Next Article in Journal
Effect of Printing Process Parameters on the Shape Transformation Capability of 3D Printed Structures
Previous Article in Journal
Thermal Behaviour of Microgels Composed of Interpenetrating Polymer Networks of Poly(N-isopropylacrylamide) and Poly(acrylic acid): A Calorimetric Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

ABC-Type Triblock Copolyacrylamides via Copper-Mediated Reversible Deactivation Radical Polymerization

by
Fehaid M. Alsubaie
1,*,
Othman Y. Alothman
2,*,
Hassan Fouad
3,4 and
Abdel-Hamid I. Mourad
5,6,7
1
National Center for Chemical Catalysis Technology, King Abdulaziz City for Science and Technology (KACST), P.O. Box 6086, Riyadh 11442, Saudi Arabia
2
Department of Chemical Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
3
Applied Medical Science Department, Community College, King Saud University, P.O. Box 10219, Riyadh 11433, Saudi Arabia
4
Biomedical Engineering Department, Faculty of Engineering, Helwan University, Cairo 11792, Egypt
5
Mechanical and Aerospace Engineering Department, College of Engineering, United Arab Emirate University, Al Ain P.O. Box 15551, United Arab Emirates
6
National Water and Energy Centre, United Arab Emirate University, Al Ain P.O. Box 15551, United Arab Emirates
7
Mechanical Design Department, Faculty of Engineering, Helwan University, Cairo 11795, Egypt
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(1), 116; https://doi.org/10.3390/polym14010116
Submission received: 8 November 2021 / Revised: 12 December 2021 / Accepted: 20 December 2021 / Published: 29 December 2021

Abstract

:
The aqueous Cu(0)-mediated reversible deactivation radical polymerization (RDRP) of triblock copolymers with two block sequences at 0.0 °C is reported herein. Well-defined triblock copolymers initiated from PHEAA or PDMA, containing (A) 2-hydroxyethyl acrylamide (HEAA), (B) N-isopropylacrylamide (NIPAM) and (C) N, N-dimethylacrylamide (DMA), were synthesized. The ultrafast one-pot synthesis of sequence-controlled triblock copolymers via iterative sequential monomer addition after full conversion, without any purification steps throughout the monomer additions, was performed. The narrow dispersities of the triblock copolymers proved the high degree of end-group fidelity of the starting macroinitiator and the absence of any significant undesirable side reactions. Controlled chain length and extremely narrow molecular weight distributions (dispersity ~1.10) were achieved, and quantitative conversion was attained in as little as 52 min. The full disproportionation of CuBr in the presence of Me6TREN in water prior to both monomer and initiator addition was crucially exploited to produce a well-defined ABC-type triblock copolymer. In addition, the undesirable side reaction that could influence the living nature of the system was investigated. The ability to incorporate several functional monomers without affecting the living nature of the polymerization proves the versatility of this approach.

1. Introduction

The homo-polymers and copolymers of acrylamide-based monomers have been employed in various applications [1,2,3,4,5]. Although the free-radical polymerization of acrylamide derivatives has been thoroughly investigated, developed controlled free-radical polymerizations have recently been utilized to produce new polymeric materials with unique properties. A few years ago, reversible deactivation radical polymerization (RDRP) started to attract great attention. The RDRP of acrylamide-based monomers has proven challenging when water is used as the solvent. Usually, in aqueous solutions, the rate of the RDRP process is rapid, resulting in termination events. This has been attributed to the lack of equilibrium between the dormant species and the active radical species. During the polymerization processes and the activation and deactivation steps, equilibrium is key to attaining the linear polymeric chains’ growth. The polymer research community has developed techniques to achieve fast equilibrium between dormant and active chain species, including nitroxide-mediated polymerization (NMP) [6,7], reversible addition–fragmentation chain transfer (RAFT) [8,9], atom transfer radical polymerization (ATRP) [10,11,12], in situ polymerization [13,14] and single-electron transfer living radical polymerization (SET-LRP) [15].
The mechanism of copper-mediated living radical polymerization (CM-LRP) proposes the equilibrium between propagating species and alkyl halides that can be mediated by Cu–ligand complexes [16,17,18,19,20]. To achieve better control of CM-LRP, a mixture of water and methanol media should be used [21]. It was reported that the polymerization of acrylamide-based monomers in water is rather difficult due to undesirable termination reactions that cause a loss of chain-end fidelity [21,22,23]. The aqueous Cu-mediated LRP of acrylamides has been shown to be problematic with regard to the control of the polymerization when water is employed as the only solvent at ambient temperature [24,25]. Furthermore, in the literature, few publications have reported controlled block copolymerization, and they are limited to just diblock. Brittain et al. [21] used different copper salts to support the aqueous ATRP of dimethylacrylamide. They concluded that the addition of the Cu salt complex to the amide group of the chain ends stabilized the radicals, resulting in an increase in the radical concentration and side reactions.
SET-LRP is an RDRP process whereby polymers with high end-group functionality can be synthesized. In this process, the key step is the disproportionation of Cu(I)Br in the presence of nitrogen-containing ligands, as this produces both the activator Cu(0) and the deactivator Cu(ΙΙ)Br2 [15]. The equilibrium constant for the disproportionation depends mainly on the nature of the solvent and ligand. The equilibrium constant is noticeably higher in the presence of Me6TREN aqueous media than in organic solvents [26,27,28,29].
Whittaker et al. [29] applied Cu(0)-mediated polymerization to formulate high-order multiblock copolymers at atmospheric temperature in DMSO. The copolymers (consisting of four acrylates) were synthesized via sequential monomer addition to build multiblock copolymers model P(MA-b-MA) homo-polymer and P(MA-b-nBuA-b-EA-b-2EHA-b-EA-b-nBuA) to allow each block to have very discreet blocks (ideally two monomer units). Interestingly, this shows that full conversion of the monomer can be reached, which illustrates the effectiveness of this technique. The molecular weight distributions remained fairly narrow (Ð ~1.2) after monomer additions (24 h per block), proving that the polymerization is highly controlled without any purification steps. Later, the authors employed a similar approach to synthesize a decablock copolymer. However, they obtained molecular weight distributions (1.72) without quantitative conversion [30].
Later, the Whittaker group [31,32] utilized a similar technique to synthesize higher-molecular-weight block homo-polymers and copolymers of different acrylates (each block DPn ≈ 100). Narrow dispersities (<1.2) were achieved up to the sixth block, while the monomer conversion rate ranged from 92 to 100%. Moreover, multiblock glycopolymers (the degree of polymerization (DP) = 2 for each block, (mannose)2-(glucose)2-(mannose)2-(glucose)2-(mannose)2-(glucose)2) were prepared [33]. The copolymerization was performed in DMSO at room temperature, with a total reaction time of 46 h, resulting in good control over the molecular weight distributions.
Another approach to the synthesis of multiblock copolymers is to employ RAFT as a means of polymerization. The synthesis of an icosablock (20 blocks) has been achieved [34,35], with each block consisting of an average of three monomer units composed of three types of acrylamide monomer, namely, N, N-dimethylacrylamide (DMA); 4-acryloylmorpholine (NAM); and N, N-diethylacrylamide (DEA). Although this is a significant addition to the copolymerization field, its utility is offset by the high reaction temperature (70 °C).
The advantage of the full disproportionation of Cu(I)Br/Me6TREN in water prior to monomer and initiator addition to synthesize both homo- and multi-block copolymers was previously exploited [36,37]. The most obvious advantage of these works is that the quantitative conversion was achieved in minutes, with final dispersity (Ð < 1.15). Interestingly, suppression of the hydrolysis reaction by employing lower reaction temperatures (i.e., ice/water bath) allowed for the copolymerization to proceed without appreciable loss of the polymerization control [37,38]. However, the synthesis of a multiblock copolymer (macromolecular alkyl bromide) was only chain extended from a NIPAM monomer, as this is the model monomer. In addition, the rate of ω-Br chain-end loss was pronounced in tertiary acrylamides relative to secondary acrylamides. Since N-hydroxyethyl acrylamide has been widely utilized in industrial and biological applications [39,40,41,42], the ability of poly N-hydroxyethyl acrylamide (PHEAA) as a macroinitiator to initiate other acrylamide monomers has recently attracted more attention. Narumi and his co-workers [43] employed the ATRP technique to produce block copolymers from PHEAA as a prepolymer. Although narrow MWDs were achieved, they were limited to just diblock. In addition, the controlled polymerizations were conducted in ethanol/water mixtures at ambient temperature, which could prevent their use in applications where pure water as the solvent is required. Furthermore, the integrity of the diblock compositions and one-pot procedure were not reported.
Herein, the one-pot polymerization procedure was exploited to synthesize a well-defined triblock copolymer from PHEAA as a macromolecular model ABC, containing (A) 2-hydroxyethyl acrylamide (HEAA); (B) N-isopropylacrylamide (NIPAM); and (C) N, N-dimethylacrylamide (DMA) segments, utilizing Cu(0)-mediated RDRP. PDMA was also employed as a macroinitiator to initiate another sequence of triblock copolymers (CBA). Pure water was utilized as the solvent to produce the different block sequences below ambient temperature. The copolymerizations proceeded in a controlled manner (Ð = 1.10), and full conversions were obtained within ~1 h, as evidenced by 1H NMR and SEC analyses. Likewise, prior disproportionation in situ was also exploited toward the ultrafast one-pot synthesis of sequence-controlled triblock (ABC model) copolymers of different acrylamides, allowing successive chain extensions and end-group fidelity. In addition, the undesirable side reaction was investigated.

2. Materials and Methods

2.1. Materials

HEAA (97%), NIPAM (97%) and DMA (99%) were obtained from Sigma-Aldrich (St. Louis, MO, USA). Prior to use, the monomers were filtered by injection over a column filled with alumina to eliminate inhibitors. HPLC-grade H2O (VWR international, LLC, Radnor, PA, USA) was used as the solvent for all reaction processes.
The water-soluble initiator (WSI), namely, dihydroxypropane-3-oxy-(2-bromo-2-methylpropionyl), and tris [2-(dimethylamino) ethyl]amine (Me6TREN) were synthesized according to previously reported procedures [34]. Copper bromide (98%, Sigma-Aldrich) was repeatedly purified with acetic acid and ethanol.

2.2. Instruments and Analysis

The proton nuclear magnetic resonance spectrum was recorded on a JEOL NMR spectrometer (600 MHz), and the NMR solvent was D2O as a locking agent. The monomer conversions for the reaction of HEAA, NIPAM and DMA polymerization were obtained by evaluating the integrals of the peaks corresponding to the vinyl protons, as well as those attributed to the N-methylene signal, the isopropyl methine proton and the methyl signal of DMA, respectively.
Size-exclusion chromatography (SEC) was utilized in the Agilent N29812 system (Agilent, Santa Clara, CA, USA), using dimethylformamide (DMF) containing 5 mM NH4BF4 as the mobile phase at 45 °C. The system was equipped with ultraviolet, refractive index and viscometer detectors, as well as an autosampler, a PLgel 5 μm MIXED-D column (300 × 7.5 mm2) and a PLgel 5 μm Guard column (50 × 7.5 mm2). In addition, an Agilent 1260 infinity isocratic pump with a maximum pressure of 600 bar was used.
Commercially available narrow linear polymers (PMMA) in the range of 200 to 1.0 × 106 g mol−1 were used to calibrate the system of reactions. All samples were passed through a 0.45 μm polytetrafluoroethylene (PTFE) filter. All reactions were performed under a N2 blanket, and dissolved air in the solution was removed by bubbling for 15 min. Then, the removed air left the solution throughput point existing in the top of the glassware, and we employed the Schlenk technique.

2.3. Synthesis of ABC-Type Triblock Copolymer by Aqueous Cu(0)-Mediated Polymerization at 0.0 °C

The disproportionation reaction of CuBr in the presence of Me6TREN ligand was conducted as reported in the literature [30]. H2O (1 mL), dihydroxypropane-3-oxy-(2-bromo-2-methylpropionyl) (60 mg, 0.00025 mole) and acrylamide monomer (2.5 mmol) were added to a clean Schlenk tube, and the solution was deoxygenated by bubbling with nitrogen for 15 min. The deoxygenated initiator/monomer solution was then injected into the disproportionation reaction Schlenk tube, which had already been immersed in the ice/water bath. The polymerization Schlenk tube was then stirred at 0.0 °C for a certain period of time. A sample was taken for conversion analysis before the addition of deoxygenated aqueous solutions of the second and third monomers at full conversion of the first and second block. Samples of the reaction mixtures were removed by a degassed syringe for 1H NMR and SEC analysis.

3. Results and Discussion

Well-defined multiblock copolymers of different acrylamide derivatives were synthesized via SET-LRP in water. However, the multiblock copolymers were only initiated from PNIPAM, as this monomer is the model monomer [36,37]. To the best of our knowledge, only very few publications have reported the controlled diblock copolymerization of acrylamides initiated from PHEAA or PDMA via RDRP techniques. In addition, ABC-type triblock copolymers initiated from a PHEA-based macroinitiator have not yet been reported. Changing the sequence of blocks is particularly important due to the chemical structures and properties. Therefore, the ability of PHEAA and PDMA to initiate block copolymers was investigated.

3.1. Investigating the Potential for ABC-Type Triblock Copolymer Synthesis via Chain Extension of PHEAA or PDMA Macroinitiator

The synthesis of two different block sequences of model triblock copolymers P(HEAA)10-b-(NIPAM)10-b-(DMA)10 and P(DMA)10-b-(NIPAM)10-b-(HEAA)10, with each block comprising ten monomer units (DP 10), was targeted. The Cu(0)-mediated polymerization method was utilized in pure water at 0.0 °C, offering a new route to biological applications. The direct disproportionation of CuBr in the presence of Me6TREN was exploited to generate Cu(0) and CuBr2 prior to iterative Cu(0)-mediated polymerization processes, as presented in Scheme 1.
After full disproportionation of CuBr, the first aliquot of monomer was allowed to polymerize under nitrogen protection. Interestingly, the full monomer conversion of the first block was attained in 20 min according to the total disappearance of vinyl groups confirmed by 1H NMR analysis. The molecular weight distributions of the first block were extremely narrow, with dispersity (Ð = 1.09) (Table 1, Figure 1), and no issues with the purity of both the initiator and ligand to conduct the polymerization were found.
In the direction of our target and in order to avoid any side reaction to the end chains of the first block during NMR and SEC measurements, we repeated the polymerization, and the second aliquot of the NIPAM monomer was directly added after 20 min. As hoped, 1H NMR analysis confirmed that the quantitative conversion of the second block was attained in 40 min, and SEC showed successful chain extension with dispersity (Ð = 1.08). The fast, controlled and quantitative conversion of the first and second block without the need for a purification step encouraged us to proceed with copolymerization. Importantly, prior to proceeding further, every chain extension cycle was precisely identified in terms of time and conversion. Consequently, the triblock copolymer was obtained via iterative chain extension in a very short reaction time (~1.5 h), negating the purification between monomer additions.
Quantitative conversions were confirmed by the 1H NMR analysis of each block according to the integration of the vinyl protons (~6.50–5.70 ppm) of the monomer with the N-methylene signal of HEAA (NH(–CH2–) (~3.3 ppm)), in the case of the second block, with the isopropyl protons of NIPAM (-CH (CH3)2) (~3.50–3.90 ppm) and with the methyl signal of DMA (N(CH3)2 (~3.0 ppm)) (Figure 2). Moreover, the molecular weight distributions of the triblock copolymers remained extremely narrow throughout the monomer additions, with a final dispersity of <1.09. These data demonstrate the level of sequence control obtained in one pot of such a new ABC-type triblock copolymer, reflecting the robustness and versatility of this method.
In line with the PHEAA macroinitiator investigation, DMA was also chosen to initiate the triblock copolymer (Scheme 2). Similarly, fast, controlled polymerization was achieved (less than 1 h) (Table 2), while the molecular weight distributions were narrow over block copolymerization. Moreover, 1H NMR and SEC confirmed the successive chain extensions, and narrow dispersities were retained throughout (final Ð = 1.08) (Figure 3 and Figure 4), implying the potential for precise control over discrete monomer sequences within the final polymer composition. In terms of the scale-up of this reaction, it is worth conducting more studies with the consideration of optimizing the reaction conditions in future research projects.

3.2. Investigating the Potential for Utilizing Halogen Exchange to Improve Resistance against Hydrolysis

In a previous work [36], hydrolysis was found in a polyacrylamide polymerized at 0.0 °C using 600 MHz NMR, and it was concluded that the hydrolysis of bromine is the predominant termination reaction. In addition, the hydrolysis reaction was found to be more obvious subsequent to full conversion, but lower reaction temperatures suppress the rate. It is well known that the alkyl halides (R-X) reactivities follow the order I > Br > Cl. Knowledge of the R-X reactivity can be exploited for the selection of appropriate initiators, particularly for the efficient synthesis of block copolymers [44]. To stabilize the carbon–halogen bond against facial hydrolysis, the bromo-based initiator/CuBr system was replaced with a chloro-based initiator/CuCl (Scheme 3). The method’s description is similar to that in Section 2.3.
The disproportionation of CuCl into Cu(0) and [Cu(Me6TREN)]Cl2 needed to be examined prior to using CuCl as a catalyst (Scheme 4). Disproportionation was monitored by UV–vis absorbance (λmax ~900 nm) corresponding to the in situ-generated [Cu(Me6TREN)]Cl2 complex and was recorded to be ~100% (Figure 5).
Utilizing the same ratio [chloro-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4 and employing NIPAM as a model acrylamide in H2O at 0.0 °C resulted in a slower polymerization rate (100% conversion, within 60 min not 11 min) [38] in comparison with the bromo-initiator/CuBr and broad MWDs (Ð = 1.60) (Figure 6 and Figure 7, blue trace). The slow polymerization rate was expected, as the R-Cl was less active than R-Br, but the broad dispersity could be attributed to either an inefficient initiation step or/and the stronger Cu(II)–Cl2 bond.
To circumvent this, the halide exchange process was investigated. Halogen exchange during ATRP using mixed halide initiation systems, (R-X/Cu-Y (X, Y = Br, Cl), has been proven to improve the control of ATRP [45,46,47,48,49]. Therefore, the bromo-initiator/CuCl system was employed under the same conditions and procedure that were previously used (Scheme 5). A fast polymerization rate was observed (100% conversion within 22 min) due to the efficient initiation (R-Br); however, the dispersity was still uncontrolled (Figure 7, red trace). Interestingly, when the catalytic ratio was increased to [CuCl]:[Me6TREN] = [1]:[1], the polymerization proceeded in a controlled manner (Ð = 1.15) (Figure 7, green trace). However, the monomer conversion was 80%, reflecting the slow polymerization rate. There is a general need to optimize the conditions in the future; nonetheless, these investigations were intended as a feasibility study, and, thus, it is satisfactory that we were able to utilize a mixed halide initiation system of aqueous Cu(0)-mediated polymerization.

4. Conclusions

The synthesis of thermosensitive homo-polymers and multiblock copolymers in aqueous media at or below ambient temperature has always been a challenge with respect to the control of the polymerization. This work proposed a simple and highly efficient route to synthesize a well-defined triblock copolymer, model ABC, initiated from PHEAA or utilizing Cu(0)-mediated RDRP in pure water at or below room temperature. A PDMA-macroinitiator was employed to synthesize another triblock copolymer in a similar manner but with a different sequence of monomer addition. The ultrafast one-pot synthesis of sequence-controlled triblock copolymers via iterative sequential monomer addition after full conversion, without any purification steps throughout the monomer additions, was exploited. The successive chain extensions of the block copolymers were demonstrated by SEC analysis, and the compositions of triblock copolymers were confirmed via 1H NMR. The robust process showed its suitability for the generation of ABC-type block copolymers with different block sequences from different acrylamides. Moreover, the potential for monomer–copper interactions, which could influence the living nature of the system, was investigated. The bromo-based initiator/CuBr system was switched with a chloro-based initiator/CuCl in order to stabilize the carbon–halogen bond. Finally, we may propose this approach as a potential method for industrial applications that specifically require pure water as the solvent at or below ambient temperature.

Author Contributions

Data curation, F.M.A., A.-H.I.M. and H.F.; formal analysis, F.M.A.; funding acquisition, O.Y.A.; methodology, F.M.A., A.-H.I.M. and H.F.; resources, O.Y.A.; writing—original draft, F.M.A.; writing—review and editing, O.Y.A., A.-H.I.M. and H.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at King Saud University through research group no. (RGP-133).

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Saud University for funding this work through research group no. (RGP-133). They also gratefully thank King Abdulaziz City for Science and Technology (KACST) for technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, Y.H.; Shiu, C.C.; Chen, T.L.; Chen, H.L.; Tsai, J.C. Solubilization behavior of homopolymer in its blend with the block copolymer displaying the feature of lower critical ordering transition. Polymers 2021, 13, 3415. [Google Scholar] [CrossRef] [PubMed]
  2. Manouras, T.; Platania, V.; Georgopoulou, A.; Chatzinikolaidou, M.; Vamvakaki, M. Responsive quaternized PDMAEMA copolymers with antimicrobial action. Polymers 2021, 13, 3051. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, Y.F.; Hong, J.W.; Chang, J.H.; Junisu, B.A.; Sun, Y.S. Influence of osmotic pressure on nanostructures in thin films of a weakly-segregated block copolymer and its blends with a homopolymer. Polymers 2021, 13, 2480. [Google Scholar] [CrossRef] [PubMed]
  4. Wever, D.A.Z.; Picchioni, F.; Broekhuis, A.A. Polymers for enhanced oil recovery: A paradigm for structure-property relationship in aqueous aolution. Prog. Polym. Sci. 2011, 36, 1558. [Google Scholar] [CrossRef]
  5. Xiong, B.; Loss, R.D.; Shields, D.; Pawlik, T.; Hochreiter, R.; Zydney, A.L.; Kumar, M. Polyacrylamide degradation and its implications in environmental systems. NPJ Clean Water 2018, 1, 17. [Google Scholar] [CrossRef]
  6. Corrigan, N.; Jung, K.; Moad, G.; Hawker, C.J.; Matyjaszewski, K.; Boyer, C. Reversible-deactivation radical polymerization (controlled/living radical polymerization): From discovery to materials design and applications. Prog. Polym. Sci. 2020, 111, 101311. [Google Scholar] [CrossRef]
  7. Nicolas, J.; Guillaneuf, Y.; Lefay, C.; Bertin, D.; Gigmes, D.; Charleux, B. Nitroxide-mediated polymerization. Prog. Polym. Sci. 2013, 38, 63. [Google Scholar] [CrossRef]
  8. Chiefari, J.; Chong, Y.K.; Ercole, F.; Krstina, J.; Jeffery, J.; Le, T.P.T.; Mayadunne, R.T.A.; Meijs, G.F.; Moad, C.L.; Moad, G.; et al. Living free-radical polymerization by reversible addition-fragmentation chain transfer: The RAFT process. Macromolecules 1998, 31, 5559. [Google Scholar] [CrossRef]
  9. Moad, G.; Rizzardo, E.; Thang, S.H. Living radical polymerization by the RAFT process. Aust. J. Chem. 2009, 62, 1402. [Google Scholar] [CrossRef]
  10. Parkatzidis, K.; Wang, H.S.; Truong, N.P.; Anastasaki, A. Recent developments and future challenges in controlled radical polymerization: A 2020 update. Chem 2020, 6, 1575. [Google Scholar] [CrossRef]
  11. Wang, J.S.; Matyjaszewski, K.J. Controlled/“living” radical polymerization. atom transfer radical polymerization in the presence of transition-metal complexes. J. Am. Chem. Soc. 1995, 117, 5614. [Google Scholar] [CrossRef]
  12. Matyjaszewski, K. Atom Transfer Radical Polymerization (ATRP): Current status and future perspectives. Macromolecules 2012, 45, 4015. [Google Scholar] [CrossRef]
  13. Cherupurakal, N.; Mozumder, M.S.; Mourad, A.-H.I.; Lalwani, S. Recent advances in superhydrophobic polymers for antireflective self-cleaning solar panels. Renew. Sust. Energ. Rev. 2021, 151, 111538. [Google Scholar] [CrossRef]
  14. Abbas, M.; Hachemaoui, A.; Yahiaoui, A.; Mourad, A.-H.I.; Belfedal, A.; Cherupurakal, N. Chemical synthesis of nanocomposites via in-situ polymerization of aniline and iodoaniline using exchanged montmorillonite. Polym. Polym. Compos. 2020, 29, 982. [Google Scholar] [CrossRef]
  15. Percec, V.; Guliashvili, T.; Ladislaw, J.S.; Wistrand, A.; Stjerndahl, A.; Sienkowska, M.J.; Monteiro, M.J.; Sahoo, S. Ultrafast synthesis of ultrahigh molar mass polymers by metal-catalyzed living radical polymerization of acrylates, methacrylates, and vinyl chloride mediated by SET at 25 °C. J. Am. Chem. Soc. 2006, 128, 14156. [Google Scholar] [CrossRef]
  16. Hawker, C.J.; Bosman, A.W.; Harth, E. New polymer synthesis by nitroxide mediated living radical polymerizations. Chem. Rev. 2001, 101, 3661. [Google Scholar] [CrossRef]
  17. Kwak, Y.; Magenau, A.; Matyjaszewski, K. ARGET ATRP of Methyl Acrylate with Inexpensive Ligands and ppm Concentrations of Catalyst. Macromolecules 2011, 44, 811–819. [Google Scholar] [CrossRef]
  18. Lligadas, G.; Percec, V. Ultrafast SET-LRP of methyl acrylate at 25 °C in alcohols. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 2745. [Google Scholar] [CrossRef]
  19. Nguyen, N.H.; Percec, V. Disproportionating versus nondisproportionating solvent effect in the SET-LRP of methyl acrylate during catalysis with nonactivated and activated Cu(0) wire. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 4756. [Google Scholar] [CrossRef]
  20. Iddon, P.D.; Robinson, K.L.; Armes, S.P. Polymerization of sodium 4-styrenesulfonate via atom transfer radical polymerization in protic media. Polymer 2004, 45, 759. [Google Scholar] [CrossRef]
  21. Rademacher, J.T.; Baum, M.; Pallack, M.E.; Brittain, W.J.; Simonsick, W.J. Atom transfer radical polymerization of N,N-Dimethylacrylamide. Macromolecules 2000, 33, 284. [Google Scholar] [CrossRef]
  22. Teodorescu, M.; Matyjaszewski, K. Atom transfer radical polymerization of (Meth)acrylamides. Macromolecules 1999, 32, 4826. [Google Scholar] [CrossRef]
  23. Teodorescu, M.; Matyjaszewski, K. Controlled polymerization of (Meth)crylamides by atom transfer radical polymerization. Macromol. Rapid Commun. 2000, 21, 190. [Google Scholar] [CrossRef]
  24. Ye, J.; Narain, R. Water-assisted atom transfer radical polymerization of n-isopropylacrylamide: Nature of solvent and temperature. J. Phys. Chem. B 2009, 113, 676. [Google Scholar] [CrossRef]
  25. Wever, D.A.Z.; Raffa, P.; Picchioni, F.; Broekhuis, A.A. Acrylamide homopolymers and acrylamide–N-isopropylacrylamide block copolymers by atomic transfer radical polymerization in water. Macromolecules 2012, 45, 4040. [Google Scholar] [CrossRef] [Green Version]
  26. Lligadas, G.; Grama, S.; Percec, V. Single-electron transfer living radical polymerization platform to practice, develop, and invent. Biomacromolecules 2017, 18, 2981. [Google Scholar] [CrossRef]
  27. Ding, L.; Li, J.; Jiang, R.Y.; Wang, L.; Song, W.; Zhu, L. Cu(0) wire-mediated single-electron transfer-living radical polymerization of oligo(ethylene oxide) methyl ether acrylate by selecting the optimal reaction conditions. Chin. J. Polym. Sci. 2019, 37, 1130. [Google Scholar] [CrossRef]
  28. Jiang, X.; Fleischmann, S.; Nguyen, N.H.; Rosen, B.M.; Percec, V. Cooperative and synergistic solvent effects in SET-LRP of MA. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 5591. [Google Scholar] [CrossRef]
  29. Boyer, C.; Derveaux, A.; Zetterlund, P.B.; Whittaker, M.R. Synthesis of multi-block copolymer stars using a simple iterative Cu(0)-mediated radical polymerization technique. Polym. Chem. 2012, 3, 117. [Google Scholar] [CrossRef]
  30. Duong, H.T.T.; Marquis, C.P.; Whittaker, M.; Davis, T.P.; Boyer, C. Acid degradable and biocompatible polymeric nanoparticles for the potential codelivery of therapeutic agents. Macromolecules 2011, 44, 8008. [Google Scholar] [CrossRef]
  31. Soeriyadi, A.H.; Boyer, C.; Nystrom, F.; Zetterlund, P.B.; Whittaker, M.R. High order multi-block copolymers via iterative Cu(0)-mediated radical polymerizations (SET-LRP): Towards biological precision. J. Am. Chem. Soc. 2011, 133, 11128. [Google Scholar] [CrossRef]
  32. Anastasaki, A.; Waldron, C.; Wilson, P.; Boyer, C.; Zetterlund, P.B.; Whittaker, M.R.; Haddleton, D.M. High molecular weight block copolymers by sequential monomer addition via Cu(0)-mediated living radical polymerization (SET-LRP): An optimized approach. ACS Macro Lett. 2013, 2, 896. [Google Scholar] [CrossRef]
  33. Zhang, Q.; Anastasaki, A.; Li, G.Z.; Haddleton, A.J.; Wilson, P.; Haddleton, D.M. Multiblock sequence-controlled glycopolymers via Cu (0)-LRP following efficient thiol–halogen, thiol–epoxy and CuAAC reactions. Polym. Chem. 2014, 5, 3876. [Google Scholar] [CrossRef]
  34. Gody, G.; Maschmeyer, T.; Zetterlund, P.B.; Perrier, S. Rapid and quantitative one-pot synthesis of sequence-controlled polymers by radical polymerization. Nat. Commun. 2013, 4, 2505. [Google Scholar] [CrossRef]
  35. Gody, G.; Maschmeyer, T.; Zetterlund, P.B.; Perrier, S. Exploitation of the degenerative transfer mechanism in RAFT polymerization for synthesis of polymer of high livingness at full monomer conversion. Macromolecules 2014, 47, 639. [Google Scholar] [CrossRef]
  36. Zhang, Q.; Wilson, P.; Li, Z.; McHale, R.; Godfrey, J.; Anastasaki, A.; Waldron, C.; Haddleton, D.M. Aqueous copper-mediated living polymerization: Exploiting rapid disproportionation of CuBr with Me6TREN. J. Am. Chem. Soc. 2013, 135, 7355. [Google Scholar] [CrossRef]
  37. Alsubaie, F.; Alothman, O.; Alshammari, B.; Fouad, H. Facile synthesis of hydrophilic homo-polyacrylamides via Cu(0)-mediated reversible deactivation radical polymerization. Polymers 2021, 13, 1947. [Google Scholar] [CrossRef]
  38. Alsubaie, F.; Anastasaki, A.; Wilson, P.; Haddleton, D.M. Sequence-controlled multi-block copolymerization of acrylamides via aqueous SET-LRP at 0 °C. Polym. Chem. 2015, 6, 406. [Google Scholar] [CrossRef]
  39. Simò-Alfonso, E.; Gelfi, C.; Lucisano, M.; Righetti, P.G. Performance of a series of novel n-substituted acrylamides in capillary electrophoresis of DNA fragments. J. Chromatogr. A 1996, 756, 255. [Google Scholar] [CrossRef] [Green Version]
  40. Albarghouthi, M.N.; Buchholz, B.A.; Doherty, E.A.S.; Bogdan, F.M.; Zhou, H.; Barron, A.E. Impact of polymer hydrophobicity on the properties and performance of DNA sequencing matrices for capillary electrophoresis. Electrophoresis 2002, 23, 1429. [Google Scholar] [CrossRef]
  41. Albarghouthi, M.N.; Stein, T.M.; Barron, A.E. Poly-N-hydroxyethylacrylamide as a novel, adsorbed coating for protein separation by capillary electrophoresis. Electrophoresis 2003, 24, 1166. [Google Scholar] [CrossRef]
  42. Tanigami, T.; Iwata, H.; Mori, T. Ion-exchange membrane based on poly(styrene sulfonic acid-co-n-(2-hydroxyethyl) acrylamide). J. Appl. Polym. Sci. 2007, 103, 2788. [Google Scholar] [CrossRef]
  43. Narumi, A.; Chen, Y.; Sone, M.; Fuchise, K.; Sakai, R.; Satoh, T.; Duan, Q.; Kawaguchi, S.; Kakuchi, T. Poly(N-hydroxyethylacrylamide) prepared by atom transfer radical polymerization as a nonionic, water-soluble, and hydrolysis-resistant polymer and/or segment of block copolymer with a well-defined molecular weight. Macromol. Chem. Phys. 2009, 210, 349. [Google Scholar] [CrossRef]
  44. Lad, J.; Harrisson, S.; Mantovani, G.; Haddleton, D.M. Copper mediated living radical polymerisation: Interactions between monomer and catalyst. Dalton Trans. 2003, 21, 4175. [Google Scholar] [CrossRef]
  45. Kamigaito, M.; Ando, T.; Sawamoto, M. Metal-catalyzed living radical polymerization. Chem. Rev. 2001, 101, 3689. [Google Scholar] [CrossRef]
  46. Ishikawa, T.; Takenaka, A.; Kikuchi, M.; Kobayashi, M.; Takahara, A. Effective Addition of Organic Chloride Salts on Atom Transfer Radical Polymerization in Fluoroalcohols. Macromolecules 2013, 46, 9189–9196. [Google Scholar] [CrossRef]
  47. Matyjaszewski, K.; Shipp, D.A.; Wang, J.L.; Grimaud, T.; Patten, T. Utilizing halide exchange to improve control of atom transfer radical polymerization. Macromolecules 1998, 31, 6836. [Google Scholar] [CrossRef]
  48. Tsarevsky, N.V.; Pintauer, T.; Matyjaszewski, K. Deactivation efficiency and degree of control over polymerization in ATRP in protic solvents. Macromolecules 2004, 37, 9768. [Google Scholar] [CrossRef]
  49. Peng, C.; Kong, J.; Seeliger, F.; Matyjaszewski, K. Mechanism of halogen exchange in ATRP. Macromolecules 2011, 44, 7546. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of triblock copolymers composed of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation. Conditions: [WSI]:[monomer]:[CuBr]:[Me6TREN] = 1:10:0.4:0.4.
Scheme 1. Synthesis of triblock copolymers composed of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation. Conditions: [WSI]:[monomer]:[CuBr]:[Me6TREN] = 1:10:0.4:0.4.
Polymers 14 00116 sch001
Figure 1. The evolution of molecular weights of successive block extensions of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization.
Figure 1. The evolution of molecular weights of successive block extensions of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization.
Polymers 14 00116 g001
Figure 2. 1H nuclear magnetic resonance (NMR) spectra (D2O) for triblock copolymers of HEAA, NIPAM and DMA and by iterative Cu(0)-mediated polymerization.
Figure 2. 1H nuclear magnetic resonance (NMR) spectra (D2O) for triblock copolymers of HEAA, NIPAM and DMA and by iterative Cu(0)-mediated polymerization.
Polymers 14 00116 g002
Scheme 2. Synthesis of triblock copolymers composed of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation. Conditions: [WSI]:[monomer]:[CuBr]: [Me6TREN] = 1:10:0.4:0.4.
Scheme 2. Synthesis of triblock copolymers composed of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation. Conditions: [WSI]:[monomer]:[CuBr]: [Me6TREN] = 1:10:0.4:0.4.
Polymers 14 00116 sch002
Figure 3. The evolution of molecular weights of successive block extensions of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization.
Figure 3. The evolution of molecular weights of successive block extensions of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization.
Polymers 14 00116 g003
Figure 4. 1H nuclear magnetic resonance (NMR) spectra (D2O) for triblock copolymers of DMA, NIPAM and HEAA and by iterative Cu(0)-mediated polymerization.
Figure 4. 1H nuclear magnetic resonance (NMR) spectra (D2O) for triblock copolymers of DMA, NIPAM and HEAA and by iterative Cu(0)-mediated polymerization.
Polymers 14 00116 g004
Scheme 3. Synthesis of PNIPAM by Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation of CuCl/ Me6TREN. Conditions: [chloro-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Scheme 3. Synthesis of PNIPAM by Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation of CuCl/ Me6TREN. Conditions: [chloro-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Polymers 14 00116 sch003
Scheme 4. Disproportionation of CuCl in the presence of Me6TREN. Conditions: [CuCl]:[Me6TREN] = 1:1 in 2 mL H2O.
Scheme 4. Disproportionation of CuCl in the presence of Me6TREN. Conditions: [CuCl]:[Me6TREN] = 1:1 in 2 mL H2O.
Polymers 14 00116 sch004
Figure 5. UV–vis spectra of CuCl2 solution in the presence of Me6TREN in 2 mL of H2O. The red line represents the UV–vis spectrum of the disproportionation of CuCl (conditions: [CuCl]:[Me6TREN] = 1:1 in 2 mL H2O). The sample was diluted before analysis into degassed H2O.
Figure 5. UV–vis spectra of CuCl2 solution in the presence of Me6TREN in 2 mL of H2O. The red line represents the UV–vis spectrum of the disproportionation of CuCl (conditions: [CuCl]:[Me6TREN] = 1:1 in 2 mL H2O). The sample was diluted before analysis into degassed H2O.
Polymers 14 00116 g005
Figure 6. DMF SEC molecular weight distributions for PNIPAM via Cu(0)-mediated polymerization in H2O at 0.0 °C. [chloro-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Figure 6. DMF SEC molecular weight distributions for PNIPAM via Cu(0)-mediated polymerization in H2O at 0.0 °C. [chloro-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Polymers 14 00116 g006
Figure 7. DMF SEC molecular weight distributions for PNIPAM via Cu(0)-mediated polymerization in H2O at 0.0 °C (blue trace) [bromo-initiator]:[CuBr]:[Me6TREN] = 1:0.4:0.4. and (red trace) [bromo-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4. (green trace) [bromo-initiator]:[CuCl]:[Me6TREN] = 1:1:1.
Figure 7. DMF SEC molecular weight distributions for PNIPAM via Cu(0)-mediated polymerization in H2O at 0.0 °C (blue trace) [bromo-initiator]:[CuBr]:[Me6TREN] = 1:0.4:0.4. and (red trace) [bromo-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4. (green trace) [bromo-initiator]:[CuCl]:[Me6TREN] = 1:1:1.
Polymers 14 00116 g007
Scheme 5. Synthesis of PNIPAM by Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation of [Cu(Me6TREN)Cl]. Conditions: [bromo-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Scheme 5. Synthesis of PNIPAM by Cu(0)-mediated polymerization in H2O at 0.0 °C via prior disproportionation of [Cu(Me6TREN)Cl]. Conditions: [bromo-initiator]:[CuCl]:[Me6TREN] = 1:0.4:0.4.
Polymers 14 00116 sch005
Table 1. Summary of characteristics of triblock copolymers of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization.
Table 1. Summary of characteristics of triblock copolymers of HEAA, NIPAM and DMA by iterative Cu(0)-mediated polymerization.
Sequence DistributionConv. (%)Time Per Block (Min)Mn,thMn,SECÐ
Block 110020140046001.09
Block 210040250065001.08
Block 39840350097001.09
Table 2. Summary of characteristics of triblock copolymers of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization.
Table 2. Summary of characteristics of triblock copolymers of DMA, NIPAM and HEAA by iterative Cu(0)-mediated polymerization.
Sequence DistributionConv. (%)Time Per
Block (Min)
Mn,thMn,SECÐ
Block 11002120031001.08
Block 210025240051001.08
Block 310025350095001.08
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alsubaie, F.M.; Alothman, O.Y.; Fouad, H.; Mourad, A.-H.I. ABC-Type Triblock Copolyacrylamides via Copper-Mediated Reversible Deactivation Radical Polymerization. Polymers 2022, 14, 116. https://doi.org/10.3390/polym14010116

AMA Style

Alsubaie FM, Alothman OY, Fouad H, Mourad A-HI. ABC-Type Triblock Copolyacrylamides via Copper-Mediated Reversible Deactivation Radical Polymerization. Polymers. 2022; 14(1):116. https://doi.org/10.3390/polym14010116

Chicago/Turabian Style

Alsubaie, Fehaid M., Othman Y. Alothman, Hassan Fouad, and Abdel-Hamid I. Mourad. 2022. "ABC-Type Triblock Copolyacrylamides via Copper-Mediated Reversible Deactivation Radical Polymerization" Polymers 14, no. 1: 116. https://doi.org/10.3390/polym14010116

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