Preprint
Review

This version is not peer-reviewed.

The Art of PEGylation: "From Simple Polymer to Sophisticated Drug Delivery System"

A peer-reviewed article of this preprint also exists.

Submitted:

03 February 2025

Posted:

04 February 2025

You are already at the latest version

Abstract

The development of effective drug delivery systems is a major challenge in cancer therapy, gene therapy, and infectious disease treatment because of its low bioavailability, rapid clearance, and toxicity towards non-targeted healthy tissues. This review discusses how PEGylation, the covalent attachment of poly(ethylene glycol) (PEG), enhances the pharmacokinetic profiles of the drug-containing nanosystems through the "stealth effect" that avoids immune system detection and improves circulation times in different nano-delivery systems. The review provides an overview of the synthetic methods of PEG derivatives, their conjugation with nanoparticles, proteins, and drugs, and their characterization using modern analytical tools. The paper explores various PEGylation strategies, including covalent conjugation and self-assembly, and discusses the influence of PEG chain length, density, and conformation on drug delivery efficiency. Despite its advantages, there are several challenges associated with PEGylation such as the immunogenicity of anti-PEG responses, the potential for accelerated clearance of PEGylated drugs, reduced therapeutic efficacy, and the possibility of allergic reactions. Consequently, the balance between the benefits of PEGylation and its immunogenic risks remains a critical area of investigation.

Keywords: 
;  ;  ;  ;  

1. Introduction

Today, cancer remains one of the biggest challenges in medicine, the cure of which is associated with great difficulties, despite the current development in chemotherapy, radiation and targeted therapy. The major obstacle to cancer treatment is the effective delivery of therapeutic drugs so that healthy cells are not damaged. Many anticancer drugs are characterised by poor bioavailability, and rapid clearance, which limits their therapeutic effect. Consequently, nano-based drug delivery systems have gained attention which offers efficient drug delivery and are characterised by non-toxicity. Nanoparticles have attracted special attention due to their biocompatibility, biodegradability and ability to encapsulate hydrophobic drugs which overcomes drug limitations. However, when nanosystems are in a physiological environment, they interact with the surrounding biological components, such as proteins, lipids and other biomolecules, which significantly influence their overall behaviour [1]. PEGylation improves the pharmacokinetic properties of NPs by enhancing their stability [2], circulation time, and avoids immune system detection by a phenomenon known as the "stealth effect". Many anticancer reagents, especially those of a hydrophobic nature, such as docetaxel, paclitaxel, and doxorubicin (DOX) are characterised by low bioavailability, rapid metabolism and high systemic toxicity, which prohibits their therapeutic activity. NP PEGylation [3] or direct drug PEGylation [4] has been shown to enhance the effectiveness and safety of those drugs. Recent reviews have discussed important elements of PEGylation. For instance, Gao et al. report the clinical significance of PEGylation technology and highlight the benefits and uses of PEGylated treatments in medicine [5]. Belen et al. provide an analysis of protein PEGylation techniques, emphasising site-specific PEGylation for proteins [6]. Similarly, Li et al. discuss recent progress in the development of PEGylated therapeutic proteins and peptides [7]. Moreover, recent works have focused on the challenges of PEGylation, such as the immunogenicity in PEGylated products [8,9]. PEGylation offers numerous benefits, however, in some cases, it also has notable drawbacks, one significant issue is the immunogenic response to PEG, referred to as 'anti-PEG antibodies' (APA) which can limit NPs efficacy and safety. The review provides a comprehensive analysis of PEGylation for drug delivery systems starting from synthetic strategies of various PEG derivatives, their applications in nanoparticle and protein PEGylation, and their pharmacokinetic implications. In general, a combination of nanoparticle-based delivery systems with PEGylation results in better drug delivery profiles and the goal of this review is to demonstrate the benefits which is associated with PEGylation generally for cancer treatment.

2. Functionalization on PEG

PEG is a hydrophilic, biocompatible polymer widely used in biomedical and pharmaceutical applications because of its unique physicochemical properties, such as high solubility in water and organic solvents, non or low toxicity, and low immunogenicity, cryoprotective [10], etc. The terminal hydroxyl group of PEG OH-(CH2CH2O)n-H, can be modified with various chemical moieties (Table 1). Unmodified PEG is biologically inert and lacks functional groups that enable it to interact specifically with biological targets, while functionalized PEG can be covalently attached to drugs [11], proteins [12], and other biomolecules that are widely used for constructing hydrogels, nanoparticles, and other materials that benefit from PEG’s antifouling and stabilizing properties. Functional PEGs can be divided into homobifunctional and heterobifunctional types. Homofunctional PEG can refer to both monofunctional (one functional group at one end) or bifunctional PEG derivatives. The key characteristic of homobifunctional PEGs is identical reactive groups at both ends of the polymer chain, which enables crosslinking, essential in hydrogel preparation [13,14]. Heterobifunctional PEGs, on the other hand, have different reactive groups present on each side that allow selective coupling to different substrates or biomolecules, which is particularly useful for site-specific conjugation.

2.1. Synthesis and Functionalization of Amine and Thiol-Terminated PEGs

Amine PEGs contain NH₂ groups at one or both ends, the polymer is suitable for conjugation with drugs, proteins, or other molecules. The first step of NH2-PEG synthesis is the activation of terminal hydroxyl groups which is performed using halogenation [11,15] or sulfonylation [15,16,17,18] by tosylation or mesylation through reacting with tosyl (Ts) and mesyl (Ms) chlorides, producing activated PEGs which are a precursor for functionalized PEG synthesis, those contain effective leaving groups for nucleophilic substitution with amines and thiols (Figure 1). The sulfonylation reaction requires bases such as pyridine [15] and triethylamine (TEA) [15,19] as a hydrochloric acid acceptor. Activated PEGs are treated with ammonia [20,21,22,23] forming NH2-PEGs. However, by-products such as secondary/tertiary amines can be produced as side reactions, which reduce the yield of PEG-amine and complicate the purification of the final product. On the other hand, sulfonylated PEGs can then be further modified into PEG azide by reacting with NaN3, which is an intermediate in the synthesis of NH2-PEGs using the Staudinger reaction [20,24,25] generally leading to a product with higher purity and yield. Zhang et al. reported the synthesis of homofunctional and bifunctional amino-PEGs using 3 step methodologies: 1) mesylation 2) the substitution reaction of activated mPEG-OMs with (Boc)2NH to obtain mPEG-N(Boc)2 followed by deprotection of the Boc groups with trifluoroacetic acid to obtain mPEG-NH2 finally, the reaction yielded in 95–99% purity [26]. Thiol-functionalized PEGs are useful for creating self-assembling structures on gold surfaces and other metal surfaces and for protein PEGylation through Michael-addition reactions. The simplest method to introduce a thiol group into PEG is through nucleophilic substitution between PEG-OTs and a thiol nucleophile namely sodium hydrosulfide [27] or through an esterification reaction between mercapto acids, such as mercaptoacetic acid [28], mercaptohexadecanoic acid [29], 3-mercaptopropionic acid and a PEG-OH in the presence of sulfuric acid [30,31,32]. Mulay et al. described an environmentally friendly approach for PEG-thiol synthesis using an enzymatic process by transesterification reaction with methyl 3-mercapto propionate in the presence of Candida antarctica lipase B enzyme, the reaction yields were approximately 100% for both PEG1000 and PEG2050 monothiols [33].

2.2. Alkyne Functionalized PEGs

Alkyne PEGs (Figure 2) are widely used in PEGylation reactions, several alkyne functional groups can be attached to PEG molecules like propargyl or strained alkynes, such as dibenzocyclooctyne or bicyclononyne. However, Propargyl PEGs are commonly used in copper-catalyzed azide-alkyne cycloaddition, while strained alkynes are employed for copper-free azide-alkyne cycloaddition also known as SPAAC, which is excellent for in vivo applications as it reduces the potential cytotoxicity associated with copper [34]. SPAAC is a bio-orthogonal reaction and minimizes interference with biological processes, the reaction occurs rapidly under physiological conditions useful for imaging [35,36,37] hydrogel [38,39] and drug delivery systems preparation. During the synthesis of propargyl derivatives, the hydroxyl group of PEG is first activated by a strong base (sodium hydride, potassium hydroxide) to deprotonate the alcohol group followed by a nucleophilic substitution reaction with propargyl bromide [40,41]. Herzberger et al. developed a novel procedure where glycidyl propargyl ether (GPgE) was used as a monomer in anionic copolymerization with ethylene oxide [42], which incorporated propargyl groups into the polymer chain without the need for protection of the alkyne group. Alternatively, Lu et al. used α-hydroxy-ω-carboxyl PEGs for alkyne-PEG synthesis, initially, the carboxyl group was converted to a propargyl group using propargyl bromide, whereas hydroxy ends were modified with amine, mercapto, and hydrazide groups at the other end to synthesize heterobifunctional PEGs. Those heterobifunctional PEG derivatives are useful for bioconjugation applications [43], particularly with azides in copper-catalyzed reactions.
Aldehyde-modified PEGs are often used in conjugating with biologically active molecules such as proteins [44], and antibodies, during the conjugation Schiff bases are formed that can revert to the original aldehyde and amine in the presence of water or under acidic conditions. A common strategy to overcome this problem is to reduce the Schiff base to an amine by reducing the reagent which forms more stable C-N bonds. Only a few literature sources are available for the synthesis of aldehyde-functionalized PEG because of the challenges associated with the low reactivity of the hydroxyl groups. The most commonly used method is the oxidation of hydroxyl groups [45,46]. In 2015, Mauri et al. reported the two-step synthesis of PEG-aldehydes, oxidation methods often struggle with the low reactivity of PEG’s hydroxyl groups and result in low yields with unwanted by-products. In contrast, in the new method (Figure 3), PEG was functionalized with terminal alkenes, followed by ozonolysis and subsequent reduction with dimethyl sulfide which forms chemically stable, reactive PEG-aldehyde derivatives without changing the structural integrity of the polymer [47].
In addition to aldehyde-functionalized PEGs (Figure 4), nitrophenyl-functionalized PEGs are commonly employed in the development of PEGylated systems, which are synthesized by a reaction between nitrophenyl chloroformate (NPC) and PEG hydroxy ends in organic solvents like THF, DMSO, and DCM in the presence of base catalysts, triethylamine [48,49,50,51] or dimethylamino pyridine (DMAP) [52]. Generally, NPC is taken in a slightly excess amount relative to PEG to ensure a high conversion of hydroxyl groups. Acylation reaction between PEG and NPC is typically performed at low temperatures (0–5 °C) which reduces the side reactions, however, the presence of water in the reaction system can trigger secondary reactions resulting in the formation of 4-nitrophenol which is a major by-product of hydrolysis and the purification presents difficulties. The NPC-PEG product is often conjugated with various molecules such as dendrimers [53,54], proteins or micelles, that contain amine groups forming a stable urethane linkage.

3. PEGylation Strategies for Nanosystems

PEGylation of nanosystems can generally be achieved through different methods, including covalent conjugation, physical adsorption and self-assembly. Non-covalent PEGylation is performed by incubating nanoparticles in a PEG solution. However, this simple incubation creates weak interactions between PEG and NPs surfaces and PEG can detach over time, especially under physiological conditions when parameters can fluctuate. In contrast, the covalent attachment of PEG provides a more stable modification. Covalent PEGylation can be accomplished by direct conjugation of PEG molecules to the surface of the nanoparticles, for example, PEG molecules with thiol groups can react with the metal nanoparticle surface like gold or silver with high affinity, forming stable Au-S [70,71], Ag-S [72] bonds. Often PEGylation is performed through a non-direct approach, when PEG is firstly modified into amphiphilic forms as part of di-block copolymers like PEG-PLA, PEG-PCL or even tri-block copolymers [73] also known as pluronics, which forms self-assembled structures on NPs. In our previous work, we reported the synthesis of PEG-grafted poly(ester amide) (PEG-PEA) using a thiol-Michael addition reaction, the NPs were prepared by a solvent-displacement method where the PEG-PEA was used as a surfactant, forming self-assembling structures (Figure 5) which were confirmed by transmission electron microscopy [74].
As a surfactant, PEG can stabilize emulsions, suspensions, and other systems. However, only PEG can not assemble on NP's surface because of its hydrophilic nature, except when in combination with hydrophobic segments (Figure 6). PEG-containing copolymers can self-assemble into well-organized structures due to molecular interactions. In aqueous environments, hydrophobic segments of amphiphilic PEGs tend to avoid water and interact with the core, through Van der Waals forces or electrostatic interactions for the charged polymers [75]. The ratio of hydrophilic and hydrophobic blocks and their molecular weight can be adjusted to control the stability and biological properties of NPs. Ueya et al. showed that increasing the PEG ratio in a PEG-PLGA[78–80 block copolymer decreased the stability of micellar NPs due to the increased hydrophilic nature, by increasing the molecular weight of PLGA, water diffusion into the core was hindered, resulting in enhanced stability [76]. The ratio of hydrophilic/hydrophobic (H/H) blocks of PEG-PLA was shown to affect the cellular internalization of self-assembled NPs in HepG2 tumour cells. The results for the HepG2 line show that clathrin-mediated endocytosis is the predominant pathway for micelles with a 5:5 H/H ratio whereas micelles with H/H ratios closer to 2:8 showed the lowest cellular uptake, which was attributed to the formation of less dense protein corona, which in turn facilitates better cellular uptake [77]. PEG content can significantly influence the drug release profile from nanoparticles, high PEG generally leads to faster and more extensive drug release, due to the increased hydrophilicity of the nanoparticle surface, which can lead to enhanced water uptake and facilitate faster drug release, especially for hydrophilic drugs [78,79,80]. Certain structures, like micelles, form self-assembly structures above a certain concentration known as the critical micelle concentration (CMC), which is the level of surfactant molecules in a solution above which they spontaneously form organized structures, high-CMC surfactants tend to form micelles that are less stable and more susceptible to dissociation when diluted [81]. The PEG-block copolymer surfactants tend to have lower CMC values compared to conventional surfactants. Long hydrophobic blocks lower the CMC [82] because they enhance the tendency for self-assembly by reducing solubility in aqueous environments while in turn a long PEG block increases the CMC, however, a balanced ratio is important to decrease CMC and maintains other physical properties which are necessary for drug delivery purposes.

4. Flory Radius, PEG Chain Length, and Density: Influence on PEG Conformation and Biological Interactions

PEGylation can lead to mushroom and brush structures on the NP surfaces depending on factors such as solvent, molecular weight and density of PEG chains (Figure 7). The conformational regime of PEG can be predicted from the Flory radius using the following relation: Rf​=αN3/5 where N is the number of monomeric units in the PEG polymer chain and α is the length of monomeric unit (0.35 nm), and the exponent value (3/5) serves only for a good solvent, that can change based on factors like molecular weight and solvent type. The grafting density or the average distance between adjacent PEG chains is calculated by the following formula D=(A/π)1/2, where A is an area occupied in nanoparticle occupied per PEG chain, using these two formulas one can predict PEG conformation, whether the PEG adopts a brush-like (D<2Rf) or mushroom-like (D>2Rf) structures. At a higher PEG density, in a brush configuration, the PEG chains are extended and occupy more space due to higher Rf relative to D and vice versa in a mushroom configuration, the chains are more compact and the effective size is smaller. High PEG surface density is generally preferred for biological applications which show biocompatibility, and low protein adsorption [83] inhibiting phagocytic uptake [84,85]. However, as brush conformation avoids phagocytic uptake it does not necessarily guarantee full protection from macrophage uptake, it was shown that a minimal brush PEG coating could not evade THP-1 macrophage endocytosis and sometimes dense-brush conformation is required for an effective reduction in protein adsorption [86] and macrophage uptake [87]. Wang et al. showed the importance of PEG chain size in the comparative study while the terminal PEG density was constantly regulated by changing the PEG-PCL block copolymer molecular weight in PEGn-PCLn/PCL3500 core-shell NPs, increasing PEG length from 10.7 nm (3.4 kDa) to 13.8 nm (5 kDa) improved tumour growth inhibition from 73.4% to 88%, however, further increase in PEG length to 24.5 nm reduced efficacy to 54%. PEG 5 kDa, particularly between PEGs 3.4 kDa and 8 kDa, showed enhanced stealth properties, longer circulation time and reduced macrophage interactions, which indicates that optimizing PEG length is crucial for designing effective drug delivery systems [88]. Gref et al. studied the impact of PEG Mw in PEG-PLA 45 kDa/PLA 40 kDa core-shell NPs on protein adsorption. A steep decrease was observed when increasing PEG Mw from 2000 to 5000 g/mole. PEG 5 kDa was found to be the most optimal for protein resistance and no significant benefits were observed when increasing PEG Mw above 5000 compared to bare PLA 40 kDa NPs. The reduced protein adsorption for PEG content in PEG 5 kDa-PLA 20 kDa/PLA 40 kDa NPs was between 2-5 weight %, however, no complete protection against plasma protein adsorption was achieved [89]. While long PEG chains improve stability and circulation, they can also inhibit cellular uptake of the nanoparticles [90,91,92].

5. Quantification of PEG Surface Density on Nanoparticles

There are several techniques available to study and evaluate PEG density around the NP surface. Quantification of PEG surface density is important to determine the PEG conformational regime, especially in the area of drug delivery, which, as shown before, directly influences NP properties such as stability, solubility, and interaction with the immune system. One of the methods to study the PEG surface density is a fluorescence-based assay, PEG molecules are labelled with a fluorescent dye, and the PEG surface density is measured by the fluorescence intensity of the dye [82,93]. NPs are typically incubated with fluorescent dyes, followed by a washing step to remove unbound labels. The signal is measured, and PEG surface density is calculated from the standard calibration curve.
X-ray photoelectron spectroscopy (XPS) examines the chemical and elemental states of a material surface, and it is commonly used to confirm the successful PEGylation of the NP surface. The existence and relative abundance of oxygen and carbon atoms indicate the presence of PEG molecules on the NP surface to analyze the PEG cloud around the NPs [25,28,29,32,94,95]. XPS is particularly useful for verifying the modification of NPs with PEG and its efficacy in reducing protein adsorption [83,96]. The surface density of PEG can be calculated by comparing the intensity of the PEG-related peaks to the total surface area of the NPs [97]. However, XPS is a surface-responsive technique, that usually covers up to 10 nanometers of the sample, therefore experimental results correspond to the outer region of the material, and PEG coverage below this surface region can not be detected.
Thermogravimetric analysis is a useful technique for evaluating the PEG density on a nanoparticle surface. The PEG number is quantified by analyzing the weight loss as a function of temperature, [94,95,98,99] based on the calibration of the pure PEG molecule which has been used for NP preparation. However, TGA provides information about the total PEG content but it does not directly reveal how PEG molecules are distributed on the surface of the nanoparticles [100].
NMR spectroscopy provides quantitative information about the surface density of PEG on NPs. The average number of PEG molecules for each NP and chains per surface area can be calculated from 1H-NMR spectra by integration of the CH2 groups of PEG peak typically observed at ~3.5 ppm which can be compared to internal [94] or the external standard signal [101], the conformation of PEG chains can vary from mushroom to brush conformations that impact the mobility and dynamics of the polymer chains, moreover, the measurement of the spin relaxation time promotes a better understanding of PEG chain conformation around the NPs, with the increased mobility of the polymer chains the relaxation time decreases [101].

6. Protein PEGylation

6.1. N-terminus and Lysine PEGylation

PEG molecules can be conjugated to functional groups in proteins and peptides which are presented in the amino acid sequence (Figure 8). The PEGylation of amino groups is the most commonly used technique for protein PEGylation that can be divided into two categories: site-specific N-terminal PEGylation, targeting the amino terminus (N-terminus) and Lysine PEGylation. N-terminus PEGylation is often employed by maleimide or NHS ester-activated PEGs. Peng et al. showed covalent attachment of PEG 5 kDa molecules to immunostimulatory peptide Thymosin alpha 1 (Tα1), peptides with PEG attached to the N-terminus exhibited a longer half-life and improved immune activity compared to those modified with a cysteine thiol (Cys-Tα1) group. The authors suggest that the degradation of Cys-Tα1 over time could be attributed to interactions between the cysteine-modified Tα1 peptide and proteins, resulting in the breakdown of the structure [66]. Ming Yu et al. reported N-terminal PEGylation of recombinant human interleukin (rh-PEG-IL11) using PEG aldehyde, which significantly increased the half-life of the drug during subcutaneous injection, approximately from 3 hours for non-PEGylated rh-IL11 to 67 hours for PEGylated-IL11 in monkeys [59]. On the other hand, PEGylation of lysine residues occurs through a nucleophilic substitution reaction, where the amine group of the lysine side chain reacts with activated PEG derivatives, however, the reaction is non-specific and can lead to the modification of other moieties of proteins. PEG conjugation to lysine has proven to be effective in reducing proteolytic degradation of the fibronectin (FN) protein which plays a major role in cell adhesion, growth, migration, and differentiation, the authors studied the influence of PEG molecular weight on the stability and biological activity of FN. Cell properties (adhesion, cell spreading) were shown to deteriorate by increasing PEG molecular weight [102]. PEG attachment to lysine side chains is a popular technique for uricase PEGylation, which diminishes blood uric acid levels by converting them into allantoin. It provides multiple benefits, including increased solubility, reduced immunogenicity, and improved enzymatic stability [103]. Lysine group PEGylation is often employed for modification interferon alfa-2a with PEG which is used in combination with the anti-viral medicine ribavirin, for a treatment for chronic hepatitis C [104]. Lysine PEGylation is commonly used to modify nitrophenyl-PEGylated liposomes and micelles for targeted drug delivery. Amine groups of proteins readily react with PEG, facilitated by the excellent leaving properties of the nitrophenol group, which forms stable urethane linkages, the reaction is conducted at a slightly basic pH (7.5-8.5) [105,106,107] and the residual nitrophenol and unbound proteins are removed by dialysis.

6.2. Cysteine (Thiol) PEGylation

Cysteine PEGylation has been effectively used in modifying proteins such as enzymes and monoclonal antibodies [108,109] for improved tumour targeting which is based on maleimide chemistry. In neutral or slightly basic buffers, maleimide reacts with the thiol groups of cysteine creating strong thioether bonds. Thiol PEGylation was successfully used for site-specific modification of granulocyte colony-stimulating factor (G-CSF) which demonstrated prolonged biological activity during neutropenia in a mice model, whereas increasing PEG molecular weight improved leukocyte proliferation [68]. In another study, authors demonstrated that PEGylated G-CSF showed significantly longer circulating half-lives compared to unmodified G-CSF, as observed in the previous work [63], the half-life of cysteine PEGylated proteins was enhanced by increasing the molecular mass of the PEG [69]. Furthermore, thiol PEGylation was used to crosslink L-asparaginase, an enzyme used in leukemia treatment. Researchers introduced recombinant cysteine residues on the enzyme surface and conjugated them with different molecular weights PEGs. The study revealed that intramolecular crosslinking of L-asparaginase with PEG stabilized the enzyme active site resulting in improved catalytic activity. Additionally, PEGylation increased the enzyme hydrodynamic radius, reducing the probability of glomerular filtration and thereby extending the therapeutic half-life [110].

6.3. Carboxyl PEGylation

Carboxyl PEGylated products have better solubility and reduced immunogenicity [111], PEGylation targets carboxyl groups typically found in aspartic and glutamic amino acids. Carboxyl group PEGylation is employed by EDC/NHS activation, firstly, the carboxyl group is activated to react with a 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC), which converts the carboxyl group into O-acylisourea reactive form reacting with the amine groups of PEG forming an amide bond between PEG and the protein. However, the reaction is limited to buffers that do not contain carboxyl or amine groups. In 1997 study demonstrates the first carboxyl-specific PEGylation of BDNF neuropeptide which plays a crucial role in neuron survival, growth, and differentiation. PEGylation improved the pharmacokinetic properties of the peptide in terms of decreased plasma and hepatic clearance, however, authors emphasize that PEGylation increased the renal clearance of BDNF-PEG2000 compared to non-PEGylated BDNF, attributed to the fact that PEGylation enhanced the cationic properties of the peptide followed by kidney clearance, but BDNF-PEG5000 showed a decrease in removal because of the increased size of the peptide [112]. In conclusion, PEGylation is a powerful tool used in preparing therapeutic proteins because of its ability to enhance half-life, solubility, and immunogenicity, and at the same time, reduce the clearance of proteins (Figure 8). However, PEGylation might alter the bioactivity or functional properties of the protein because PEG could interfere with the active site or overall structure of the protein [113,114,115,116], therefore, selecting the appropriate modification method is important to maintain the protein functional integrity.

7. PEG Immunogenicity

PEGylation enhances drug properties by increasing circulation time, improving stability, and solubility, however, the repeated use of PEG has led to an unexpected consequence, the development of anti-PEG antibodies (APA). The recognition of PEG by immune system seems to be a complex mechanism. When the body encounters PEGylated compounds, immune cells can generate antibodies specifically targeting PEG epitopes. These antibodies may be of different isotypes, predominantly IgM [117,118,119] and IgG [120] followed by a quick removal and immune responses. Studies have revealed the prevalence of APA in the general population, 25% of healthy blood donors possessed it [121]. In another study, APA were detected in 44.3% of healthy donors both antibodies IgM (27.1%) and IgG (25.7%), antibodies were significantly more common in females as compared to males, which is in accordance with the separate study [122], likely due to cumulative exposure to PEG containing products in daily life. The presence of these antibodies has been associated with accelerated blood clearance (ABC) phenomenon, reduced therapeutic efficacy, and potential hypersensitivity reactions in patients receiving PEGylated medications [8,123]. Several key factors influence the development of anti-PEG immunity and ABC phenomenon such as previous exposure to PEG, administration routines, PEG characteristics, etc. Münter et al. compared the antibody generation against PEGylated mRNA-carrying NPs by different administration routes. Intramuscular injections in mice were found to generate overall low and dose-independent levels of antibodies, while intravenous and subcutaneous injections generated substantial levels of IgG and IgM antibodies [124]. In separate research by Takata et al., all tested routes for mPEG2000-DSPE induced anti-PEG IgM production, however, the maximum production of anti-PEG IgM antibodies was observed after intravenous administration and anti-PEG IgM production was significantly reduced after splenectomy for all administration routes, indicating the spleen key role in antibody generation [125]. Opinions vary on how PEG chain density affects ABC phenomena, for example, Li et al. reported that liposomes with a higher PEG grafting density (e.g., 9%) tend to induce enhanced ABC phenomenon compared to those with a lower PEG density (e.g., 3%).
Both high and low PEG formulations induced similar levels of anti-PEG IgM following the first dose. However, the higher PEG density results in stronger recognition and neutralization during subsequent doses, leading to faster clearance [126]. On the contrary, Ishida et al. showed that a PEG surface density of 5 mol% on mPEG2000-DSPE liposomes was found to induce ABC phenomena at the first dose, while the PEG density way below or higher this threshold reduced the ABC phenomenon, authors suggest that a high density of PEG reduced the activity of splenic B cells, resulting in less clearance of subsequent dose and increasing PEG Mw from 2000 up to 5000 g/mol, did not affect ABC phenomenon [127]. In addition to PEG Mw, the immunogenicity of PEG is influenced by its terminal group [128,129]. The molecular basis of anti-PEG antibody formation involves complex immunological mechanisms. Research by Ishida et al. has shown that empty PEGylated liposomes induce an anti-PEG IgM response in rats and mice, even in T-cell-deficient mice, indicating that the response is T-cell-independent [130].
APA detection methodologies have evolved significantly to name some of them immunosorbent assay ELISA, flow cytometry, and surface plasmon resonance. Ehlingeret et al. demonstrated high sensitivity in detecting clinically relevant anti-PEG antibody levels in a large analysis involving 200 healthy patients the ELISA method detected 97.5% pre-existing antibodies in human serum [131]. The existence of anti-PEG antibodies remains challenging in the vaccination process, cases of immediate hypersensitivity reactions including anaphylaxis were reported following the administration of PEG-containing COVID-19 vaccines, particularly those employing PEGylated lipid nanoparticles (LNPs), in the study 130 adults received either the BNT162b2 (Pfizer-BioNTech) and mRNA-1273 (Moderna) mRNA vaccines against SARS-CoV-2, ELISA analysis showed anti-PEG IgG was detectable in 71% of subjects prior to vaccination. However, APA were boosted after 2 doses vaccination: Pfizer-BioNTech showed tiny 1.78-fold change for IgG and 2.64 for IgM, while Moderna vaccine showed 13.1 for IgG and 68.5 for IgM and no increase was observed in unvaccinated controls [122]. In 2023, Kozma et al. studied the potential role of APA in allergic reactions to PEG COVID-19 vaccines, results showed elevated levels of IgG and IgM antibodies indicating that individuals with higher levels are more likely to experience allergic reactions upon receiving PEG-containing vaccines [132].

8. PEGylated Nanocarriers in Cancer Therapy

8.1. PEGylated liposomes

PEGylated liposomes are lipid bilayers that are coated by PEG on the surface, which reduces aggregation and enhances the colloidal stability of this formulation. Due to the enhanced permeability and retention effect, PEGylated liposomes accumulate in tumour tissues more effectively than conventional drug formulations (Table 2). One of the first stealth liposomal formulations coated with PEG was reported by Klibanov et al. who demonstrated the positive effect of liposome PEGylation on prolonged circulation time in the blood [133], which were capable of evading the RES, since then new strategies have been developed to increase the efficacy of that system, such as pH-sensitive design, which contains acid labile bonds these liposomes remain stable under physiological pH characteristic of healthy tissues and the bloodstream, but they degrade in acidic conditions which are found in tumour microenvironment where encapsulated drugs are released [134,135,136,137].
One of the most widely used PEGylated liposomal formulation against cancer is DOXIL which is the carrier of the chemotherapeutic drug doxorubicin (DOX), approved for the treatment of breast cancer, ovarian cancer, and multiple myeloma. Safra et al. reported that 500 mg/m2 of liposomal DOXIL significantly reduced the cardiomyopathy risk which is generally associated with the use of free doxorubicin [138]. Several methods have been used for the preparation of liposomal formulations, such as the pre-insertion method, where PEGylated lipids are incorporated into the formulation before the liposome formation. Since the PEGylated lipids are present throughout the formulation process, the resulting liposomes have a consistent surface modification, but a lower PEG surface density [139]. Meanwhile, in the post-insertion method, liposomes are first formed without PEGylation. Afterwards, PEG-lipid conjugates are added to the pre-formed liposomes, and the PEGylated lipids spontaneously integrate into the liposomal bilayer due to hydrophobic interactions. However, repeated administration of several PEGylated liposomes was reported to lead to the generation of anti-PEG antibodies, resulting in faster clearance [140,141,142,143]. Additionally, liposomes are limited by their drug-loading ability, purification steps and high production costs.

8.2. PEGylated Micelles

Micelles are self-assembled structures formed by amphiphilic surfactants or block copolymers. PEG block copolymers PEG-PLA and PEG-PCL form polymeric micelles by self-assembling processes, they consist of a hydrophilic outer PEG shell and a hydrophobic polymer core, that can encapsulate poorly soluble drugs (Figure 9). Polymeric micelles like liposomes are effective in solid tumours, PEGylated micelles typically range from 20 to 100 nm in size, and are ideal for passive targeting via the enhanced permeability and retention (EPR) effect. On the other hand, micelles are used for active targeting, which is achieved by functionalizing with ligands (e.g., antibodies, peptides, or small molecules) that specifically bind to overexpressed receptors on the cancer cells. PEGylated micelles are prepared by thin-film hydration: generally, the drug and polymer are dissolved in an organic solvent typically chloroform, or methanol, the organic solvents are removed by the rotary evaporation and a produced thin film is hydrated forming micelles with the drug trapped inside. The thin film method has been successfully used for the preparation of transferrin-targeted PEGylated liposomes for treating ovarian carcinoma in-vivo models [144,145].

8.3. PEGylated Dendrimers

Dendrimers are highly branched, nano-sized three-dimensional polymeric structures with unique properties, they increase the solubility of poorly soluble drugs and their half-life [146,147,148]. Positively charged dendrimers are widely used in drug delivery reacting with the negatively charged cell membrane; however, the charge on the surface also causes cytotoxic effects [149] and rapid clearance. The interaction of dendrimers with biological membranes can lead to membrane disruption by creating small holes or pores in the membrane [150]. PEGylation of cationic dendrimers is an effective way to overcome cytotoxicity problems [151,152,153]. PEGylation of poly(amidoamine) dendrimer (PAMAM) was shown to improve the solubility and the anticancer activity of piperlongumine drug in human colon cancer with sustained drug release [154]. In another study, PEGylated DOX-loaded dendrimers showed high accumulation in C26 murine colon tumour cells (around 20 times) and reduced accumulation of DOX/dendrimers in the spleen compared to the free drug [155]. Besides reducing toxicity, PEGylation enables the further modification of dendrimers with different ligands for targeted delivery [156]. Dual-responsive PEGylated dendrimers have shown promise in glioblastoma treatment by improving the delivery of therapeutic agents across the blood-brain barrier (BBB), studies in the C6 glioma cell line have demonstrated that ligand-modified PAMAM dendrimers conjugated with drugs like doxorubicin and tamoxifen can increase their half-life, accumulation in brain tumour and reduce non-specific toxicity [157,158,159,160].

8.4. PEGyled Polymeric Nanoparticles

Polymeric polyester nanoparticles such as PLGA, PLA, PCL or chitosan are often used as drug delivery systems, polymer matrices can encapsulate both hydrophobic and hydrophilic drugs, protecting them from degradation and providing sustained release. In addition, the NP's PEG surface can be modified with various ligands, such as antibodies or proteins, that can specifically target receptors overexpressed on cancer cells. However, “naked” nanoparticles are easily recognised by the immune system, when foreign bodies, including NPs, enter the bloodstream, the immune system quickly detects and removes them, by the process known as opsonization. During opsonization, the surface of nanoparticles is covered with opsonin proteins, signalling macrophages to neutralize them by phagocytosis [161]. PEGylation of the NP surface overcomes this problem by forming a hydrophilic "shield" around NPs and improves the nanoparticles stability, biocompatibility, and circulation time [162]. PEGylation was shown to increase the water solubility and bioavailability of hydrophobic drugs [163]. The size of PEGylated polymeric nanoparticles is one of the most important factors that determine NP behaviour. Nanoparticles in the 10 nm to 200 nm size range are optimal for efficient cellular uptake, tissue penetration, and circulation in the bloodstream, as most nanoparticles around 10 nm are filtered by the kidneys. As it is known the stability of non-modified nanoparticles highly depends on a surface charge [164]. PEGylation of NPs seems to reduce their zeta-potential [165,166,167] however, this does not present a significant issue as PEG prevents nanoparticle aggregation through steric hindrance, rather than the electrostatic stabilization mechanism. PEGylation affects interactions between nanoparticles and cell membranes and their subsequent cellular internalization [168]. Generally speaking, amphiphilic nanoparticles are internalized by cells through endocytosis, which is a major pathway for cellular uptake, for particles with diameters up to 200 nm clathrin-mediated endocytosis has been a predominant pathway for cellular uptake [169].

8.4.1. Methods for Preparation of PEGylated Polymeric Nanoparticles

Nanoprecipitation

Nanoprecipitation, also known as the solvent displacement method, is the most commonly used technique for the preparation of PEGylated polymer nanoparticles (Table 3), which is very useful for encapsulating hydrophobic drugs; however, less effective for hydrophilic drugs as precipitation is performed in the aqueous phase. NPs are prepared by adding an organic to the water phase, therefore, the organic solvent must be miscible with water. The polymer and the drug are dissolved in an organic solvent and slowly added to the aqueous phase under continuous stirring, the solvent can be removed by evaporation or dialysis. The particle size of PEGylated NPs can be controlled by modifying several key parameters such as polymer concentration, antisolvent volume-flow rate, and stirring rate. The concentration of the polymer in the organic solvent impacts the size of the resulting NPs, generally speaking, the higher the polymer concentrations the larger particles are formed, and increasing the polymer concentration too much can lead to aggregation and finally precipitation. Normally, the nanoprecipitation is performed without the use of an external stabilizer as PEG itself can stabilize suspension, however, when NPs are freeze-dried, it is often associated with structural collapse and difficulty in reconstitution or changes in particles physiochemical properties, adding a small amount of stabilizer can prevent this issue [167,170,171,172,173]. The nanoprecipitation method is generally used for preparing self-assembled NPs, which consists of supersaturation, nucleation, growth, and coagulation steps [174].

Single (O/W) and Double Emulsion (W/O/W) Solvent Evaporation Techniques

The single-emulsion solvent evaporation method can mainly encapsulate hydrophobic drugs in PEGylated polymeric nanoparticles (Table 4). The PEG-block-copolymer and hydrophobic drug are dissolved in a volatile organic solvent, which is added to the stabilizer-containing water under high-speed stirring or sonication, creating an oil-in-water emulsion (O/W). The organic solvent can be removed by stirring at room temperature; however, the boiling point should be less than water to ensure a complete evaporation process. As the solvent evaporates, the polymer precipitates and forms solid, drug-encapsulated NPs, resulting nanoparticles can be collected by centrifugation, filtration or lyophilization. NP size like in the nanoprecipitation method depends on several factors, such as polymer concentration that affects the viscosity of the organic phase, higher polymer concentrations typically result in larger nanoparticles. The choice of organic solvent plays a significant role in determining the final size of the PEGylated NPs, particularly, the evaporation rate of the solvent affects the rate at which the polymer precipitates to form nanoparticles. Solvents like DCM or chloroform are commonly used as they evaporate relatively quickly, promoting the rapid formation of nanoparticles, slower evaporation rates, on the other hand, allow the polymer to precipitate more slowly, leading to the formation of larger particles. Matsumoto et al. demonstrated that the release of the drug from nanoparticles was affected by PEG content in PLA-PEG-PLA and its molecular weight, copolymers with higher PEG content (25.8%), showed faster drug release (95% in 24 hours). In contrast, copolymers with lower PEG content (5.2%) exhibited slower release (44% in 24 hours), additionally, lower MW of copolymers and higher PEG Mw enhanced the initial and overall drug release [181]. In a different study, Xu et al. demonstrated the difference in paclitaxel drug release profiles between micellar PLGA 45 kDa-PEG 5kDa NPs and the core-shell PLGA 45 kDa/PLGA 20 kDa-PEG 5 kDa NPs which showed sustained release, whereas it seems that PEG content and matrix structure of the nano-system is detrimental for the drug release [182]
Meanwhile, O/W is typically used for encapsulating hydrophobic drugs, The double-emulsion or water-in-oil-in-water (W/O/W) technique is useful for encapsulating both hydrophilic and hydrophobic drugs (Table 5). They are dissolved in water and emulsified in an organic phase containing polymer creating a water-in-oil (W/O) which is a primary emulsion, dispersed in water phase forming a water-in-oil-in-water emulsion. The double-emulsion technique allows a more sustained and slow release of therapeutic reagents [165].
Figure 10. Single and double-emulsion solvent evaporation methods for nanoparticle preparation.
Figure 10. Single and double-emulsion solvent evaporation methods for nanoparticle preparation.
Preprints 148174 g010

Emulsion Polymerization

Emulsion polymerization (EP) takes place in the aqueous phase, where the PEG-block copolymers are initially presented as a “monomer droplet”, those block copolymers form micelles above the CMC. Monomers and initiators from the aqueous phase diffuse into the micelles where the polymerization process occurs. As more monomers diffuse inward, polymer chains grow inside the micelle. At the final step, chain growth is terminated and polymer particles are formed (Figure 11). Gurnani et al. synthesized biocompatible core-shell PEGylated butyl acrylate nanoparticles by RAFT emulsion polymerization technique with very small particle size, high biocompatibility, long circulation time and without a need for purification; however, authors report high NP accumulation in the liver in a CD1 mouse [193]. Colombo et al. reported the synthesis of surfactant-free PEGylated PCL NPs loaded with the anticancer drug paclitaxel (TXL) using EP without cremphor, which is widely used for NP preparation, as it increases drug solubility, however, cremophor causes serious side effects including hypersensitivity and toxicity. NPs prepared by EP showed similar antitumor activity to free TXL and the pharmacokinetic behaviour of TXL encapsulated in PCL nanoparticles. It shows no significant difference in drug recovery from urine and faeces compared to free paclitaxel. This formulation offers advantages in terms of safety and potential targeting modifications for enhanced cancer therapy compared to the free drugs [194]. Lupi et al. prepared PEGylated Rhodamine B labelled PCL NPs using an environmentally friendly two-step approach combining ring-opening polymerization and water-based semi-batch emulsion polymerization, which is a surfactant-free process and no organic solvents are required, PEGylation provided stealth properties, fast cellular uptake (within 1-2 hours) and maintained steady concentration in tumours for 120 hours [195]. Similarly, Ferrari et al. prepared PEGylated PLA NPs through surfactant-free and aided copolymerization of (hydroxyethyl)methacrylate (HEMA)-LA4 with HEMA-PEGm macromonomers, after surfactant removal NPs with short PEG chain length of less than 8.3 nm precipitated; however, when PEG chain length was greater than 8.3, the PEGylated NPs remained stable in PBS solution, and no aggregation was observed, the surfactant addition reduced the final NP size to 30 nm. In the absence of surfactant, as the PEG average chain length increased, the diameter of the NPs tended to decrease at lower copolymer ratios, indicating that longer PEG chains give rise to smaller particles [196].

9. Conclusions

PEGylation technology has revolutionized drug delivery systems by overcoming some of the challenges in drug administration in cancer treatment, infectious diseases, gene therapy and so on. In this review, we examined the evolution of PEG from a simple aliphatic polymer to a sophisticated drug delivery platform, which improves pharmacokinetic properties- such as solubility, stability, and bioavailability, while reducing the toxicity of the drug. The PEGylation improves those properties by stealth effect, thus the nanocarriers evade immune system detection and reach extended circulation times while maintaining safety and biocompatibility. The molecular weight, architecture, and density of PEG chains significantly impact the behaviour of PEGylated systems. Properly optimized chain size and conformation enhance pharmacokinetics, immune evasion, and controlled drug release. Despite these advances, anti-PEG immunogenicity and the accelerated clearance of PEGylated products remain critical challenges. Current research into optimizing PEG molecular architecture, as well as developing new conjugation strategies are important to overcome these limitations.

Author Contributions

Conceptualization, D.M. and J.P.; formal analysis, D.M., L.d.V. and J.P.; writing—original draft preparation, D.M and R.K..; writing—review and editing, D.M., R.K., L.d.V. and J.P.; supervision, L.d.V. and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministry of Science, Innovation and Universities through PID2022-140302OB-I00 and by the Generalitat de Catalunya under the project 2021-SGR-01042. This work is part of Maria de Maeztu Units of Excellence Programme CEX2023-001300-M / funded by MCIN/AEI / 10.13039/501100011033.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

References

  1. OWENSIII, D.; PEPPAS, N. Opsonization, Biodistribution, and Pharmacokinetics of Polymeric Nanoparticles. Int J Pharm 2006, 307, 93–102. [Google Scholar] [CrossRef] [PubMed]
  2. Tobı́o, M.; Sánchez, A.; Vila, A.; Soriano, I.; Evora, C.; Vila-Jato, J.L.; Alonso, M.J. The Role of PEG on the Stability in Digestive Fluids and in Vivo Fate of PEG-PLA Nanoparticles Following Oral Administration. Colloids Surf B Biointerfaces 2000, 18, 315–323. [Google Scholar] [CrossRef] [PubMed]
  3. Korake, S.; Bothiraja, C.; Pawar, A. Design, Development, and in-Vitro/in-Vivo Evaluation of Docetaxel-Loaded PEGylated Solid Lipid Nanoparticles in Prostate Cancer Therapy. European Journal of Pharmaceutics and Biopharmaceutics 2023, 189, 15–27. [Google Scholar] [CrossRef]
  4. Luo, T.; Loira-Pastoriza, C.; Patil, H.P.; Ucakar, B.; Muccioli, G.G.; Bosquillon, C.; Vanbever, R. PEGylation of Paclitaxel Largely Improves Its Safety and Anti-Tumor Efficacy Following Pulmonary Delivery in a Mouse Model of Lung Carcinoma. Journal of Controlled Release 2016, 239, 62–71. [Google Scholar] [CrossRef] [PubMed]
  5. Gao, Y.; Joshi, M.; Zhao, Z.; Mitragotri, S. PEGylated Therapeutics in the Clinic. Bioeng Transl Med 2024, 9. [Google Scholar] [CrossRef]
  6. Belén, L.H.; Rangel-Yagui, C. de O.; Beltrán Lissabet, J.F.; Effer, B.; Lee-Estevez, M.; Pessoa, A.; Castillo, R.L.; Farías, J.G. From Synthesis to Characterization of Site-Selective PEGylated Proteins. Front Pharmacol 2019, 10. [Google Scholar] [CrossRef]
  7. Li, C.; Li, T.; Tian, X.; An, W.; Wang, Z.; Han, B.; Tao, H.; Wang, J.; Wang, X. Research Progress on the PEGylation of Therapeutic Proteins and Peptides (TPPs). Front Pharmacol 2024, 15. [Google Scholar] [CrossRef]
  8. Ibrahim, M.; Ramadan, E.; Elsadek, N.E.; Emam, S.E.; Shimizu, T.; Ando, H.; Ishima, Y.; Elgarhy, O.H.; Sarhan, H.A.; Hussein, A.K.; et al. Polyethylene Glycol (PEG): The Nature, Immunogenicity, and Role in the Hypersensitivity of PEGylated Products. Journal of Controlled Release 2022, 351, 215–230. [Google Scholar] [CrossRef]
  9. Chen, B.-M.; Cheng, T.-L.; Roffler, S.R. Polyethylene Glycol Immunogenicity: Theoretical, Clinical, and Practical Aspects of Anti-Polyethylene Glycol Antibodies. ACS Nano 2021, 15, 14022–14048. [Google Scholar] [CrossRef]
  10. Patel, M.; Park, J.K.; Jeong, B. Rediscovery of Poly(Ethylene Glycol)s as a Cryoprotectant for Mesenchymal Stem Cells. Biomater Res 2023, 27. [Google Scholar] [CrossRef]
  11. Zalipsky, S.; Gilon, C.; Zilkha, A. Attachment of Drugs to Polyethylene Glycols. Eur Polym J 1983, 19, 1177–1183. [Google Scholar] [CrossRef]
  12. Jevševar, S.; Kunstelj, M.; Porekar, V.G. PEGylation of Therapeutic Proteins. Biotechnol J 2010, 5, 113–128. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, C.-C.; Anseth, K.S. PEG Hydrogels for the Controlled Release of Biomolecules in Regenerative Medicine. Pharm Res 2009, 26, 631–643. [Google Scholar] [CrossRef] [PubMed]
  14. Lin, C.-C.; Raza, A.; Shih, H. PEG Hydrogels Formed by Thiol-Ene Photo-Click Chemistry and Their Effect on the Formation and Recovery of Insulin-Secreting Cell Spheroids. Biomaterials 2011, 32, 9685–9695. [Google Scholar] [CrossRef]
  15. Harris, J.M.; Struck, E.C.; Case, M.G.; Paley, M.S.; Yalpani, M.; Van Alstine, J.M.; Brooks, D.E. Synthesis and Characterization of Poly(Ethylene Glycol) Derivatives. Journal of Polymer Science: Polymer Chemistry Edition 1984, 22, 341–352. [Google Scholar] [CrossRef]
  16. Edward Semple, J.; Sullivan, B.; Vojkovsky, T.; Sill, K.N. Synthesis and Facile End-group Quantification of Functionalized PEG Azides. J Polym Sci A Polym Chem 2016, 54, 2888–2895. [Google Scholar] [CrossRef]
  17. González-Fernández, D.; Torneiro, M.; Lazzari, M. Some Guidelines for the Synthesis and Melting Characterization of Azide Poly(Ethylene Glycol) Derivatives. Polymers (Basel) 2020, 12, 1269. [Google Scholar] [CrossRef]
  18. Woghiren, C.; Sharma, B.; Stein, S. Protected Thiol-Polyethylene Glycol: A New Activated Polymer for Reversible Protein Modification. Bioconjug Chem 1993, 4, 314–318. [Google Scholar] [CrossRef]
  19. Zhang, X.; Chen, X.; Chen, X.; Wang, S.; Wang, M.; Geng, C.; Xu, G.; Wang, S. A Scalable and Efficient Approach to High-Fidelity Amine Functionalized Poly(Ethylene Glycol) Derivatives. Polym Chem 2023, 14, 3352–3356. [Google Scholar] [CrossRef]
  20. Mahou, R.; Wandrey, C. Versatile Route to Synthesize Heterobifunctional Poly(Ethylene Glycol) of Variable Functionality for Subsequent Pegylation. Polymers (Basel) 2012, 4, 561–589. [Google Scholar] [CrossRef]
  21. Elbert, D.L.; Hubbell, J.A. Conjugate Addition Reactions Combined with Free-Radical Cross-Linking for the Design of Materials for Tissue Engineering. Biomacromolecules 2001, 2, 430–441. [Google Scholar] [CrossRef] [PubMed]
  22. Herman, S.; Hooftman, G.; Schacht, E. Poly(Ethylene Glycol) with Reactive Endgroups: I. Modification of Proteins. J Bioact Compat Polym 1995, 10, 145–187. [Google Scholar] [CrossRef]
  23. Nguyen, P.K.; Snyder, C.G.; Shields, J.D.; Smith, A.W.; Elbert, D.L. Clickable Poly(Ethylene Glycol)-Microsphere-Based Cell Scaffolds. Macromol Chem Phys 2013, 214, 948–956. [Google Scholar] [CrossRef]
  24. Menger, F.M.; Zhang, H. Self-Adhesion among Phospholipid Vesicles. J Am Chem Soc 2006, 128, 1414–1415. [Google Scholar] [CrossRef] [PubMed]
  25. Das, M.; Bandyopadhyay, D.; Singh, R.P.; Harde, H.; Kumar, S.; Jain, S. Orthogonal Biofunctionalization of Magnetic Nanoparticles via “Clickable” Poly(Ethylene Glycol) Silanes: A “Universal Ligand” Strategy to Design Stealth and Target-Specific Nanocarriers. J Mater Chem 2012, 22, 24652. [Google Scholar] [CrossRef]
  26. Zhang, X.; Chen, X.; Chen, X.; Wang, S.; Wang, M.; Geng, C.; Xu, G.; Wang, S. A Scalable and Efficient Approach to High-Fidelity Amine Functionalized Poly(Ethylene Glycol) Derivatives. Polym Chem 2023, 14, 3352–3356. [Google Scholar] [CrossRef]
  27. Mahou, R.; Wandrey, C. Versatile Route to Synthesize Heterobifunctional Poly(Ethylene Glycol) of Variable Functionality for Subsequent Pegylation. Polymers (Basel) 2012, 4, 561–589. [Google Scholar] [CrossRef]
  28. Du, Y.J.; Brash, J.L. Synthesis and Characterization of Thiol-terminated Poly(Ethylene Oxide) for Chemisorption to Gold Surface. J Appl Polym Sci 2003, 90, 594–607. [Google Scholar] [CrossRef]
  29. Lisboa, P.; Valsesia, A.; Colpo, P.; Gilliland, D.; Ceccone, G.; Papadopoulou-Bouraoui, A.; Rauscher, H.; Reniero, F.; Guillou, C.; Rossi, F. Thiolated Polyethylene Oxide as a Non-Fouling Element for Nano-Patterned Bio-Devices. Appl Surf Sci 2007, 253, 4796–4804. [Google Scholar] [CrossRef]
  30. Lüchow, M.; Fortuin, L.; Malkoch, M. Modular, Synthetic, Thiol-ene Mediated Hydrogel Networks as Potential Scaffolds for <scp>3D</Scp> Cell Cultures and Tissue Regeneration. Journal of Polymer Science 2020, 58, 3153–3164. [Google Scholar] [CrossRef]
  31. Nie, T.; Baldwin, A.; Yamaguchi, N.; Kiick, K.L. Production of Heparin-Functionalized Hydrogels for the Development of Responsive and Controlled Growth Factor Delivery Systems. Journal of Controlled Release 2007, 122, 287–296. [Google Scholar] [CrossRef]
  32. Unsworth, L.D.; Tun, Z.; Sheardown, H.; Brash, J.L. Chemisorption of Thiolated Poly(Ethylene Oxide) to Gold: Surface Chain Densities Measured by Ellipsometry and Neutron Reflectometry. J Colloid Interface Sci 2005, 281, 112–121. [Google Scholar] [CrossRef] [PubMed]
  33. Mulay, P.; Shrikhande, G.; Puskas, J.E. Synthesis of Mono- and Dithiols of Tetraethylene Glycol and Poly(Ethylene Glycol)s via Enzyme Catalysis. Catalysts 2019, 9, 228. [Google Scholar] [CrossRef]
  34. Hodgson, S.M.; McNelles, S.A.; Abdullahu, L.; Marozas, I.A.; Anseth, K.S.; Adronov, A. Reproducible Dendronized PEG Hydrogels via SPAAC Cross-Linking. Biomacromolecules 2017, 18, 4054–4059. [Google Scholar] [CrossRef]
  35. Spampinato, A.; Kužmová, E.; Pohl, R.; Sýkorová, V.; Vrábel, M.; Kraus, T.; Hocek, M. Trans -Cyclooctene- and Bicyclononyne-Linked Nucleotides for Click Modification of DNA with Fluorogenic Tetrazines and Live Cell Metabolic Labeling and Imaging. Bioconjug Chem 2023. [Google Scholar] [CrossRef]
  36. Wilding, K.M.; Smith, A.K.; Wilkerson, J.W.; Bush, D.B.; Knotts, T.A.; Bundy, B.C. The Locational Impact of Site-Specific PEGylation: Streamlined Screening with Cell-Free Protein Expression and Coarse-Grain Simulation. ACS Synth Biol 2018, 7, 510–521. [Google Scholar] [CrossRef] [PubMed]
  37. Macias-Contreras, M.; He, H.; Little, K.N.; Lee, J.P.; Campbell, R.P.; Royzen, M.; Zhu, L. SNAP/CLIP-Tags and Strain-Promoted Azide–Alkyne Cycloaddition (SPAAC)/Inverse Electron Demand Diels–Alder (IEDDA) for Intracellular Orthogonal/Bioorthogonal Labeling. Bioconjug Chem 2020, 31, 1370–1381. [Google Scholar] [CrossRef]
  38. Hodgson, S.M.; McNelles, S.A.; Abdullahu, L.; Marozas, I.A.; Anseth, K.S.; Adronov, A. Reproducible Dendronized PEG Hydrogels via SPAAC Cross-Linking. Biomacromolecules 2017, 18, 4054–4059. [Google Scholar] [CrossRef]
  39. Hodgson, S.M.; Bakaic, E.; Stewart, S.A.; Hoare, T.; Adronov, A. Properties of Poly(Ethylene Glycol) Hydrogels Cross-Linked via Strain-Promoted Alkyne–Azide Cycloaddition (SPAAC). Biomacromolecules 2016, 17, 1093–1100. [Google Scholar] [CrossRef]
  40. Lin, C.-C. Recent Advances in Crosslinking Chemistry of Biomimetic Poly(Ethylene Glycol) Hydrogels. RSC Adv 2015, 5, 39844–39853. [Google Scholar] [CrossRef]
  41. Pan, X.; Gao, H.; Fu, G.; Gao, Y.; Zhang, W. Synthesis, Characterization and Chondrocyte Culture of Polyhedral Oligomeric Silsesquioxane (POSS)-Containing Hybrid Hydrogels. RSC Adv 2016, 6, 23471–23478. [Google Scholar] [CrossRef]
  42. Herzberger, J.; Leibig, D.; Langhanki, J.; Moers, C.; Opatz, T.; Frey, H. “Clickable PEG” via Anionic Copolymerization of Ethylene Oxide and Glycidyl Propargyl Ether. Polym Chem 2017, 8, 1882–1887. [Google Scholar] [CrossRef]
  43. Lu, C.; Zhong, W. Synthesis of Propargyl-Terminated Heterobifunctional Poly(Ethylene Glycol). Polymers (Basel) 2010, 2, 407–417. [Google Scholar] [CrossRef]
  44. Wolkersdorfer, A.M.; Jugovic, I.; Scheller, L.; Gutmann, M.; Hahn, L.; Diessner, J.; Lühmann, T.; Meinel, L. PEGylation of Human Vascular Endothelial Growth Factor. ACS Biomater Sci Eng 2024, 10, 149–155. [Google Scholar] [CrossRef]
  45. Harris, J.M. LABORATORY SYNTHESIS OF POLYETHYLENE GLYCOL DERIVATIVES. Journal of Macromolecular Science, Part C 1985, 25, 325–373. [Google Scholar] [CrossRef]
  46. Topchieva, I.N.; Kuzaev, A.I.; Zubov, V.P. Modification of Polyethylene Glycol. Eur Polym J 1988, 24, 899–904. [Google Scholar] [CrossRef]
  47. Mauri, E.; Rossi, F.; Sacchetti, A. Simple and Efficient Strategy to Synthesize PEG-aldehyde Derivatives for Hydrazone Orthogonal Chemistry. Polym Adv Technol 2015, 26, 1456–1460. [Google Scholar] [CrossRef]
  48. Pan, G.; Lemmouchi, Y.; Akala, E.O.; Bakare, O. Studies on PEGylated and Drug-Loaded PAMAM Dendrimers. J Bioact Compat Polym 2005, 20, 113–128. [Google Scholar] [CrossRef]
  49. Guo, N.; Jiang, D.; Wang, L.; You, X.; Teng, Y.-O.; Yu, P. Synthesis and Biological Evaluation of Novel Water-Soluble Poly-(Ethylene Glycol)-10-Hydroxycamptothecin Conjugates. Molecules 2015, 20, 9393–9404. [Google Scholar] [CrossRef]
  50. Schiavon, O.; Pasut, G.; Moro, S.; Orsolini, P.; Guiotto, A.; Veronese, F.M. PEG–Ara-C Conjugates for Controlled Release. Eur J Med Chem 2004, 39, 123–133. [Google Scholar] [CrossRef]
  51. Ahmed, R.; Aucamp, M.; Ebrahim, N.; Samsodien, H. Supramolecular Assembly of Rifampicin and PEGylated PAMAM Dendrimer as a Novel Conjugate for Tuberculosis. J Drug Deliv Sci Technol 2021, 66, 102773. [Google Scholar] [CrossRef]
  52. KITO, M.; MIRON, T.; WILCHEK, M.; KOJIMA, N.; OHISHI, N.; YAGI, K. A Simple and Efficient Method for Preparation of Monomethoxypolyethylene Glycol Activated with P-Nitrophenylchloroformate and Its Application to Modification of L-Asparaginase. J Clin Biochem Nutr 1996, 21, 101–111. [Google Scholar] [CrossRef]
  53. Pan, G.; Lemmouchi, Y.; Akala, E.O.; Bakare, O. Studies on PEGylated and Drug-Loaded PAMAM Dendrimers. J Bioact Compat Polym 2005, 20, 113–128. [Google Scholar] [CrossRef]
  54. Bai, S.; Ahsan, F. Synthesis and Evaluation of Pegylated Dendrimeric Nanocarrier for Pulmonary Delivery of Low Molecular Weight Heparin. Pharm Res 2009, 26, 539–548. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, Y.-J.; Hao, S.-J.; Liu, Y.-D.; Hu, T.; Zhang, G.-F.; Zhang, X.; Qi, Q.-S.; Ma, G.-H.; Su, Z.-G. PEGylation Markedly Enhances the in Vivo Potency of Recombinant Human Non-Glycosylated Erythropoietin: A Comparison with Glycosylated Erythropoietin. Journal of Controlled Release 2010, 145, 306–313. [Google Scholar] [CrossRef]
  56. Nojima, Y.; Suzuki, Y.; Yoshida, K.; Abe, F.; Shiga, T.; Takeuchi, T.; Sugiyama, A.; Shimizu, H.; Sato, A. Lactoferrin Conjugated with 40-KDa Branched Poly(Ethylene Glycol) Has an Improved Circulating Half-Life. Pharm Res 2009, 26, 2125–2132. [Google Scholar] [CrossRef]
  57. Santos, J.H.P.M.; Feitosa, V.A.; Meneguetti, G.P.; Carretero, G.; Coutinho, J.A.P.; Ventura, S.P.M.; Rangel-Yagui, C.O. Lysine-PEGylated Cytochrome C with Enhanced Shelf-Life Stability. Biosensors (Basel) 2022, 12, 94. [Google Scholar] [CrossRef]
  58. Javid, A.; Ahmadian, S.; Saboury, A.A.; Kalantar, S.M.; Rezaei-Zarchi, S.; Shahzad, S. Biocompatible APTES–PEG Modified Magnetite Nanoparticles: Effective Carriers of Antineoplastic Agents to Ovarian Cancer. Appl Biochem Biotechnol 2014, 173, 36–54. [Google Scholar] [CrossRef]
  59. Yu, K.-M.; Lau, J.Y.-N.; Fok, M.; Yeung, Y.-K.; Fok, S.-P.; Zhang, S.; Ye, P.; Zhang, K.; Li, X.; Li, J.; et al. Preclinical Evaluation of the Mono-PEGylated Recombinant Human Interleukin-11 in Cynomolgus Monkeys. Toxicol Appl Pharmacol 2018, 342, 39–49. [Google Scholar] [CrossRef]
  60. Soderquist, R.G.; Milligan, E.D.; Sloane, E.M.; Harrison, J.A.; Douvas, K.K.; Potter, J.M.; Hughes, T.S.; Chavez, R.A.; Johnson, K.; Watkins, L.R.; et al. PEGylation of Brain-derived Neurotrophic Factor for Preserved Biological Activity and Enhanced Spinal Cord Distribution. J Biomed Mater Res A 2009, 91A, 719–729. [Google Scholar] [CrossRef]
  61. Li, Z.; Zhou, D. Acrylate-Based PEG Hydrogels with Ultrafast Biodegradability for 3D Cell Culture. Biomacromolecules 2024, 25, 6195–6202. [Google Scholar] [CrossRef] [PubMed]
  62. Almany, L.; Seliktar, D. Biosynthetic Hydrogel Scaffolds Made from Fibrinogen and Polyethylene Glycol for 3D Cell Cultures. Biomaterials 2005, 26, 2467–2477. [Google Scholar] [CrossRef]
  63. Zhang, X.; Yang, D.; Nie, J. Chitosan/Polyethylene Glycol Diacrylate Films as Potential Wound Dressing Material. Int J Biol Macromol 2008, 43, 456–462. [Google Scholar] [CrossRef]
  64. Lan, Z.; Kar, R.; Chwatko, M.; Shoga, E.; Cosgriff-Hernandez, E. High Porosity <scp>PEG</Scp> -based Hydrogel Foams with Self-tuning Moisture Balance as Chronic Wound Dressings. J Biomed Mater Res A 2023, 111, 465–477. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, X.; Yan, Z.; Guan, G.; Lu, Z.; Yan, S.; Du, A.; Wang, L.; Li, Q. Polyethylene Glycol Diacrylate Scaffold Filled with Cell-Laden Methacrylamide Gelatin/Alginate Hydrogels Used for Cartilage Repair. J Biomater Appl 2022, 36, 1019–1032. [Google Scholar] [CrossRef]
  66. Peng, G.; Pan, X.; Hu, H.; Xu, Y.; Wu, C. N-Terminal Site-Specific PEGylation Enhances the Circulation Half-Life of Thymosin Alpha 1. J Drug Deliv Sci Technol 2019, 49, 405–412. [Google Scholar] [CrossRef]
  67. Pan, L.-Q.; Wang, H.-B.; Lai, J.; Xu, Y.-C.; Zhang, C.; Chen, S.-Q. Site-Specific PEGylation of a Mutated-Cysteine Residue and Its Effect on Tumor Necrosis Factor (TNF)-Related Apoptosis-Inducing Ligand (TRAIL). Biomaterials 2013, 34, 9115–9123. [Google Scholar] [CrossRef] [PubMed]
  68. Kumari, M.; Sahni, G.; Datta, S. Development of Site-Specific PEGylated Granulocyte Colony Stimulating Factor With Prolonged Biological Activity. Front Bioeng Biotechnol 2020, 8. [Google Scholar] [CrossRef]
  69. Doherty, D.H.; Rosendahl, M.S.; Smith, D.J.; Hughes, J.M.; Chlipala, E.A.; Cox, G.N. Site-Specific PEGylation of Engineered Cysteine Analogues of Recombinant Human Granulocyte-Macrophage Colony-Stimulating Factor. Bioconjug Chem 2005, 16, 1291–1298. [Google Scholar] [CrossRef]
  70. Zhu, J.; Waengler, C.; Lennox, R.B.; Schirrmacher, R. Preparation of Water-Soluble Maleimide-Functionalized 3 Nm Gold Nanoparticles: A New Bioconjugation Template. Langmuir 2012, 28, 5508–5512. [Google Scholar] [CrossRef]
  71. Schulz, F.; Vossmeyer, T.; Bastús, N.G.; Weller, H. Effect of the Spacer Structure on the Stability of Gold Nanoparticles Functionalized with Monodentate Thiolated Poly(Ethylene Glycol) Ligands. Langmuir 2013, 29, 9897–9908. [Google Scholar] [CrossRef]
  72. Chang, W.-C.; Tai, J.-T.; Wang, H.-F.; Ho, R.-M.; Hsiao, T.-C.; Tsai, D.-H. Surface PEGylation of Silver Nanoparticles: Kinetics of Simultaneous Surface Dissolution and Molecular Desorption. Langmuir 2016, 32, 9807–9815. [Google Scholar] [CrossRef]
  73. Endres, T.K.; Beck-Broichsitter, M.; Samsonova, O.; Renette, T.; Kissel, T.H. Self-Assembled Biodegradable Amphiphilic PEG–PCL–LPEI Triblock Copolymers at the Borderline between Micelles and Nanoparticles Designed for Drug and Gene Delivery. Biomaterials 2011, 32, 7721–7731. [Google Scholar] [CrossRef] [PubMed]
  74. Makharadze, D.; Kantaria, T.; Yousef, I.; del Valle, L.J.; Katsarava, R.; Puiggalí, J. PEGylated Micro/Nanoparticles Based on Biodegradable Poly(Ester Amides): Preparation and Study of the Core–Shell Structure by Synchrotron Radiation-Based FTIR Microspectroscopy and Electron Microscopy. Int J Mol Sci 2024, 25, 6999. [Google Scholar] [CrossRef] [PubMed]
  75. Thierry, B.; Zimmer, L.; McNiven, S.; Finnie, K.; Barbé, C.; Griesser, H.J. Electrostatic Self-Assembly of PEG Copolymers onto Porous Silica Nanoparticles. Langmuir 2008, 24, 8143–8150. [Google Scholar] [CrossRef] [PubMed]
  76. Ueya, Y.; Umezawa, M.; Kobayashi, Y.; Ichihashi, K.; Kobayashi, H.; Matsuda, T.; Takamoto, E.; Kamimura, M.; Soga, K. Effects of Hydrophilic/Hydrophobic Blocks Ratio of PEG-b-PLGA on Emission Intensity and Stability of over-1000 Nm near-Infrared (NIR-II) Fluorescence Dye-Loaded Polymeric Micellar Nanoparticles. Analytical Sciences 2022, 38, 199–205. [Google Scholar] [CrossRef]
  77. Zhang, Z.; Qu, Q.; Li, J.; Zhou, S. The Effect of the Hydrophilic/Hydrophobic Ratio of Polymeric Micelles on Their Endocytosis Pathways into Cells. Macromol Biosci 2013, 13, 789–798. [Google Scholar] [CrossRef]
  78. Cardoso, M.M.; Peça, I.N.; Bicho, A. Impact of PEG Content on Doxorubicin Release from PLGA-Co-PEG Nanoparticles. Materials 2024, 17, 3544. [Google Scholar] [CrossRef]
  79. Caliceti, P.; Salmaso, S.; Elvassore, N.; Bertucco, A. Effective Protein Release from PEG/PLA Nano-Particles Produced by Compressed Gas Anti-Solvent Precipitation Techniques. Journal of Controlled Release 2004, 94, 195–205. [Google Scholar] [CrossRef]
  80. Zhang, X.G.; Teng, D.Y.; Wu, Z.M.; Wang, X.; Wang, Z.; Yu, D.M.; Li, C.X. PEG-Grafted Chitosan Nanoparticles as an Injectable Carrier for Sustained Protein Release. J Mater Sci Mater Med 2008, 19, 3525–3533. [Google Scholar] [CrossRef]
  81. Torchilin, V.P. Structure and Design of Polymeric Surfactant-Based Drug Delivery Systems. Journal of Controlled Release 2001, 73, 137–172. [Google Scholar] [CrossRef]
  82. Ukawala, M.; Rajyaguru, T.; Chaudhari, K.; Manjappa, A.S.; Pimple, S.; Babbar, A.K.; Mathur, R.; Mishra, A.K.; Murthy, R.S.R. Investigation on Design of Stable Etoposide-Loaded PEG-PCL Micelles: Effect of Molecular Weight of PEG-PCL Diblock Copolymer on the in Vitro and in Vivo Performance of Micelles. Drug Deliv 2012, 19, 155–167. [Google Scholar] [CrossRef] [PubMed]
  83. Michel, R.; Pasche, S.; Textor, M.; Castner, D.G. Influence of PEG Architecture on Protein Adsorption and Conformation. Langmuir 2005, 21, 12327–12332. [Google Scholar] [CrossRef]
  84. Li, M.; Jiang, S.; Simon, J.; Paßlick, D.; Frey, M.-L.; Wagner, M.; Mailänder, V.; Crespy, D.; Landfester, K. Brush Conformation of Polyethylene Glycol Determines the Stealth Effect of Nanocarriers in the Low Protein Adsorption Regime. Nano Lett 2021, 21, 1591–1598. [Google Scholar] [CrossRef] [PubMed]
  85. Walkey, C.D.; Olsen, J.B.; Guo, H.; Emili, A.; Chan, W.C.W. Nanoparticle Size and Surface Chemistry Determine Serum Protein Adsorption and Macrophage Uptake. J Am Chem Soc 2012, 134, 2139–2147. [Google Scholar] [CrossRef] [PubMed]
  86. Du, H.; Chandaroy, P.; Hui, S.W. Grafted Poly-(Ethylene Glycol) on Lipid Surfaces Inhibits Protein Adsorption and Cell Adhesion. Biochimica et Biophysica Acta (BBA) - Biomembranes 1997, 1326, 236–248. [Google Scholar] [CrossRef]
  87. Yang, Q.; Jones, S.W.; Parker, C.L.; Zamboni, W.C.; Bear, J.E.; Lai, S.K. Evading Immune Cell Uptake and Clearance Requires PEG Grafting at Densities Substantially Exceeding the Minimum for Brush Conformation. Mol Pharm 2014, 11, 1250–1258. [Google Scholar] [CrossRef]
  88. Wang, J.-L.; Du, X.-J.; Yang, J.-X.; Shen, S.; Li, H.-J.; Luo, Y.-L.; Iqbal, S.; Xu, C.-F.; Ye, X.-D.; Cao, J.; et al. The Effect of Surface Poly(Ethylene Glycol) Length on in Vivo Drug Delivery Behaviors of Polymeric Nanoparticles. Biomaterials 2018, 182, 104–113. [Google Scholar] [CrossRef]
  89. Gref, R.; Lück, M.; Quellec, P.; Marchand, M.; Dellacherie, E.; Harnisch, S.; Blunk, T.; Müller, R.H. ‘Stealth’ Corona-Core Nanoparticles Surface Modified by Polyethylene Glycol (PEG): Influences of the Corona (PEG Chain Length and Surface Density) and of the Core Composition on Phagocytic Uptake and Plasma Protein Adsorption. Colloids Surf B Biointerfaces 2000, 18, 301–313. [Google Scholar] [CrossRef]
  90. Chithrani, D.B. Polyethylene Glycol Density and Length Affects Nanoparticle Uptake by Cancer Cells. J Nanomed Res 2014, 1. [Google Scholar] [CrossRef]
  91. Pozzi, D.; Colapicchioni, V.; Caracciolo, G.; Piovesana, S.; Capriotti, A.L.; Palchetti, S.; De Grossi, S.; Riccioli, A.; Amenitsch, H.; Laganà, A. Effect of Polyethyleneglycol (PEG) Chain Length on the Bio–Nano-Interactions between PEGylated Lipid Nanoparticles and Biological Fluids: From Nanostructure to Uptake in Cancer Cells. Nanoscale 2014, 6, 2782. [Google Scholar] [CrossRef] [PubMed]
  92. Mishra, S.; Webster, P.; Davis, M.E. PEGylation Significantly Affects Cellular Uptake and Intracellular Trafficking of Non-Viral Gene Delivery Particles. Eur J Cell Biol 2004, 83, 97–111. [Google Scholar] [CrossRef]
  93. Xia, X.; Yang, M.; Wang, Y.; Zheng, Y.; Li, Q.; Chen, J.; Xia, Y. Quantifying the Coverage Density of Poly(Ethylene Glycol) Chains on the Surface of Gold Nanostructures. ACS Nano 2012, 6, 512–522. [Google Scholar] [CrossRef]
  94. Zhu, J.; Waengler, C.; Lennox, R.B.; Schirrmacher, R. Preparation of Water-Soluble Maleimide-Functionalized 3 Nm Gold Nanoparticles: A New Bioconjugation Template. Langmuir 2012, 28, 5508–5512. [Google Scholar] [CrossRef]
  95. von Baeckmann, C.; Kählig, H.; Lindén, M.; Kleitz, F. On the Importance of the Linking Chemistry for the PEGylation of Mesoporous Silica Nanoparticles. J Colloid Interface Sci 2021, 589, 453–461. [Google Scholar] [CrossRef]
  96. Lin, Y.S.; Hlady, V.; Gölander, C.-G. The Surface Density Gradient of Grafted Poly(Ethylene Glycol): Preparation, Characterization and Protein Adsorption. Colloids Surf B Biointerfaces 1994, 3, 49–62. [Google Scholar] [CrossRef] [PubMed]
  97. Tanuma, S.; Powell, C.J.; Penn, D.R. Calculations of Electron Inelastic Mean Free Paths. V. Data for 14 Organic Compounds over the 50-2000 EV Range. Surface and Interface Analysis 1994, 21, 165–176. [Google Scholar] [CrossRef]
  98. Sebby, K.B.; Mansfield, E. Determination of the Surface Density of Polyethylene Glycol on Gold Nanoparticles by Use of Microscale Thermogravimetric Analysis. Anal Bioanal Chem 2015, 407, 2913–2922. [Google Scholar] [CrossRef]
  99. Rahme, K.; Chen, L.; Hobbs, R.G.; Morris, M.A.; O’Driscoll, C.; Holmes, J.D. PEGylated Gold Nanoparticles: Polymer Quantification as a Function of PEG Lengths and Nanoparticle Dimensions. RSC Adv. 2013, 3, 6085–6094. [Google Scholar] [CrossRef]
  100. Benoit, D.N.; Zhu, H.; Lilierose, M.H.; Verm, R.A.; Ali, N.; Morrison, A.N.; Fortner, J.D.; Avendano, C.; Colvin, V.L. Measuring the Grafting Density of Nanoparticles in Solution by Analytical Ultracentrifugation and Total Organic Carbon Analysis. Anal Chem 2012, 84, 9238–9245. [Google Scholar] [CrossRef]
  101. Li, M.; Jiang, S.; Simon, J.; Paßlick, D.; Frey, M.-L.; Wagner, M.; Mailänder, V.; Crespy, D.; Landfester, K. Brush Conformation of Polyethylene Glycol Determines the Stealth Effect of Nanocarriers in the Low Protein Adsorption Regime. Nano Lett 2021, 21, 1591–1598. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, C.; Desai, R.; Perez-Luna, V.; Karuri, N. PEGylation of Lysine Residues Improves the Proteolytic Stability of Fibronectin While Retaining Biological Activity. Biotechnol J 2014, 9, 1033–1043. [Google Scholar] [CrossRef] [PubMed]
  103. Sherman, M.R.; Saifer, M.G.P.; Perez-Ruiz, F. PEG-Uricase in the Management of Treatment-Resistant Gout and Hyperuricemia. Adv Drug Deliv Rev 2008, 60, 59–68. [Google Scholar] [CrossRef]
  104. Bailon, P.; Palleroni, A.; Schaffer, C.A.; Spence, C.L.; Fung, W.-J.; Porter, J.E.; Ehrlich, G.K.; Pan, W.; Xu, Z.-X.; Modi, M.W.; et al. Rational Design of a Potent, Long-Lasting Form of Interferon: A 40 KDa Branched Polyethylene Glycol-Conjugated Interferon α-2a for the Treatment of Hepatitis C. Bioconjug Chem 2001, 12, 195–202. [Google Scholar] [CrossRef]
  105. Abouzeid, A.H.; Patel, N.R.; Sarisozen, C.; Torchilin, V.P. Transferrin-Targeted Polymeric Micelles Co-Loaded with Curcumin and Paclitaxel: Efficient Killing of Paclitaxel-Resistant Cancer Cells. Pharm Res 2014, 31, 1938–1945. [Google Scholar] [CrossRef]
  106. Torchilin, V.P.; Levchenko, T.S.; Lukyanov, A.N.; Khaw, B.A.; Klibanov, A.L.; Rammohan, R.; Samokhin, G.P.; Whiteman, K.R. P-Nitrophenylcarbonyl-PEG-PE-Liposomes: Fast and Simple Attachment of Specific Ligands, Including Monoclonal Antibodies, to Distal Ends of PEG Chains via p-Nitrophenylcarbonyl Groups. Biochimica et Biophysica Acta (BBA) - Biomembranes 2001, 1511, 397–411. [Google Scholar] [CrossRef]
  107. Sriraman, S.K.; Salzano, G.; Sarisozen, C.; Torchilin, V. Anti-Cancer Activity of Doxorubicin-Loaded Liposomes Co-Modified with Transferrin and Folic Acid. European Journal of Pharmaceutics and Biopharmaceutics 2016, 105, 40–49. [Google Scholar] [CrossRef]
  108. Dunne, M.; Zheng, J.; Rosenblat, J.; Jaffray, D.A.; Allen, C. APN/CD13-Targeting as a Strategy to Alter the Tumor Accumulation of Liposomes. Journal of Controlled Release 2011, 154, 298–305. [Google Scholar] [CrossRef]
  109. Zhu, L.; Kate, P.; Torchilin, V.P. Matrix Metalloprotease 2-Responsive Multifunctional Liposomal Nanocarrier for Enhanced Tumor Targeting. ACS Nano 2012, 6, 3491–3498. [Google Scholar] [CrossRef]
  110. Ramirez-Paz, J.; Saxena, M.; Delinois, L.J.; Joaquín-Ovalle, F.M.; Lin, S.; Chen, Z.; Rojas-Nieves, V.A.; Griebenow, K. Thiol-Maleimide Poly(Ethylene Glycol) Crosslinking of L-Asparaginase Subunits at Recombinant Cysteine Residues Introduced by Mutagenesis. PLoS One 2018, 13, e0197643. [Google Scholar] [CrossRef]
  111. Luo, S.; Lu, X.; Liu, C.; Zhong, J.; Zhou, L.; Chen, T. Site Specific PEGylation of β-Lactoglobulin at Glutamine Residues and Its Influence on Conformation and Antigenicity. Food Research International 2019, 123, 623–630. [Google Scholar] [CrossRef]
  112. Sakane, T.; Pardridge, W.M. Carboxyl-Directed Pegylation of Brain-Derived Neurotrophic Factor Markedly Reduces Systemic Clearance with Minimal Loss of Biologic Activity. Pharm Res 1997, 14, 1085–1091. [Google Scholar] [CrossRef] [PubMed]
  113. Wang, B.; Cao, Y.; Chi, S.; Lou, D. A PEGylation Technology of L-Asparaginase with Monomethoxy Polyethylene Glycol-Propionaldehyde. Zeitschrift für Naturforschung C 2012, 67, 312–318. [Google Scholar] [CrossRef] [PubMed]
  114. Rodríguez-Martínez, J.A.; Rivera-Rivera, I.; Solá, R.J.; Griebenow, K. Enzymatic Activity and Thermal Stability of PEG-α-Chymotrypsin Conjugates. Biotechnol Lett 2009, 31, 883–887. [Google Scholar] [CrossRef] [PubMed]
  115. Pettit, D.K.; Bonnert, T.P.; Eisenman, J.; Srinivasan, S.; Paxton, R.; Beers, C.; Lynch, D.; Miller, B.; Yost, J.; Grabstein, K.H.; et al. Structure-Function Studies of Interleukin 15 Using Site-Specific Mutagenesis, Polyethylene Glycol Conjugation, and Homology Modeling. Journal of Biological Chemistry 1997, 272, 2312–2318. [Google Scholar] [CrossRef]
  116. Zaghmi, A.; Mendez-Villuendas, E.; Greschner, A.A.; Liu, J.Y.; de Haan, H.W.; Gauthier, M.A. Mechanisms of Activity Loss for a Multi-PEGylated Protein by Experiment and Simulation. Mater Today Chem 2019, 12, 121–131. [Google Scholar] [CrossRef]
  117. Yang, Q.; Lai, S.K. Anti- <scp>PEG</Scp> Immunity: Emergence, Characteristics, and Unaddressed Questions. WIREs Nanomedicine and Nanobiotechnology 2015, 7, 655–677. [Google Scholar] [CrossRef]
  118. Ichihara, M.; Shimizu, T.; Imoto, A.; Hashiguchi, Y.; Uehara, Y.; Ishida, T.; Kiwada, H. Anti-PEG IgM Response against PEGylated Liposomes in Mice and Rats. Pharmaceutics 2010, 3, 1–11. [Google Scholar] [CrossRef]
  119. Suzuki, T.; Suzuki, Y.; Hihara, T.; Kubara, K.; Kondo, K.; Hyodo, K.; Yamazaki, K.; Ishida, T.; Ishihara, H. PEG Shedding-Rate-Dependent Blood Clearance of PEGylated Lipid Nanoparticles in Mice: Faster PEG Shedding Attenuates Anti-PEG IgM Production. Int J Pharm 2020, 588, 119792. [Google Scholar] [CrossRef]
  120. Sylvestre, M.; Lv, S.; Yang, L.F.; Luera, N.; Peeler, D.J.; Chen, B.-M.; Roffler, S.R.; Pun, S.H. Replacement of L-Amino Acid Peptides with D-Amino Acid Peptides Mitigates Anti-PEG Antibody Generation against Polymer-Peptide Conjugates in Mice. Journal of Controlled Release 2021, 331, 142–153. [Google Scholar] [CrossRef]
  121. Garay, R.P.; El-Gewely, R.; Armstrong, J.K.; Garratty, G.; Richette, P. Antibodies against Polyethylene Glycol in Healthy Subjects and in Patients Treated with PEG-Conjugated Agents. Expert Opin Drug Deliv 2012, 9, 1319–1323. [Google Scholar] [CrossRef]
  122. Ju, Y.; Lee, W.S.; Pilkington, E.H.; Kelly, H.G.; Li, S.; Selva, K.J.; Wragg, K.M.; Subbarao, K.; Nguyen, T.H.O.; Rowntree, L.C.; et al. Anti-PEG Antibodies Boosted in Humans by SARS-CoV-2 Lipid Nanoparticle MRNA Vaccine. ACS Nano 2022, 16, 11769–11780. [Google Scholar] [CrossRef] [PubMed]
  123. Bianchi, A.; Bottau, P.; Calamelli, E.; Caimmi, S.; Crisafulli, G.; Franceschini, F.; Liotti, L.; Mori, F.; Paglialunga, C.; Saretta, F.; et al. Hypersensitivity to Polyethylene Glycol in Adults and Children: An Emerging Challenge. Acta Biomed 2021, 92, e2021519. [Google Scholar] [CrossRef]
  124. Münter, R.; Christensen, E.; Andresen, T.L.; Larsen, J.B. Studying How Administration Route and Dose Regulates Antibody Generation against LNPs for MRNA Delivery with Single-Particle Resolution. Mol Ther Methods Clin Dev 2023, 29, 450–459. [Google Scholar] [CrossRef] [PubMed]
  125. Takata, H.; Shimizu, T.; Yamade, R.; Elsadek, N.E.; Emam, S.E.; Ando, H.; Ishima, Y.; Ishida, T. Anti-PEG IgM Production Induced by PEGylated Liposomes as a Function of Administration Route. Journal of Controlled Release 2023, 360, 285–292. [Google Scholar] [CrossRef] [PubMed]
  126. Li, C.; Cao, J.; Wang, Y.; Zhao, X.; Deng, C.; Wei, N.; Yang, J.; Cui, J. Accelerated Blood Clearance of Pegylated Liposomal Topotecan: Influence of Polyethylene Glycol Grafting Density and Animal Species. J Pharm Sci 2012, 101, 3864–3876. [Google Scholar] [CrossRef] [PubMed]
  127. Ishida, T.; Harada, M.; Wang, X.Y.; Ichihara, M.; Irimura, K.; Kiwada, H. Accelerated Blood Clearance of PEGylated Liposomes Following Preceding Liposome Injection: Effects of Lipid Dose and PEG Surface-Density and Chain Length of the First-Dose Liposomes. Journal of Controlled Release 2005, 105, 305–317. [Google Scholar] [CrossRef]
  128. Saifer, M.G.P.; Williams, L.D.; Sobczyk, M.A.; Michaels, S.J.; Sherman, M.R. Selectivity of Binding of PEGs and PEG-like Oligomers to Anti-PEG Antibodies Induced by MethoxyPEG-Proteins. Mol Immunol 2014, 57, 236–246. [Google Scholar] [CrossRef] [PubMed]
  129. Sherman, M.R.; Williams, L.D.; Sobczyk, M.A.; Michaels, S.J.; Saifer, M.G.P. Role of the Methoxy Group in Immune Responses to MPEG-Protein Conjugates. Bioconjug Chem 2012, 23, 485–499. [Google Scholar] [CrossRef]
  130. Ishida, T.; Wang, X.; Shimizu, T.; Nawata, K.; Kiwada, H. PEGylated Liposomes Elicit an Anti-PEG IgM Response in a T Cell-Independent Manner. Journal of Controlled Release 2007, 122, 349–355. [Google Scholar] [CrossRef]
  131. Ehlinger, C.; Spear, N.; Doddareddy, R.; Shankar, G.; Schantz, A. A Generic Method for the Detection of Polyethylene Glycol Specific IgG and IgM Antibodies in Human Serum. J Immunol Methods 2019, 474, 112669. [Google Scholar] [CrossRef] [PubMed]
  132. Kozma, G.T.; Mészáros, T.; Berényi, P.; Facskó, R.; Patkó, Z.; Oláh, C.Zs.; Nagy, A.; Fülöp, T.G.; Glatter, K.A.; Radovits, T.; et al. Role of Anti-Polyethylene Glycol (PEG) Antibodies in the Allergic Reactions to PEG-Containing Covid-19 Vaccines: Evidence for Immunogenicity of PEG. Vaccine 2023, 41, 4561–4570. [Google Scholar] [CrossRef]
  133. Klibanov, A.L.; Maruyama, K.; Torchilin, V.P.; Huang, L. Amphipathic Polyethyleneglycols Effectively Prolong the Circulation Time of Liposomes. FEBS Lett 1990, 268, 235–237. [Google Scholar] [CrossRef]
  134. Kanamala, M.; Palmer, B.D.; Jamieson, S.M.; Wilson, W.R.; Wu, Z. Dual PH-Sensitive Liposomes with Low PH-Triggered Sheddable PEG for Enhanced Tumor-Targeted Drug Delivery. Nanomedicine 2019, 14, 1971–1989. [Google Scholar] [CrossRef]
  135. Hong, M.-S.; Lim, S.-J.; Oh, Y.-K.; Kim, C.-K. PH-Sensitive, Serum-Stable and Long-Circulating Liposomes as a New Drug Delivery System. Journal of Pharmacy and Pharmacology 2002, 54, 51–58. [Google Scholar] [CrossRef]
  136. Xia, Y.; Fang, M.; Dong, J.; Xu, C.; Liao, Z.; Ning, P.; Zeng, Q. PH Sensitive Liposomes Delivering Tariquidar and Doxorubicin to Overcome Multidrug Resistance of Resistant Ovarian Cancer Cells. Colloids Surf B Biointerfaces 2018, 170, 514–520. [Google Scholar] [CrossRef]
  137. Kale, A.A.; Torchilin, V.P. “Smart” Drug Carriers: PEGylated TATp-Modified PH-Sensitive Liposomes. J Liposome Res 2007, 17, 197–203. [Google Scholar] [CrossRef] [PubMed]
  138. Safra, T.; Muggia, F.; Jeffers, S.; Tsao-Wei, D.D.; Groshen, S.; Lyass, O.; Henderson, R.; Berry, G.; Gabizon, A. Pegylated Liposomal Doxorubicin (Doxil): Reduced Clinical Cardiotoxicity in Patients Reaching or Exceeding Cumulative Doses of 500 Mg/M2. Annals of Oncology 2000, 11, 1029–1034. [Google Scholar] [CrossRef] [PubMed]
  139. Nosova, A.S.; Koloskova, O.O.; Nikonova, A.A.; Simonova, V.A.; Smirnov, V. V.; Kudlay, D.; Khaitov, M.R. Diversity of PEGylation Methods of Liposomes and Their Influence on RNA Delivery. Medchemcomm 2019, 10, 369–377. [Google Scholar] [CrossRef]
  140. Ishida, T.; Kiwada, H. Accelerated Blood Clearance (ABC) Phenomenon upon Repeated Injection of PEGylated Liposomes. Int J Pharm 2008, 354, 56–62. [Google Scholar] [CrossRef]
  141. Ishida, T.; Ichihara, M.; Wang, X.; Yamamoto, K.; Kimura, J.; Majima, E.; Kiwada, H. Injection of PEGylated Liposomes in Rats Elicits PEG-Specific IgM, Which Is Responsible for Rapid Elimination of a Second Dose of PEGylated Liposomes. Journal of Controlled Release 2006, 112, 15–25. [Google Scholar] [CrossRef] [PubMed]
  142. El Sayed, M.M.; Takata, H.; Shimizu, T.; Kawaguchi, Y.; Abu Lila, A.S.; Elsadek, N.E.; Alaaeldin, E.; Ishima, Y.; Ando, H.; Kamal, A.; et al. Hepatosplenic Phagocytic Cells Indirectly Contribute to Anti-PEG IgM Production in the Accelerated Blood Clearance (ABC) Phenomenon against PEGylated Liposomes: Appearance of an Unexplained Mechanism in the ABC Phenomenon. Journal of Controlled Release 2020, 323, 102–109. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, X.; Ishida, T.; Kiwada, H. Anti-PEG IgM Elicited by Injection of Liposomes Is Involved in the Enhanced Blood Clearance of a Subsequent Dose of PEGylated Liposomes. Journal of Controlled Release 2007, 119, 236–244. [Google Scholar] [CrossRef] [PubMed]
  144. Sawant, R.R.; Jhaveri, A.M.; Koshkaryev, A.; Zhu, L.; Qureshi, F.; Torchilin, V.P. Targeted Transferrin-Modified Polymeric Micelles: Enhanced Efficacy in Vitro and in Vivo in Ovarian Carcinoma. Mol Pharm 2014, 11, 375–381. [Google Scholar] [CrossRef]
  145. Sarisozen, C.; Abouzeid, A.H.; Torchilin, V.P. The Effect of Co-Delivery of Paclitaxel and Curcumin by Transferrin-Targeted PEG-PE-Based Mixed Micelles on Resistant Ovarian Cancer in 3-D Spheroids and in Vivo Tumors. European Journal of Pharmaceutics and Biopharmaceutics 2014, 88, 539–550. [Google Scholar] [CrossRef]
  146. Gajbhiye, V.; Vijayaraj Kumar, P.; Tekade, R.K.; Jain, N.K. PEGylated PPI Dendritic Architectures for Sustained Delivery of H2 Receptor Antagonist. Eur J Med Chem 2009, 44, 1155–1166. [Google Scholar] [CrossRef]
  147. Kesharwani, P.; Jain, K.; Jain, N.K. Dendrimer as Nanocarrier for Drug Delivery. Prog Polym Sci 2014, 39, 268–307. [Google Scholar] [CrossRef]
  148. Sahoo, R.K.; Gothwal, A.; Rani, S.; Nakhate, K.T.; Ajazuddin; Gupta, U. PEGylated Dendrimer Mediated Delivery of Bortezomib: Drug Conjugation versus Encapsulation. Int J Pharm 2020, 584, 119389. [Google Scholar] [CrossRef]
  149. Hong, S.; Bielinska, A.U.; Mecke, A.; Keszler, B.; Beals, J.L.; Shi, X.; Balogh, L.; Orr, B.G.; Baker, J.R.; Banaszak Holl, M.M. Interaction of Poly(Amidoamine) Dendrimers with Supported Lipid Bilayers and Cells: Hole Formation and the Relation to Transport. Bioconjug Chem 2004, 15, 774–782. [Google Scholar] [CrossRef]
  150. Jain, K.; Kesharwani, P.; Gupta, U.; Jain, N.K. Dendrimer Toxicity: Let’s Meet the Challenge. Int J Pharm 2010, 394, 122–142. [Google Scholar] [CrossRef]
  151. Somani, S.; Laskar, P.; Altwaijry, N.; Kewcharoenvong, P.; Irving, C.; Robb, G.; Pickard, B.S.; Dufès, C. PEGylation of Polypropylenimine Dendrimers: Effects on Cytotoxicity, DNA Condensation, Gene Delivery and Expression in Cancer Cells. Sci Rep 2018, 8, 9410. [Google Scholar] [CrossRef]
  152. Luo, D.; Haverstick, K.; Belcheva, N.; Han, E.; Saltzman, W.M. Poly(Ethylene Glycol)-Conjugated PAMAM Dendrimer for Biocompatible, High-Efficiency DNA Delivery. Macromolecules 2002, 35, 3456–3462. [Google Scholar] [CrossRef]
  153. Wang, W.; Xiong, W.; Wan, J.; Sun, X.; Xu, H.; Yang, X. The Decrease of PAMAM Dendrimer-Induced Cytotoxicity by PEGylation via Attenuation of Oxidative Stress. Nanotechnology 2009, 20, 105103. [Google Scholar] [CrossRef] [PubMed]
  154. Jangid, A.K.; Patel, K.; Joshi, U.; Patel, S.; Singh, A.; Pooja, D.; Saharan, V.A.; Kulhari, H. PEGylated G4 Dendrimers as a Promising Nanocarrier for Piperlongumine Delivery: Synthesis, Characterization, and Anticancer Activity. Eur Polym J 2022, 179, 111547. [Google Scholar] [CrossRef]
  155. Guillaudeu, S.J.; Fox, M.E.; Haidar, Y.M.; Dy, E.E.; Szoka, F.C.; Fréchet, J.M.J. PEGylated Dendrimers with Core Functionality for Biological Applications. Bioconjug Chem 2008, 19, 461–469. [Google Scholar] [CrossRef] [PubMed]
  156. Sideratou, Z.; Kontoyianni, C.; Drossopoulou, G.I.; Paleos, C.M. Synthesis of a Folate Functionalized PEGylated Poly(Propylene Imine) Dendrimer as Prospective Targeted Drug Delivery System. Bioorg Med Chem Lett 2010, 20, 6513–6517. [Google Scholar] [CrossRef]
  157. Xu, Z.; Wang, Y.; Ma, Z.; Wang, Z.; Wei, Y.; Jia, X. A Poly(Amidoamine) Dendrimer-Based Nanocarrier Conjugated with Angiopep-2 for Dual-Targeting Function in Treating Glioma Cells. Polym Chem 2016, 7, 715–721. [Google Scholar] [CrossRef]
  158. He, H.; Li, Y.; Jia, X.-R.; Du, J.; Ying, X.; Lu, W.-L.; Lou, J.-N.; Wei, Y. PEGylated Poly(Amidoamine) Dendrimer-Based Dual-Targeting Carrier for Treating Brain Tumors. Biomaterials 2011, 32, 478–487. [Google Scholar] [CrossRef]
  159. Zhang, L.; Zhu, S.; Qian, L.; Pei, Y.; Qiu, Y.; Jiang, Y. RGD-Modified PEG–PAMAM–DOX Conjugates: In Vitro and in Vivo Studies for Glioma. European Journal of Pharmaceutics and Biopharmaceutics 2011, 79, 232–240. [Google Scholar] [CrossRef]
  160. Li, Y.; He, H.; Jia, X.; Lu, W.-L.; Lou, J.; Wei, Y. A Dual-Targeting Nanocarrier Based on Poly(Amidoamine) Dendrimers Conjugated with Transferrin and Tamoxifen for Treating Brain Gliomas. Biomaterials 2012, 33, 3899–3908. [Google Scholar] [CrossRef]
  161. OWENSIII, D.; PEPPAS, N. Opsonization, Biodistribution, and Pharmacokinetics of Polymeric Nanoparticles. Int J Pharm 2006, 307, 93–102. [Google Scholar] [CrossRef] [PubMed]
  162. Shan, X.; Yuan, Y.; Liu, C.; Tao, X.; Sheng, Y.; Xu, F. Influence of PEG Chain on the Complement Activation Suppression and Longevity in Vivo Prolongation of the PCL Biomedical Nanoparticles. Biomed Microdevices 2009, 11, 1187–1194. [Google Scholar] [CrossRef] [PubMed]
  163. Khalil, N.M.; Nascimento, T.C.F. do; Casa, D.M.; Dalmolin, L.F.; Mattos, A.C. de; Hoss, I.; Romano, M.A.; Mainardes, R.M. Pharmacokinetics of Curcumin-Loaded PLGA and PLGA–PEG Blend Nanoparticles after Oral Administration in Rats. Colloids Surf B Biointerfaces 2013, 101, 353–360. [Google Scholar] [CrossRef]
  164. Behzadi, S.; Serpooshan, V.; Tao, W.; Hamaly, M.A.; Alkawareek, M.Y.; Dreaden, E.C.; Brown, D.; Alkilany, A.M.; Farokhzad, O.C.; Mahmoudi, M. Cellular Uptake of Nanoparticles: Journey inside the Cell. Chem Soc Rev 2017, 46, 4218–4244. [Google Scholar] [CrossRef] [PubMed]
  165. Li, Y.-P.; Pei, Y.-Y.; Zhang, X.-Y.; Gu, Z.-H.; Zhou, Z.-H.; Yuan, W.-F.; Zhou, J.-J.; Zhu, J.-H.; Gao, X.-J. PEGylated PLGA Nanoparticles as Protein Carriers: Synthesis, Preparation and Biodistribution in Rats. Journal of Controlled Release 2001, 71, 203–211. [Google Scholar] [CrossRef]
  166. Ahmed, O.A.A.; Badr-Eldin, S.M. Biodegradable Self-Assembled Nanoparticles of PEG-PLGA Amphiphilic Diblock Copolymer as a Promising Stealth System for Augmented Vinpocetine Brain Delivery. Int J Pharm 2020, 588, 119778. [Google Scholar] [CrossRef]
  167. Layre, A.; Couvreur, P.; Chacun, H.; Richard, J.; Passirani, C.; Requier, D.; Benoit, J.P.; Gref, R. Novel Composite Core-Shell Nanoparticles as Busulfan Carriers. Journal of Controlled Release 2006, 111, 271–280. [Google Scholar] [CrossRef]
  168. Pelaz, B.; del Pino, P.; Maffre, P.; Hartmann, R.; Gallego, M.; Rivera-Fernández, S.; de la Fuente, J.M.; Nienhaus, G.U.; Parak, W.J. Surface Functionalization of Nanoparticles with Polyethylene Glycol: Effects on Protein Adsorption and Cellular Uptake. ACS Nano 2015, 9, 6996–7008. [Google Scholar] [CrossRef]
  169. Mazumdar, S.; Chitkara, D.; Mittal, A. Exploration and Insights into the Cellular Internalization and Intracellular Fate of Amphiphilic Polymeric Nanocarriers. Acta Pharm Sin B 2021, 11, 903–924. [Google Scholar] [CrossRef]
  170. Ayen, W.Y.; Kumar, N. A Systematic Study on Lyophilization Process of Polymersomes for Long-Term Storage Using Doxorubicin-Loaded (PEG)3–PLA Nanopolymersomes. European Journal of Pharmaceutical Sciences 2012, 46, 405–414. [Google Scholar] [CrossRef]
  171. Zheng, Z.; Wu, Y.; Ahmad, A.; Ramzan, N.; Kayitmazer, A.B.; Zhou, X.; Xu, Y. Lyophilizable Polymer–Lipid Hybrid Nanoparticles with High Paclitaxel Loading. ACS Appl Nano Mater 2024, 7, 19194–19210. [Google Scholar] [CrossRef]
  172. Bonaccorso, A.; Musumeci, T.; Carbone, C.; Vicari, L.; Lauro, M.R.; Puglisi, G. Revisiting the Role of Sucrose in PLGA-PEG Nanocarrier for Potential Intranasal Delivery. Pharm Dev Technol 2018, 23, 265–274. [Google Scholar] [CrossRef] [PubMed]
  173. Xu, H.; Fang, G.; Gou, J.; Wang, S.; Yao, Q.; Zhang, Y.; Tang, X.; Zhao, Y. Lyophilization of Self-Assembled Polymeric Nanoparticles Without Compromising Their Microstructure and Their &lt;I&gt;In&lt;/I&gt; &lt;I&gt;Vivo&lt;/I&gt; Evaluation: Pharmacokinetics, Tissue Distribution and Toxicity. J Biomater Tissue Eng 2015, 5, 919–929. [Google Scholar] [CrossRef]
  174. Joye, I.J.; McClements, D.J. Production of Nanoparticles by Anti-Solvent Precipitation for Use in Food Systems. Trends Food Sci Technol 2013, 34, 109–123. [Google Scholar] [CrossRef]
  175. Farokhzad, O.C.; Cheng, J.; Teply, B.A.; Sherifi, I.; Jon, S.; Kantoff, P.W.; Richie, J.P.; Langer, R. Targeted Nanoparticle-Aptamer Bioconjugates for Cancer Chemotherapy in Vivo. Proceedings of the National Academy of Sciences 2006, 103, 6315–6320. [Google Scholar] [CrossRef] [PubMed]
  176. Xu, J.; Chen, Y.; Jiang, X.; Gui, Z.; Zhang, L. Development of Hydrophilic Drug Encapsulation and Controlled Release Using a Modified Nanoprecipitation Method. Processes 2019, 7, 331. [Google Scholar] [CrossRef]
  177. Zou, W.; Liu, C.; Chen, Z.; Zhang, N. Preparation and Characterization of Cationic PLA-PEG Nanoparticles for Delivery of Plasmid DNA. Nanoscale Res Lett 2009, 4, 982. [Google Scholar] [CrossRef]
  178. Yadav, A.K.; Mishra, P.; Jain, S.; Mishra, P.; Mishra, A.K.; Agrawal, G.P. Preparation and Characterization of HA–PEG–PCL Intelligent Core–Corona Nanoparticles for Delivery of Doxorubicin. J Drug Target 2008, 16, 464–478. [Google Scholar] [CrossRef]
  179. Dhar, S.; Gu, F.X.; Langer, R.; Farokhzad, O.C.; Lippard, S.J. Targeted Delivery of Cisplatin to Prostate Cancer Cells by Aptamer Functionalized Pt(IV) Prodrug-PLGA–PEG Nanoparticles. Proceedings of the National Academy of Sciences 2008, 105, 17356–17361. [Google Scholar] [CrossRef]
  180. Sánchez-López, E.; Egea, M.A.; Cano, A.; Espina, M.; Calpena, A.C.; Ettcheto, M.; Camins, A.; Souto, E.B.; Silva, A.M.; García, M.L. PEGylated PLGA Nanospheres Optimized by Design of Experiments for Ocular Administration of Dexibuprofen—in Vitro, Ex Vivo and in Vivo Characterization. Colloids Surf B Biointerfaces 2016, 145, 241–250. [Google Scholar] [CrossRef]
  181. Matsumoto, J.; Nakada, Y.; Sakurai, K.; Nakamura, T.; Takahashi, Y. Preparation of Nanoparticles Consisted of Poly(l-Lactide)–Poly(Ethylene Glycol)–Poly(l-Lactide) and Their Evaluation in Vitro. Int J Pharm 1999, 185, 93–101. [Google Scholar] [CrossRef] [PubMed]
  182. Xu, Q.; Ensign, L.M.; Boylan, N.J.; Schön, A.; Gong, X.; Yang, J.-C.; Lamb, N.W.; Cai, S.; Yu, T.; Freire, E.; et al. Impact of Surface Polyethylene Glycol (PEG) Density on Biodegradable Nanoparticle Transport in Mucus Ex Vivo and Distribution in Vivo. ACS Nano 2015, 9, 9217–9227. [Google Scholar] [CrossRef] [PubMed]
  183. Jaiswal, J.; Kumar Gupta, S.; Kreuter, J. Preparation of Biodegradable Cyclosporine Nanoparticles by High-Pressure Emulsification-Solvent Evaporation Process. Journal of Controlled Release 2004, 96, 169–178. [Google Scholar] [CrossRef] [PubMed]
  184. Vangara, K.K.; Liu, J.L.; Palakurthi, S. Hyaluronic Acid-Decorated PLGA-PEG Nanoparticles for Targeted Delivery of SN-38 to Ovarian Cancer. Anticancer Res 2013, 33, 2425–2434. [Google Scholar]
  185. Essa, S.; Rabanel, J.M.; Hildgen, P. Characterization of Rhodamine Loaded PEG-g-PLA Nanoparticles (NPs): Effect of Poly(Ethylene Glycol) Grafting Density. Int J Pharm 2011, 411, 178–187. [Google Scholar] [CrossRef]
  186. Pillai, J.J.; Thulasidasan, A.K.T.; Anto, R.J.; Devika, N.C.; Ashwanikumar, N.; Kumar, G.S.V. Curcumin Entrapped Folic Acid Conjugated PLGA–PEG Nanoparticles Exhibit Enhanced Anticancer Activity by Site Specific Delivery. RSC Adv 2015, 5, 25518–25524. [Google Scholar] [CrossRef]
  187. Vila, A.; Sánchez, A.; Pérez, C.; Alonso, M.J. PLA-PEG Nanospheres: New Carriers for Transmucosal Delivery of Proteins and Plasmid DNA. Polym Adv Technol 2002, 13, 851–858. [Google Scholar] [CrossRef]
  188. Sánchez-López, E.; Egea, M.A.; Davis, B.M.; Guo, L.; Espina, M.; Silva, A.M.; Calpena, A.C.; Souto, E.M.B.; Ravindran, N.; Ettcheto, M.; et al. Memantine-Loaded PEGylated Biodegradable Nanoparticles for the Treatment of Glaucoma. Small 2018, 14. [Google Scholar] [CrossRef]
  189. Massadeh, S.; Alaamery, M.; Al-Qatanani, S.; Alarifi, S.; Bawazeer, S.; Alyafee, Y. Synthesis of Protein-Coated Biocompatible Methotrexate-Loaded PLA-PEG-PLA Nanoparticles for Breast Cancer Treatment. Nano Rev Exp 2016, 7, 31996. [Google Scholar] [CrossRef]
  190. Tobío, M.; Gref, R.; Sánchez, A.; Langer, R.; Alonso, M.J. Stealth PLA-PEG Nanoparticles as Protein Carriers for Nasal Administration. Pharm Res 1998, 15, 270–275. [Google Scholar] [CrossRef]
  191. Babos, G.; Biró, E.; Meiczinger, M.; Feczkó, T. Dual Drug Delivery of Sorafenib and Doxorubicin from PLGA and PEG-PLGA Polymeric Nanoparticles. Polymers (Basel) 2018, 10. [Google Scholar] [CrossRef] [PubMed]
  192. Wang, H.; Zhao, Y.; Wu, Y.; Hu, Y.; Nan, K.; Nie, G.; Chen, H. Enhanced Anti-Tumor Efficacy by Co-Delivery of Doxorubicin and Paclitaxel with Amphiphilic Methoxy PEG-PLGA Copolymer Nanoparticles. Biomaterials 2011, 32, 8281–8290. [Google Scholar] [CrossRef] [PubMed]
  193. Gurnani, P.; Sanchez-Cano, C.; Abraham, K.; Xandri-Monje, H.; Cook, A.B.; Hartlieb, M.; Lévi, F.; Dallmann, R.; Perrier, S. RAFT Emulsion Polymerization as a Platform to Generate Well-Defined Biocompatible Latex Nanoparticles. Macromol Biosci 2018, 18. [Google Scholar] [CrossRef] [PubMed]
  194. Colombo, C.; Morosi, L.; Bello, E.; Ferrari, R.; Licandro, S.A.; Lupi, M.; Ubezio, P.; Morbidelli, M.; Zucchetti, M.; D’Incalci, M.; et al. PEGylated Nanoparticles Obtained through Emulsion Polymerization as Paclitaxel Carriers. Mol Pharm 2016, 13, 40–46. [Google Scholar] [CrossRef]
  195. Lupi, M.; Colombo, C.; Frapolli, R.; Ferrari, R.; Sitia, L.; Dragoni, L.; Bello, E.; Licandro, S.A.; Falcetta, F.; Ubezio, P.; et al. A Biodistribution Study of PEGylated PCL-Based Nanoparticles in C57BL/6 Mice Bearing B16/F10 Melanoma. Nanotechnology 2014, 25, 335706. [Google Scholar] [CrossRef]
  196. Ferrari, R.; Yu, Y.; Lattuada, M.; Storti, G.; Morbidelli, M.; Moscatelli, D. Controlled PEGylation of PLA-Based Nanoparticles. Macromol Chem Phys 2012, 213, 2012–2018. [Google Scholar] [CrossRef]
Figure 1. Synthetic pathways for NH₂-PEG preparation.
Figure 1. Synthetic pathways for NH₂-PEG preparation.
Preprints 148174 g001
Figure 2. Synthesis of alkyne-functionalized PEG.
Figure 2. Synthesis of alkyne-functionalized PEG.
Preprints 148174 g002
Figure 3. Synthesis of PEG-aldehyde by ozonolysis.
Figure 3. Synthesis of PEG-aldehyde by ozonolysis.
Preprints 148174 g003
Figure 4. Synthesis of p-nitrophenyl-activated PEG via acyl substitution reaction.
Figure 4. Synthesis of p-nitrophenyl-activated PEG via acyl substitution reaction.
Preprints 148174 g004
Figure 5. TEM micrographs core-shell PEGylated NPs. Reprinted with permission from Ref. [74].
Figure 5. TEM micrographs core-shell PEGylated NPs. Reprinted with permission from Ref. [74].
Preprints 148174 g005
Figure 6. Comparison of PEGylation strategies for nanoparticle surface modification.
Figure 6. Comparison of PEGylation strategies for nanoparticle surface modification.
Preprints 148174 g006
Figure 7. PEG conformational regime on NP surface.
Figure 7. PEG conformational regime on NP surface.
Preprints 148174 g007
Figure 8. PEGylation improves the pharmacokinetics of proteins.
Figure 8. PEGylation improves the pharmacokinetics of proteins.
Preprints 148174 g008
Figure 9. Polymeric micelle formation and conjugation strategies.
Figure 9. Polymeric micelle formation and conjugation strategies.
Preprints 148174 g009
Figure 11. PEG-block copolymer micelle formation and drug encapsulation mechanism.
Figure 11. PEG-block copolymer micelle formation and drug encapsulation mechanism.
Preprints 148174 g011
Table 1. PEG Derivatives and their application.
Table 1. PEG Derivatives and their application.
PEG Derivative Primary Application Reaction Benefits
PEG-Amine Protein, peptide conjugation N-terminal, Lysine PEGylation Increased stability and half-life [55,56,57]
PEG-Silane Surface functionalization, drug delivery Highly reactive with hydroxyl groups on surfaces Stable surface functionalization, increased circulation time [58]
PEG-Aldehyde Protein conjugation Nucleophilic addition with hydroxyl or amine groups Improved half-life [59,60]
PEG-Azide Click chemistry Click reaction with alkyne-functionalized molecules High specificity, bioorthogonality
PEG-Acrylate Tissue engineering and hydrogel preparation Michael addition, radical Polymerization Hydrogel scaffolds for 3D cell culture​[61,62], wound dressing [63,64],tissue engineering [65]
PEG-Maleimide Protein, drug, NP conjugation Reacts with thiols (cysteine) in proteins, gold NP surface Increased stability, half-life [66,67,68,69]
PEG-Nitrophenyl Protein conjugation, crosslinking Nucleophilic substitution reactions Rapid and simple modification of nanocarriers for protein conjugation
Table 2. List of FDA-approved PEGylated liposomal formulations.
Table 2. List of FDA-approved PEGylated liposomal formulations.
Brand Name Active Ingredient Cancer Type Mechanism of Action
Onivyde Irinotecan hydrochloride trihydrate Pancreatic cancer, small cell lung cancer, colon cancer Topoisomerase I inhibitor
DaunoXome Daunorubicin Kaposi’s sarcoma, Non-Hodgkin lymphoma Topoisomerase II inhibitor
Doxil, Zolsketil , Caelyx, Myocet Doxorubicin hydrochloride Ovarian cancer, Kaposi’s sarcoma, myeloid melanoma, Breast cancer Topoisomerase II inhibitor
Mepact Mifamurtide Osteosarcoma Activating immune cells
Marqibo Vincristine sulfate Acute lymphoblastic leukemia (ALL), Hodgkin's disease, neuroblastoma Microtubule synthesis Inhibition
Vyxeos Daunorubicin, cytarabine (in combination) Leukemia Topoisomerase II, Nucleoside Inhibitor
DepoCyt Cytarabine Lymphomatous meningitis Nucleoside Inhibitor
Table 3. Characteristics of PEGylated NPs prepared by nanoprecipitation method.
Table 3. Characteristics of PEGylated NPs prepared by nanoprecipitation method.
Polymer Encapsulated Drug Organic Solvent Antisolvent EE or DL (%) Stabilizer Particle Size (nm) Activity Refs.
PEG-PLGA Vinpocetine Acetone Water 60-90 PVA 30-290 Cerebrovascular disorder [166]
PEG-PLGA Docetaxel Acetonitrile Water - Without stabilizer 153.3 Antineoplastic agent [175]
PLGA-PEG Ciprofloxacin DMSO Water 1.70-3.29 * Without stabilizer 174-205 Antibiotic [176]
PEG-PLA pEGFP Acetone Water 95.56 CTAB/Tween
80
128.9 Plasmid vector [177]
HA–PEG–PCL DOXORUBICIN Acetone Water 92 Pluronic F-68 95 Antineoplastic agent [178]
Poly(isobutylcyanoacrylate/ PCL–PEG Busulfan Acetone Water 17.0 Without stabilizer 152 Antineoplastic agent [167]
PEG-PLGA Platinum(IV) Acetonitrile Water 18.4 * Without stabilizer 172 Antineoplastic agent [179]
PEG-PLGA Dexibuprofen Acetone Water 85-100 PVA 201-226 Anti-inflammatory drug [180]
∗ Drug loading (DL), ● Encapsulation efficiency (EE).
Table 4. Characteristics of PEGylated NPs prepared by single emulsion-solvent evaporation method.
Table 4. Characteristics of PEGylated NPs prepared by single emulsion-solvent evaporation method.
Polymer Encapsulated
Drug
Organic
solvent
Emulsifier Particle size
(nm)
Encapsulation Efficiency
(%)
Refs.
PLGA-PEG Cyclosporine DCM PVA 212 91.90 [183]
PLA-PEG-PLA Progesterone DCM PVA 193-335 65-71 [181]
PLGA-PEG Curcumin Ethyl Acetate/DCM PVA 152.37 73.22 [163]
PLGA-PEG SN-38 DCM PVA 249.2 81.85 [184]
PEG-PLA Rhodamine B DCM PVA 169.5 67.79 [185]
PLGA–PEG Curcumin DCM PVA 100-200 52.2 [186]
Table 5. Characteristics of PEGylated NPs prepared by double emulsion-solvent evaporation method.
Table 5. Characteristics of PEGylated NPs prepared by double emulsion-solvent evaporation method.
Polymer Encapsulated Drug Organic solvent Emulsifier Particle size
(nm)
Encapsulation
Efficiency
(%)
Refs.
PLA-PEG Tetanus toxoid Ethyl Acetate Sodium cholate 196 33.4 [187]
PLGA-PEG Memantine Ethyl Acetate PVA 193-224 77-80 [188]
PLA-PEG-PLA Methotrexate Chloroform PVA 100-173 23-48 [189]
PLA-PEG Tetanus toxoid Ethyl Acetate Gelatin 136.8 35.3 [190]
PLGA-PEG BSA DCM PVA 198.1 48.6 [165]
PLGA-PEG Sorafenib,
Doxorubicin
DCM/Acetone PVA 177.2 88
69
[191]
PLGA-PEG Paclitaxel,
Doxorubicin
DCM PVA 243.63 70.13
57.5
[192]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2025 MDPI (Basel, Switzerland) unless otherwise stated