced by IR spectra by the observation of the disappearance of the characteristic band of the acrylate double bonds at 1620 cm−1 (Figure 1C) [10,49].
Interestingly, the addition of NVK strongly improves the polymerization profiles, i.e., both the polymerization rate and the conversion are higher in the presence of NVK (the increase of the temperature is directly proportional to the monomer conversion [29,48]) (Figure 2). A decrease of the tack free time is also noted (i.e., ~1 min vs. 3 min with and without NVK, respectively). Remarkably, no polymerization is observed when using NVK/AgSbF6 system in the same conditions (i.e., room temperature and under air). Consequently, the different three-component system (Silane/NVK/Silver salt) can be considered as excellent initiating system for thermal free radical polymerization under air and at RT (Figure 2).
Figure 1. (A) Role of the atmosphere in the free radical polymerization of TMPTA in the presence of DPSi/AgSbF6 (2%/1% w/w): (1) under air and (2) under argon; (B) IR band of the decreasing of the DPSi recorded during the polymerization; and (C) real-time FT-IR spectra in the 1600–1670 cm−1 wavenumber range of sample containing 1% (wt) of AgSbF6 and 2% (wt) of DPSi in TMPTA under air (1- before and 2- after polymerization).
Figure 1. (A) Role of the atmosphere in the free radical polymerization of TMPTA in the presence of DPSi/AgSbF6 (2%/1% w/w): (1) under air and (2) under argon; (B) IR band of the decreasing of the DPSi recorded during the polymerization; and (C) real-time FT-IR spectra in the 1600–1670 cm−1 wavenumber range of sample containing 1% (wt) of AgSbF6 and 2% (wt) of DPSi in TMPTA under air (1- before and 2- after polymerization).
Applsci 05 00241 g001aApplsci 05 00241 g001b
Figure 2. Comparison of the polymerization initiating ability under air and at RT of sample containing 1% (wt) of AgSbF6 and 2% (wt) of DPSi in TMPTA: (1) without NVK and (2) with 2% (wt) of NVK.
Figure 2. Comparison of the polymerization initiating ability under air and at RT of sample containing 1% (wt) of AgSbF6 and 2% (wt) of DPSi in TMPTA: (1) without NVK and (2) with 2% (wt) of NVK.
Applsci 05 00241 g002
However, the great interest of the proposed approach lies on the fact that the initiating system is efficient using simple glass vials and unpurified TMPTA, even in the presence of water or oxygen, which can act as polymerization inhibitors with other synthetic approaches [50]. This can open new opportunities for mass production and applications of PTMPTA.
Interestingly, when using a styrene monomer (St) instead TMPTA in the same experimental conditions, FRP occurs but the process is less efficient (e.g., inhibition time is high and tack free time ~4 h using St vs. no inhibition time and 1 min using TMPTA) (Figure 3A). This can be related to the low viscosity of this monomer (0.762 cp) compared to that of TMPTA (70–100 cp), which increases the oxygen inhibition [33,51]. The formation of polystyrene (PSt) is well evidenced by IR spectra of the obtained polymer (Figure 3B), which are fully consistent with the spectra that show the disappearance of the characteristic band of the acrylate double bonds at 1630 cm−1. Remarkably, the consumption of the silanes Si-H is almost complete (~100%) (Figure 3C). This is also found in agreement with the proposed mechanism (Scheme 1 and Scheme 2).
Figure 3. (A) Role of the monomer in the thermal free radical polymerization of Styrene (St) under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w); (B) real-time FT-IR spectra in the 1600–1650 cm−1 wavenumber range of sample containing 1% (wt) of AgSbF6, 2% (wt) of NVK and 2% (wt) of DPSi in Styrene under air (1) before and (2) after polymerization; and (C) IR band of the decreasing of the DPSi recorded during the polymerization.
Figure 3. (A) Role of the monomer in the thermal free radical polymerization of Styrene (St) under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w); (B) real-time FT-IR spectra in the 1600–1650 cm−1 wavenumber range of sample containing 1% (wt) of AgSbF6, 2% (wt) of NVK and 2% (wt) of DPSi in Styrene under air (1) before and (2) after polymerization; and (C) IR band of the decreasing of the DPSi recorded during the polymerization.
Applsci 05 00241 g003

3.2. Cationic Polymerization (CP)

Interestingly, the proposed new system (DPSi/NVK/AgSbF6) is quite versatile since they are also able to initiate the polymerization of epoxy monomers, such as (3,4-Epoxycyclohexane) methyl 3,4-epoxycyclohexylcarboxylate (EPOX), which indicates that the formation of the cationic species is also efficient in this case (Figure 4A). A quite good polymerization is always obtained at RT and under air leading to tack free surface coatings (gel time ~5 min and tack free time ~1 h). It is evident the important decrease of the typical epoxy band at 770–830 cm−1 and the high increase of the band at 1080 cm−1 due to ether group formed during the ring-opening polymerization (Figure 4B) [30,31,32,33,34,35,36,37]. Almost no inhibition is noted, as a short period is already requested for the solubilization of the silane into the formulation.
Figure 4. (A) Role of the monomer in the thermal free radical polymerization of EPOX under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w). (B) IR spectra recorded during the thermal cationic polymerization in the range between 760–1140 cm−1 of sample containing 1% (wt) of AgSbF6, 2% (wt) of DPSi and 2% (wt) of NVK in EPOX (1) before and (2) after polymerization.
Figure 4. (A) Role of the monomer in the thermal free radical polymerization of EPOX under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w). (B) IR spectra recorded during the thermal cationic polymerization in the range between 760–1140 cm−1 of sample containing 1% (wt) of AgSbF6, 2% (wt) of DPSi and 2% (wt) of NVK in EPOX (1) before and (2) after polymerization.
Applsci 05 00241 g004

3.3. DPSi/NVK/AgSbF6 for the Synthesis of Interpenetrated Polymer Networks (IPNs)

The new system (DPSi/NVK/AgSbF6) that simultaneously and efficiently generates initiating radicals and cations (see below, r4) can allow an easy synthesis of interpenetrated polymer network from TMPTA/EPOX blend (50%/50%) by thermal polymerization at RT and under air (Figure 5A). They ensure good to excellent polymerizations are always obtained leading to tack free surface coatings after only 3 min. A concomitant increase of the band at 1080 cm−1 and decrease of the band at 1620 cm−1 is also observed due to the formation of the polyether network and polyacrylate, respectively (Figure 5B) [30,31,32]. A high Si-H consumption is also found in agreement with the proposed mechanisms (Scheme 1 and Scheme 2) (Figure 5C). Remarkably, a decrease of the concentration of DPSi reduces the efficiency of polymerization: a decrease of the temperature and an increase of the tack free time is also noted (i.e., 3 and 5 min for 2 and 1% w/w of DPSi, respectively) (Figure 5D). This demonstrates that the silane concentration is directly related to the overall efficiency of the system.
Figure 5. (A) Role of the organosilane in the thermal interpenetrated polymer networks (IPNs) of TMPTA/EPOX blend (50%/50%) under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w). (B) IR spectra recorded during the thermal IPN polymerization of the TMPTA/EPOX blend (50%/50%) of sample containing 1% (wt) of AgSbF6, 2% (wt) of DPSi and 2% (wt) of NVK (1) before and (2) after polymerization. (C) IR band of the decreasing of the DPSi recorded during the polymerization. (D) Role of the concentration of organosilane in the thermal interpenetrated polymer networks (IPNs) of TMPTA/EPOX blend (50%/50%) under air in the presence of NVK/AgSbF6 (2%/1% w/w): (1) with 1% (wt) DPSi and (2) 2% (wt) DPSi.
Figure 5. (A) Role of the organosilane in the thermal interpenetrated polymer networks (IPNs) of TMPTA/EPOX blend (50%/50%) under air in the presence of DPSi/NVK/AgSbF6 (2%/2%/1% w/w). (B) IR spectra recorded during the thermal IPN polymerization of the TMPTA/EPOX blend (50%/50%) of sample containing 1% (wt) of AgSbF6, 2% (wt) of DPSi and 2% (wt) of NVK (1) before and (2) after polymerization. (C) IR band of the decreasing of the DPSi recorded during the polymerization. (D) Role of the concentration of organosilane in the thermal interpenetrated polymer networks (IPNs) of TMPTA/EPOX blend (50%/50%) under air in the presence of NVK/AgSbF6 (2%/1% w/w): (1) with 1% (wt) DPSi and (2) 2% (wt) DPSi.
Applsci 05 00241 g005aApplsci 05 00241 g005b

3.4. In-situ Incorporation of Silver Nanoparticules (Ag(0)NPs)

The in-situ formation of silver nanoparticles (Ag(0)NPs) during the polymerization process is demonstrated by UV-visible spectroscopy (Figure 6). The addition of DPSi to AgSbF6 in TMPTA, St or EPOX monomers leads to the formation of Ag(0) nanoparticles clearly evidenced by UV-visible spectroscopy (i.e., surface plasmon resonance (SPR) band typical of silver nanoparticles observed at about 405 nm (Figure 6)). Unprotected Ag(0) nanoparticles characterized by the size domain < 15 nm display a SPR band at a similar wavelength [52,53].
Figure 6. UV-visible absorption spectra of a polymerized TMPTA film using AgSbF6 (1% w/w) and NVK (2% w/w) in the absence (1) and in the presence (2) of DPSi (2% w/w); Insert: Evolution of Ag(0)NPs colors during the synthesis in TMPTA at RT and under air.
Figure 6. UV-visible absorption spectra of a polymerized TMPTA film using AgSbF6 (1% w/w) and NVK (2% w/w) in the absence (1) and in the presence (2) of DPSi (2% w/w); Insert: Evolution of Ag(0)NPs colors during the synthesis in TMPTA at RT and under air.
Applsci 05 00241 g006

3.5. Reaction Mechanisms

The formation of a silyl radicals (R3Si) and/or silylium cations (R3Si+) through the interaction of the silane with silver salt is assumed (Scheme 2) by analogy with the reported behavior of R3SiH/AgOCl4 in organic chemistry [54,55,56,57]. These reactions might proceed by an attack of the electrophile on a Si-H bond [54,55,56,57]. On the other side, the formation of silylium cations when using chlorosilanes (R3SiCl) instead of R3SiH with silver salts (r’1) was well characterized [58]. The good polymerization ability of the R3SiH/AgSbF6 and R3SiH/AgSbF6/NVK systems appears here as in full agreement with the formation of R3Si, NVK, R3Si+, NVK+ and H+, which are largely known [16,17,43,44,45,59,60,61,62] as efficient initiating species for both thermal and photochemical radical and/or cationic polymerization of TMPTA and EPOX.
Scheme 2, however, does not account for the favorable oxygen effect. The mechanism (Scheme 4) involves the oxidation of the silyl radical into a peroxyl, a rearrangement of the peroxyl (as already observed elsewhere) [63] and the further re-generation of a silyl radical together with the formation of R-SiH(O-R)2. Such a process was already reported in [63,64].
Scheme 4. Kinetics of rearrangement of the peroxyl radicals into silyl radicals.
Scheme 4. Kinetics of rearrangement of the peroxyl radicals into silyl radicals.
Applsci 05 00241 g013
Upon addition of NVK, R3Si react with the NVK double bond leading to R3Si-NVK (r’2), as known in other systems [59,60,61,62]. As previously discussed [59,60,61,62], R3Si-NVK radicals are easily oxidized by Ag+ leading to the formation of R3Si-NVK+ (r3) [65,66]. R3Si-NVK is characterized by a very low ionization potential [60]; this is fully consistent with the associated fast oxidation process.
The R3Si-NVK and R3Si-NVK+ being a highly efficient initiating structure for the FRP and ROP of an acrylate and an epoxy [59,65,66,67], it is not surprising that an efficient polymerization is found in the presence of NVK.
In these reactions, radicals and cations are probably generated to initiate the free radical polymerization (FRP), ring opening polymerization (ROP) and interpenetrated polymer network (IPN) of the acrylate, epoxy and acrylate/epoxy blend monomers (r4), respectively, with in-situ formation of silver nanoparticles (Ag(0)NPs) during the polymerization process (see above). Other elegant initiating systems based on other radical oxidation processes by Ag+ were proposed in ref [68,69,70,71,72,73,74,75].
R3SiCl + AgSbF6 →→→ R3Si+SbF6 + Ag(0) + products
R3Si + NVK → R3Si-NVK
R3Si-NVK + Ag+ → R3Si-NVK+ + Ag(0)
R3Si-NVK, R3Si + acrylate → Polymer
R3Si-NVK+, R3Si+, H+ + epoxy → Polymer

3.6. Characterization of Ag(0) Nanoparticles (NPs) and Polymer Films

Functionalized Ag(0)NPs were characterized by TEM and EDX. The bright field transmission electron microscopy images (TEM) of films cured using AgSbF6/NVK/DPSi (1%/2%/2% w/w) system dissolved in different monomers (TMPTA, St or EPOX) are reported in Figure 7. The TEM micrograph (Figure 7) of Ag(0)NPs obtained after thermal polymerization at RT shows that the Ag(0)NPs remain well-dispersed in the polymer matrix and are well dispersed with no significant agglomeration. A spherical geometry was noticed of the particles whose diameters are in the 5 to 25 nanometer range. The TEM images suggest the possibility of a homogeneous polymeric coating on the nanoparticle surface. These morphological results are in agreement with ref [10,62,72,76].
Figure 7. TEM analysis of silver nanoparticles obtained after thermal polymerization using AgSbF6/NVK/DPSi (1%/2%/2% w/w) system in (A) TMPTA, (B) St, (C) EPOX and (D) IPN (TMPTA/EPOX).
Figure 7. TEM analysis of silver nanoparticles obtained after thermal polymerization using AgSbF6/NVK/DPSi (1%/2%/2% w/w) system in (A) TMPTA, (B) St, (C) EPOX and (D) IPN (TMPTA/EPOX).
Applsci 05 00241 g007
The formation of Ag(0)NPs created in situ in this hybrid polymer matrix during thermal radical and/or cationic polymerization at room temperature were monitored using energy dispersive X-ray (EDS) spectroscopy analysis, as shown in Figure 8. Indeed, in the resulting EDS pattern, C, O, F, Si, Au, Ag and Sb peaks are clearly shown, demonstrating that silver/polymer composites were successfully prepared by this one-step process. Remarkably, the presence of Au(0) particules is due to the phenomenon of metallization of the polymer matrix before making the measurement.
Thermal behavior of the cured films was evaluated by DSC measurements. The DSC curves are reported in Figure 9. Remarkably, a clear increase of the glass-transition-temperature (Tg) values after addition of N-vinylcarbazole (NVK) in the formulations was noted (i.e., 226 °C for curve 2 vs. 212.5 °C for curve 1, Figure 9). The reason for such an increase is due to the formation of polyvinylcarbazole (PVK) in the film of polymer, which leads to (1) a decrease of the free volume; and (2) the impediment of the mobility crosslinked epoxide and/or an acrylate polymer network.
Figure 8. Energy-Dispersive X-ray Spectrometry (EDS) analysis recorded from Ag(0) nanoparticles in EPOX.
Figure 8. Energy-Dispersive X-ray Spectrometry (EDS) analysis recorded from Ag(0) nanoparticles in EPOX.
Applsci 05 00241 g008
Figure 9. DSC thermograms for the cured samples containing (1) DPSi/AgSbF6 (2%/1% w/w) and (2) DPSi/NVK/AgSbF6 (2%/2%/1% w/w) in TMPTA.
Figure 9. DSC thermograms for the cured samples containing (1) DPSi/AgSbF6 (2%/1% w/w) and (2) DPSi/NVK/AgSbF6 (2%/2%/1% w/w) in TMPTA.
Applsci 05 00241 g009

4. Conclusions

In the present study, in situ synthesis of silver-acrylate, silver-epoxy and silver-IPN nanocomposites was achieved by thermal polymerization through a simultaneous radical and/or cationic polymerization processes using a silane/silver salt couple at room temperature and under air. Remarkably, and unlike published works so far, heating is not required to activate the reaction. The new redox systems were proposed to initiate in two-component systems (DPSi/silver salt) the ring-opening polymerization (ROP) of (3,4-Epoxycyclohexane) methyl 3,4-epoxycyclohexylcarboxylate (EPOX) and the free radical polymerization (FRP) of trimethylolpropane triacrylate (TMPTA) and styrene (St). Interestingly, the addition of N-Vinylcarbazole (NVK) to the silane/silver salt systems leads to a drastic increase of polymerization reactivity and efficiency, respectively. This three-component system generates free radicals (R3Si and NVK) or ions (R3Si+, NVK+ and H+). Remarkably, fast ROP and FRP are observed. Interestingly, these initiators can also be used for the synthesis of interpenetrated polymer networks (IPN), e.g., for the polymerization of TMPTA/EPOX blends. Highly crosslinked network were achieved and the nanocomposite obtained, contains near spherical geometry for most of the nanoparticles, homogeneously dispersed in the polymer network without significant agglomeration as a result of a surrounding effect of polymer chains formed by the rapid cationic and/or radical chain growth. Interestingly, an increase on Tg values was observed after addition of N-vinylcarbazole (NVK) in the formulations. This was ascribed to a decrease of free volume of the polymer network because the formation of PVK in the film of polymer leads in the impediment of the mobility crosslinked epoxide and/or an acrylate polymer network. The mechanical and optical properties of the obtained new organic polymer/silver nanoparticles will be investigated in near future.

Author Contributions

Hereby, the three authors of the paper testify that we have contributed substantially to the study design, data analysis and manuscript writing and revision. Study conception and design were done by M.-A.T., S.E. and M.R. Experiemental part was carried out by M.-A.T. Data analysis and interpretation were performed by M.-A.T., S.E. and M.R. Drafting of manuscript and revision were carried out by M.-A.T., S.E. and M.R.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Carotenuto, G.; Martorana, B.; Perlo, P.B.; Nicolais, L. A universal method for the synthesis of metal and metal sulfide clusters embedded in polymer matrices. J. Mater. Chem. 2003, 13, 2927–2930. [Google Scholar] [CrossRef]
  2. Biswas, A.; Aktas, O.C.; Schurmann, U.; Saeed, U.; Zaporojtchenko, V.; Faupel, F. Controlled syntheses of Ag-polytetrafluoroethylene nanocomposite thin films by co-sputtering from two magnetron sources. Appl. Phys. Lett. 2004, 84, 2655–2657. [Google Scholar]
  3. Sih, B.C.; Wolf, M.O. Metal nanoparticle—Conjugated polymer nanocomposites. Chem. Commun. 2005, 27, 3375–3384. [Google Scholar] [CrossRef] [PubMed]
  4. Fan, F.R.F.; Bard, A.J. Chemical, electrochemical, gravimetric, and microscopic studies on antimicrobial silver films. J. Phys. Chem. B 2002, 106, 279–287. [Google Scholar] [CrossRef]
  5. Ouyang, J.Y.; Chu, C.W.; Szmanda, C.R.; Ma, L.P.; Yang, Y. Programmable polymer thin film and non-volatile memory device. Nat. Mater. 2004, 3, 918–922. [Google Scholar] [CrossRef] [PubMed]
  6. Chang, L.T.; Yen, C.C. Studies on the preparation and properties of conductive polymers. VIII. Use of heat treatment to prepare metallized films from silver chelate of PVA and PAN. J. Appl. Polym. Sci. 1995, 55, 371–374. [Google Scholar] [CrossRef]
  7. Shanmugam, S.; Viswanathan, B.; Varadarajan, T.K. A novel single step chemical route for noble metal nanoparticles embedded. Mater. Chem. Phys. 2006, 95, 51–55. [Google Scholar] [CrossRef]
  8. Lin, W.C.; Yang, M.C. Novel Silver/Poly(vinyl alcohol) Nanocomposites for Surface-Enhanced Raman Scattering-Active Substrates. Macromol. Rapid Commun. 2005, 26, 1942–1947. [Google Scholar] [CrossRef]
  9. Balan, L.; Burget, D. Synthesis of metal/polymer nanocomposite by UV-radiation curing. Eur. Polym. J. 2006, 42, 3180–3189. [Google Scholar] [CrossRef]
  10. Balan, L.; Schneider, R.; Lougnot, D.J. A new and convenient route to polyacrylate/silver nanocomposites by light-induced cross-linking polymerization. Prog. Org. Coat. 2008, 62, 351–357. [Google Scholar] [CrossRef]
  11. Selvan, S.T.; Spatz, J.P.; Klok, H.A.; Möller, M. Gold–Polypyrrole Core–Shell Particles in Diblock Copolymer Micelles. Adv. Mater. 1998, 10, 132–134. [Google Scholar] [CrossRef]
  12. Corbierre, M.K.; Cameron, N.S.; Sutton, M.; Mochrie, S.G.; Lurio, L.B.; Rühm, A.; Lennox, R.B. Polymer-Stabilized Gold Nanoparticles and Their Incorporation into Polymer Matrices. J. Am. Chem. Soc. 2001, 123, 10411–10412. [Google Scholar] [CrossRef] [PubMed]
  13. Matyjaszewski, K.; Gnanou, Y.; Leibler, L. Macromolecular Engineering: From Precise Macromolecular Synthesis to Macroscopic Materials Properties and Applications; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  14. Pappas, S.P. UV Curing: Science and Technology; Technology Marketing Corp: Aachen, Germany, 1985. [Google Scholar]
  15. Dietliker, K. A Compilation of Photoinitiators Commercially Available for UV Today; Sita Technology Ltd.: Edinburgh, UK; London, UK, 2002. [Google Scholar]
  16. Crivello, J.V. Photoinitiators for Free Radical, Cationic and Anionic Photopolymerization, 2nd ed.; Bradley, G., Ed.; Wiley & Sons: New York, NY, USA, 1998. [Google Scholar]
  17. Crivello, J.V. Ring-Opening Polymerization; Brunelle, D.J., Ed.; Hanser: Munich, Germany, 1993. [Google Scholar]
  18. Sangermano, M.; Crivello, J.V. Photoinitiated Polymerization; Belfied, K.D., Crivello, J.V., Eds.; ACS Symposium Series 847; Washington, DC, USA, 2003. [Google Scholar]
  19. Fouassier, J.P. Photoinitiation Photopolymerization and Photocuring: Fundamental and Applications; Hanser Publishers: New York, NY, USA, 1995. [Google Scholar]
  20. Radiation Curing in Polymer Science and Technology; Fouassier, J.P.; Rabek, J.F. (Eds.) Elsevier Science Publishers Ltd.: London, UK, 1993.
  21. El-Roz, M.; Lalevée, J.; Allonas, X.; Fouassier, J.P. Mechanistic Investigation of the Silane, Germane, and Stannane Behavior When Incorporated in Type I and Type II Photoinitiators of Polymerization in Aerated Media. Macromolecules 2009, 42, 8725–8732. [Google Scholar] [CrossRef]
  22. Hoyle, C.E.; Kim, K.-J. Photoinitiated polymerization of 1,6-hexanediol diacrylate-amine synergists. J. Rad. Curing 1985, 12, 9–15. [Google Scholar]
  23. Decker, C.; Jenkins, A.D. Kinetic approach of oxygen inhibition in ultraviolet- and laser-induced polymerizations. Macromolecules 1985, 18, 1241–1244. [Google Scholar] [CrossRef]
  24. Gou, L.; Opheim, B.; Scranton, A.B. Photochemistry and UV Curing: New Trends; Fouassier, J.P., Ed.; Research Signpost: Trivandrum, India, 2006. [Google Scholar]
  25. Crivello, J.V.; Lam, J.H.W. Redox cationic polymerization: The diaryliodonium salt/ascorbate redox couple. J. Polym. Sci. Part A: Polym. Chem. 1981, 19, 539–548. [Google Scholar] [CrossRef]
  26. Crivello, J.V.; Lee, J. Redox initiators for cationic polymerization: The diaryliodonium Salt/Sn(II) redox couple. Makromol. Chem. 1983, 184, 463–473. [Google Scholar] [CrossRef]
  27. Onen, A.; Yagci, Y. Redox-initiated cationic polymerization: The pyridinium salt/ascorbate redox couple. Polymer 1997, 38, 1423–1425. [Google Scholar] [CrossRef]
  28. Onen, A.; Denizligil, S.; Yagci, Y. Cationic polymerization using allyl sulfonium salt. 3. Allyl sulfonium salt/ascorbate redox couple. Angew. Makromol. Chem. 1997, 245, 149–154. [Google Scholar] [CrossRef]
  29. Crivello, J.V. Redox initiated cationic polymerization. J. Polym. Sci. Part A: Polym. Chem. 2009, 47, 1825–1835. [Google Scholar] [CrossRef]
  30. Souane, R.; Tehfe, M.-A.; Lalevée, J.; Gigmes, D.; Fouassier, J.-P. New Initiating Systems for Thermal Cationic Polymerization at Ambient Temperature with in situ Formation of Ag(0) Nanoparticles: A Silane/Silver Salt Combination. Macromol. Chem. Phys. 2010, 211, 1441–1445. [Google Scholar] [CrossRef]
  31. Tehfe, M.-A.; Jamois, R.; Cousin, P.; Elkoun, S.; Robert, M. In situ synthesis and characterization of silver/polymer nanocomposites by thermal cationic polymerization processes at room temperature: Initiating systems based on organosilanes and starch nanocrystals. Langmuir 2015, 31, 4305–4315. [Google Scholar] [CrossRef] [PubMed]
  32. Yagci, Y.; Sahin, O.; Ozturk, T.; Marchi, S.; Grassini, S.; Sangermano, M. Synthesis of silver/epoxy nanocomposites by visible light sensitization using highly conjugated thiophene derivatives. React. Funct. Polym. 2011, 71, 857–862. [Google Scholar] [CrossRef]
  33. Lalevée, J.; Tehfe, M.A.; Morlet-Savary, F.; Allonas, X.; Graff, B.; Fouassier, J.P. Oxygen mediated and wavelength tunable cationic photopolymerization reactions under air and low intensity: A new concept. Prog. Org. Coat. 2011, 70, 23–31. [Google Scholar] [CrossRef]
  34. Lalevée, J.; Blanchard, N.; Tehfe, M.A.; Morlet-Savary, F.; Fouassier, J.P. Green Bulb Light Source Induced Epoxy Cationic Polymerization under Air Using Tris(2,20-bipyridine)ruthenium(II) and Silyl Radicals. Macromolecules 2010, 43, 10191–10195. [Google Scholar] [CrossRef]
  35. Lalevée, J.; Blanchard, N.; Tehfe, M.A.; Peter, M.; Morlet-Savary, F.; Gigmes, D.; Fouassier, J.P. Efficient dual radical/cationic photoinitiator under visible light: A new concept. Polym. Chem. 2011, 2, 1986–1991. [Google Scholar] [CrossRef]
  36. Lalevée, J.; Blanchard, N.; Tehfe, M.A.; Peter, M.; Morlet-Savary, F.; Fouassier, J.P. A Novel Photopolymerization Initiating System Based on an Iridium Complex Photocatalyst. Macromol. Rapid Commun. 2011, 32, 917–920. [Google Scholar] [CrossRef] [PubMed]
  37. Lalevée, J.; Blanchard, N.; Tehfe, M.A.; Peter, M.; Morlet-Savary, F.; Fouassier, J.P. Household LED irradiation under air: Cationic polymerization using iridium or ruthenium complex photocatalysts. Polym. Bull. 2011, 68, 341–347. [Google Scholar] [CrossRef]
  38. Durmaz, Y.Y.; Moszner, N.; Yagci, Y. Visible Light Initiated Free Radical Promoted Cationic Polymerization Using Acylgermane Based Photoinitiator in the Presence of Onium Salts. Macromolecules 2008, 41, 6714. [Google Scholar] [CrossRef]
  39. Crivello, J.V. A new visible light sensitive photoinitiator system for the cationic polymerization of epoxides. J. Polym. Sci. Part A: Polym. Chem. 2009, 47, 866–875. [Google Scholar] [CrossRef]
  40. Crivello, J.V. Radical-Promoted Visible Light Photoinitiated Cationic Polymerization of Epoxides. J. Macromol. Sci. Part A 2009, 46, 474–483. [Google Scholar] [CrossRef]
  41. Tunc, D.; Yagci, Y. Thioxanthone-ethylcarbazole as a soluble visible light photoinitiator for free radical and free radical promoted cationic polymerizations. Polym. Chem. 2011, 2, 2557–2563. [Google Scholar] [CrossRef]
  42. Yilmaz, G.; Beyazit, S.; Yagci, Y. Visible light induced free radical promoted cationic polymerization using thioxanthone derivatives. J. Polym. Sci. Part A: Polym. Chem. 2011, 49, 1591–1596. [Google Scholar] [CrossRef]
  43. Lalevée, J.; Blanchard, N.; El-Roz, M.; Graff, B.; Allonas, X.; Fouassier, J.P. New Photoinitiators Based on the Silyl Radical Chemistry: Polymerization Ability, ESR Spin Trapping, and Laser Flash Photolysis Investigation. Macromolecules 2008, 41, 4180–4186. [Google Scholar] [CrossRef]
  44. Lalevée, J.; El-Roz, M.; Allonas, X.; Fouassier, J.P. Free-radical-promoted cationic photopolymerization under visible light in aerated media: New and highly efficient silane-containing initiating systems. J. Polym. Sci. Part A: Polym. Chem. 2008, 46, 2008–2014. [Google Scholar] [CrossRef]
  45. Tehfe, M.A.; Lalevée, J.; Allonas, X.; Fouassier, J.P. Long Wavelength Cationic Photopolymerization in Aerated Media: A Remarkable Titanocene/Tris(trimethylsilyl)silane/Onium Salt Photoinitiating System. Macromolecules 2009, 42, 8669–8674. [Google Scholar] [CrossRef]
  46. Uygun, M.; Kahveci, M.U.; Odaci, D.; Timur, S.; Yagci, Y. Antibacterial Acrylamide Hydrogels Containing Silver Nanoparticles by Simultaneous Photoinduced Free Radical Polymerization and Electron Transfer Processes. Macromol. Chem. Phys. 2009, 210, 1867–1875. [Google Scholar] [CrossRef]
  47. Dawson-Andoh, B.; Matuana, L. High density polyethylene-wood flour composite lumber: Efficacy of two proprietary biocides in the control of fungal colonization and discoloration. Eur. J. Wood Wood Products 2007, 65, 331–334. [Google Scholar] [CrossRef]
  48. Crivello, J.V. Redox initiated cationic polymerization: Reduction of diaryliodonium salts by 9-BBN. J. Polym. Sci. Part A: Polym. Chem. 2009, 47, 5639–5651. [Google Scholar] [CrossRef]
  49. Tehfe, M.-A.; Dumur, F.; Telitel, S.; Gigmes, D.; Contal, E.; Bertin, D.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; Lalevée, J. Zinc-based metal complexes as new photocatalysts in polymerization initiating systems. Eur. Polym. J. 2013, 49, 1040–1049. [Google Scholar] [CrossRef]
  50. Kunioka, M.; Wang, Y.; Onozawa, S-Y. Polymerization of Poly(ε-caprolactone) Using Yttrium Triflate. Polym. J. 2003, 35, 422–429. [Google Scholar] [CrossRef]
  51. Lalevée, J.; Dirani, A.; El Roz, M.; Graff, B.; Allonas, X.; Fouassier, J.P. Silanes as New Highly Efficient Co-initiators for Radical Polymerization in Aerated Media. Macromolecules 2008, 41, 2003–2010. [Google Scholar] [CrossRef]
  52. Taleb, A.; Petit, C.; Pileni, M.P. Optical Properties of Self-Assembled 2D and 3D Superlattices of Silver Nanoparticles. J. Phys. Chem. B 1998, 102, 2214–2220. [Google Scholar] [CrossRef]
  53. He, S.; Yao, J.; Jiang, P.; Shi, D.; Zhang, H.; Xie, S.; Pang, S.; Gao, H. Formation of Silver Nanoparticles and Self-Assembled Two-Dimensional Ordered Superlattice. Langmuir 2001, 17, 1571–1575. [Google Scholar] [CrossRef]
  54. Eaborn, C. The reaction between trisubstituted silanes and silver perchlorate. Organosilicon Compounds. XIV. J. Chem. Soc. 1955, 3, 2517–2519. [Google Scholar]
  55. Marciniec, B. Mechanism of competitive-consecutive reaction of trisubstituted silanes with silver perchlorate in benzene. Z. Anorg. Allg. Chem. 1982, 495, 232–240. [Google Scholar] [CrossRef]
  56. Ouellette, R.J.; Marks, D.L.; Miller, D.; Kesatie, D. Oxidation of silanes. III. Reaction of aryldimethylsilanes with mercuric acetate and thallium triacetate. J. Org. Chem. 1969, 34, 1769–1771. [Google Scholar] [CrossRef]
  57. Ouellette, R.J.; Marks, D.L. Oxidation of silanes II. The reaction of ozone with aryldimethylsilanes. J. Organomet. Chem. 1968, 11, 407–413. [Google Scholar] [CrossRef]
  58. Lambert, J.B.; Kania, L.; Zhang, S. Modern Approaches to Silylium Cations in Condensed Phase. Chem. Rev. 1995, 95, 1191–1201. [Google Scholar] [CrossRef]
  59. Lalevée, J.; Tehfe, M.-A.; Zein-Fakih, A.; Ball, B.; Telitel, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.P. N-Vinylcarbazole: An Additive for Free Radical Promoted Cationic Polymerization upon Visible Light. ACS Macro Lett. 2012, 1, 802–806. [Google Scholar] [CrossRef]
  60. Telitel, S.; Dumur, F.; Faury, T.; Graff, B.; Tehfe, M.-A.; Gigmes, D.; Fouassier, J.-P.; Lalevée, J. New core-pyrene π structure organophotocatalysts usable as highly efficient photoinitiators. Beilstein J. Org. Chem. 2013, 9, 877–890. [Google Scholar] [CrossRef] [PubMed]
  61. Lalevée, J.; Dumur, F.; Mayer, C.-R.; Gigmes, D.; Nasr, G.; Tehfe, M.-A.; Telitel, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P. Photopolymerization of N-Vinylcarbazole Using Visible-Light Harvesting Iridium Complexes as Photoinitiators. Macromolecules 2012, 45, 4134–4141. [Google Scholar] [CrossRef]
  62. Mokbel, H.; Dumur, F.; Telitel, S.; Vidal, L.; Xiao, P.; Versace, D.-L.; Tehfe, M.-A.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; et al. Photoinitiating systems of polymerization and in situ incorporation of metal nanoparticles into polymer matrices upon exposure to visible light: Push–pull malonate and malononitrile based dyes. Polym. Chem. 2013, 4, 5679–5687. [Google Scholar] [CrossRef]
  63. Lalevée, J.; Blanchard, N.; Graff, B.; Allonas, X.; Fouassier, J.P. Tris(trimethylsilyl)silyl versus tris(trimethylsilyl)germyl: Radical reactivity and oxidation ability. J. Organomet. Chem. 2008, 693, 3643–3649. [Google Scholar] [CrossRef]
  64. Zaborovskiy, A.B.; Lutsyk, D.S.; Prystansky, R.E.; Kopylets, V.I.; Timokhin, V.I.; Chatgilialoglu, C. A mechanistic investigation of (Me3Si)3SiH oxidation. J. Organomet. Chem. 2004, 689, 2912–2919. [Google Scholar] [CrossRef]
  65. Hua, Y.; Crivello, J.V. Synergistic interaction of epoxides and N-vinylcarbazole during photoinitiated cationic polymerization. J. Polym. Sci. A: Polym. Chem. 2000, 38, 3697–3709. [Google Scholar] [CrossRef]
  66. Abdoul-Rasoul, F.A.M.; Ledwith, A.; Yagci, Y. Photochemical and thermal cationic polymerizations promoted by free radical initiators. Polymer 1978, 19, 1219–1222. [Google Scholar] [CrossRef]
  67. Xiao, P.; Lalevée, J.; Zhao, J.; Stenzel, M.H. N-Vinylcarbazole as Versatile Photoinaddimer of Photopolymerization under Household UV LED Bulb (392 nm). Macromol. Rapid Commun. 2015. [Google Scholar] [CrossRef] [PubMed]
  68. Scaiano, J.C.; Billone, P.; Gonzalez, C.M.; Maretti, L.; Marin, M.L.; McGilvray, K.L.; Yuan, N. Photochemical routes to silver and gold nanoparticles. Pure Appl. Chem. 2009, 81, 635–647. [Google Scholar] [CrossRef]
  69. McGilvray, K.L.; Decan, M.R.; Wang, D.; Scaiano, J.C. Facile Photochemical Synthesis of Unprotected Aqueous Gold Nanoparticles. J. Am. Chem. Soc. 2006, 128, 15980–15981. [Google Scholar] [CrossRef] [PubMed]
  70. Marin, M.L.; McGilvray, K.L.; Scaiano, J.C. Photochemical Strategies for the Synthesis of Gold Nanoparticles from Au(III) and Au(I) Using Photoinduced Free Radical Generation. J. Am. Chem. Soc. 2008, 130, 16572–16584. [Google Scholar] [CrossRef] [PubMed]
  71. Stamplecoskie, K.G.; Scaiano, J.C. Kinetics of the Formation of Silver Dimers: Early Stages in the Formation of Silver Nanoparticles. J. Am. Chem. Soc. 2011, 133, 3913–3920. [Google Scholar] [CrossRef] [PubMed]
  72. Sangermano, M.; Yagci, Y.; Rizza, G. In Situ Synthesis of Silver-Epoxy Nanocomposites by Photoinduced Electron Transfer and Cationic Polymerization Processes. Macromolecules 2007, 40, 8827–8829. [Google Scholar] [CrossRef]
  73. Yagci, Y.; Sangermano, M.; Rizza, G. A visible light photochemical route to silver-epoxy nanocomposites by simultaneous polymerization–reduction approach. Polymer 2008, 49, 5195–5198. [Google Scholar] [CrossRef]
  74. Yagci, Y.; Sangermano, M.; Rizza, G. In situ synthesis of gold-cross-linked poly(ethylene glycol) nanocomposites by photoinduced electron transfer and free radical polymerization processes. Chem. Commun. 2008, 24, 2771–2773. [Google Scholar]
  75. Nehlig, E.; Schneider, R.; Vidal, L.; Clavier, G.; Balan, L. Silver Nanoparticles Coated with Thioxanthone Derivative as Hybrid Photoinitiating Systems for Free Radical Polymerization. Langmuir 2012, 28, 17795–17802. [Google Scholar] [CrossRef] [PubMed]
  76. Buruiana, T.; Melinte, V.; Jitaru, F.; Buruiana, E.-C.; Balan, L. Preparation of siloxane-based urethane dimethacrylates carrying carboxylic groups and the effect of silver nanoparticles on the properties of composite polymer films. Polym. Chem. Part A 2012, 50, 874–883. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Tehfe, M.-A.; Elkoun, S.; Robert, M. N-Vinylcarbazole: As an Additive for Thermal Polymerization at Room Temperature with in situ Formation of Ag(0) Nanoparticules. Appl. Sci. 2015, 5, 241-258. https://doi.org/10.3390/app5030241

AMA Style

Tehfe M-A, Elkoun S, Robert M. N-Vinylcarbazole: As an Additive for Thermal Polymerization at Room Temperature with in situ Formation of Ag(0) Nanoparticules. Applied Sciences. 2015; 5(3):241-258. https://doi.org/10.3390/app5030241

Chicago/Turabian Style

Tehfe, Mohamad-Ali, Saïd Elkoun, and Mathieu Robert. 2015. "N-Vinylcarbazole: As an Additive for Thermal Polymerization at Room Temperature with in situ Formation of Ag(0) Nanoparticules" Applied Sciences 5, no. 3: 241-258. https://doi.org/10.3390/app5030241

Article Metrics

Back to TopTop