Next Article in Journal
Thermoplastic Starch (TPS) Films Added with Mucilage from Opuntia Ficus Indica: Mechanical, Microstructural and Thermal Characterization
Next Article in Special Issue
Low Ppm Atom Transfer Radical Polymerization in (Mini)Emulsion Systems
Previous Article in Journal
Chitosan-Functionalized Graphene Nanocomposite Films: Interfacial Interplay and Biological Activity
Previous Article in Special Issue
Ultrasound-Mediated Atom Transfer Radical Polymerization (ATRP)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Surface Modifications of Poly(Ether Ether Ketone) via Polymerization Methods—Current Status and Future Prospects

Department of Physical Chemistry, Faculty of Chemistry, Rzeszow University of Technology, Al. Powstańców Warszawy 6, 35-959 Rzeszów, Poland
*
Author to whom correspondence should be addressed.
Materials 2020, 13(4), 999; https://doi.org/10.3390/ma13040999
Submission received: 24 January 2020 / Revised: 14 February 2020 / Accepted: 20 February 2020 / Published: 23 February 2020

Abstract

:
Surface modification of poly(ether ether ketone) (PEEK) aimed at applying it as a bone implant material aroused the unflagging interest of the research community. In view of the development of implantology and the growing demand for new biomaterials, increasing biocompatibility and improving osseointegration are becoming the primary goals of PEEK surface modifications. The main aim of this review is to summarize the use of polymerization methods and various monomers applied for surface modification of PEEK to increase its bioactivity, which is a critical factor for successful applications of biomedical materials. In addition, the future directions of PEEK surface modifications are suggested, pointing to low-ppm surface-initiated atom transfer radical polymerization (SI-ATRP) as a method with unexplored capacity for flat surface modifications.

1. Introduction

Currently, the production of bone implants is limited only to metal materials (stainless steel, cobalt–chromium, titanium). However, in the production of personalized bone implants, there is an alternative synthetic polymer named poly(ether ether ketone) (PEEK) [1,2,3]. The chemical structure of PEEK can be defined as an alternating combination of aryl rings through ketone and ether groups; therefore, it belongs to the family of polyaryletherketone polymers. The Yang’s modulus of pure PEEK is close to the elastic modulus of the bones and is about 3.6 GPa, which is an unquestionable advantage of this material over ceramic or pure steel implants [4]. Furthermore, PEEK has high abrasion resistance, low friction coefficient, and low sensitivity to temperature change [5]. The glass transition temperature of amorphous PEEK is about 145 °C, and the melting temperature of pristine polymer is about 343 °C [6]. The synthetic process of obtaining PEEK material is based on the dialkylation of bisphenolate salts, which results in a chemically resistant, non-biodegradable product. In the case of the production of permanent implants, the lack of biodegradability is a desired feature and determines the usefulness of the material. Therefore, since the United States (US) Food and Drug Administration (FDA) certified PEEK as a suitable bone implant material in the 1990s, the application potential of the material in the orthopedic engineering is constantly increasing [7].
Implants made from poly(ether ether ketone) can be used in complicated reconstructions, even in areas with difficult access [8]. The fastening of an implant to surrounding tissues can be performed by using standard screws chosen by a surgeon. Moreover, PEEK implants are transparent to X-rays, which is a very beneficial factor in the analysis of treatment progress by computed tomography or nuclear magnetic resonance in the postoperative recovery phase. Additionally, PEEK can be sterilized with popular methods such as the use of steam under pressure, gamma radiation, or ethylene oxide.
Typically, the manufacturing of an implant made from PEEK is based on structuring it by milling from the block of material. Due to its rather high hydrophobicity, it shows good bioinert properties, but not sufficient ingrowth in the surrounding tissue. This is the reason why it is mainly used for injuries where no shear forces will act on the surgery site, for example, with screw-fixated vertebral bodies in the spine. Therefore, there is no real adhesion between the implant and surrounding tissue, as it is only tightly surrounded by the freshly formed bone tissue [9].
The above-mentioned problem with the adhesion of living tissue with an implant made of PEEK drives scientists to search for effective methods and chemicals to modify the surface in order to improve its bioactivity. Among the methods for surface modification of the implant material, a number of mechanical, chemical, or physical methods can be distinguished. The high application potential of PEEK as a bone implant material intensified the research over the last decade, where various techniques of PEEK surface modification were described [10,11,12,13,14].
In modern times, surface functionalization is a driving tool for the development of new biomaterials with improved biological properties. Surface functionalization of inorganic and organic materials (e.g., PEEK) is mainly based on the absorption of substances, reaction with small molecules, or just grafting of polymer chains on the substrate surface, as presented in this review paper.
The application of controlled radical polymerization for surface modification of materials is based on the formation of covalent bonds between the modified surface and synthesized polymer brushes. In general, there are three synthetic approaches that enable the organic polymer layer to be permanently bonded to the modified material. The first approach called “grafting from” uses a previously functionalized surface with a covalently attached polymerization initiator containing a halogenated atom [15]. In the second approach, “grafting onto”, functional groups on the surface of the functionalized material are compatible with the chain-end groups of polymer brushes, resulting in the covalent attachment of the polymer chains [16]. In turn, “grafting through” is a two-stage process. At first, a polymerized monomer unit is attached to the modified surface; then, the created macromonomer is subjected to (co)polymerization with single monomer units [17]. The advantages and disadvantages of each of the mentioned synthetic approaches were pointed out in our previous paper about the creations of hybrid nanomaterials via surface-initiated atom transfer radical polymerization (SI-ATRP) [18,19].
The main objective of this review is to summarize the role of the polymerization process in the modification of the PEEK surface. The article presents the effects of using the traditional photo-induced polymerization process and classical surface-initiated atom transfer radical polymerization (SI-ATRP) to enhance the bioactivity of the PEEK surface, additionally indicating new possibilities offered by the application of the low-ppm or metal-free ATRP approach.

2. Photo-Induced Polymerization for Surface Modification of PEEK

Low cost, no complicated equipment, easy application in industry, and mild reaction conditions are just a few of the advantages that make photo-induced polymerization outweigh plasma-induced or radiation-induced polymerization. In 2009, Kyomoto and Ishihara [20] put forward hypotheses that the diphenylketone group present in the PEEK backbone acts as a photoinitiator in polymerization reactions, similar to benzophenone previously described in the literature. Self-induced grafting of polymer chains is powered by the formation of semibenzopinacol-containing radicals of benzophenone units in the PEEK molecular structure, as shown in Figure 1. The successful polymerization confirmed the assumptions and paved the way for the use of ultraviolet (UV)-initiated polymerization in PEEK surface treatment. The list of PEEK surface modifications carried out with the usage of photo-initiated polymerization is summarized in Table 1.
The first papers (2009, 2010, and 2013) on the successful self-initiated surface graft polymerization of 2-methacryloyloxyethyl phosphorylcholine (MPC) on PEEK by photoirradiation were published by Kyomoto [10,20,22]. A variety of performed tests revealed significant improvements in the water wettability, frictional properties, and wear resistance of modified PEEK surfaces. Moreover, the impact of the monomer concentration and UV exposure time on the extent of the grafted PMPC layer was investigated. The result showed that the grafting density and thickness of the polymer layer can be controlled by adjusting the concentration of the monomer and the length of irradiation time [10]. Furthermore, as an application to improving the durability of artificial hips, the nanometer-scale photo-induced grafting of poly(2-methacryloyloxyethyl phosphorylcholine) (PMPC) onto PEEK and carbon-fiber-reinforced PEEK orthopedic bearing surfaces and interfaces was demonstrated.
The continuation of research conducted by Ishihara et al. resulted in the publication of an article presenting the biological performance of the PMPC-grafted PEEK and its comparison with the unmodified PEEK surface [24,25]. It was concluded that an increase in the thickness of the PMPC layer caused a significant decrease in the amount of fibrinogen adsorption in comparison with the results obtained in the case of unmodified PEEK. Moreover, bacterial adhesion was reduced dramatically on the PMPC-grafted PEEK, which means that the antithrombogenicity of potential bone implant material was improved.
A similar concept presenting a PEEK surface modified with methacryloyl-functionalized MPC polymers (MAMA) was described by Kawasaki in 2014 [23]. At first, poly(MPC-co-AEMA) (PMA) was synthesized by conventional free-radical polymerization using 2,2′-azobisisobutyronitrile as an initiator; then, in contrast to previous studies, 1-butanol solution containing the PMAMA was dropped onto the PEEK plates, and they were immediately exposed to UV light. Various protic solvents (e.g., water, methanol, ethanol) for making polymer solutions were tried, but they did not spread well on PEEK. Then, fibrinogen adsorption patterns of a polymer surface after contact with human plasma were observed using scanning electron microscopy (SEM) pictures. Results of the blood compatibility test indicated that an appropriate layer of PMAMA was immobilized on PEEK.
Subsequent studies were aimed at optimizing polymerization conditions and presented high-efficiency preparation of the PMPC layer on PEEK in the presence of inorganic salt additive [26]. It turned out that the addition of inorganic salt (LiCl, NaCl, KCl) increases the polymerization rate and it may play a significant role by concentrating MPC in the solution through ionic hydration.
The next step in the development of PEEK surface modification using photopolymerization was the grafting of polyelectrolyte brushes (2-(methacryloyloxy) ethyltrimethylammonium chloride (MTAC), 3-sulfopropyl methacrylate potassium salt (SPMK)), providing excellent antifouling properties and a low friction coefficient under wet condition [28]. Such modification opened up a variety of PEEK surface design possibilities for novel applications in harsh working and water-lubricated conditions. In contrast to the modifications previously described in the literature, the polymerization was carried out at room temperature, providing even milder reaction conditions, which are undoubtedly an advantage of the surface modification approach used. Another benefit of carrying out room-temperature polymerization emphasized by Yameen et al. [27] is the avoidance of monomer polymerization or its crosslinking. In the work mentioned above, a UV-light-mediated single-step procedure was successfully used for a variety of monomers and polymers with completely different chemical natures (e.g., n-butyl acrylate (BA), styrene (St), vinyl phosphonic acid (VPA), oligo(ethylene glycol) methacrylate (MeOEGMA), and acrylic acid (AA)). Then, the surface modification of PEEK was effectively evaluated by atomic force microscopy (AFM) measurement (Figure 2).
The surface characteristics of natural cartilage are often an inspiration to search for new bone implant materials with improved biological properties. The brush-like structures present on the surface of cartilage are aimed at supporting stress and impact, as well as adsorbing synovial fluid, which has an influence on properties such as wettability, lubrication, and friction. Unmodified PEEK is a bioinert but hydrophobic material, which is an important limitation in its use for the production of artificial joints. Therefore, a change in wettability leading to increased surface bioactivity is the main purpose of conducted modifications. In addition, the formation of a hydrophilic layer on the surface of the PEEK leads to improved wear resistance for aqueous lubrication. Referring to those facts, AA-enriched PEEK surfaces were prepared via UV-initiated polymerization [29]. As an effect, a material with a lower contact angle value was obtained, which clearly indicates an increase in surface hydrophilicity.
Recently, bioactive PEEK, surface-modified simultaneously with electrospun titanium dioxide (TiO2) and methacrylated hyaluronic acid (MeHA) (PEEK-MeHA-TiO2) by ultraviolet irradiation, was created [30]. Similarly to the previously described modifications, the roughness and hydrophobicity of PEEK were greatly changed. With regard to properties of titanium dioxide (TiO2), which is especially well suited for functionalizing the material interface of medical devices, the authors presented conception based on light-inducible electrospun TiO2 that imitated the fiber topology of cortical bone, and they used it to fabricate a composite with PEEK (Figure 3). In turn, HA forms a polymer chain backbone with which mesenchymal stem cells (MSCs) can interact via receptors expressed on the cell surface; therefore, it was used to cover the external part of the created composite. As a result, the modified surface showed increased adhesion, proliferation, and osteogenic differentiation capacity.
More recently Zheng et al. [31] applied single-step ultraviolet-initiated graft polymerization for introduction of sulfonate groups onto the PEEK surface. The use of vinylsulfonic acid sodium (VSA) and vinylphosphonic acid (VPA) to improve the bioactivity of the treated PEEK surface opened a new modification path intensively applied in the latest research [31,32,33]. In vitro studies using the MC3T3-E1 line of osteoblasts showed that initial adhesion, spreading, proliferation, and osteogenic differentiation were greatly improved after PEEK surface modification. As proven in earlier studies, in comparison to hydrophobic surfaces, hydrophilic surfaces were more favorable for cell adhesion and proliferation [34]. Therefore, the surface-sulfonated PEEK implant material may improve bone–implant integration in comparison with pristine PEEK.
The main conclusion confirming the validity of VPA modification to improve the bioactivity of PEEK was the result of tests carried out with an animal model (Figure 4). Higher bone–implant contact and a smaller space between the bone tissue and the surface-phosphorylated PEEK implants were observed. Interestingly, it was first shown by in vivo tests that PVPA had excellent bone tissue compatibility.

3. Surface Modification of PEEK via ATRP Method

ATRP is one of the most versatile techniques of reversible-deactivation radical polymerization (RDRP), which enables the synthesis of well-defined macromolecules with predefined structure, predetermined properties, and chain-end functionality [35,36,37,38]. In this case, the reaction mechanism is based on reversible activation and deactivation of propagating chains and catalyst complex molecules [39]. An equilibrium between active and dormant species is hardly shifted on the deactivation side, to maintain low active radical concentration, causing the equable growth of polymer chains [40,41]. In contrast to conventional free-radical polymerization, ATRP allows the extension of the lifetime of growing chains from seconds to hours or days [42]. To ensure the controlled nature of the process, the initiation stage is significantly faster than the propagation stage, which enables the simultaneous growth of all polymer chains. Moreover, control of the reaction is provided by an extension of the lifetime of propagating chains, which permits the synthesis of polymers with different topologies as well-defined block copolymers [43,44,45,46].
The application of ATRP in the synthesis of polymer brushes grafted from the flat surfaces creates the possibility of obtaining polymers characterized by narrow molecular weight distribution (MWD) and controlled architecture [47,48,49,50,51]. Additionally, SI-ATRP allows the growing functionality of polymer brushes from different surfaces [41,52,53,54,55] with a high degree of synthetic flexibility, enabling the introduction of a variety of functional groups [18,56,57]. Despite the high potential of SI-ATRP for the enhancement of PEEK surface properties, only limited research was conducted in this area. Nevertheless, this chapter summarizes the successful use of SI-ATRP for the modification of the bone–implant material. Surface modifications of poly(ether ether ketone) via the ATRP method are presented in Table 2.
For the first time, the unexplored capacity of SI-ATRP as a versatile methodology for controlling the surface properties of PEEK was presented by Yameen et al. [58]. Successful grafting of three different monomers, namely, potassium 3-(methacryloyloxy) propane-1-sulfonate (MPS), MeOEGMA, and N-isopropylacrylamide (NIPAM), was carried out according to Figure 5.
In the case of using the SI-ATRP approach, it is necessary to attach an initiator in the first step. For this purpose, the reduction of PEEK surface carbonyl groups to the hydroxy groups was performed, and then 2-bromoisobutyryl groups were covalently anchored at the PEEK surface as the polymerization initiator. Subsequently, for each of the grafted monomers, a different catalytic system was applied, i.e., 25,000 ppm of CuIICl2/2,2′-bipyridine (BPY) for MPS around, 2270 ppm of CuIIBr2/BPY for MeOEGA, and 10,000 ppm of CuIIBr2/N,N,N′,N″,N″-pentamethyldiethylenetriamine (PMDETA) for NIPAM. Successful PEEK surface modification was monitored by ATR infrared (IR) spectroscopy. The variety of properties of used monomers allowed examining the electrostatic interaction of PEEK-g-PMPS and rhodamine 6G, the antifouling evaluation performed for PEEK-g-PMeOEGMA, and the thermally responsive switching between hydrophilicity and hydrophobicity of PEEK-g-PNIPAM. From the implantology point of view, the most interesting result was presented for the PEEK-g-PMeOEGMA material. Surface modification was designed by the inspiration of materials coated with polyethylene glycol (PEG) or oligo(ethylene glycol), which prevent bioadhesion of proteins or living cells and bacteria [60]. The antifouling properties were demonstrated by exposing PEEK-g-PMeOEGMA to a culture medium of Escherichia coli bacteria. Interestingly, a significant number of bacteria attached to the surface of the pristine PEEK were observed, but there were no bacteria adhered to the PEEK-g-PMeOEGMA surface [58].
A similar approach with a significantly higher amount of catalytic complex (56,977 ppm of CuIICl2/BPY) was applied for the synthesis of PPEGMA on the PEEK surface via SI-ATRP. The concept of hydrophilization of bone–implant material using PPEGMA has a double benefit. Firstly, as in the examples described earlier, the hydrophilic layer causes the material to exhibit strongly limited protein adsorption. In addition, it allows for the deposition of a metallic layer, which, in the case of inert polymeric material, cannot be electrolessly metallized without proper surface treatment. In this case, one functionalization enables two different applications: electronics and medical [11].

4. Future Prospects and Conclusion

It is well known that surface characteristics and physico-chemical properties are key elements in modulating cell and tissue interactions with bone implant material. PEEK surface modifications can contribute to making it more favorable for human cell adhesion, growth, and differentiation, leading to enhanced implant–tissue integration. Over the past decade, two strategies (UV-initiated polymerization and traditional (high ppm) SI-ATRP) were mainly employed to improve PEEK surface biocompatibility.
A definite advantage of using the UV-initiated polymerization method for PEEK surface improvement is the one-step modification path. The lack of the need for additional actions related to the introduction of the initiator onto the modified surface makes this method simple and useful. Nevertheless, despite the successful application of UV-initiated polymerization, it still remains one of the methods of conventional radical polymerization not allowing for precise control of the architecture of the synthesized polymer. Thus, some limitations arise in the synthesis of, e.g., block (co)polymer coatings or gradient layers with extended topography. Therefore, the development of the UV-initiated technique in the context of flat surface modification is partially limited.
Despite numerous studies focusing on improving the biological properties of PEEK surfaces, there are still no published papers presenting the use of low-ppm RDRP solutions to improve the bioactivity of PEEK surfaces. Due to the fact that conventional ATRP requires a high catalyst concentration, leading to higher cost of purification of the final polymer product, low ppm ATRP techniques were proposed. In new ATRP methods, catalyst molecules are regenerated during the reaction progress [45,61,62,63,64]. There are many ways to achieve catalyst regeneration in ATRP, which can be distinguished as physical or chemical factors. Physical factors includes electrochemically mediated ATRP (eATRP) [39,46,65,66,67,68,69,70], photo-initiated ATRP (π-ATRP) [71,72,73,74], mechanically induced ATRP (mechano-ATRP) [75,76,77], microwave-assisted ATRP (MW-assisted ATRP) [78,79], and novel ultrasonication-induced ATRP (sono-ATRP) [73,75,78,80,81]. Chemical factors include initiators for continuous activator regeneration (ICAR) ATRP [74,82,83], activator regeneration by electron transfer (ARGET) ATRP [84,85,86], and supplemental activator and reducing agent (SARA) ATRP [87,88,89,90].
Nevertheless, many low-ppm ATRP techniques possess some disadvantageous, such as a lack of control over the amount of copper present in the experimental system (SARA ATRP) or necessity for application of expensive devices (eATRP). However, the ARGET ATRP approach eliminates the defects and provides simplification of the reaction system [91]. A potential advantage of this innovative solution is the possibility of a significant reduction in copper catalyst concentration (under 100 ppm), in comparison to the traditional ATRP reaction system. Furthermore, the application of ARGET ATRP enables carrying out the reaction under aerobic conditions [92]. Another profitable aspect is the opportunity for substantial process cost minimalization by elimination of expensive potentiostat devices according to the eATRP approach [68]. The use of ascorbic acid as a reducing agent is one of the cheapest (cheaper than an ultrasonic cleaner (needed in sono-ATRP) and a typical electrochemical set-up) solutions leading to the reduction of the amount of catalytic complex. In addition, ascorbic acid is approved for use by the US FDA, which confirms the legitimacy of its use in the syntheses leading to the production of biomaterials (e.g., PEEK surface modifications) [93].
An additional convenience of the low-ppm ATRP approach is the resignation of organic solvent usage, due to the possibility of conducting the reaction in a miniemulsion system. The biphasic nature of water–oil dispersion simplifies catalyst removal, while also providing increased contact surface area and facilitating mass and heat transport. ATRP in a miniemulsion system is an environmentally friendly solution, as it requires water as solvent, which is less toxic and less expensive than most organic solvents [41].
On the other hand, contamination of even small amounts of inorganic catalysts in biological systems can be slightly problematic. However, a metal-free ATRP based on regeneration of the deactivator with application of UV light is a promising prospect for the future research in this area [94,95]. For this reason, the focus on the use of organic photo-reduced catalysts in ATRP reactions might open up a new doorway for the design of biologically compatible systems.
In view of the possibility of using a wide range of various monomers, ATRP with diminished catalyst concentration can be a highly versatile and powerful tool for controlling the surface properties of inorganic and organic materials (e.g., PEEK). Firstly, effective optimization of SI-ATRP can lead to an ultra-high thickness of polymer brushes grafted from the flat surface. Secondly, following the 12 principles of green chemistry, in order to design a more environmentally friendly experimental setup, the use of organic solvents can be reduced by performing syntheses in water or using a miniemulsion system. Moreover, the low concentration of the catalytic complex simplifies the purification of the final product. Additionally, the concept of grafting from a polymer also allows an initiator gradient, enabling variation in the grafting density, resulting in materials with increased functionality. Therefore, future work should focus on involving an attempt to graft biocompatible (co)polymer brushes from the PEEK surface via a low-ppm RDRP approach.
An ideal bone implant material remains to be developed; thus, there is an intensive need for research on alternative materials with improved biological properties. Only limited monomers were grafted from the PEEK surface; however, the composition of the surface of the material guaranteeing ideal conditions for the adhesion of the implant with bone tissue remains to be determined. Nevertheless, there are literature reports of positive results from in vitro and in vivo tests performed on PEEK modified via surface-initiated polymerization. The increase in the implant’s adhesion capacity with living tissue confirmed by in vivo tests shows that research in the right direction already started. Nevertheless, it is worthwhile to use a combination of various surface modification methods. At the same time, it should be noted that special emphasis is placed on carrying out synthetic procedures according to environmentally friendly principles. Therefore, modern methods of polymerization (e.g., light-induced controlled polymerization) can be a useful and versatile tool for the synthesis and modification of new biomaterials. Another challenge for future science is to find new biologically compatible monomers, which can be successfully polymerized with, e.g., the low-ppm SI-ATRP method, as well as metal-free controlled radical polymerization, leading to a breakthrough in the bone implant material industry.

Abbreviations (Alphabetical Order)

AAacrylic acid
AEMA2-aminoethyl methacrylate hydrochloride
AFMatomic force microscopy
BAn-butyl acrylate
BPY2,2′-bipyridine
HAhyaluronic acid
MAMApoly(MPC-co-AEMA-co-N-methacryloyl methacrylamide)
MeHAmethacrylated hyaluronic acid
MeOEGMAoligo(ethylene glycol) methacrylate
MPC2-methacryloyloxyethyl phosphorylcholine
MPSpotassium 3-(methacryloyloxy) propane-1-sulfonate
MSCmesenchymal stem cell
MTAC2-(methacryloyloxy) ethyltrimethylammonium chloride
NIPAMN-isopropylacrylamide
PEEKpoly(ether ether ketone)
PMApoly(MPC-co-AEMA)
PMDETAN,N,N′,N″,N″-pentamethyldiethylenetriamine
RDRPreversible-deactivation radical polymerization
SEMscanning electron microscopy
SI-ATRPsurface-initiated atom transfer radical polymerization
SPMK3-sulfopropyl methacrylate potassium salt
Ststyrene
TiO2titanium dioxide
UVultraviolet
VPAvinylphosphonic acid
VSAvinylsulfonic acid sodium

Funding

This research received no external funding.

Acknowledgments

P.C. acknowledges the Minister of Science and Higher Education scholarship for outstanding young scientists (0001/E-363/STYP/13/2018).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kurtz, S.M.; Devine, J.N. PEEK biomaterials in trauma, orthopedic, and spinal implants. Biomaterials 2007, 28, 4845–4869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Najeeb, S.; Zafar, M.S.; Khurshid, Z.; Siddiqui, F. Applications of polyetheretherketone (PEEK) in oral implantology and prosthodontics. J. Prosthodont. Res. 2016, 60, 12–19. [Google Scholar] [CrossRef] [PubMed]
  3. He, X.H.; Deng, Y.; Yu, Y.; Lyu, H.; Liao, L. Drug-loaded/grafted peptide-modified porous PEEK to promote bone tissue repair and eliminate bacteria. Colloid Surf. B Biointerfaces 2019, 181, 767–777. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, J.; Shi, L.; Pei, Y.J.; Yang, D.; Gao, P.; Xiao, X.; Guo, S.; Li, M.H.; Li, X.K.; Guo, Z. Comparative effectiveness of PEEK rods versus titanium alloy rods in cervical fusion in a new sheep model. Eur. Spine J. 2020, in press. [Google Scholar] [CrossRef]
  5. Vaezi, M.; Yang, S.F. Extrusion-based additive manufacturing of PEEK for biomedical applications. Virtual Phys. Prototyp. 2015, 10, 123–135. [Google Scholar] [CrossRef]
  6. Gladson, T.S.F.; Ramesh, R.; Kavitha, C. Experimental investigation of mechanical, tribological and dielectric properties of alumina nano wire-reinforced PEEK/PTFE composites. Mater. Res. Express 2019, 6, 10. [Google Scholar]
  7. Graham, J.; Peck, J. PEEK Biomaterials Handbook; Elsevier: Amsterdam, The Netherlands, 2019. [Google Scholar]
  8. Panayotov, I.V.; Orti, V.; Cuisinier, F.; Yachouh, J. Polyetheretherketone (PEEK) for medical applications. J. Mater. Sci. Mater. Med. 2016, 27, 1–11. [Google Scholar] [CrossRef]
  9. Almasi, D.; Iqbal, N.; Sadeghi, M.; Sudin, I.; Kadir, M.R.A.; Kamarul, T. Preparation methods for improving PEEK’s bioactivity for orthopedic and dental application: A review. Int. J. Biomater. 2016, 2016, 1–12. [Google Scholar] [CrossRef] [Green Version]
  10. Kyomoto, M.; Moro, T.; Takatori, Y.; Kawaguchi, H.; Nakamura, K.; Ishihara, K. Self-initiated surface grafting with poly (2-methacryloyloxyethyl phosphorylcholine) on poly(ether-ether-ketone). Biomaterials 2010, 31, 1017–1024. [Google Scholar] [CrossRef]
  11. Fristrup, C.J.; Jankova, K.; Hvilsted, S. Hydrophilization of poly (ether ether ketone) films by surface-initiated atom transfer radical polymerization. Polym. Chem. 2010, 1, 1696–1701. [Google Scholar] [CrossRef]
  12. Riveiro, A.; Soto, R.; Comesana, R.; Boutinguiza, M.; del Val, J.; Quintero, F.; Lusquinos, F.; Pou, J. Laser surface modification of PEEK. Appl. Surf. Sci. 2012, 258, 9437–9442. [Google Scholar] [CrossRef]
  13. Zhao, Y.; Wong, H.M.; Wang, W.H.; Li, P.H.; Xu, Z.S.; Chong, E.Y.W.; Yan, C.H.; Yeung, K.W.K.; Chu, P.K. Cytocompatibility, osseointegration, and bioactivity of three-dimensional porous and nanostructured network on polyetheretherketone. Biomaterials 2013, 34, 9264–9277. [Google Scholar] [CrossRef] [PubMed]
  14. Toita, R.; Rashid, S.A.N.; Tsuru, K.; Ishikawa, K. Modulation of the osteoconductive property and immune response of poly (ether ether ketone) by modification with calcium ions. J. Mat. Chem. B 2015, 3, 8738–8746. [Google Scholar] [CrossRef]
  15. Matyjaszewski, K.; Dong, H.C.; Jakubowski, W.; Pietrasik, J.; Kusumo, A. Grafting from surfaces for “Everyone”: ARGET ATRP in the presence of air. Langmuir 2007, 23, 4528–4531. [Google Scholar] [CrossRef]
  16. Harrisson, S.; Drisko, G.L.; Malmstrom, E.; Hult, A.; Wooley, K.L. Hybrid rigid/soft and biologic/synthetic materials: Polymers grafted onto cellulose microcrystals. Biomacromolecules 2011, 12, 1214–1223. [Google Scholar] [CrossRef]
  17. Henze, M.; Madge, D.; Prucker, O.; Ruhe, J. “Grafting through”: Mechanistic aspects of radical polymerization reactions with surface-attached monomers. Macromolecules 2014, 47, 2929–2937. [Google Scholar] [CrossRef]
  18. Hui, C.M.; Pietrasik, J.; Schmitt, M.; Mahoney, C.; Choi, J.; Bockstaller, M.R.; Matyjaszewski, K. Surface-initiated polymerization as an enabling tool for multifunctional (nano-)engineered hybrid materials. Chem. Mat. 2014, 26, 745–762. [Google Scholar] [CrossRef]
  19. Flejszar, M.; Chmielarz, P. Surface-initiated atom transfer radical polymerization for the preparation of well-defined organic-inorganic hybrid nanomaterials. Materials 2019, 12, 3030. [Google Scholar] [CrossRef] [Green Version]
  20. Kyomoto, M.; Ishihara, K. Self-initiated surface graft polymerization of 2-methacryloyloxyethyl phosphorylcholine on poly (ether ether ketone) by photoirradiation. ACS Appl. Mater. Interfaces 2009, 1, 537–542. [Google Scholar] [CrossRef]
  21. Kyomoto, P.M.; Moro, M.D.P.T.; Yamane, M.S.; Watanabe, B.S.K.; Takatori, M.D.P.Y.; Tanaka, M.D.P.S.; Ishihara, P.K. Smart PEEK modified by self-initiated surface graft polymerization for orthopedic bearings. Reconstr. Rev. 2014, 4, 36–45. [Google Scholar]
  22. Kyomoto, M.; Moro, T.; Yamane, S.; Hashimoto, M.; Takatori, Y.; Ishihara, K. Poly(ether-ether-ketone) orthopedic bearing surface modified by self-initiated surface grafting of poly (2-methacryloyloxyethyl phosphorylcholine). Biomaterials 2013, 34, 7829–7839. [Google Scholar] [CrossRef] [PubMed]
  23. Kawasaki, Y.; Iwasaki, Y. Surface modification of poly (ether ether ketone) with methacryloyl-functionalized phospholipid polymers via self-initiation graft polymerization. J. Biomater. Sci. Polym. Ed. 2014, 25, 895–906. [Google Scholar] [CrossRef] [PubMed]
  24. Tateishi, T.; Kyomoto, M.; Kakinoki, S.; Yamaoka, T.; Ishihara, K. Reduced platelets and bacteria adhesion on poly (ether ether ketone) by photoinduced and self-initiated graft polymerization of 2-methacryloyloxyethyl phosphorylcholine. J. Biomed. Mater. Res. Part A 2014, 102, 1342–1349. [Google Scholar] [CrossRef] [PubMed]
  25. Ishihara, K.; Inoue, Y.; Kyomoto, M. Surface modification on poly (ether ether ketone) with phospholipid polymer via photoinduced self-initiated grafting. Macromol. Symp. 2015, 354, 230–236. [Google Scholar] [CrossRef]
  26. Shiojima, T.; Inoue, Y.; Kyomoto, M.; Ishihara, K. High-efficiency preparation of poly(2-methacryloyloxyethyl phosphorylcholine) grafting layer on poly(ether ether ketone) by photoinduced and self-initiated graft polymerization in an aqueous solution in the presence of inorganic salt additives. Acta Biomater. 2016, 40, 38–45. [Google Scholar] [CrossRef]
  27. Yousaf, A.; Farrukh, A.; Oluz, Z.; Tuncel, E.; Duran, H.; Dogan, S.Y.; Tekinay, T.; Rehman, H.U.; Yameen, B. UV-light assisted single step route to functional PEEK surfaces. React. Funct. Polym. 2014, 83, 70–75. [Google Scholar] [CrossRef]
  28. Chouwatat, P.; Hirai, T.; Higaki, K.; Higaki, Y.; Sue, H.J.; Takahara, A. Aqueous lubrication of poly (etheretherketone) via surface-initiated polymerization of electrolyte monomers. Polymer 2017, 116, 549–555. [Google Scholar] [CrossRef]
  29. Zhao, X.D.; Xiong, D.S.; Wang, K.; Wang, N. Improved biotribological properties of PEEK by photo-induced graft polymerization of acrylic acid. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 75, 777–783. [Google Scholar] [CrossRef]
  30. Liu, S.D.; Zhu, Y.T.; Gao, H.N.; Ge, P.; Ren, K.L.; Gao, J.W.; Cao, Y.P.; Han, D.; Zhang, J.H. One-step fabrication of functionalized poly(etheretherketone) surfaces with enhanced biocompatibility and osteogenic activity. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 88, 70–78. [Google Scholar] [CrossRef]
  31. Zheng, Y.Y.; Liu, L.H.; Ma, Y.; Xiao, L.; Liu, Y. Enhanced osteoblasts responses to surface-sulfonated polyetheretherketone via a single-step ultraviolet-initiated graft polymerization. Ind. Eng. Chem. Res. 2018, 57, 10403–10410. [Google Scholar] [CrossRef]
  32. Liu, H.; Zheng, Y.Y.; Zhang, Q.Y.; Yu, L.; Hu, Z.L.; Liu, Y. Surface phosphonation treatment shows dose-dependent enhancement of the bioactivity of polyetheretherketone. RSC Adv. 2019, 9, 30076–30086. [Google Scholar] [CrossRef] [Green Version]
  33. Zheng, Y.Y.; Liu, L.H.; Xiao, L.; Zhang, Q.Y.; Liu, Y. Enhanced osteogenic activity of phosphorylated polyetheretherketone via surface-initiated grafting polymerization of vinylphosphonic acid. Colloid Surf. B Biointerfaces 2019, 173, 591–598. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, H.Y.; Lu, T.; Meng, F.H.; Zhu, H.Q.; Liu, X.Y. Enhanced osteoblast responses to poly ether ether ketone surface modified by water plasma immersion ion implantation. Colloid Surf. B Biointerfaces 2014, 117, 89–97. [Google Scholar] [CrossRef] [PubMed]
  35. Matyjaszewski, K. Atom transfer radical polymerization (ATRP): Current status and future perspectives. Macromolecules 2012, 45, 4015–4039. [Google Scholar] [CrossRef]
  36. Matyjaszewski, K.; Tsarevsky, N.V. Macromolecular engineering by atom transfer radical polymerization. J. Am. Chem. Soc. 2014, 136, 6513–6533. [Google Scholar] [CrossRef]
  37. Pan, X.C.; Fantin, M.; Yuan, F.; Matyjaszewski, K. Externally controlled atom transfer radical polymerization. Chem. Soc. Rev. 2018, 47, 5457–5490. [Google Scholar] [CrossRef]
  38. Zaborniak, I.; Chmielarz, P. Ultrasound-mediated atom transfer radical polymerization (ATRP). Materials 2019, 12, 3600. [Google Scholar] [CrossRef] [Green Version]
  39. Park, S.; Chmielarz, P.; Gennaro, A.; Matyjaszewski, K. Simplified electrochemically mediated atom transfer radical polymerization using a sacrificial anode. Angew. Chem. Int. Ed. 2015, 54, 2388–2392. [Google Scholar] [CrossRef]
  40. Matyjaszewski, K. Atom transfer radical polymerization: From mechanisms to applications. Isr. J. Chem. 2012, 52, 206–220. [Google Scholar] [CrossRef]
  41. Zaborniak, I.; Chmielarz, P. Temporally controlled ultrasonication-mediated atom transfer radical polymerization in miniemulsion. Macromol. Chem. Phys. 2019, 220, 1900285. [Google Scholar] [CrossRef]
  42. Chmielarz, P. Synthesis of cationic star polymers by simplified electrochemically mediated ATRP. Express Polym. Lett. 2016, 10, 810–821. [Google Scholar] [CrossRef]
  43. Bussels, R.; Bergman-Gottgens, C.; Meuldijk, J.; Koning, C. Multiblock copolymers synthesized in aqueous dispersions using multifunctional RAFT agents. Polymer 2005, 46, 8546–8554. [Google Scholar] [CrossRef]
  44. Chmielarz, P.; Yan, J.J.; Krys, P.; Wang, Y.; Wang, Z.Y.; Bockstaller, M.R.; Matyjaszewski, K. Synthesis of nanoparticle copolymer brushes via surface-initiated seATRP. Macromolecules 2017, 50, 4151–4159. [Google Scholar] [CrossRef]
  45. Chmielarz, P.; Sobkowiak, A. Ultralow ppm seATRP synthesis of PEO-b-PBA copolymers. J. Polym. Res. 2017, 24, 77. [Google Scholar] [CrossRef]
  46. Chmielarz, P. Synthesis of multiarm star block copolymers via simplified electrochemically mediated ATRP. Chem. Pap. 2017, 71, 161–170. [Google Scholar] [CrossRef]
  47. Turan, E.; Caykara, T. Construction of hydroxyl-terminated poly(N-isopropylacrylamide) brushes on silicon wafer via surface-initiated atom transfer radical polymerization. J. Polym. Sci. Pol. Chem. 2010, 48, 3880–3887. [Google Scholar] [CrossRef]
  48. Li, X.; Wang, M.M.; Wang, L.; Shi, X.J.; Ya, J.X.; Song, B.; Chen, H. Block copolymer modified surfaces for conjugation of biomacromolecules with control of quantity and activity. Langmuir 2013, 29, 1122–1128. [Google Scholar] [CrossRef]
  49. Khabibullin, A.; Mastan, E.; Matyjaszewski, K.; Zhu, S.P. Surface-initiated atom transfer radical polymerization. In Controlled Radical Polymerization at and from Solid Surfaces; Vana, P., Ed.; Springer: Berlin/Heidelberg, Germany, 2016; Volume 270, pp. 29–76. [Google Scholar]
  50. Chmielarz, P.; Krys, P.; Wang, Z.Y.; Wang, Y.; Matyjaszewski, K. Synthesis of well-defined polymer brushes from silicon wafers via surface-initiated seATRP. Macromol. Chem. Phys. 2017, 218, 1700106. [Google Scholar] [CrossRef]
  51. Wei, Y.; Zhang, J.X.; Feng, X.T.; Liu, D.Y. Bioactive zwitterionic polymer brushes grafted from silicon wafers via SI-ATRP for enhancement of antifouling properties and endothelial cell selectivity. J. Biomater. Sci. Polym. Ed. 2017, 28, 2101–2116. [Google Scholar] [CrossRef]
  52. Ejaz, M.; Tsujii, Y.; Fukuda, T. Controlled grafting of a well-defined polymer on a porous glass filter by surface-initiated atom transfer radical polymerization. Polymer 2001, 42, 6811–6815. [Google Scholar] [CrossRef]
  53. Chen, M.; Briscoe, W.H.; Armes, S.P.; Cohen, H.; Klein, J. Robust, biomimetic polymer brush layers grown directly from a planar mica surface. Chem. Phys. Chem. 2007, 8, 1303–1306. [Google Scholar] [CrossRef] [PubMed]
  54. Gong, R.; Maclaughlin, S.; Zhu, S.P. Surface modification of active metals through atom transfer radical polymerization grafting of acrylics. Appl. Surf. Sci. 2008, 254, 6802–6809. [Google Scholar] [CrossRef]
  55. Cheesman, B.T.; Smith, E.G.; Murdoch, T.J.; Guibert, C.; Webber, G.B.; Edmondson, S.; Wanless, E.J. Polyelectrolyte brush pH-response at the silica-aqueous solution interface: A kinetic and equilibrium investigation. Phys. Chem. Chem. Phys. 2013, 15, 14502–14510. [Google Scholar] [CrossRef] [PubMed]
  56. Qiu, J.H.; Zhang, Y.W.; Shen, Y.B.; Zhang, Y.T.; Zhang, H.Q.; Liu, J.D. Hydrophilic modification of microporous polysulfone membrane via surface-initiated atom transfer radical polymerization of acrylamide. Appl. Surf. Sci. 2010, 256, 3274–3280. [Google Scholar] [CrossRef]
  57. Yang, Y.; Liu, X.G.; Ye, G.; Zhu, S.; Wang, Z.; Huo, X.M.; Matyjaszewski, K.; Lu, Y.X.; Chen, J. Metal-free photoinduced electron transfer-atom transfer radical polymerization integrated with bioinspired polydopamine chemistry as a green strategy for surface engineering of magnetic nanoparticles. ACS Appl. Mater. Interfaces 2017, 9, 13637–13646. [Google Scholar] [CrossRef] [PubMed]
  58. Yameen, B.; Alvarez, M.; Azzaroni, O.; Jonas, U.; Knoll, W. Tailoring of poly (ether ether ketone) surface properties via surface-initiated atom transfer radical polymerization. Langmuir 2009, 25, 6214–6220. [Google Scholar] [CrossRef] [PubMed]
  59. Yameen, B.; Khan, H.U.; Knoll, W.; Forch, R.; Jonas, U. Surface initiated polymerization on pulsed plasma deposited polyallylamine: A polymer substrate-independent strategy to soft surfaces with polymer brushes. Macromol. Rapid Commun. 2011, 32, 1735–1740. [Google Scholar] [CrossRef]
  60. Langer, R.; Tirrell, D.A. Designing materials for biology and medicine. Nature 2004, 428, 487–492. [Google Scholar] [CrossRef]
  61. Elsen, A.M.; Burdynska, J.; Park, S.; Matyjaszewski, K. Active ligand for low ppm miniemulsion atom transfer radical polymerization. Macromolecules 2012, 45, 7356–7363. [Google Scholar] [CrossRef]
  62. Chmielarz, P. Synthesis of inositol-based star polymers through low ppm ATRP methods. Polym. Adv. Technol. 2017, 28, 1804–1812. [Google Scholar] [CrossRef]
  63. Chmielarz, P. Synthesis of pyridoxine-based eagle-shaped asymmetric star polymers through seATRP. Polym. Adv. Technol. 2017, 28, 1787–1793. [Google Scholar] [CrossRef]
  64. Wang, Y.; Fantin, M.; Matyjaszewski, K. Synergy between electrochemical ATRP and RAFT for polymerization at low copper loading. Macromol. Rapid Commun. 2018, 39, 1–5. [Google Scholar] [CrossRef]
  65. Chmielarz, P.; Sobkowiak, A.; Matyjaszewski, K. A simplified electrochemically mediated ATRP synthesis of PEO-b-PMMA copolymers. Polymer 2015, 77, 266–271. [Google Scholar] [CrossRef] [Green Version]
  66. Fantin, M.; Lorandi, F.; Isse, A.A.; Gennaro, A. Sustainable electrochemically-mediated atom transfer radical polymerization with inexpensive non-platinum electrodes. Macromol. Rapid Commun. 2016, 37, 1318–1322. [Google Scholar] [CrossRef] [PubMed]
  67. Chmielarz, P.; Paczesniak, T.; Rydel-Ciszek, K.; Zaborniak, I.; Biedka, P.; Sobkowiak, A. Synthesis of naturally-derived macromolecules through simplified electrochemically mediated ATRP. Beilstein J. Org. Chem. 2017, 13, 2466–2472. [Google Scholar] [CrossRef] [Green Version]
  68. Chmielarz, P.; Fantin, M.; Park, S.; Isse, A.A.; Gennaro, A.; Magenau, A.J.D.; Sobkowiak, A.; Matyjaszewski, K. Electrochemically mediated atom transfer radical polymerization (eATRP). Prog. Polym. Sci. 2017, 69, 47–78. [Google Scholar] [CrossRef]
  69. Sun, Y.; Lathwal, S.; Wang, Y.; Fu, L.Y.; Olszewski, M.; Fantin, M.; Enciso, A.E.; Szczepaniak, G.; Das, S.; Matyjaszewski, K. Preparation of well-defined polymers and DNA-polymer bioconjugates via small-volume eATRP in the presence of air. ACS Macro Lett. 2019, 8, 603–609. [Google Scholar] [CrossRef]
  70. Guo, J.K.; Zhou, Y.N.; Luo, Z.H. Electrochemically mediated ATRP process intensified by ionic liquid: A “flash” polymerization of methyl acrylate. Chem. Eng. J. 2019, 372, 163–170. [Google Scholar] [CrossRef]
  71. Pan, X.; Malhotra, N.; Simakova, A.; Wang, Z.; Konkolewicz, D.; Matyjaszewski, K. Photoinduced atom transfer radical polymerization with ppm-level Cu catalyst by visible light in aqueous media. J. Am. Chem. Soc. 2015, 137, 15430–15433. [Google Scholar] [CrossRef]
  72. Aydogan, C.; Yilmaz, G.; Yagci, Y. Synthesis of hyperbranched polymers by photoinduced metal-free ATRP. Macromolecules 2017, 50, 9115–9120. [Google Scholar] [CrossRef]
  73. Wang, Y.; Dadashi-Silab, S.; Matyjaszewski, K. Photoinduced miniemulsion atom transfer radical polymerization. ACS Macro Lett. 2018, 7, 720–725. [Google Scholar] [CrossRef]
  74. Vasu, V.; Kim, J.S.; Yu, H.S.; Bannerman, W.I.; Johnson, M.E.; Asandei, A.D. Normal, ICAR and photomediated butadiene-ATRP with iron complexes. Polym. Chem. 2018, 9, 2389–2406. [Google Scholar] [CrossRef]
  75. Wang, Z.H.; Pan, X.C.; Li, L.C.; Fantin, M.; Yan, J.J.; Wang, Z.Y.; Wang, Z.H.; Xia, H.S.; Matyjaszewski, K. Enhancing mechanically induced ATRP by promoting interfacial electron transfer from piezoelectric nanoparticles to Cu catalysts. Macromolecules 2017, 50, 7940–7948. [Google Scholar] [CrossRef]
  76. Wang, Z.; Pan, X.; Yan, J.; Dadashi-Silab, S.; Xie, G.; Zhang, J.; Wang, Z.; Xia, H.; Matyjaszewski, K. Temporal control in mechanically controlled atom transfer radical polymerization using low ppm of Cu catalyst. ACS Macro Lett. 2017, 6, 546–549. [Google Scholar] [CrossRef]
  77. Mohapatra, H.; Kleiman, M.; Esser-Kahn, A.P. Mechanically controlled radical polymerization initiated by ultrasound. Nat. Chem. 2017, 9, 135–139. [Google Scholar] [CrossRef]
  78. Cheng, Z.P.; Zhu, X.L.; Zhou, N.C.; Zhu, J.; Zhang, Z.B. Atom transfer radical polymerization of styrene under pulsed microwave irradiation. Radiat. Phys. Chem. 2005, 72, 695–701. [Google Scholar] [CrossRef]
  79. Xie, Z.K.; Guo, J.K.; Luo, Z.H. Assessment of microwave effect on polymerization conducted under ARGET ATRP conditions. Macromol. React. Eng. 2018, 12, 1700032. [Google Scholar] [CrossRef]
  80. Fantin, M.; Chmielarz, P.; Wang, Y.; Lorandi, F.; Isse, A.A.; Gennaro, A.; Matyjaszewski, K. Harnessing the interaction between surfactant and hydrophilic catalyst to control eATRP in miniemulsion. Macromolecules 2017, 50, 3726–3732. [Google Scholar] [CrossRef]
  81. Wang, Z.H.; Wang, Z.H.; Pan, X.C.; Fu, L.Y.; Lathwal, S.; Olszewski, M.; Yan, J.J.; Enciso, A.E.; Wang, Z.Y.; Xia, H.S.; et al. Ultrasonication-induced aqueous atom transfer radical polymerization. ACS Macro Lett. 2018, 7, 275–280. [Google Scholar] [CrossRef]
  82. Konkolewicz, D.; Magenau, A.J.D.; Averick, S.E.; Simakova, A.; He, H.K.; Matyjaszewski, K. ICAR ATRP with ppm Cu catalyst in water. Macromolecules 2012, 45, 4461–4468. [Google Scholar] [CrossRef]
  83. Wang, G.W.; Wang, Z.Y.; Lee, B.; Yuan, R.; Lu, Z.; Yan, J.J.; Pan, X.C.; Song, Y.; Bockstaller, M.R.; Matyjaszewski, K. Polymerization-induced self-assembly of acrylonitrile via ICAR ATRP. Polymer 2017, 129, 57–67. [Google Scholar] [CrossRef]
  84. Wang, Y.; Lorandi, F.; Fantin, M.; Chmielarz, P.; Isse, A.A.; Gennaro, A.; Matyjaszewski, K. Miniemulsion ARGET ATRP via interfacial and ion-pair catalysis: From ppm to ppb of residual copper. Macromolecules 2017, 50, 8417–8425. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, Z.; Wang, X.S.; Tam, K.C.; Sebe, G. A comparative study on grafting polymers from cellulose nanocrystals via surface-initiated atom transfer radical polymerization (ATRP) and activator re-generated by electron transfer ATRP. Carbohydr. Polym. 2019, 205, 322–329. [Google Scholar] [CrossRef] [PubMed]
  86. Navarro, L.A.; Enciso, A.E.; Matyjaszewski, K.; Zauscher, S. Enzymatically degassed surface-initiated atom transfer radical polymerization with real-time monitoring. J. Am. Chem. Soc. 2019, 141, 3100–3109. [Google Scholar] [CrossRef] [PubMed]
  87. Konkolewicz, D.; Wang, Y.; Zhong, M.J.; Krys, P.; Isse, A.A.; Gennaro, A.; Matyjaszewski, K. Reversible-deactivation radical polymerization in the presence of metallic copper. A critical assessment of the SARA ATRP and SET-LRP mechanisms. Macromolecules 2013, 46, 8749–8772. [Google Scholar] [CrossRef]
  88. Konkolewicz, D.; Wang, Y.; Krys, P.; Zhong, M.J.; Isse, A.A.; Gennaro, A.; Matyjaszewski, K. SARA ATRP or SET-LRP. End of controversy? Polym. Chem. 2014, 5, 4396–4417. [Google Scholar] [CrossRef]
  89. Chmielarz, P.; Krys, P.; Park, S.; Matyjaszewski, K. PEO-b-PNIPAM copolymers via SARA ATRP and eATRP in aqueous media. Polymer 2015, 71, 143–147. [Google Scholar] [CrossRef] [Green Version]
  90. Oliveira, A.S.R.; Mendonca, P.V.; Serra, A.C.; Coelho, J.F.J. Self-degassing SARA ATRP mediated by Na2S2O4 with no external additives. J. Polym. Sci. Pol. Chem. 2020, 58, 145–153. [Google Scholar]
  91. Król, P.; Chmielarz, P. Synthesis of PMMA-b-PU-b-PMMA tri-block copolymers through ARGET ATRP in the presence of air. Express Polym. Lett. 2013, 7, 249–260. [Google Scholar] [CrossRef]
  92. Kang, H.; Jeong, W.; Hong, D. Antifouling surface coating using droplet-based SI-ARGET ATRP of carboxybetaine under open-air conditions. Langmuir 2019, 35, 7744–7750. [Google Scholar] [CrossRef]
  93. Assaf, C.E.; De Clercq, N.; Van Poucke, C.; Vlaemynck, G.; Van Coillie, E.; Van Pamel, E. Effects of ascorbic acid on patulin in aqueous solution and in cloudy apple juice. Mycotoxin Res. 2019, 35, 341–351. [Google Scholar] [CrossRef] [PubMed]
  94. Discekici, E.H.; Anastasaki, A.; de Alaniz, J.R.; Hawker, C.J. Evolution and future directions of metal-free atom transfer radical polymerization. Macromolecules 2018, 51, 7421–7434. [Google Scholar] [CrossRef] [Green Version]
  95. Liu, Q.R.; Jian, L.H.; Liu, R.Q.; Yang, H.X.; Kong, J.M.; Zhang, X.J. Metal-free photoinduced atom transfer radical polymerization for highly sensitive detection of lung cancer DNA. Chem. Eur. J. 2020, 26, 1633–1639. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The scheme of poly(ether ether ketone) (PEEK) surface modifications via ultraviolet (UV)-induced polymerization of 2-methacryloyloxyethyl phosphorylcholine (MPC). Reprinted with permission from Reference [21].
Figure 1. The scheme of poly(ether ether ketone) (PEEK) surface modifications via ultraviolet (UV)-induced polymerization of 2-methacryloyloxyethyl phosphorylcholine (MPC). Reprinted with permission from Reference [21].
Materials 13 00999 g001
Figure 2. Atomic force microscopy (AFM) images of (a) unmodified PEEK sheet, (b) PEEK-g-poly(oligo(ethylene glycol) methacrylate) (PMeOEGMA), (c) PEEK-g-poly(acrylic acid) (PAA), (d) PEEK-g-poly(n-butyl acrylate) (PBA), (e) PEEK-g-polystyrene (PS), and (f) PEEK-g-poly(vinyl phosphonic acid) (PVPA). Reprinted with permission from Reference [27], 2014 Elsevier B.V.
Figure 2. Atomic force microscopy (AFM) images of (a) unmodified PEEK sheet, (b) PEEK-g-poly(oligo(ethylene glycol) methacrylate) (PMeOEGMA), (c) PEEK-g-poly(acrylic acid) (PAA), (d) PEEK-g-poly(n-butyl acrylate) (PBA), (e) PEEK-g-polystyrene (PS), and (f) PEEK-g-poly(vinyl phosphonic acid) (PVPA). Reprinted with permission from Reference [27], 2014 Elsevier B.V.
Materials 13 00999 g002
Figure 3. Schematically illustrated conception of PEEK surface modification based on light-inducible electrospinning. Reprinted with permission from Reference [30], 2018 Elsevier B. V.
Figure 3. Schematically illustrated conception of PEEK surface modification based on light-inducible electrospinning. Reprinted with permission from Reference [30], 2018 Elsevier B. V.
Materials 13 00999 g003
Figure 4. The photos were taken after 12 weeks of implantation in adult male New Zealand White rabbits. Direct contact between bone tissue and (A) pristine PEEK and (B) PEEK-g-PVPA is marked with white arrows. Reprinted with permission from Reference [33], 2019 Elsevier B. V.
Figure 4. The photos were taken after 12 weeks of implantation in adult male New Zealand White rabbits. Direct contact between bone tissue and (A) pristine PEEK and (B) PEEK-g-PVPA is marked with white arrows. Reprinted with permission from Reference [33], 2019 Elsevier B. V.
Materials 13 00999 g004
Figure 5. Schematic illustration of PEEK surface functionalization. (A) At first, keto group reduction was performed. (B) Then, atom transfer radical polymerization (ATRP) initiator (2-BiBr) was attached. (C) Subsequently, SI-ATRP of MPS, MeOEGMA, and N-isopropylacrylamide (NIPAM) was carried out. Reprinted with permission from Reference [58], ACS 2009.
Figure 5. Schematic illustration of PEEK surface functionalization. (A) At first, keto group reduction was performed. (B) Then, atom transfer radical polymerization (ATRP) initiator (2-BiBr) was attached. (C) Subsequently, SI-ATRP of MPS, MeOEGMA, and N-isopropylacrylamide (NIPAM) was carried out. Reprinted with permission from Reference [58], ACS 2009.
Materials 13 00999 g005
Table 1. Surface modifications of poly(ether ether ketone) via photo-induced polymerization method.
Table 1. Surface modifications of poly(ether ether ketone) via photo-induced polymerization method.
MonomerSolventT 1 (°C)Time (min)Wavelength (nm)Light Intensity (mW/cm2)Reference
MPCH2O605–90350 ± 505[10,20,21,22] 2
MPC1-butanolNIA 3NIA350 ± 501[23]
MPCH2O25–6090350 ± 501.5–9[24]
MPCH2O6090NIA2.5–10[25]
MPCH2O605–90350 ± 5020[26]
StNIArt 4270315–400NIA[27]
AANIArt270315–400NIA[27]
MeOEGMANIArt270315–400NIA[27]
VPANIArt270315–400NIA[27]
BANIArt270315–400NIA[27]
MTACH2Ort5–903655[28]
SPMKH2Ort5–903655[28]
AAH2ONIA30, 45, 60, 90350 ± 50NIA[29]
MeHANIANIA0.5, 1, 1.5, 2, 4NIA5[30]
VSAH2Ort40365NIA[31]
VPAH2Ort20, 50, 90365NIA[32]
VPAH2Ort40365NIA[33]
1 Temperature; 2 all cited papers were based on the same synthetic procedure; 3 no information available; 4 synthesis was carried out at room temperature.
Table 2. Surface modifications of poly(ether ether ketone) via ATRP method.
Table 2. Surface modifications of poly(ether ether ketone) via ATRP method.
InitiatorMonomerCatalyst ComplexSolventT (°C)Catalyst ConcentrationReference
ppm (Catalyst/Monomer)ppm (by Weight)
2-BiBrMPSCuIICl2/BPYMeOH/H2Ort 125,0713533[58]
2-BiBrMeOEGMACuIIBr2/BPYH2O3022702251[58,59] 2
2-BiBrNIPAMCuIIBr2/PMDETAMeOH/H2Ort10,0002753[58]
2-BiBrPEGMACuIICl2/BPYH2O3056,9776130[11]
1 Synthesis was carried out at room temperature; 2 all cited papers were based on the same synthetic procedure.

Share and Cite

MDPI and ACS Style

Flejszar, M.; Chmielarz, P. Surface Modifications of Poly(Ether Ether Ketone) via Polymerization Methods—Current Status and Future Prospects. Materials 2020, 13, 999. https://doi.org/10.3390/ma13040999

AMA Style

Flejszar M, Chmielarz P. Surface Modifications of Poly(Ether Ether Ketone) via Polymerization Methods—Current Status and Future Prospects. Materials. 2020; 13(4):999. https://doi.org/10.3390/ma13040999

Chicago/Turabian Style

Flejszar, Monika, and Paweł Chmielarz. 2020. "Surface Modifications of Poly(Ether Ether Ketone) via Polymerization Methods—Current Status and Future Prospects" Materials 13, no. 4: 999. https://doi.org/10.3390/ma13040999

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