Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3235 2023-06-29 12:05:23 |
2 only format change Meta information modification 3235 2023-06-30 04:53:02 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Parcheta, M.; Sobiesiak, M. Applications of Polymeric Materials with Antibacterial Properties. Encyclopedia. Available online: https://encyclopedia.pub/entry/46214 (accessed on 04 July 2024).
Parcheta M, Sobiesiak M. Applications of Polymeric Materials with Antibacterial Properties. Encyclopedia. Available at: https://encyclopedia.pub/entry/46214. Accessed July 04, 2024.
Parcheta, Monika, Magdalena Sobiesiak. "Applications of Polymeric Materials with Antibacterial Properties" Encyclopedia, https://encyclopedia.pub/entry/46214 (accessed July 04, 2024).
Parcheta, M., & Sobiesiak, M. (2023, June 29). Applications of Polymeric Materials with Antibacterial Properties. In Encyclopedia. https://encyclopedia.pub/entry/46214
Parcheta, Monika and Magdalena Sobiesiak. "Applications of Polymeric Materials with Antibacterial Properties." Encyclopedia. Web. 29 June, 2023.
Applications of Polymeric Materials with Antibacterial Properties
Edit

The presence of antibiotic-resistant bacteria in people's environment is a matter of growing concern. The issue of multidrug-resistant bacteria urged the need to elaborate the novel antipathogen agents. The polymers and copolymers modified with bioactive compounds have emerged as a group of highly effective antimicrobial agents that find usage in many fields. The natural polymers have a great advantage over the synthetic ones due to their non-toxicity, biocompatibility, non-immunogenicity, and high stability. On the other hand, they are less effective in biomedical applications in comparison to synthetic polymers. The modifications that provide the natural polymers with desirable industrial activity include chemical treatment processes such as hydroxylation, carboxylation and epoxidation, or in vitro enzyme treatment.

polymeric materials antibacterial activity antibiotic-resistant bacteria

1. Antibiotic Resistant Bacteria in Environment

In recent years, the growing amount of antibiotics in wastewaters constitute a great challenge to the wastewater treatment plants, as the conventional methods—e.g., flocculation, sedimentation, filtration, or coagulation—are not sufficient to remove these pollutants from environment sewage [1]. The most abundant antibiotics found in sewage are trimethoprim, sulfonamides (SA), quinolones, and macrolides, the frequent occurrence of which results from their stability and wide application in the treatment of bacterial diseases in humans and animals [2].
Trimethoprim (TMP) is an antibiotic; its moieties contain electron-rich aromatic rings and a deprotonated amine group, and it is susceptible to the oxidation process, which is proposed as one of the ways to eliminate this compound from aqueous systems [3].
SA are a class of antibiotics that include sulfadiazine, sulfamethazine, and sulfamethoxazole (SMX). One of the possible pathways for the removal of SA is bioaugmentation, which leads to anaerobic degradation of these antibiotics [4].
Quinolones (ofloxacin, ciprofloxacin, norfloxacin) and macrolides (clarithromycin, erythromycin, azithromycin) were recorded in Asia and Europe with occurrence frequencies between 6–30% and 6–10%, respectively [5].
The widespread use of antibiotics contributes significantly to the resistance of bacteria to their bactericidal properties, which is a great challenge for modern medicine [6]. An excessive amount of prescribed antibiotics and their limited metabolism in human cells (30–90% of consumed antibiotics are not metabolized in the human body and are excreted into wastewater systems) leads to antibiotic exposure and accelerates resistance in bacteria [7]. The scheme of antibiotic distribution in the natural environment is presented in Figure 1.
Figure 1. Antibiotic distribution in environment.
Antibiotic resistant bacteria (ARB) acquire resistivity by producing antibiotic resistance genes (ARGs) through the cellular expression process [8]. The capacity of the horizontal transmission of genes in aquatic environments provides an easiness in the spread of antibiotic resistance among humans and animals, which poses a significant risk to health [9]. Antibiotics may enter the environment in the form of metabolites with the retained activity of the original drugs, or they may be excreted from the human/animal body as more polar derivatives of the original antibiotic, which can be then converted by the bacteria into the original drug [10]. ARBs are divided into multidrug-resistant (MDR), extensively-drug resistant (XDR) and pan-drug-resistant (PDR) bacteria, and the criterion for this classification is the number of classes of antibiotics to which the bacteria are resistant. MDRs are bacteria that are resistant to at least one drug belonging to three or more classes of antibiotics at the same time, XDRs are bacteria resistant to at least one antibiotic of each class, except for two or less antimicrobial categories, and finally, PDRs are bacteria resistant to all antibiotics in all antimicrobial categories. MDRs, which have become resistant due to high-volume and long-term use of antibiotics, are of particular concern in Chinese and European intensive care units, where they were responsible for 1.27 million deaths in 2019 [11]. Due to the ability of PDR to effectively withstand all forms of antibiotic therapy, test bacteria pose a particular threat to health facilities. Among the PDR, Gram-negative bacteria can be distinguished, i.e., E. coli, P. aeruginosa, K. pneumoniae, and A. baumannii. P. aeruginosa is the bacterium responsible for pneumonia, which is particularly common in intensive care units, and poses a particular threat to people suffering from cystic fibrosis and the formation of biofilms. Most nosocomial infections are caused by K. pneumoniae, also called the super bacteria, because it has become resistant to all the beta-lactams that make it difficult to treat diseases caused by this bacterium. Some E. coli strains may affect the urinary tract, digestive tract, spinal cord, and brain. A. baumannii can lead to pneumonia and infections of wounds and the intra-abdomen [12]. Carbapenem-antibiotics-resistant XDR bacteria such as P. aeruginosa, K. pneumoniae, and A. baumannii cause bloodstream infections with high mortality [13].

2. Polymeric Materials with Antibacterial Properties—Mechanism of Action and Medical Applications

It is predicted that antibiotic-resistant bacteria can result in 10 million deaths by 2050 [14]. The issue of multidrug-resistant bacteria urged the need to elaborate the novel antipathogen agents [15]. The polymers and copolymers modified with bioactive compounds have emerged as a group of highly effective antimicrobial agents [16] that find usage in many fields (Figure 2).
Figure 2. The industrial application of polymers with antibacterial activity.
The factors that are of the greatest influence on the antimicrobial properties of bioactive polymers are their low toxicity towards human cells and high activity in fighting bacteria cells [17]. The mechanism of antibacterial activity of these compounds is provided by active and passive ways of interacting with pathogens [18]. The active mechanism of disrupting the function of bacteria cells consists of the destabilisation of bacteria cells through electrostatic interactions between the predominantly hydrophobic and negatively charged bacteria plasma membrane, and the positively charged surface of the modified cationic polymer [19]. The most popular active substances used to modify the surface of polymers are quaternary ammonium salts, which interact with the negatively charged membranes of bacteria, causing leakage of components out of the bacterial cell, and consequently, the cell’s death [20]. Similar mechanisms of active functional disturbance by electrostatic interaction are exhibited by polyethylenimines. Another example of modifiers on the polymer surface are polyguanidines, which inhibits the bacterial growth due to it breaking the Ca2+ salt bridges and N-halamine, which disrupts the function of the amino cell receptors in bacteria by generating the oxidative halogen [21]. The passive mechanism of fighting bacteria cells relys on the synthesis of the passive polymer layer, which prevents the adhesion of bacteria on the modified polymer surface, thereby repelling the bacteria without any active interaction with them [22].
The natural polymers have a great advantage over the synthetic ones due to their non-toxicity, biocompatibility, non-immunogenicity, and high stability. On the other hand, they are less effective in biomedical applications in comparison to synthetic polymers [23]. The modifications that provide the natural polymers with desirable industrial activity include chemical treatment processes such as hydroxylation, carboxylation and epoxidation, or in vitro enzyme treatment [24].
Synthetic polymers frequently used in the synthesis of polymers with antimicrobial activity are based on poly(lactic acid) (PLA), polyethylene glycol (PEG), and polyamides [25].

2.1. The Medical Application of Polymeric Materials with Antibacterial Activity

The polymers with antibacterial activity are applied in medicine as drug carriers. The encapsulation of the drugs into micelles, nanogels, or vesicles [26] not only allows it to curb the bacterial resistance to antibiotics, but also increases the bioavailability of the drug compared to the same conventional antibiotic.
For medical purposes, natural (alginate), artificial (CTS, ethyl cellulose (ET)), and synthetic (PCL—poly(epsilon-caprolactone), PDLA—poly(D-lactide), PGA—poly(glycolide), PLA, PLGA—poly(lactic-co-glycolic acid) polymers are used. The choice of polymer applied as a drug carrier is determined by required residence time and administration site in human cells [27]. In addition, the toxicity and tolerance of the polymer carriers in the relevant cell type is assessed.
The polymers applied in drug delivery systems should have hydrolytically or enzymatically cleavable chemical bonds that provide biodegradability in the body, although the non-biodegradable polymers such as polymethacrylates also constitute a promising alternative [28][29][30].
One of the most innovative drug delivery approaches involve the polymer nanoparticles (NPs) [31]. The NPs, due to their nanometric dimensions (1–100 nm), are easily accessible to cells and tissues, and deliver a drug straight to the site of action in the human body [32]. The NPs are synthetised in the form of nanospheres or nanocapsules. The main difference between them relies on the placement of the carried drug and the mechanism of drug incorporation. The nanospheres are colloidal particles, which adsorb the drug molecule on the particle surface, while the nanocapsules take the form of surrounded polymer shell vesicles with the core filled with aqueous or oily liquid in which the drug is dissolved [33]. Among the nanocapsules, dendrimers, micelles, liposomes, and polymersomes are used as nanoparticles to deliver drugs, including antimicrobials [34]. Lipid constructs called liposomes are composed of bilayers made of amphipathic lipids. Natural liposomes can be found and isolated from the cells, but synthetic liposomes also are commercially available. Due to the presence of the aqueous phase inside and between the lipid bilayers, they can deliver both lipophilic and hydrophilic drugs to human cells [35].
Dendrimers, synthetised for the first time in 1978 by Vögtle et al., are the smallest among the NPs, with a diameter between 1–10 nm. They are obtained in the reaction of protection–deprotection synthesis of the hyperbranched macromolecules, followed by the elongation of the bioactive site from the multifunctional core molecules [36]. Dendrimers are particularly interesting as drug carriers due to their amphiphilic structure, globular shape, low dispersity, and highly branched three-dimensional structure [37]. The bioactive sites of dendrimers are formed by their surface functional groups, and can be modified with biologically active antimicrobial groups, which provide antimicrobial activity to the polymer [38]. The interaction between the bacterial cell and the modified dendrimer surface take place through the electrostatic interactions. Negatively charged bacterial cells interact with positively charged dendrimer functional groups, increasing the permeability of the cell membrane and contributing to the biocidal effect [39]. PAMAM—poly(amido)amine, dendritic polylysine, and polypropylenimine (PPI) [40] are the most popular, commercially available dendrimers with a cationic surface.
Polymersomes are amphiphilic bilayer vesicles made of tri- or di-copolymer blocks, whose properties are crucial for the overall features of the obtained vesicle. In comparison to liposomes, the polymersomes exhibit greater structural and mechanical stability [41]. However, the mechanism of drug transportation is similar—the water-soluble molecules are carried in the inner space of vesicle while the hydrophobic molecules are transferred in the bilayer [42].
Polymer micelles are vesicles with a lipophilic core in which only the hydrophobic drugs can be encapsulated, and a hydrophilic shell ensures water solubility of the entire particle. On the contrary to polymersomes, micelles are not able to transport the hydrophilic drugs [43]. Figure 3 shows a schematic representation of various drug nanocarriers.
Figure 3. The schemes of polymer nanoparticles. Blue colour represents the hydrophilic part of the NPs; red colour represents the hydrophobic one. The space in which the drugs are introduced is marked in yellow.
Polymer nanoparticles in combination with antibiotics can also be used as synergistic or additive agents to chemically or physically weaken the bacteria via the use of elevated temperatures or the formation of reactive oxygen species. Gold NPs, thanks to their high photothermal efficiency in the presence of near-infrared radiation, are of particular clinical interest because exposure of bacteria to temperatures in the range of 45–50 °C causes a strong antibacterial effect in the body in the form of an increase in the level of cytokines and the body’s cellular immune response. The photothermal NPs may be incorporated in the structures of microneedle (MN) arrays, enhancing the antibiotic delivery directly to the site of infection. MNs are obtained from soluble polymers, making them suitable for delivering antibiotics in a humid environment, and providing high local concentrations of antibiotics to infected cells of the human body. Among the antibiotics successfully delivered by the MNs are vancomycin, polymyxin, tetracycline, chloramphenicol, clindamycin, cephalexin, doxycycline, and gentamicin. Further examples of NPs exhibiting a synergistic effect with antibiotics are tetracycline, chloramphenicol, and rifampicin, which are N-alkylaminated CTS NPs that showed high efficiency against Gram-negative bacteria (E. coli, S. typhimurium). In comparison to metal NPs, these natural mucopolysaccharide NPs are considered more biocompatible and biodegradable, but at a concentration higher than 200 mg/L, chitin nanoparticles exhibit cytotoxic properties [44]. Photothermal antibacterial treatment gained recognition due to the reduction of side effects in tissues, low toxicity, high selectivity, and the lack of drug resistance [45]. One of the main disadvantages of photothermal treatment of bacterial infections is the necessity for application at a high temperature to make this treatment efficient against drug-resistant bacteria. In order to eliminate this issue, the photothermal treatment can be replaced with chemodynamic therapy with a synergistic effect. The environment of the infection site is characterized by a low pH and overexpression of H2O2, which allows for precise targeting of the drug. The application of the silver-doped polyoxometalate (AgPOM) injectable in situ hydrogels are one of the most direct infection-targeting methods, featuring good tissue adhesion, a long-lasting effect, good repeatability, and great photothermal performance [46]. Wounds infected with bacteria can also be successfully treated by using hydrogels prepared from poly(aspartic acid) modified with a quaternary ammonium compound/boronic acid cross-linked with poly(vinyl alcohol) polymers. These hydrogels reduced epidermal bacterial survival to 2.3% with an optimal healing rate of 92% after 7 days [47].

2.2. Polymer Materials as Antifouling Agents

Apart from drug delivery and antibacterial treatments, the natural and synthetic nanocomposites are applied in cancer therapy, dental applications, and tissues engineering [48].
Bacterial infections, apart from mechanical damage, are one of the main causes of transplant failures. Polymeric biomaterials are often used as an antibacterial surface in regenerative medicine, and as the coating for medical implants that prevent bacterial biofilm generation. Bacterial biofilm formation is a defence mechanism against host immune cells, ensures chronicity of infection, and is initiated by a bacterial recognition known as quorum sensing. According to statistics, as much as 80% of clinical infections in humans are caused by biofilms. The presence of biofilm is often observed on the surface of orthopedic screws made of stainless steel and titanium. A mechanism of biofilm formation is presented in Figure 4. The first step is the bacteria;s adhesion to the surface. At a distance of about 50 nm from the implant surface, bacterial cells are attracted by Van der Waals forces. At a distance of 20 nm, between the bacteria and the implant, electrostatic repulsive forces occur depending on the interaction between the surface and the usually negatively charged bacteria. At 5 nm from the surface, the strongest Van der Waals and electrostatic forces, as well as hydrophobic and site-specific interactions, begin to occur. After adhesion to the surface, bacteria start to proliferate and grow, producing extracellular polymeric substances that help them capture nutrients and improve their survivability. Due to the cell–to–cell communication in biofilm, bacteria are able to adapt to environmental conditions and colonize new surfaces. After biofilm maturation, some of it dissipates, releasing floating bacteria that can redeposit on the surface [49][50][51].
Figure 4. The scheme of biofilm formation. Arrow describes the circulation of bacteria cells.
Among the polymeric materials that can be helpful in preventing biofilm formation are antifouling polymers, which repel the bacteria from the surface with chemical or physical mechanisms, and antibacterial ones, e.g., peptide mimetic polymers and cationic polymers [52]. Within the antifouling agents, the surfaces functionalised with hydrophilic, zwitterionic, and superhydrophobic polymers should be listed [53]. The feature of hydrophilic polymers is their favourable interaction with water, which provides them good solubility and swellability [54]. In transplantology, hydrophilic polymers are of great interest due to their ability to mimic the properties of natural cartilage [55]. Poly (ethylene glycol) (PEG) and poly (acrylamide) (PAM) (Figure 5) are popular representatives of such hydrophilic polymers [56].
Figure 5. The condensed formulas of PEG (a) and PAM (b).
An alternative to hydrophilic polymers such as PEG is zwitterionic polymers, which shows better antifouling properties [57]. Zwitterionic polymers also exhibit significant chemical and thermal stability, and excellent biocompatibility even in complex surroundings, e.g., serum or blood [58]. The structure of these polymers mimics natural compounds occurring in human cells, such as glycine betaine [59]. The repeating constitutional units of zwitterionic polymers contain both negative and positive charges which make them electrically neutral and hydrophilic; furthermore, the entire network of such a polymer exhibits the same characteristics (electrical neutrality and hydrophilicity) [60]. On the other hand, the hydrophilicity of zwitterionic polymers is one of their greatest disadvantages, as it leads to a strong absorption of water. High solubility in water and susceptibility to hydrolysis limit their ability to form a film, and as a result, it curbs the use of these polymers as antifouling agents [61]. To overcome this difficulty, cross-linking molecules such as polydimethylsiloxane (PDMS) are used to form thin zwitterionic films [62].
The most popular zwitterionic polymers are polybetaines. The positive charge in their monomeric units is provided by a quaternary ammonium group, while the negative one is related to the presence of the anionic groups such as sulfonates, carboxylates, phosphonates, phosphates, and phosphinates. According to the charge distribution mode, apart from polybetaines, among the zwitterionic materials, the polyampholytes are also distinguished. The main difference between polybetaines and polyampholytes is the position of the charge. Polybetaines have both cationic and anionic groups located on the same monomer unit separated by an alkyl chain, while polyampholytes have their negative and positive charges situated on different monomer units [63].
Another type of bacteria-repelling molecule is superhydrophobic polymers, inspired by lotus leaves, covered by hydrophobic wax. The fluorinated silica-colloid-based surfaces are an example of superhydrophobic polymers exhibiting antiadhesive activity towards S. aureus and P. aeruginosa [64]. There is also growing interest around titanium-based materials. Their superhydrophobic properties, bioavailability, and favourable mechanical properties make them useful for cardiac implants [65]. In orthopaedics and dentistry, the magnesium alloy coated by hydroxyapatite (HA) and stearic acid confer great antibacterial adhesion capacity [66]. The polymers applied in regenerative medicine and tissue engineering must cope with the changes of the extracellular environment that accompany physiological and pathological processes.
Chemically synthesized materials that mimic the extracellular matrix (ECM) appear to be a promising approach to imitate the biological activity of cells [67]. The ECM are mostly composed of proteins that perform essential functions in biological processes such as enzymatic reactions, immunological response, cells motility, or signal transduction [68]. Thus, protein-mimetic polymers offer hope for accessing complex natural mechanisms. The amphiphilic polymers imitating antimicrobial peptides (AMPs) are highly efficient in preventing biofilm formation. One example of AMPs is photoresponsive AMP based on the N-substituted glycine skeleton, which—due to its efficiency, controllability, and high selectivity—has been used in hydrogels and antifouling surfaces [69].
Another kind of polymeric material used as an antifouling agent are the cationic polymers mentioned above, which have been proved to exhibit excellent antibacterial properties. In implantology, polyurethane catheters are often used as implantable medical instruments. Unfortunately, their surface is susceptible to the adhesion of bacteria, which necessitates their frequent replacement in order to prevent bacterial infection. To thwart the formation of a biofilm on polyurethane catheters, the surface modification with cationic polymers can be applied. For this purpose, quaternary ammonium compounds or metal ions are used [70].
Some examples of a quaternized compound are benzophenone-based esters and benzophenone quaternary amides, which can be cross-linked on surfaces upon UV radiation. These coatings are efficient against the methicillin-resistant Staphylococcus aureus (MRSA), fluconazole-resistant Candida albicans spp., and influenza virus with 100% efficiency [71]. A modified (quaternized or alkylated) polyethyleneimine (PEI) is a cationic polymer containing amino- and imino-groups, also known for its antibacterial properties. PEI is positively charged in neutral and basic solutions, having a high zero potential point at pH values up to 10 [72].
The cationic polymers can also be obtained in the innovative reaction of photopolymerization. This method is applied to prepare Sulphur-containing polymers, whose monomers are ionized giving a positive charge, which provides the polymer with antifouling properties [73].

References

  1. Nasrollahi, N.; Vatanpour, V.; Khataee, A. Removal of antibiotics from wastewater by membrane technology: Limitations; successes; and future improvements. Sci. Total Environ. 2022, 838, 156010.
  2. Korzeniewska, E.; Harnisz, M. Sources; occurence and envrionmental risk assessment of antibiotics and antimibrobial-resistant bacteria in aquatic environments of Poland. Pol. River Basins Lakes—Part II 2020, 87, 179–193.
  3. Mpatani, M.M.; Aryee, A.A.; Kani, A.N.; Han, R.; Li, Z.; Dovi, E.; Qu, L. A review of treatement techniques applied for selective removal of emerging pollutant-trimethoprim from aqueous systems. J. Clean. Prod. 2021, 308, 127359–127379.
  4. Chen, J.F.; Yang, Y.Y.; Ke, Y.C.; Chen, X.L.; Jiang, X.S.; Chen, C.; Xie, S.G. Anaerobic sulfamethoxazole-degrading bacterial consortia in antibiotic-contaminated wetland sediments identified by DNS-stable isotope probing and metagenomics analysis. Environ. Microbiol. 2022, 24, 3751–3763.
  5. Omuferen, L.O.; Maseko, B.; Olowoyo, J.O. Occurrence of antibiotics in wastewater from hospial and convectional wastewater treatment plants and their impack on the effluent receiving rivers: Current knowledge between 2010 and 2019. Environ. Monit. Assess. 2022, 194, 306–331.
  6. Swaminathan, P.; Sen, S.; Mandira, M.A.; Prasad, A.A. Compounds and methods to resesitize Antibiotic-resistant bacteria Mediterr. J. Infect. Microbes Antimicrob. 2021, 10, 46–53.
  7. Russel, J.N.; Yost, C.K. Anternative; environmentally conscious approaches for removing antibiotics from wastewater treatment systems. Chemosphere 2021, 263, 1281177–1281187.
  8. Balloudj, J.; BhuvaneswariAssadi, I.; Nasrallah, N.; Jery, A.L.; Khezami, L.; Assadi, A.A. Simultaneus removal of antibiotics and inactivation of antibiotic-resistant bacteria by photocatalysis: A review. J. Water Process. Eng. 2021, 42, 102089–1021100.
  9. Cheng, D.; Ngo, H.H.; Guo, W.; Chang, S.W.; Nguyen, D.D.; Liu, Y.; Shan, X.; Nghiem, L.D.; Nguyen, L.N. Removal process of antibiotics during anaerobic treatment of swine wastewater. Bioresour. Technol. 2020, 300, 122707–122739.
  10. Monohan, C.; Nag, R.; Morris, D.; Cummins, E. Antibiotic residues in aquatic environment-current perspective and risk consideration. J. Environ. Sci. Health A 2021, 56, 733–751.
  11. Wu, C.; Ruan, L.; Yau, L. Tracking Epidemiological Characteristics and Risk Factors of Multi-Drug Resistant Bacteria in Intensive Care Units. Infect. Drug Resist. 2023, 16, 1499–1509.
  12. Fahad, K.H.; Al-Muhana, B.M.M.; Sadiq, J.N.N. Beneficial antimicrobial treatment options forpan-drug-resistant bacterial species. J. Agric. Biol. Sci. 2023, 14, 45–51.
  13. Geetha, M.; Rajendran, I.; Jayakumar, T.; Dhayalan, S. Prevalence and Detection of Multidrug Resistance Bacterial Strains Isolated from the Different Inanimate Surfaces of the Hospital Environment. Uttar Pradesh J. Zool 2023, 44, 95–104.
  14. De Kraker, M.A.D.; Stewardson, A.J.; Harbarth, S. Will 10 Million People Die a Year due to Antimicrobial Resistance by 2050? PLoS Med. 2016, 13, 1–6.
  15. Chen, J.; Wang, F.; Liu, Q.; Du, J. Antibacterial polymeric nanostructures for biomedical applications. Chem. Commun. 2014, 93, 14482–14493.
  16. Rofeal, M.; Abdelmalek, F.; Steinbüchel, A. Naturally—Sourced Antibacterial Polymeric Nonomaterials with Special Reference to Modified Polymer Variants. Int. J. Mol. Sci. 2022, 23, 4101.
  17. Arora, A.; Mishra, A. Antibacterial Polymers—A Mini Review. Mater. Today Proc. 2018, 5, 17156–17161.
  18. Yu, K.; Mei, Y.; Hadjesfandiari, N.; Kizhakkedathu, J.N. Engineering biometrials surfaces to modulate the host response. Colloids Surf. B 2014, 124, 69–79.
  19. Deka, S.R.; Sharma, A.K.; Kumar, P. Cationic polymers and their self assembly for antibacterial applications. Curr. Top Med. Chem. 2015, 15, 1179–1195.
  20. Xue, Y.; Xiao, H.; Zhang, Y. Antimicrobial Polymeric Materials with Quaternary Ammonium and phosphonium salts. Int. J. Mol. Sci. 2015, 16, 3626–3655.
  21. Jain, A.; Duvvuri, L.S.; Farah, S.; Beyth, N.; Domb, A.J.; Khan, W. Antimicrobial Polymers. Adv. Healthc. Mater. 2014, 3, 1969–1985.
  22. Zhang, H.; Chiao, M. Anti-fouling Coatings of Poly(dimethylosiloxane) Devices for Biological and Biomedical Applications. J. Med. Biol. Eng. 2015, 35, 143–155.
  23. Jeepery, I.F.; Sudesh, K.; Abe, H. Miscibility and enzymatic degradability of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate)-based polyester blends by PHB depolymerase and lipase. Polym. Degrad. Stab. 2021, 192, 109692–109704.
  24. El-malek, F.A.; Steinbüchel, A. Post-synthetic enzymatic and chemical modifications for novel sustainable polyesters. Front. Bioeng. Biotechnol. 2022, 9, 1460.
  25. Zhao, H.; Lin, Z.Y.; Yildirimer, L.; Dhinakar, A.; Zhao, X.; Wu, J. Polymer-based nanoparticles for protein delivery: Design; strategies and applications. J. Mater. Chem. 2016, 4, 4060–4071.
  26. Liu, Y.; van der Mei, H.; Zhao, B.; Zhai, Y.; Cheng, T.; Li, Y.; Zhang, Z.; Busscher, H.; Ren, Y.; Shi, L. Eradication of Multidrug—Resistant Staphylococcal Infections by Light—Activable Micelar Nanocarriers in a Murine Model. Adv. Funct. Mater. 2017, 27, 1701974.
  27. Imperiale, J.; Acosta, G.; Sosnik Alejandro, B. Polymer-based carriers for ophthalmic drug delivery. J. Control. Release 2018, 285, 106–141.
  28. Pignatello, R.; Bucolo, C.; Puglisi, G. Ocular tollerability of Eudragit@ and RL100@ nanosuspensions as carriers for ophthalamic controlled drug delivery. J. Pharm. Sci. 2002, 91, 2636–2641.
  29. Adibkia, K.; Shadbad, M.R.S.; Nokhodchi, A.; Javadzedeh, A.; Barzegar-Jalali, M.; Barar, J.; Omidi, Y. Piroxicam nanoparticles for ocular delivery: Physicochemical characterization and implemantation in endotoxin—Induced uveitis. J. Drug Target. 2007, 15, 407–416.
  30. Rathod, L.V.; Kapadia, R.; Sawant, K.K. A novel nanoparticles impregnated ocular insert for enhanced bioavailability to posterior segment of eye: In vitro; in vivo and stability studies. Mater. Sci. Eng. C 2017, 71, 529–540.
  31. Calzoni, E.; Caseretii, A.; Polchi, A.; Di Michele, A.; Tancini, B.; Emiliani, C. Biocompatible Polymer Nanoparticles for Drug Delivery Applications in Cancer and Neurodegenerative Disorder Therapies. J. Funct. Biomater 2018, 10, 4.
  32. Torchilin, V.P. Structure and design of polymeric surfactant-based drug delivery systems. J. Control. Release 2001, 73, 137–172.
  33. Letchford, K.; Burt, H. A review of the formation and classification of amphiphilic block copolymer nanoparticulate structures: Micelles; nanospheres; nanocapsules and polymersomes. Eur. J. Pharm. Biopharm. 2007, 65, 259–269.
  34. Musyanovytch, A.; Landefster, K. Polymer Micro- and Nanocapsules as biological carriers with multifunctional properties. Macomol. Biosci. 2014, 14, 458–477.
  35. Hallaj-Nezhadi, S.; Hassan, M. Nanoliposome—Based antibacterial drug delivery. Drug Deliv. 2015, 22, 581–589.
  36. Buhleier, E.; Wehner, W.; Vögtle, F. Cascade and Nonskid—Chain—Like Syntheses of Molecular Cavity Topologies. Synthesis 1978, 2, 155–158.
  37. Svenson, S.; Tomalia, D.A. Dendrimers in biomedical applications—Reflections on the field. Adv. Drug Deliv. Rev. 2005, 15, 2106–2129.
  38. Chen, C.Z.S.; Cooper, S.L. Recent Advances in Antimicrobial Dendrimers. Adv. Mater. 2000, 12, 843–846.
  39. Neu, H.C. The crisis in antibiotic resistance. Science 1992, 257, 1064–1073.
  40. Mintzer, M.A.; Dane, E.L.; O’Toole, G.A.; Grinstaff, M.W. Exploiting Dendrimer Multivalency to Combat Emerging and Re-Emerging Infectious Diseases. Mol. Pharm. 2012, 9, 342–354.
  41. Thiele, A.R.; Abate, H.C.; Shum, S.; Bachtler, S.; Feorster, D.A. Weitz Fabrication of Polymersomes using Double-Emulsion Templates in Glass-Coated Stamped Microfluidic Devices. Small 2010, 6, 1723–1727.
  42. Mueller, W.; Koyonov, K.; Fischer, K.; Hartmann, S.; Pierrat, S.; Baschee, T.; Maskos, M. Hydrophobic Shell Loading of PB-b-PEO Vesicles. Macromolecules 2008, 42, 357–361.
  43. Wang, T.; Qin, J.; Cheng, J.; Li, C.; Du, J. Inteligent design of polymersomes for antibacterial and anticancer applications. Wiley Interdiscip. Rev. Nanomed. Nanotechnol. 2022, 14, e1822.
  44. Kaur, M.; Cohen, Y.; Poverenov, E.; Eltzov, E. Binding selectivity of N-alkylaminated modified chitosan nanoparticles produce a synergistic antibacterial effect against gram-negative strains. React. Funct. Polym. 2023, 186, 105567.
  45. Ziesmer, J.; Larsson, J.V.; Sotiriou, G.A. Hybrid microneedle arrays for antibiotic and near-IR photothermal synergistic antimicrobial effect against Methicillin-Resistant Staphylococcus aureus. J. Chem. Eng. 2023, 462, 142127.
  46. Juang, H.; Su, Y.; Wang, C.; Lei, B.; Song, X.; Wang, W.; Wu, P.; Liu, X.; Dong, X.; Zhong, L. Injectable Tissue-Adhesive Hydrogel for Photothermal/ Chemodynamic Synergistic Antibacterial and Wound Healing Promotion. Appl. Mater Interfaces 2023, 15, 2714–2724.
  47. Li, W.; Cai, J.; Zhou, W.; Zhao, X.; Wang, M.; Zhou, X.; Ren, L. Poly(aspartic acid)-based self-healing hydrogel with precise antibacterial ability for rapid infected-wound repairing. Colloids Surf. B 2023, 221, 112982.
  48. Feldman, D. Polymer nanocomposytes in medicine. J. Macromol. Sci A 2016, 53, 55–62.
  49. Escobar, A.; Muzzio, N.; Moya, S. Antibacterial Lyer—By—Lyer Coating for Medical Implants. Pharmaceutics 2021, 13, 16.
  50. Quinn, J.; McFadden, R.; Chan, C.-W.; Carson, L. Titanium for Orthopedic Applications: An Overview of Surface Modification to Improve Biocompatibility and Prevent Bacterial Biofilm Formation. IScience 2020, 23, 101745.
  51. Goudarzi, M.; Navidinia, M.; Khadembashi, N.; Rasouli, R. Biofilm Matrix Formation in Human: Clinical Significance, Diagnostic Techniques, and Therapeutic Drugs. Arch. Clin. Infect. Dis. 2021, 16, 107919–107929.
  52. Francolini, I.; Piozzi, A. Polymeric Systems as Antimicrobial or Antifouling Agents. Int. J. Mol. Sci. 2019, 20, 4866.
  53. Phuong, T.; Nguyen, T. Polymer and Surface Modifications for Antibacterial Purposes. Ph.D. Thesis, Université Paris Saclay, Paris, France, 2019.
  54. Shmidt Bernhard, V.K.J. Hydrophilic polymers. Polymers 2019, 11, 693.
  55. Zhang, C.; Chen, J.; Liu, M.; Liu, Y.; Liu, Z.; Chu, H.; Cheng, Q.; Wang, J. Regulation mechanism of biomolecule interaction behaviors on the superlubricity of hydroplilic coatings. Friction 2022, 10, 94–109.
  56. Brunzel, M.; Majdanski, T.C.; Vitz, J.; Nischang, I.; Schubert, U.S. Fast Screening of Diol Impurities in Methoxy Poly(Ethylene Glycol)s (mPEG)s by Liquid Chromatography on Monolithic Silica Rods. Polymers 2018, 10, 1395.
  57. Wu, C.; Zhou, Y.; Wang, H.; Hu, J. P4VP Modified Zwitterionic Polymer for the Preparation of Antifouling Functionalized Materials. Nanomaterials 2019, 9, 706.
  58. Zheng, J.; Li, L.; Tsao, H.K.; Sheng, Y.J.; Chen, S.; Jiang, S. Strong Repulsive Forces between Protein and Oligo (Ethylene Glycol) Self-Assembled Monolayers: A Molecular Simulation Study. Biophys. J. 2005, 89, 158–166.
  59. Chen, C.H.; Chen, S.H.; Mao, S.H.; Tsai, M.J.; Chou, P.Y.; Liao, C.H.; Chen, J.P. Injectable thermosensitive hydrogel containing hyaluronic acid and chitosan as a barrier for prevention of postoperative peritoneal adhesion. Carbohydr. Polym. 2017, 173, 721–731.
  60. Sun, Q.; Su, Y.; Ma, X.; Wang, Y.; Jiang, Z. Improved antifouling property of zwitterionic ultrafiltration membrane composed of acrylonitrile and sulfobetaine copolymer. J. Membr. Sci. 2006, 285, 299–305.
  61. Shafi, H.Z.; Khan, Z.; Yang, R.; Gleason, K.K. Surface modification of reverse osmosis membranes with zwitterionic coating for improved resistance to fouling. Desalination 2015, 362, 93–103.
  62. Leigh, B.L.; Cheng, E.; Xu, L.; Derk, A.; Hansen, M.R.; Guymon, C.A. Antifouling Photograftable Zwitterionic Coatings on PDMS Substrates. Langmuir 2019, 35, 1100–1110.
  63. Racovita, S.; Trofin, M.A.; Loghin, D.F.; Zaharia, M.M.; Bucatariu, F.; Mihai, M.; Vasiliu, S. Polybetaines in Biomedical Applications. Mol. Sci. 2021, 22, 9321.
  64. Privett, B.J.; Youn, J.; Hong, S.A.; Lee, J.; Han, J.; Shin, J.H.; Schoenfich, M.H. Antibacterial Fluorinated Silica Colloid Superhydrophobic Surfaces. Langmuir 2011, 27, 9597–9601.
  65. Manivasagam, V.K.; Perumal, G.; Arora, H.S.; Popat, K.C. Enhanced antibacterial properties on superhydrophobic micronano structured titanium surface. J. Biomed. Mater. 2022, 110, 1314–1328.
  66. Qianqian, Q.; Xiaogang, B.; Jine, S.; Shu, C.; Yao, X.; Yuan, Y.; Jia, L.; Guohua, X. Fabrication of superhydrophobic composite coating of hydroxyapatite/stearic acid on magnesium alloy and its corrosion resistance; antibacterial adhesion. J. Mater. Sci. 2020, 56, 5233–5249.
  67. Oki, Y.; Kirita, K.; Ohta, S.; Ohba, S.; Horiguchi, I.; Sakai, Y.; Ito, T. Switching of Cell Proliferation/Differentiation in Thiol−Maleimide Clickable Microcapsules Triggered by in Situ Conjugation of Biomimetic Peptides. Biomacromolecules 2019, 20, 2350–2359.
  68. Hasoyama, K.; Lazurko, C.; Muñoz, M.; Mc Tiernan, C.D.; Alarcon, E.I. Peptide Based Functional for Soft-Tissue Repair. Front. Bioeng. Biotechnol. 2019, 7, 205.
  69. Lin, M.; Ding, J.; Sun, J. Photo-triggered polymeric antimicrobial peptide mimics with excellent selectivity and antifouling and antimicrobial hydrogels. Giant 2022, 10, 100097.
  70. Liu, F.; Qu, W.; Zhang, J.; Liu, J.; Zhu, Q.; Yue, T.; Xu, X.; Ma, N.; Ma, J.; Sun, Y.; et al. Cationic Alternating Polypeptide Fixed on Polyurethane at Multiple Sites for Excellent Antibacterial and Antifouling Properties. Langmuir 2021, 37, 10657–10667.
  71. Ghosh, S.; Mukherjee, R.; Basak, D.; Haldar, J. One-Step Curable; Covalently Immobilized Coating for Clinically Relevant Surfaces That Can Kill Bacteria; Fungi; and Influenza Virus. Appl. Mater. Interfaces 2020, 12, 27853–27865.
  72. Bai, Z.; Liu, Q.; Zhang, H.; Liu, J.; Yu, J.; Wang, J. A novel 3D reticular anti-fouling bio-adsorbent for uranium extraction from seawater: Polyethylenimine and guanidyl functionalized hemp fibers. J. Chem. Eng. 2019, 382, 122555.
  73. Xu, X.; Wang, Q.; Chang, Y.; Peng, H.; Whittaker, A.K.; Fu, C. Antifouling and Antibacterial Surfaces Grafted with Sulfur-Containing Copolymers. Appl. Mater. Interfaces 2022, 14, 41400–41411.
More
Information
Subjects: Polymer Science
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 369
Revisions: 2 times (View History)
Update Date: 04 Jul 2023
1000/1000
Video Production Service